Skip to main content

Full text of "Fishery bulletin"

See other formats


U.S.  Department 
of  Commerce 

Volume  102 
Number  1 
January  2004 


Fishery 
Bulletin 


U.S.  Department 
of  Commerce 

Donaid  L.  Evans 
Secretary 


National  Oceanic 
and  Atmospheric 
Administration 

Vice  Admiral 

Conrad  C.  Lautenbacher  Jr., 

USN  (ret.) 

Under  Secretary  for 
Oceans  and  Atmosphere 


National  Marine 
Fisheries  Service 

William  T.  Hogarth 

Assistant  Administrator 
for  Fisheries 


.^TOFCo. 


X 


K1^  / 


The  Fishery'  Bulletin  (ISSN  0090-0656) 
is  published  quarterly  by  the  Scientific 
Publications  Office,  National  Marine  Fish- 
N(  >AA,  7600  Sand  Point  Way 
NE,  BIN  C15700,  Seattle.  WA  981 15-0070. 
Periodicals  postage  is  paid  at  Seattle,  WA, 
and  at  additional  mailing  offices.  POST- 
MASTER: Send  address  changes  for  sub- 
scription- to  Fishery  Bulletin.  Superin- 
tendent of  Documents,  Attn.:  Chief.  .Mail 
List  Branch,  Mail  Stop  SSOM,  Washing- 
ton. DC  20402-9373. 

Although  the  contents  of  this  publica- 
tion have  not  been  copyrighted  and  may 
be  reprinted  entirely,  reference  to  source 
is  appreciated. 

The  Set  if  Commerce  has  deter- 

mined that  the  publication  of  tin 

ording  to  law  for 

the  transaction  of  public  business  of  this 

Department.  Use  of  funds  for  printing  of 

nodical  has  been  approved  by  the 

oroftheOffii  cement  and 

Budget. 

For  sale  by  the  Superintendent  of 
nuts.  US.  Government  Printing 
I  mice,  Washington,  DC  20402.  Subscrip- 
tion pi  i  it:  $55.00  domestic  and 
$68.75  foreign.  Cost  per  single  issue: 
$28.00  dome  ,5.00  foreign.  See 
back  for  order  form. 


Scientific  Editor 

Dr.  Norman  Bartoo 

Associate  Editor 

Sarah  Shoffler 

National  Marine  Fisheries  Service,  NOAA 
8604  La  Jolla  Shores  Drive 
La  Jolla,  California  92037 

Managing  Editor 

Sharyn  Matriotti 

National  Marine  Fisheries  Service 
Scientific  Publications  Office 
7600  Sand  Point  Way  NE,  BIN  C15700 
Seattle,  Washington  981 15-0070 


Editorial  Committee 

Dr.  Harlyn  O.  Halvorson 
Dr.  Ronald  W.  Hardy 
Dr.  Richard  D.  Methot 
Dr.  Theodore  W.  Pietsch 
Dr.  Joseph  E.  Powers 
Dr.  Harald  Rosenthal 
Dr.  Fredric  M.  Serchuk 
Dr.  George  Watters 


University  of  Massachusetts,  Boston 
University  of  Idaho,  Hagerman 
National  Marine  Fisheries  Service 
University  of  Washington,  Seattle 
National  Marine  Fisheries  Service 
Universitat  Kiel,  Germany 
National  Marine  Fisheries  Service 
National  Marine  Fisheries  Service 


Fishery  Bulletin  web  site:  www.fishbull.noaa.gov 


The  Fishery  Bulletin  carries  original  research  reports  and  technical  notes  on  investigations  in 
fishery  scien  ring,  and  economics.  It  began  as  the  Bulletin  of  the  United  States  Pish 

Commission  in  1881;  it  became  the  Bulletin  of  the  Bureau  of  Fisheries  in  1904  and  the  Fishery 
Bulletin  of  the  Fish  and  Wildlife  Service  in  1941.  Separates  were  issued  as  documents  tl 
volume  46;  the  last  document  was  No.  1103.  Beginning  with  volume  47  in  1931  and  continuing 
through  volume  62  in  196.1  ired  as  a  numbered  bulletin.  A  new  system 

began  in  1963  with  volume  6:3  in  which  papers  are  bound  together  in  a  single  issue  of  the 
bulletin.  Beginning  with  volume  70.  number  1.  January  1972,  the  Fishery  Bulletin  became  a 
lieal,  issued  quarterly.  In  this  form,  it  is  available  by  subscription  from  the  Superintendent 
of  Documents.  U.S.  Government  Printing  Office,  Washington,  DC  20402.  It  is  also  available  free  in 
limited  numbers  to  libl  irch  institutions.  State  and  Federal  agencies,  and  in  exi 

for  other  scientific  publications. 


U.S.  Department 
of  Commerce 

Seattle,  Washington 

Volume  102 
Number  1 
January  2004 


Fishery 
Bulletin 


Contents 


ary 

MAR      5  2004 


The  conclusions  and  opinions  expressed 
in  Fisher)'  Bulletin  are  solely  those  of  the 
authors  and  do  not  represent  the  official 
position  of  the  National  Marine  Fisher- 
ies Service  (NOAA)  or  any  other  agency 
or  institution. 

The  National  Marine  Fisheries  Service 
(NMFS)  does  not  approve,  recommend,  or 
endorse  any  proprietary  product  or  pro- 
prietary material  mentioned  in  this  pub- 
lication. No  reference  shall  be  made  to 
NMFS.  or  to  this  publication  furnished  by 
NMFS,  in  any  advertising  or  sales  pro- 
motion which  would  indicate  or  imply 
that  NMFS  approves,  recommends,  or 
endorses  any  proprietary  product  or  pro- 
prietary material  mentioned  herein,  or 
which  has  as  its  purpose  an  intent  to 
cause  directly  or  indirectly  the  advertised 
product  to  be  used  or  purchased  because 
of  this  NMFS  publication. 


Articles 

1-13  Alonzo,  Suzanne  H.,  and  Marc  Mangel 

The  effects  of  size-selective  fisheries  on  the  stock  dynamics  of 
and  sperm  limitation  in  sex-changing  fish 

14-24  Baba,  Katsuhisa,  Toshifumi  Kawajiri, 

Yasuhiro  Kuwahara,  and  Shigeru  Nakao 

An  environmentally  based  growth  model  that  uses  finite 
difference  calculus  with  maximum  likelihood  method: 
its  application  to  the  brackish  water  bivalve 
Corbicula  /aponica  in  Lake  Abashiri,  Japan 

25-46  Brodeur,  Rick  D.,  Joseph  P.  Fisher,  David  J.  Teel, 

Robert  L.  Emmett,  Edmundo  Casillas, 
and  Todd  W.  Miller 

Juvenile  salmomd  distribution,  growth,  condition,  origin, 
and  environmental  and  species  associations 
in  the  Northern  California  Current 

47-62  Garcia-Rodrfguez,  Francisco  J., 

and  David  Aurioles-Gamboa 

Spatial  and  temporal  variation  in  the  diet  of  the 
California  sea  lion  (Zalophus  californianus) 
in  the  Gulf  of  California,  Mexico 

63-77  Jung,  Sukgeun,  and  Edward  D.  Houde 

Recruitment  and  spawning-stock  biomass  distribution 
of  bay  anchovy  (Anchoa  mitchilli)  in  Chesapeake  Bay 

78-93  Kellison,  Todd  G.,  and  David  B.  Eggleston 

Coupling  ecology  and  economy:  modeling 
optimal  release  scenarios  for  summer  flounder 
(Paralichthys  dentatus)  stock  enhancement 

94-107  Kritzer,  Jacob  P. 

Sex-specific  growth  and  mortality,  spawning  season, 
and  female  maturation  of  the  stripey  bass 
(Lut/anus  carponotatus)  on  the  Great  Barrrier  Reef 


Fishery  Bulletin  102(1) 


108-117  Orr,  Anthony  J.,  Adria  S.  Banks,  Steve  Mellman,  Harriet  R.  Huber, 

Robert  L.  DeLong,  and  Robin  F.  Brown 

Examination  of  the  foraging  habits  of  Pacific  harbor  seal  (Phoca  vitulina  richardsi)  to  describe  their  use 
of  the  Umpqua  River,  Oregon,  and  their  predation  on  salmonids 
Companion  paper  with  Purcell  et  al.,  see  "Notes"  below. 

118-126  Park,  Wongyu,  R.  Ian  Perry,  and  Sung  Yun  Hong 

Larval  development  of  the  sidestriped  shrimp  (Pandalopsis  dispar  Rathbun)  (Crustacea,  Decapoda,  Pandahdae) 
reared  in  the  laboratory 

127-141  Pearson,  Donald  E.,  and  Franklin  R.  Shaw 

Sources  of  age  determination  errors  for  sablefish  (Anop/opoma  fimbria) 

142-155  Powell,  Allyn  B.,  Robin  T.  Cheshire,  Elisabeth  H.  Laban,  James  Colvocoresses,  Patrick  O  Donnell, 

and  Marie  Davidian 

Growth,  mortality,  and  hatchdate  distributions  of  larval  and  juvenile  spotted  seatrout  (Cynoscion  nebulosus)  in 
Florida  Bay,  Everglades  National  Park 

156-167  Santana,  Francisco  M.,  and  Rosangela  Lessa 

Age  determination  and  growth  of  the  night  shark  (Carcharhinus  signatus)  off  the  northeastern  Brazilian  coast 

168-178  Smith,  Keith  R„  David  A.  Somerton,  Mei-Sun  Yang,  and  Daniel  G.  Nichol 

Distribution  and  biology  of  prowfish  (Zaprora  silenus)  in  the  northeast  Pacific 

179-195  Ward,  Peter,  Ransom  A.  Myers,  and  Wade  Blanchard 

Fish  lost  at  sea:  the  effect  of  soak  time  on  pelagic  longlme  catches 

196-206  Watanabe,  Chikako,  and  Akihiko  Yatsu 

Effects  of  density-dependence  and  sea  surface  temperature  on  interannual  variation  in  length-at-age 
of  chub  mackerel  (Scomber  japonicus)  in  the  Kuroshio-Oyashio  area  during  1970-1997 


Notes 

207-212  Llanos-Rivera,  Alejandra,  and  Leonardo  R.  Castro 

Latitudinal  and  seasonal  egg-size  variation  of  the  anchoveta  (Engrauhs  nngens)  off  the  Chilean  coast 

213-220  Purcell,  Maureen,  Greg  Mackey,  Eric  LaHood,  Harriet  Huber,  and  Linda  Park 

Molecular  methods  for  the  genetic  identification  of  salmonid  prey  from  Pacific  harbor  seal 
(.Phoca  vitulina  richardsi)  scat 

Companion  paper  with  Orr  et  al.,  see  "Articles"  above. 


221-229  Weng,  Kevin  C,  and  Barbara  A.  Block 

Diel  vertical  migration  of  the  bigeye  thresher  shark  (Alopias  superciliosus),  a  species  possessing 
orbital  retia  mirabilia 


231  Subscription  form 


Abstract— Fisheries  models  have  tradi- 
tionally focused  on  patterns  of  growth, 
fecundity,  and  survival  offish.  However, 
reproductive  rates  are  the  outcome  of 
a  variety  of  interconnected  factors 
such  as  life-history  strategies,  mating 
patterns,  population  sex  ratio,  social 
interactions,  and  individual  fecundity 
and  fertility.  Behaviorally  appropriate 
models  are  necessary  to  understand 
stock  dynamics  and  predict  the  success 
of  management  strategies.  Protogynous 
sex-changing  fish  present  a  challenge 
for  management  because  size-selective 
fisheries  can  drastically  reduce  repro- 
ductive rates.  We  present  a  general 
framework  using  an  individual-based 
simulation  model  to  determine  the 
effect,  of  life-history  pattern,  sperm 
production,  mating  system,  and  man- 
agement strategy  on  stock  dynamics. 
We  apply  this  general  approach  to  the 
specific  question  of  how  size-selective 
fisheries  that  remove  mainly  males 
will  impact  the  stock  dynamics  of  a 
protogynous  population  with  fixed 
sex  change  compared  to  an  otherwise 
identical  dioecious  population.  In 
this  dioecious  population,  we  kept  all 
aspects  of  the  stock  constant  except 
for  the  pattern  of  sex  determination 
(i.e.  whether  the  species  changes  sex 
or  is  dioecious).  Protogynous  stocks 
with  fixed  sex  change  are  predicted  to 
be  very  sensitive  to  the  size-selective 
fishing  pattern.  If  all  male  size  classes 
are  fished,  protogynous  populations  are 
predicted  to  crash  even  at  relatively  low 
fishing  mortality.  When  some  male  size 
classes  escape  fishing,  we  predict  that 
the  mean  population  size  of  sex-chang- 
ing stocks  will  decrease  proportionally 
less  than  the  mean  population  size  of 
dioecious  species  experiencing  the  same 
fishing  mortality.  For  protogynous  spe- 
cies, spawning-per-recruit  measures 
that  ignore  fertilization  rates  are  not 
good  indicators  of  the  impact  of  fishing 
on  the  population.  Decreased  mating 
aggregation  size  is  predicted  to  lead  to 
an  increased  effect  of  sperm  limitation 
at  constant  fishing  mortality  and  effort. 
Marine  protected  areas  have  the  poten- 
tial to  mitigate  some  effects  of  fishing 
on  sperm  limitation  in  sex-changing 
populations. 


Manuscript  approved  for  publication 
23  July  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull  102:1-13(2004). 


The  effects  of  size-selective  fisheries 

on  the  stock  dynamics  of  and  sperm  limitation 

in  sex-changing  fish 


Suzanne  H.  Aionzo 

Institute  of  Marine  Sciences  and  the  Center  lor  Stock  Assessment  Research  (CSTAR) 

University  of  California  Santa  Cruz 

1156  High  Street 

Santa  Cruz,  California  95064 

E-mail  address  shalonzoiS'ucscedu 


Marc  Mangel 

Department  of  Applied  Mathematics  and  Statistics 

Jack  Baskin  School  of  Engineering  and  the  Center  for  Stock  Assessment  Research  (CSTAR) 

University  of  California  Santa  Cruz 

1156  High  Street 

Santa  Cruz,  California  95064 


Fisheries  models  are  generally  used 
to  predict  the  impact  of  fishing  on 
stock  dynamics  and  yield  (Quinn  and 
Deriso,  1999;  Haddon,  2001).  Classic 
models  have  focused  mainly  on  growth, 
fecundity,  and  survival  of  species,  with- 
out considering  the  impact  of  mating 
patterns  on  reproduction,  survival, 
and  recruitment.  It  is  now  recognized 
that  life-history  strategies  and  mating 
behavior  will  affect  stock  dynamics. 
Even  so,  general  quantitative  predic- 
tions regarding  the  effect  of  specific 
life-history  patterns  on  fished  popula- 
tions are  limited  and  further  theory  is 
needed  (Levin  and  Grimes.  2002).  It 
is  likely  that  management  strategies 
taking  into  account  a  species'  reproduc- 
tive behavior  will  greatly  improve  our 
ability  to  manage  stocks  (e.g.  Beets  and 
Friedlander,  1999).  We  would  also  like 
to  know  when  the  mating  behavior  and 
reproductive  strategies  of  a  stock  will 
be  worth  investigating  and  when  tradi- 
tional management  techniques  will  be 
sufficient.  For  example,  in  a  manage- 
ment context,  how  do  sex-changing 
stocks  differ  from  separate-sex  species? 
Here,  we  take  an  initial  step  toward 
generating  a  theory  of  the  combined 
effect  of  life  history  and  mating  pat- 
terns on  stock  dynamics  by  focusing 
on  the  potential  for  and  effect  of  sperm 
limitation  in  a  protogynous  (female  to 
male)  sex-changing  stock.  We  focus 
on  protogyny  for  this  article  because 


numerous  protogynous  species  are  com- 
mercially important,  namely  red  porgy 
{Pagrus  pagrus),  gag  grouper  iMyc- 
teroperca  microlepis),  and  California 
sheephead  iSemicossyphus pulcher). 

Sex-changing  fish  present  a  unique 
challenge  for  management  because  size- 
selective  fisheries  have  the  potential  to 
drastically  reduce  reproductive  rates 
and  population  size  at  levels  of  fishing 
that  would  not  pose  a  problem  for  dioe- 
cious (separate-sex)  species  (Huntsman 
and  Schaaf,  1994;  Armsworth,  2001;  Fu 
et  al.,  2001).  On  the  other  hand,  pro- 
togynous stocks  may  be  less  sensitive 
to  the  removal  of  large  individuals  if 
females  are  not  fished  and  fertilization 
rates  remain  high.  Many  commercially 
important  species  are  known  to  change 
sex  (Bannerot  et  al.,  1987;  Shapiro, 
1987;  Coleman  et  al.,  1996;  Brule  et  al., 
1999;  Adams  et  al.,  2000;  Armsworth, 
2001;  Fu  et  al.,  2001).  Previous  models 
have  shown  that  sex-changing  fish  may 
be  vulnerable  to  fishing  (Bannerot  et 
al.,  1987;  Huntsman  and  Schaaf,  1994; 
Armsworth,  2001;  Fu  et  al..  2001). 

Complications  arise  because  the  ef- 
fect of  fishing  on  a  sex-changing  spe- 
cies is  mediated  by  many  aspects  of 
their  reproductive  biology,  such  as  sex 
ratio,  size-dependent  fecundity,  spawn- 
ing aggregation  size,  and  reproductive 
skew.  Furthermore,  patterns  of  sex 
change  have  cascading  effects  on  the 
sex  ratio,  social  interactions,  population 


Fishery  Bulletin  102(1) 


fecundity,  and  male  sperm  production — all  of  which  can 
affect  stock  dynamics.  Thus,  we  cannot  treat  sex  change  as 
an  isolated  aspect  of  a  species.  Instead,  we  must  consider 
sex  change  within  the  context  of  the  mating  system  and 
the  life  history  of  the  species  to  make  general  predictions. 
Behaviorally  appropriate  models  are  required  to  gener- 
ate constructive  qualitative  and  quantitative  theory.  Past 
theory  has  indicated  that  sex-changing  populations  exhibit 
stock  dynamics  that  often  differ  from  those  of  dioecious 
populations  (Bannerot  et  al.,  1987;  Huntsman  and  Schaaf, 
1994;  Armsworth,  2001;  Fu  et  al,  2001 ).  Furthermore,  pro- 
togynous  stocks  are  predicted  to  be  sensitive  to  fishing  pat- 
tern and  may  exhibit  nonlinear  dynamics  that  could  lead 
to  population  crashes  (Armsworth,  2001).  However,  it  is  not 
known  which  aspects  of  the  mating  behavior  and  life  his- 
tory pattern  of  sex-changing  stocks  drive  these  differences. 
Here  we  focus  on  comparing  a  protogynous  stock  with  an 
otherwise  identical  dioecious  population  to  determine  the 
effect  of  mating  aggregation  size,  fertilization  rates,  and 
life  history  pattern  on  stock  dynamics. 

Size-selective  (or  age-selective)  fisheries  can  impact  a 
species  through  a  decrease  in  spawning  stock  biomass,  in 
general  and  through  the  removal  of  highly  fecund  larger 
and  older  individuals,  in  particular  (Sadovy,  2001).  How- 
ever, in  protogynous  species,  fisheries  that  preferentially 
remove  large  males  can  also  change  the  population  sex 
ratio;  however,  the  exact  effect  of  fishing  pressure  on  stock 
dynamics  in  a  protogynous  species  is  complex.  At  one 
extreme,  the  complete  removal  of  males  from  the  popula- 
tion would  cause  a  stock  to  crash,  potentially  making  sex- 
changing  species  more  vulnerable  than  dioecious  species 
in  the  face  of  high  fishing  pressures.  At  the  other  extreme, 
sex-changing  species  may  be  less  affected  by  size-selective 
fisheries  if  female  fecundity  limits  recruitment  and  males 
are  not  removed  in  such  numbers  as  to  reduce  mating 
or  fertilization  rates.  Currently,  there  is  no  theory  that 
predicts  the  potential  for  sperm  limitation  in  protogynous 
stocks  as  a  function  of  gamete  production,  fertilization 
rates,  and  mating  pattern. 

It  has  been  suggested  that  marine  reserves  may  be  a  vi- 
able management  option  for  species  where  highly  fecund 
older  individuals  are  critical  to  reproduction  (Levin  and 
Grimes,  2002).  However,  no  theory  exists  that  can  predict 
the  impact  of  marine  reserves  on  stock  dynamics  in  sex- 
changing  species.  We  consider  the  impact  of  a  no-take 
marine  reserve  on  the  stock  dynamics.  We  compare  the 
effect  of  setting  aside  0-30%  of  the  spawning  population 
in  a  reserve.  We  assume  that  larval  production  is  exported 
from  within  the  reserve  to  the  rest  of  the  population  and 
determine  whether  the  reserve  can  mediate  some  of  the  ef- 
fects of  fishing  outside  the  reserve  because  this  represents 
the  optimal  scenario  for  marine  reserves.  We  also  compare 
mean  catch  rates  in  the  presence  and  absence  of  a  reserve 
as  a  function  of  fishing  mortality. 

Spawning-per-recruit  (SPR)  measures  are  often  used  to 
estimate  the  impact  of  fishing  on  a  stock  (Parkes,  2000; 
Jennings  et  al.,  2001).  Ideally,  a  spawning-per-recruit  mea- 
sure would  keep  track  of  per-recruit  production  of  larvae 
or  eggs  (Jennings  et  al.,  2001).  However,  spawning  stock 
biomass  per  recruit  (SSBR)  is  commonly  used  to  estimate 


the  reproductive  output  per  recruit  at  different  intensities 
of  fishing.  One  assumes  that  the  biomass  of  mature  fish  is 
linearly  related  to  reproductive  output,  which  may  be  the 
case  when  egg  production  limits  biomass  and  fecundity  in- 
creases linearly  with  biomass.  In  protogynous  stocks,  over- 
fishing of  males  alone  may  decrease  fertilization  rates  and 
hence  reproductive  output  without  affecting  either  female 
biomass  or  egg  production.  Thus,  in  protogynous  stocks  or 
sex-selective  fisheries,  classic  measures  of  spawning  per  re- 
cruit may  misrepresent  the  impact  of  fishing  on  the  stock's 
reproduction  and  hence  population  stability  (Punt  et  al., 
1993).  We  examine  a  variety  of  per-recruit  measures  and 
determine  their  ability  to  predict  changes  due  to  exploita- 
tion in  mean  population  size. 

In  this  study,  we  describe  a  general  approach  using  sex- 
and  size-dependent  individual-based  simulation  models 
that  predict  reproduction,  size  distribution,  and  sex  ratio 
in  fished  populations  as  a  function  of  mating  system  and 
sex-change  pattern.  We  examine  the  case  where  sex  change 
occurs  at  a  specific  size  threshold.  We  recognize  that  plastic 
and  socially  mediated  sex-change  patterns  have  been  ob- 
served, and  our  results  will  apply  only  to  species  with  fixed 
sex  change.  We  explore  the  impact  of  mating  aggregation 
size,  sperm  production,  and  asymptotic  fertilization  rates 
on  the  predicted  stock  dynamics  in  the  presence  of  exploita- 
tion. We  make  predictions  regarding  the  effects  of  fishing 
on  population  size,  reproduction,  sex  ratio,  size  distribu- 
tion, and  fertilization  rates.  We  also  compare  our  results 
to  previous  work  and  discuss  future  directions. 


Methods 

We  used  an  individual-based  simulation  to  predict  the  size 
distribution,  individual  and  population  fecundity,  popula- 
tion sex  ratio,  fertilization  rate,  and  population  size  as  a 
function  of  fishing  mortality  (Fig.  1).  Individuals  vary  in 
age,  size,  sex,  and  mating  site.  Population  size  varies  as  a 
function  of  baseline  survival,  fishing  mortality,  reproduc- 
tion, and  larval  recruitment.  Reproduction  depends  on  the 
pattern  of  sex  change,  mating  system,  sex  ratio,  mating  site, 
and  fecundity  (or  fertility)  of  individual  males  and  females. 
For  each  annual  time  period,  we  determined  individual 
survival,  the  size  and  age  of  these  individuals  in  the  next 
time  period,  and  the  total  production  of  surviving  offspring 
by  those  individuals.  Initial  analyses  showed  that  a  station- 
ary size,  sex,  and  age  distribution  is  found  within  approxi- 
mately 50  time  periods  and  is  independent  of  the  initial 
population  conditions.  Thus,  we  simulated  100  time  periods 
prior  to  examining  the  impact  of  fishing  on  stock  dynamics 
to  ensure  that  the  population  had  already  reached  the  sta- 
tionary size  and  sex  distribution  for  that  scenario  and  set 
of  parameters.  We  then  examined  the  model  for  100  repro- 
ductive seasons  in  the  presence  of  fishing  with  a  constant 
mean  fishing  mortality.  Because  a  number  of  elements  of 
the  model  were  stochastic,  we  examined  20  simulations  for 
each  scenario  and  set  of  parameter  values.  Initial  analyses 
indicated  that  20  simulations  were  more  than  sufficient  to 
lead  to  low  variability  in  the  key  measures  of  interest.  We 
assumed  that  reproduction  occurs  at  the  level  of  the  mating 


Alonzo  and  Mangel:  The  effects  of  size-selective  fisheries  on  the  stock  dynamics  of  and  sperm  limitation  in  sex-changing  fish         3 


group  at  different  reproductive  sites.  Individual  sur- 
vival, maturation,  sex  change,  and  mating  site  were 
determined  stochastically  as  described  below. 

Fishing  and  adult  survival 

We  assumed  that  adult  survival  is  density  indepen- 
dent but  depends  on  fishing  selectivity,  fishing  mor- 
tality, and  baseline  adult  mortality  in  the  absence  of 
fishing.  For  simplicity,  we  assumed  that  age  and  size  do 
not  affect  nonfishing  adult  mortality  p.A.  We  assumed 
that  the  fishery  is  size  selective;  we  let  L  represent  fish 
size,  F  represent  annual  fishing  mortality,  Lf  represent 
the  size  at  which  there  is  50%  chance  an  individual  of 
that  size  will  be  taken,  and  r  represent  the  steepness 
of  the  selectivity  pattern.  Then  fishing  selectivity  per 
size  class  siL)  is  given  by 


siD- 


l  +  exp-HL-L,)) 


and  adult  annual  survival  becomes 


cr(L)  =  exp(-/iA-Fs(L)) 


(1) 


(2) 


We  assumed  that  fishing  does  not  differentially 
affect  the  sexes  independent  of  size.  We  recognize, 
however,  that  for  some  species  this  may  not  be  the 
case.  We  also  assumed  that  fishing  occurs  each  year 
prior  to  reproduction  and  can  represent  either  pulse  or 
continuous  fishing  with  an  annual  mortality  F.  We  let 
N  it)  represent  the  number  of  individuals  in  age  class 
a  at  time  t  so  that  population  size  N(t)=  Sa  Na(t). 

Population  dynamics 

We  assumed  that  the  number  of  larvae  that  enter  the  popu- 
lation is  determined  by  the  production  of  fertilized  eggs  Pit) 
and  the  probability  that  those  larvae  will  survive  to  recruit. 
Pit)  is  determined  by  the  adult  fecundity  and  fertilization 
rates  described  below.  For  computational  tractability,  we 
also  assumed  that  a  population  ceiling Nmax exists  (Mangel 
and  Tier,  1993, 1994 ).  However,  we  chose  NmBX  large  enough 
that  the  stable  population  size  was  below  the  ceiling.  Larval 
survival  has  both  density-independent  and  density-depen- 
dent components  (e.g.  Cowen  et  al.,  2000;  Sale,  2002).  We 
used  a  Beverton-Holt  recruitment  function  to  determine 
larval  survival  to  the  next  age  class  (Quinn  and  Deriso, 
1999;  Jennings  et  al.,  2001).  Larvae  represented  the  zero- 
age  class  N0(t)  and  thus  the  number  of  larvae  surviving  to 
recruit  in  any  year  t  is  given  by 

Nnit)  =  (oPit))/(l+pPit))  if  (ctP(t))/(l+pP(t)) 

+JjNjt)<Nmax 

(3) 


AT0(*)  =  max|  0,Nmax-^Na(t)  |  if  (aPlt))/(l+ pPit)) 


MATING  SITES 

Adult  survival  determined  by  baseline  mortality  and  fishing  pattern 

Reproduction  determined  by  group  fecundity  and  fertility 

No  migration  between  mating  sites 


Density-dependent  and 
density-independent 
larval  survival 


Recruitment  random 
across  mating  sites 

Figure  1 

Structure  and  population  dynamics  of  the  individual-based 
model.  We  assumed  that  all  mating  sites  contribute  to  a  single 
larval  pool. 


where  a  gives  density-independent  survival;  and  /3  deter- 
mines the  strength  of  the  density-dependence  in  the  larval 
phase.  In  this  function,  we  used  the  number  of  fertilized 
eggs  produced,  Pit),  rather  than  spawning  stock  size.  We 
selected  parameter  values  for  larval  survival  that  allowed 
the  mean  population  size  to  be  stationary  near  the  ceiling 
in  the  absence  of  fishing.  We  assumed  a  single  larval  pool 
and  that  larvae  recruit  to  mating  sites  at  random  (Fig.  1). 
The  population  was  open  between  mating  sites  and  we 
were  simulating  the  entire  stock.  Thus,  there  was  no  emi- 
gration to  or  immigration  from  outside  populations. 

Growth  dynamics 

We  assumed  that  all  larvae  enter  the  population  at  the 
same  size,  L0.  We  assumed  that  growth  is  deterministic 
and  independent  of  sex  or  reproductive  status.  We  used  a 
discrete  time  version  of  the  von  Bertalanffy  growth  equa- 
tion (Beverton,  1987,  1992)  to  determine  growth  between 
age  classes  of  surviving  adults  in  which  Lmf  represents  the 
asymptotic  size  and  k  is  the  growth  rate.  Then  an  indi- 
vidual of  length  Lit)  at  time  t  will  grow  in  the  next  time 
period  to  size  Lit+1)  as  follows: 


Z,(f+l)  =  Llnf(l  +  exp(-fc))  +  L(f>exp(-A). 


(4) 


Mating  system 

We  assumed  that  reproduction  occurs  at  the  level  of  the 
mating  group,  and  we  examined  the  effect  of  varying  mating 
group  size  and  the  number  of  mating  sites.  We  assumed 


Fishery  Bulletin  102(1) 


that  juveniles  and  adults  exhibit  site  fidelity  but  that  larvae 
settle  randomly  among  mating  sites.  We  also  assumed  that 
the  population  carrying  capacity  is  split  equally  among  the 
mating  sites  and  that  the  total  capacity  of  all  mating  sites 
exceeds  the  maximum  population  size  in  the  absence  of  fish- 
ing as  determined  by  adult  mortality  and  the  recruitment 
function.  Therefore,  mating  sites  do  not  limit  recruitment 
but  may  affect  reproductive  rates.  We  examined  three  cases: 
1 )  the  entire  population  mates  at  one  site  (one  mating  site 
with  up  to  1000  individuals);  2)  a  few  large  mating  groups 
exist  ( 10  sites  with  a  maximum  of  100  individuals  per  site); 
and  3)  many  small  mating  aggregations  exist  (20  mating 
sites  with  a  maximum  of  50  individuals  per  site).  For  sim- 
plicity, we  assumed  that  within  a  mating  site,  individuals 
mate  in  proportion  to  their  fertility  and  fecundity.  Therefore, 
large  males  and  females  have  higher  expected  reproductive 
success.  However,  we  assumed  that  all  males  that  are  large 
enough  to  change  sex  have  a  chance  of  reproducing  propor- 
tional to  their  fertility.  This  is  equivalent  to  assuming  that 
females  exhibit  a  mate  choice  threshold  I  Janetos,  1980)  that 
has  evolved  with  the  size-at-sex  change  and  that  females 
have  an  equal  probability  of  mating  with  males  above  this 
size  threshold.  However,  a  large  male  mating  advantage 
clearly  still  exists.  We  also  assumed  that  fishing  mortality 
remains  constant  as  mating  aggregation  size  varies.  Thus, 
we  assumed  that  fishing  effort  per  site  does  not  increase  as 
the  number  of  mating  sites  decreases.  An  alternative  would 
be  to  assume  that  total  fishing  mortality  increases  as  the 
number  of  mating  aggregations  decreases. 

Maturity 

The  probability  that  an  individual  matures  pm(L)  is  deter- 
mined by  size.  Once  an  individual  matures,  she  remains 
female  until  sex  change  (see  below).  We  let  Lm  represent 
the  length  at  which  50%  of  the  individuals  will  have 
matured. 


EiL)=aLh, 


(7) 


P,JL)- 


1 


where  a  and  b  are  constants. 

Once  an  individual  has  changed  sex  (as  determined  by 
the  sex  change  rule  described  above)  sperm  production  (in 
millions)  S(L)  is  given  by 


S{L)=cLd , 


(8) 


l  +  exp(-q(L-  Lm 


(5) 


where  c  and  d  are  constants. 

Size-dependent  fecundity  has  been  measured  in  many 
fish  species  (e.g.  Gunderson,  1997).  A  general  allometric 
relationship  between  sperm  production  and  size  has  not 
been  established.  Therefore,  we  assumed  that  male  gamete 
production  increases  with  size  at  the  same  rate  as  that  for 
females  ib=d).  We  also  assumed  that  males  produce  many 
more  sperm  at  any  body  length  than  females  produce 
eggs.  Clearly,  other  possible  patterns  exist.  We  examined 
the  case  where  males  produce  from  102  to  106  sperm  for 
every  egg  produced  by  a  female.  In  the  pelagic  spawning 
wrasse  (Thalassoma  bifasciatum ),  large  males  release  ap- 
proximately 1000  times  more  sperm  than  females  release 
eggs  (Schultz  and  Warner,  1991;  Warner  et  al.,  1995). 

We  used  recently  published  data  on  sperm  production 
and  fertilization  rates  in  the  bluehead  wrasse  (Thalas- 
soma bifasciatum)  to  generate  a  biologically  appropriate 
fertilization  function  for  our  model  (Warner  et  al.,  1995; 
Petersen  et  al.,  2001).  It  is  critical  to  consider  a  biologically 
appropriate  form  for  the  function  to  express  fertilization 
rates  when  considering  the  potential  for  sperm  limitation. 
The  probability  an  egg  will  be  fertilized  is  an  increasing 
function  of  the  number  of  sperm  available  for  that  mat- 
ing (Fig.  2).  The  number  of  eggs  released  per  mating  also 
affects  the  fertilization  rate  (Fig.  2).  For  simplicity,  we  cal- 
culated the  average  expected  fertilization  rate  per  mating 
site  based  on  the  total  production  of  sperm  and  eggs  at  the 
site.  We  let  S  represent  the  number  of  sperm  released  (in 
millions)  and  £  the  number  of  eggs  released  at  each  mating 
site.  We  assumed  that  the  proportion  of  eggs  fertilized  per 
mating  site  pF  is  given  by 


where  q  determines  the  steepness  of  the  probability 
function. 

Sex  change 

The  probability  of  sex  change,  pciL),  is  a  logistic  function 
of  absolute  size  L 


P,.(L)  = 


l  +  exp(-p(L-L,  )) 


(6) 


where  Lr  represents  the  size  at  which  50%  of  the  indi- 
viduals will  change  sex  from  female  to  male  and  p  is  a 
constant. 

Reproduction 

We  assumed  that  female  fecundity  E(L)  depends  on  indi- 
vidual size  according  to  the  allometric  relationship 


Pf 


l  +  iisE  +  X)S 


(9) 


where  k  and  %  are  constants  fitted  to  the  data. 

The  number  of  eggs  fertilized  per  group  is  ph-E  and  the 
total  production  of  fertilized  eggs.  Pit),  is  the  sum  of  the 
number  of  eggs  fertilized  in  all  mating  groups. 

Measures  of  spawning  stock  biomass  per  recruit 

To  measure  the  impact  of  fishing  on  stock  dynamics,  we 
computed  the  total  spawning  stock  biomass  per  recruit 
starting  from  the  beginning  of  fishing  for  the  next  50 
years.  We  used  the  generally  recognized  pattern  that 
fish  wet  weight  tends  to  be  approximately  proportional 
to  the  cube  offish  length  (Gunderson,  1997)  to  convert 
fish  length,  L,  into  relative  biomass,  B(L)~L\  Then  we 
calculated  total  female  and  male  spawning  stock  biomass 


Alonzo  and  Mangel:  The  effects  of  size-selective  fisheries  on  the  stock  dynamics  of  and  sperm  limitation  in  sex-changing  fish 


per  recruit  (SSBR).  We  also  kept  track  of  the 
total  fecundity  (egg  production  per  recruit  I, 
fertility  (sperm  production  per  recruit),  and 
eggs  fertilized  per  recruit. 

Marine  reserves 


^=150  (about  60  females) 


OS- 


'S     0.6 


■s     0.4 


02 


We  examined  the  effect  of  no-take  marine 
reserves  on  the  predicted  stock  dynamics  by 
comparing  the  stock  dynamics  in  the  presence 
and  absence  of  reserves.  Without  a  reserve, 
individuals  at  all  mating  sites  are  subject  to 
fishing.  In  the  presence  of  a  no-take  marine 
reserve,  we  "protect"  a  percentage  of  the 
mating  sites  (and  thus  the  population)  from 
fishing.  We  examined  cases  in  which  09c,  10%, 
20%,  and  30%  of  mating  sites  were  protected 
from  fishing.  We  assumed  that  the  population 
is  completely  open  among  mating  sites.  Thus, 
eggs  produced  from  all  mating  sites  enter  one 
larval  pool  and  recruitment  occurs  randomly 
between  mating  sites.  Clearly  other  possibili- 
ties exist  and  could  be  considered  in  future  analyses,  but 
this  case  represents  a  reasonable  baseline  situation  to  con- 
sider because  many  marine  fish  have  pelagic  larval  phases. 
We  also  recognize  that  these  analyses  ignore  the  effect  of 
interactions  between  species  within  the  reserve  on  stock 
dynamics.  We  examined  two  situations.  In  the  first  case, 
reduced  fishing  effort  occurs  when  mean  fishing  mortality 
is  decreased  in  the  presence  of  reserves  because  fishing 
mortality  (F)  at  the  unprotected  sites  remains  the  same  as 
before  the  reserve.  In  the  second  case,  the  redistribution  of 
fishing  effort  occurs  when  mean  fishing  mortality  across  all 
sites  remains  the  same  because  fishing  mortality  increases 
at  the  unprotected  sites. 

Comparison  of  sex-changing  stocks  and  dioecious  stocks 

Ideally,  we  would  like  to  distinguish  the  effects  of  sex  change 
in  isolation  from  the  confounding  effects  of  mating  pattern, 
sex  ratio,  survival,  growth,  and  population  fecundity  on 
stock  dynamics.  To  differentiate  whether  sex  change  in  iso- 
lation or  other  aspects  of  the  mating  system  determine  the 
predicted  stock  dynamics,  we  also  examined  a  version  of  the 
model  described  above  for  a  population  where  sex  is  fixed 
at  birth.  In  this  dioecious  population,  we  keep  all  aspects 
of  the  stock  constant  except  for  the  pattern  of  sex  determi- 
nation (whether  the  species  changes  sex  or  is  dioecious). 
One  would  generally  expect  a  dioecious  population  with 
no  differences  between  the  sexes  in  mortality  to  exhibit  a 
50:50  sex  ratio  ( Fisher,  1930;  Trivers,  1972;  Charnov,  1982 ). 
However,  we  wanted  to  control  for  all  differences  between 
the  dioecious  and  protogynous  stocks  other  than  the  sex- 
determination  pattern.  Therefore,  we  considered  the  same 
sex  ratio  at  maturity  (0.67=the  proportion  of  adults  that 
are  female)  as  found  in  the  sex-changing  population  in  the 
absence  of  fishing.  Assuming  no  sex-specific  differences  in 
survival  to  maturity,  this  is  the  same  as  assuming  a  0.67 
sex  ratio  at  birth.  In  this  model,  individuals  remain  one  sex 
(determined  randomly  at  birth)  throughout  their  lifetime. 


Km=1750  (about  700  females) 


K>750  (about  300  females) 


5.000 


10.000 


1 5.000 


20,000 


Sperm  number  (S) 
(in  millions,  about  1  to  100  males) 

Figure  2 

Fertilization  rate  as  a  function  of  the  number  of  eggs  and  sperm  per  mating 
site.  The  saturation  parameter  Km=\E+x  is  taken  from  Equation  9. 


Fishing  is  size  but  not  sex  selective.  We  assumed  that  males 
mature  at  the  same  size  as  females. 

Parameter  values 

We  used  previous  research  on  California  sheephead  (Lab- 
ridae,  Semicossyphus  pulcher),  a  commercially  important 
sex-changing  fish,  to  provide  evolutionarily  and  ecologi- 
cally reasonable  parameters  for  the  model.  Although  the 
growth,  survival,  and  reproduction  of  this  species  have 
been  studied,  less  is  known  about  the  factors  that  induce 
sex  change  and  mating  behavior.  In  this  species,  sex  change 
occurs  at  approximately  30  cm  although  the  exact  pattern 
varies  among  populations  (Warner,  1975;  Cowen,  1990).  It 
is  not  known  whether  sex  change  is  fixed  or  socially  medi- 
ated. Because  nothing  is  known  about  fertilization  rates  in 
the  California  sheephead,  we  generated  k  and  y  <Eq.  9)  by 
fitting  a  line  through  the  estimated  values  of  Km  for  small 
and  large  bluehead  wrasse  females  as  a  function  of  their 
mean  egg  production  (see  Table  1  and  Fig.  2;  Warner  et 
al.,  1995;  Petersen  et  al.,  2001).  For  parameter  values  and 
sensitivity  analyses  see  Table  1. 


Results 

We  present  the  average  across  20  simulations  of  the  mean 
population  measures  of  the  last  50  years  for  each  simula- 
tion. The  variation  around  the  mean  in  all  measures  con- 
sidered was  very  low  (hundredths  of  a  percent  of  the  mean 
or  less).  For  the  spawning  per  recruit  (SPR)  measures  we 
give  the  mean  value  across  the  first  50  years  of  fishing  to 
ensure  that  the  entire  cohort  had  died  before  the  end  of  the 
simulation.  When  the  ratio  of  sperm  to  eggs  is  104  to  106, 
a  single  male  can  fertilize  all  of  the  eggs  in  the  population. 
When  the  ratio  of  sperm  to  eggs  is  102,  sperm  limitation 
occurs  even  in  the  absence  of  fishing.  Therefore,  we  present 
results  for  the  case  where  the  ratio  of  sperm  to  eggs  is  103 


Fishery  Bulletin  102(1) 


Table  1 

The  following  parameters  were  used  in  the  model. 

Parameter 

Baseline  values 

Definition  and  source 

Growth 

* 

0.05 

growth  rate  (based  on  Cowen,  1990) 

*W 

90  cm 

asymptotic  size  (based  on  Cowen,  1990) 

h 

8  cm 

larval  size  at  recruitment 

Population 

N 

max 

1000 

maximum  population  size 

V-A 

0.35 

adult  mortality  (based  on  Cowen,  19901 

a 

0.0001 

density-independent  larval  mortality 

P 

a/(l-exp(- 

-H» 

))N 

,,^3.33x10--) 

larval  recruitment  function  parameter  (see  text) 

Fishing 

r 

1     (0.1) 

steepness  of  selectivity  curve 

Lf 

30     (25,35) 

length  at  which  509?  chance  a  fish  will  be  removed 

F 

0-3 

fishing  mortality 

Reproduction 

a 

7.04 

constant  in  the  fecundity  relationship  (Warner,  1975) 

6 

2.95 

exponent  in  the  fecundity  relationship  (Warner,  1975) 

c 

10-3a  (10- 

2a, 

10" 

4Q) 

constant  in  the  sperm  production  function  (measured  in  millions 
of  sperm) 

d 

b 

exponent  in  the  fertility  relationship  (Warner,  1975) 

K 

0.000003 

slope  of  fertilization  function  parameter 

X 

0.09 

intercept  of  fertilization  function  parameter  (based  on  Peterson 
et  al.,  2001)  see  text  for  details 

Maturity 

Lm 

20  cm 

length  at  which  507c  offish  mature  (Warner,  1975;  Cowen.  1990) 

Q 

1 

shape  parameter  in  the  maturity  function 

Sex  change 

h 

30  cm 

length  at  which  50%  offish  change  sex  (Warner,  1975;  Cowen.  1990) 

P 

1 

shape  parameter  in  the  sex  change  function 

and  fertilization  rates  are  100%  in  the  absence  of  fishing, 
but  the  population  must  have  multiple  males  for  high  fer- 
tilization rates.  For  all  the  results  presented  in  our  study 
we  assumed  a  fixed  sex-change  pattern,  mating  among 
males  and  females  at  each  site  proportional  to  gamete 
production,  and  larval  export  among  mating  sites.  We  also 
assumed,  unless  otherwise  noted,  a  sharp  size-selective 
fishing  pattern  (r=l)  and  that  the  probability  of  sex  change 
and  removal  of  sex-changing  fish  by  the  fishery  are  cen- 
tered at  the  same  mean  size  or  Lj=Lc.  Clearly,  the  results 
presented  in  our  study  may  not  apply  to  cases  where  these 
assumptions  are  not  met. 

General  patterns  predicted  by  the  model 

First,  we  examined  the  general  effect  of  fishing  mortality 
on  the  sex-changing  stock  for  the  case  when  one  mating 
site  exists.  When  Lf=  Lc,  eggs  produced  per  recruit  decrease 
only  slightly  with  fishing  mortality  (e.g.  a  3%  drop  as  fish- 


ing mortality  increased  from  0  to  3,  Fig.  3A).  However,  the 
mean  number  of  eggs  fertilized  (both  total  and  per  recruit) 
decreases  sharply  as  fishing  mortality  increases  (e.g.  a  30% 
drop  as  fishing  mortality  increased  from  0  to  3,  Fig.  3A). 
The  number  of  recruits  per  year  decreases  as  well.  As  fish- 
ing mortality  increases,  male  spawning  stock  biomass  per 
recruit  decreases  dramatically,  whereas  changes  in  female 
spawning  stock  biomass  would  be  practically  undetectable 
(90%  drop  for  male  SSBR,  compared  with  a  3%  drop  for 
female  SSBR  as  F  increases  from  0  to  3,  Fig.  3B).  Because  of 
the  drop  in  male  SSBR,  total  spawning  stock  biomass  (males 
and  females)  per  recruit  also  decreases  as  fishing  mortal- 
ity increases.  Sperm  production  per  recruit  is  predicted  to 
decrease  with  increasing  fishing  mortality  (Fig.  3C). 

Sensitivity  of  stock  dynamics  to  fishing  pattern 

In  general,  mean  population  size  decreases  as  fishing  pres- 
sure increases  (Fig.  4A).  The  adult  sex  ratio  (measured  as 


Alonzo  and  Mangel:  The  effects  of  size-selective  fisheries  on  the  stock  dynamics  of  and  sperm  limitation  in  sex-changing  fish 


the  percentage  of  mature  individuals  that  are  female)  also 
increases  as  fishing  mortality  increases  (Fig.  4B)  and  the 
mean  size  of  adults  in  the  population  decreases.  These  pat- 
terns depend  on  fishing  being  size  selective,  which  causes  a 
disproportional  take  of  males.  If  the  size-selectivity  of  the 
fishery  targeted  smaller  size  classes  (L<LC),  a  decline  in 
annual  biomass  removed  by  the  fishery  is  predicted  with 
increasing  F  and  the  stock  is  predicted  to  crash  at  a  rela- 
tively low  fishing  mortality  (Fig.  4C).  If  the  fishery  is  less 
selective  (r=0.1,  L,=LC),  the  population  is  also  predicted 
to  crash  for  most  fishing  mortalities.  Thus,  allowing  some 
proportion  of  mature  males  to  consistently  escape  fishing 
is  critical  even  at  low  fishing  mortality.  As  fishing  mortality 
increases,  the  predicted  biomass  removed  by  the  fishery 
increases  with  diminishing  returns  ( Fig.  4C ).  When  Lf=Lc, 
the  biomass  removed  by  the  fishery  does  not  continue  to 
increase  with  F  because  all  males  above  the  size  at  sex 
change  are  being  removed  by  the  fishery.  In  this  case,  the 
males  in  the  population  are  essentially  breeding  only  once 
before  they  are  taken  by  the  fishery.  For  the  range  of  fish- 
ing mortality  considered,  we  did  not  observe  a  decline  in 
biomass  taken  with  increasing  F  unless  L<Lt.  or  r=0.1.  If 
more  size  classes  are  allowed  to  escape  fishing  (Lf>Lc),  the 
general  patterns  remain  the  same,  but  for  the  same  fishing 
mortality  (.F),  the  effect  of  fishing  on  the  population  is  less 
(Fig.  4).  Female  biomass  does  not  decrease  much  with  fish- 
ing mortality  when  Lf=Lc  even  though  some  females  are 
removed  by  the  fishery  because  the  probability  of  a  female 
changing  sex  is  the  probability  of  it  being  fished.  Therefore, 
female  loss  due  to  the  fishery  affects  male  biomass  rather 
than  female  biomass  in  the  population. 

Sperm  limitation  and  production 

The  removal  of  large  males  from  the  population  is  pre- 
dicted to  cause  sperm  limitation  and  decreased  fertiliza- 
tion rates  (Fig.  3,  A  and  C),  leading  to  a  decrease  in  mean 
population  size  (Fig.  4A).  The  degree  to  which  the  fertiliza- 
tion rate  and  thus  the  population  size  decreases  depends 
to  a  great  extent  on  the  pattern  of  sperm  production 
and  fertilization.  We  assumed  that  only  a  few  males  are 
needed  to  fertilize  the  eggs  of  many  females  (Fig.  2).  We 
also  assumed  that  per-capita  reproduction  and  recruitment 
are  high  even  at  a  low  population  size  (Barrowman  and 
Myers,  2000).  Thus,  protogynous  populations  with  lower 
sperm  production  or  fertilization  rates  would  experience 
greater  effects  from  fishing  than  predicted  in  the  present 
study.  Similarly,  populations  with  lower  production  or  sur- 
vival would  experience  larger  decreases  in  population  size 
even  with  the  same  level  of  sperm  limitation  and  fishing. 
In  general,  however,  the  removal  of  males  alone  from  a  pro- 
togynous population  with  a  fixed  sex  change  is  predicted  to 
cause  decreased  fertilization  rates  and  lower  mean  popula- 
tion size  even  when  the  fertilization  rate  function  is  asymp- 
totic and  individual  male  sperm  production  is  high. 

Mating  aggregation  size 

As  mating  aggregation  size  decreased  and  fishing  mortality 
and  effort  remained  constant,  the  effect  of  fishing  on  the  pop- 


Eggs  produced 


1  1.5  2 

Fishing  mortality  (F) 

Figure  3 

Spawning-per-recruit  measures.  Results  are  presented  for 
the  sex-changing  stock  with  one  mating  site  when  L^=  Lc 
and  r=l.  Means  across  20  simulations  are  given.  For  details 
see  the  general  text. 


ulation  increased.  As  described  above,  we  assumed  that  fish- 
ing effort  would  not  be  concentrated  on  the  few  large  mating 
aggregations  and  thus  increase  total  fishing  mortality.  The 
sex  ratio,  mean  size,  mean  fecundity,  and  mean  fertility  all 
remained  the  same  across  different  mating  aggregation 
sizes  with  constant  fishing  mortality.  However,  the  mean 
fertilization  rate  and  number  of  fertilized  eggs  per  recruit 
decreased  with  mating  group  size  ( Fig.  5 )  even  though  male 
biomass  and  SSBR  remained  the  same.  Both  predicted 
mean  population  size  and  biomass  taken  decreased  as  fish- 
ing mortality  increased  (Fig.  5).  This  pattern  was  generated 
by  sperm  limitation  in  small  mating  groups.  Smaller  groups 
have  higher  probabilities  that  sperm  production  within  the 
group  will  not  be  sufficient  to  fertilize  the  eggs  produced 
within  the  mating  group.  Small  mating  aggregations  may 
not  only  be  sperm  limited  but  also  be  male  limited  and  fail 
to  reproduce  completely;  populations  with  small  group  sizes 
(50  individuals  or  less)  were  predicted  to  become  extinct  in 


Fishery  Bulletin  102(1) 


5-25%  of  the  simulations  as  fishing  mortality  (F)  increased 
from  0  to  1.  The  impact  of  mating  group  size  on  stock  dynam- 
ics is  thus  predicted  to  be  nonlinear.  A  threshold  mating 
aggregation  size  appeared  to  exist  below  which  sperm  limi- 
tation and  reproductive  failure  become  common. 

Spawning-per-recruit  measures 

For  size-selective  fishing,  the  spawning  stock  biomass  per 
recruit  of  females  is  not  predicted  to  decrease  significantly 
with  increased  fishing  mortality  as  long  as  some  male  size 
classes  escape  fishing  (Lr>Lv).  However,  male  biomass  per 
recruit  and  sperm  production  per  recruit  are  both  predicted 
to  decrease.  Although  egg  production  is  not  predicted  to 


900 
800 

A 

L,>LC 

CD 

n     700 

to 

^\            L,=LC 

<=     600  J 

o 

'ra     500  " 

3 

g-    400  " 

Q. 

c     300  " 

ra 

|     200  ' 

\l,<lc 

100 

0 

i 

0 

0.5              1               1.5             2              2.5             3 

1   , 

B 

L,=LC 

x  ratio 
male) 

o    o    c 
-g    03    CO 

<n  •£     °  6  ' 

g  .2     0.5  • 

ra   o     0.4  . 

3   Q. 

o.  2     0.3  . 

P   a. 

0-  —    0.2  . 

0.1   ■ 

\l,<lc 

o 

0 

0.5              1              1.5              2              2.5              3 

600.000 

C 

L,=LC 

iomass 
the  fishery 

o       o 
o       o 

o       o 

o       o 

■o   >, 

ra  £?     300.000 

D   T3 

a   a> 

c    > 

<   °     200.000 

CD 

100,000 

U/< 

0 

0 

0.5              1               15              2              2.5             3 

Fishing  mortality  (F) 

Figure  4 

The  effect  of  size-s 

ilect  ive  fishing  on  stock  dynamics.  We  present 

results  for  the  sex-changing  stock  with  one  mating  site  when 

r=l.  Means  across 

20  simulations  are  given.  For  details  see  the 

general  text. 

decrease  with  increasing  size-selective  fishing  pressure, 
the  number  of  fertilized  eggs  is  predicted  to  decrease. 
When  all  male  size  classes  are  fished  iL.>Lc),  the  stock 
is  predicted  to  crash  and  therefore  clearly  female  biomass 
and  egg  production  are  predicted  to  decrease  with  fishing 
mortality.  In  general,  the  predicted  decrease  in  mean  popu- 
lation size  and  reproduction  is  driven  for  the  most  part  by 
decreased  sperm  production  and  consequently  a  reduction 
in  the  number  of  eggs  fertilized  per  recruit.  The  relation- 
ships between  fishing  pressure  and  the  classic  spawning- 
per-recruit  measures  do  not  indicate  the  true  effect  that 
fishing  is  predicted  to  have  on  the  protogynous  population 
(Fig.  6).  When  Lf>Lc,  female  spawning  stock  biomass  per 
recruit  and  eggs  produced  per  recruit  showed  almost  no 
effect  of  fishing  on  the  population,  even  as  mean 
population  size  decreased.  Because  of  the  size-selec- 
tive fishing  pattern,  total  and  male  biomass  per  recruit 
decreased  with  fishing  mortality  and  decreasing  mean 
population  size.  However,  male  and  total  biomass  per 
recruit  did  not  reflect  the  increased  effect  of  fishing  on 
populations  with  smaller  mating  aggregations.  The 
production  of  fertilized  eggs  per  recruit  decreased  with 
increased  fishing  pressure  and  decreased  more  sharply 
for  smaller  mating  aggregations.  Only  the  number  of 
fertilized  eggs  per  recruit  could  assess  the  predicted 
effect  of  fishing  on  the  protogynous  population.  Thus, 
classic  SPR  measures  were  predicted  to  fail  in  the 
presence  of  sperm  limitation  to  assess  the  impact  of 
fishing  on  a  protogynous  stock. 

Marine  reserves  and  fishery  management 

In  the  situation  considered  in  this  study,  the  pattern 
of  fishing  is  more  important  to  stock  dynamics  than 
the  presence  of  marine  reserves.  We  assumed  a  size- 
selectivity  that  allowed  on  average  50%  of  individuals 
of  sex-changing  size  to  escape  the  fishing  gear.  Thus, 
although  the  sex  ratio  does  increase  (become  more 
female)  by  20-40%,  all  males  are  not  lost  from  the 
population  (when  Lfs.Lt.  and  r=l ).  If  fishing  selectivity 
occurs  at  a  smaller  size,  then  the  effects  on  the  popula- 
tion are  predicted  to  be  much  greater  and  the  protogy- 
nous stock  would  suddenly  become  more  affected  than 
the  dioecious  population.  For  example,  at  L^=25  cm  the 
protogynous  stock  is  predicted  to  crash  whenever  F^l. 
This  occurs  not  because  of  a  reduction  in  the  produc- 
tion of  eggs  but  rather  because  of  a  failure  to  fertilize 
the  eggs  produced  by  surviving  females.  When  males 
of  all  size  classes  are  fished,  populations  can  become 
male  limited  and  fertilization  rates  drop  drastically.  A 
decrease  in  the  production  of  fertilized  eggs  can  lead  to 
a  decrease  in  female  biomass,  but  it  is  the  removal  of 
males  rather  than  females  that  causes  this  decline. 

When  fishing  effort  is  not  redistributed  after  the 
formation  of  a  reserve,  the  impact  of  fishing  on  the 
mean  population  size  and  SPR  measures  is  predicted 
to  decrease  (e.g.  Fig.  7A).  However,  if  fishing  effort  is 
redistributed  among  unprotected  areas,  the  benefit 
of  the  reserves  to  the  protogynous  stock  decreases 
(Fig.  8A).  Protecting  some  sites  allows  large  males  to 


Alonzo  and  Mangel:  The  effects  of  size-selective  fisheries  on  the  stock  dynamics  of  and  sperm  limitation  in  sex-changing  fish        9 


escape  fishing  and  thus  increases  the  pro- 
duction of  fertilized  eggs  at  the  population 
level.  However,  yield  decreased  proportion- 
ally to  the  percentage  of  sites  protected  by 
the  reserve  unless  fishing  effort  is  redis- 
tributed among  the  remaining  sites.  We  as- 
sumed that  fish  do  not  move  between  sites 
after  the  larval  stage,  and  thus  larger  and 
older  individuals  do  not  leave  the  reserve 
and  become  exposed  to  fishing.  Although  this 
assumption  is  clearly  appropriate  for  some 
species,  it  is  important  to  realize  that  the  dy- 
namics and  predictions  would  differ  for  more 
closed  populations  or  migratory  species.  For 
the  fishing  pattern  and  biological  scenario 
examined  in  this  study,  marine  reserves  are 
not  predicted  to  increase  biomass  available 
to  the  fishery  (Figs.  7B  and  8B). 

Dynamics  of  dioecious  versus  protogynous 
stocks 

In  the  dioecious  stock  with  a  single  ran- 
domly mating  aggregation,  both  male  and 
female  biomass  per  recruit  and  fecundity 
or  fertility  per  recruit  are  predicted  to 
decrease  as  fishing  mortality  increases  ( Fig. 
6).  Because  both  egg  production  and  sperm 
production  decrease  with  increased  fishing 
pressure  in  the  dioecious  stock,  the  number 
of  eggs  fertilized  per  recruit  did  not  differ 
much  from  the  other  SPR  measures.  Thus, 
SSBR  and  eggs  per  recruit  also  indicated  the 
impact  of  fishing  on  the  stock  in  dioecious 
stocks  with  large  mating  aggregations.  The 
percent  drop  in  population  size  and  fertil- 
ized egg  production  is  predicted  to  be  much 
greater  in  dioecious  species  and  occurred 
more  quickly  than  in  the  sex-changing 
stock  because  of  a  reduction  in  overall 
population  fecundity  even  in  the  absence  of 
decreased  fertilization  rates.  However,  dioe- 
cious stocks  are  predicted  to  exhibit  larger 
mean  population  size  for  the  same  fishing 
mortality  and  to  support  a  larger  fishery 
because  of  the  additional  egg  production  of 
large  fecund  females.  At  very  small  mating 
aggregations,  sperm  limitation  is  predicted 
even  in  the  dioecious  stock  and  fertilized 
eggs  per  recruit  become  a  better  indicator 
of  stock  dynamics  in  the  presence  of  fishing. 
Dioecious  stocks  are  also  predicted  to  benefit 
from  marine  no-take  reserves  through  the 
protection  of  large  fecund  females  ( Fig.  7 ). 


Discussion 

In  this  study  we  developed  a  general  frame- 
work that  examines  the  consequences  to 


0.95 


0.9 


0.85 


Egg  production 
(per  recruit) 


Fertilized  eggs 
(per  recruit) 


Mean  population 
size 


Figure  5 

Mating  aggregation  size  affects  the  response  to  fishing.  Large  (one  large 
mating  aggregation )  and  small  ( 10  smaller  mating  aggregations  I  situations 
are  compared.  Percent  change  in  the  presence  of  fishing  (from  F=0  to  F=l> 
in  egg  production  per  recruit,  mean  fertilized  egg  production  per  recruit,  and 
mean  population  size  are  given.  Total  population  fecundity  and  mean  body 
size  are  lower  for  the  smaller  mating  aggregations. 


PROTOGYNOUS  POPULATION 


1 1 

F=3 

Eggs  produced 

F=0 

r&                            S 

0.9- 
0.8- 

Eggs  fertilized        _n 
.a 
St' 

»■" 

0.7- 

a®                       x 

o 

Eggs  produced 
and  fertilized 

DIOECIOUS  POPULATION 

0.6- 
0.5. 

ft 

ti 

* 

0.4 

F=3 

F=0 

600  650  700  750  800  850  900  950 

Mean  population  size 

Figure  6 

Spawning-per-recruit  (SPR)  measures  in  a  protogynous  (squares)  and  dioe- 
cious (triangles)  stock:  Mean  egg  production  per  recruit  (filled)  and  mean 
fertilized  eggs  per  recruit  (open)  are  shown  for  a  randomly  mating  popula- 
tion with  one  large  mating  group.  Error  bars  indicate  the  standard  error  of 
the  mean.  For  the  dioecious  population,  the  two  SPR  measures  overlap. 


10 


Fishery  Bulletin  102(1) 


fisheries  management  of  a  behaviorally  and  evolution- 
ary reasonable  life-history  and  sex-change  pattern.  We 
based  our  assumptions  and  parameter  values  on  patterns 
observed  in  natural  populations  that  have  presumably 
evolved  given  the  life  history  tradeoffs  and  expected  repro- 
ductive success  associated  with  these  behaviors.  However, 
we  made  various  assumptions  that  affect  the  predicted 
patterns  such  as  a  fixed  sex-change  pattern,  male  mating 
success  proportional  to  sperm  production,  and  a  very  resil- 
ient recruitment  function.  Despite  these  assumptions,  a 
number  of  general  patterns  emerge. 

Life-history  pattern  is  important  but  not  sufficient 
to  predict  stock  dynamics 

In  general,  we  predicted  that  a  protogynous  stock  with 
fixed  sex  change  will  respond  to  the  same  fishing  pressure 


o 
o  ^ 

fl 

Q.— ' 

°>  S3, 
-=    cr> 


X 


<D 


0)     Q 


0.75 
0.5  ■ 


0.25 


Protogynous 


t-  C\J 


Dioecious 


B 


E 
g 
a 


600,000- 
500,000' 
400,000- 
300,000- 
200,000- 
100.000 


Protogynous 


Dioecious 


Figure  7 

The  effect  of  marine  reserves  on  protogynous  and  dioecious  popula- 
tions when  fishing  effort  is  decreased  (case  1 1. 1  A)  Percent  change 
in  the  presence  of  fishing  CF=1)  in  the  production  of  fertilized  eggs 
compared  to  in  the  absence  of  fishing.  ( B)  Annual  biomass  removed 
by  the  fisheries  varies  with  marine  reserve  and  sex-change  pat- 
tern. Numbers  shown  are  for  10  mating  sites  when  F=l. 


differently  than  an  otherwise  identical  dioecious  stock. 
Understanding  the  life  history  of  the  population  is  clearly 
important  to  our  understanding  of  stock  dynamics.  How- 
ever, it  is  not  possible  to  classify  protogynous  stocks  simply 
as  more  or  less  sensitive  to  fishing.  The  differences  between 
dioecious  and  sex-changing  fish  are  relatively  complex,  and 
it  is  not  the  case  that  one  life  history  is  expected  to  be  more 
or  less  vulnerable  to  fishing.  Although  the  sex  change  and 
fishing  pattern  are  important,  they  must  be  seen  in  the 
context  of  the  mating  system,  reproductive  behavior,  and 
population  dynamics  of  the  species.  If  no  male  size  classes 
escape  fishing,  then  the  sex-changing  population  will  be 
much  more  sensitive  to  fishing  and  may  crash  even  at  low 
fishing  mortality.  When  some  male  size  classes  escape  fish- 
ing, an  identical  dioecious  stock  is  predicted  to  experience 
a  greater  decrease  in  mean  population  size  than  the  pro- 
togynous population.  However,  the  protogynous  species  is 
predicted  to  be  much  more  sensitive  to  mating  aggre- 
gation size  and  sperm  limitation.  Protogynous  stocks 
are  predicted  to  benefit  from  marine  protected  areas 
at  high  levels  of  fishing  mortality  where  sperm  limi- 
tation is  common  at  fished  mating  sites.  In  contrast, 
the  dioecious  stock  is  predicted  to  derive  a  greater 
benefit  of  marine  reserves  even  at  low  fishing  mortal- 
ity because  of  the  protection  of  large  fecund  females 
( Fig.  7 ).  Although  the  sex-changing  population  is  pre- 
dicted to  be  less  sensitive  to  fishing  mortality  overall, 
it  is  clearly  very  important  to  understand  the  exact 
details  of  the  sex-change  pattern  and  the  size-selec- 
tivity of  fishing  in  relation  to  sex  change.  It  will  also 
be  important  to  understand  the  mating  system  and 
patterns  of  fertilization  success  and  sperm  produc- 
tion in  males  when  managing  a  protogynous  stock. 
Given  the  sensitivity  of  the  sex-changing  stock  to  the 
size-selective  pattern  of  fishing,  we  recommend  the 
precautionary  approach  of  keeping  fishing  mortality 
sufficiently  low  so  that  some  males  of  all  size  classes 
always  escape  fishing  (Fig.  4C).  Clearly,  protogynous 
stocks  cannot  be  managed  as  if  they  were  dioecious. 

Sperm  limitation  and  mating  aggregation  size  affect 
stock  dynamics 

The  removal  of  large  males  from  the  population  can 
cause  sperm  limitation,  decreased  fertilization  rates, 
and  decreased  population  size  even  in  a  resilient  spe- 
cies with  high  sperm  production.  Sperm  limitation 
will  increase  as  mating  group  size  decreases.  In  the 
present  model,  even  small  males  produced  relatively 
large  amounts  of  sperm.  If  males  are  removed,  popu- 
lations with  lower  sperm  production  are  predicted  to 
be  more  sensitive  to  the  removal  of  large  fertile  males. 
Our  assumption  of  fertilization  rates  determined  by 
total  egg  and  sperm  production  per  mating  site  will, 
if  anything,  have  underestimated  the  potential  for 
sperm  limitation.  Other  mating  systems  and  repro- 
ductive behaviors  could  lead  to  greater  sperm  limita- 
tion than  predicted  in  our  study  For  example,  species 
that  have  not  evolved  under  sperm  competition  should 
be  more  affected  by  the  removal  of  large  males  than 


Alonzo  and  Mangel:  The  effects  of  size-selective  fisheries  on  the  stock  dynamics  of  and  sperm  limitation  in  sex-changing  fish 


species  with  sperm  competition  because  of  decreased 
allocation  to  sperm  production.  Pair  spawning  among 
individuals  could  also  lead  to  decreased  fertilization 
rates.  Reproductive  behaviors  often  found  in  sex- 
changing  species,  such  as  territoriality,  female  choice, 
resource-defense  polygyny,  and  mate  monopolization, 
all  lead  to  skewed  reproductive  success  for  males  and 
could  further  decrease  fertilization  rates.  Sperm  limita- 
tion is  predicted  to  occur,  and  an  understanding  of  such 
factors  as  fertilization  rate,  sperm  production,  mating 
skew,  and  mating  group  size  will  increase  our  ability  to 
understand  and  predict  stock  dynamics. 

Traditional  spawning-per-recruit  measures  can  fail 
in  the  presence  of  sperm  limitation 

Although  problems  exist  with  traditional  spawning- 
per-recruit  measures  in  general  (Parkes,  2000),  they 
are  especially  problematic  for  sex-changing  stocks.  In 
the  dioecious  stock,  the  relationship  between  female 
and  total  spawning  stock  biomass  per  recruit  exhibits  a 
roughly  linear  relationship  with  population  size.  In  the 
sex-changing  stock,  female  fecundity  does  not  reflect 
the  changes  in  mean  population  size.  Although  total  or 
male  spawning  stock  biomass  per  recruit  did  decrease 
with  decreased  population  size,  the  fit  between  these 
measures  will  depend  greatly  on  the  size-dependent 
sperm  production  of  males,  mating  aggregation  size, 
and  other  factors  determining  the  potential  for  sperm 
limitation.  Male  or  total  spawning  stock  biomass 
per  recruit  alone  cannot  predict  sperm  limitation 
and  thus  will  fail  to  predict  the  potential  population 
crashes  that  may  result.  We  conclude  that  any  mea- 
sure of  spawning  per  recruit  in  a  sex-changing  species 
that  does  not  consider  sperm  limitation  and  reduced 
fertilization  rates  has  the  potential  to  underestimate 
the  impact  of  fishing  on  the  population.  The  number 
of  eggs  produced  or  female  spawning  stock  biomass 
can  remain  relatively  unchanged  in  the  face  of  high 
fishing  mortality  even  as  the  population  is  predicted 
to  decline.  However,  the  failure  of  classic  spawning- 
per-recruit  measures  in  the  presence  of  declines  due 
to  sperm  limitation  or  decreased  fertilization  rate  will 
not  be  limited  to  protogynous  stocks.  Although  sperm 
production  patterns  and  fertilization  rates  are  not  known 
for  many  commercially  important  species,  this  information 
can  be  collected  to  develop  a  general  sense  of  how  sperm 
production  depends  on  individual  size.  We  also  have  a 
general  sense  of  the  factors  that  are  expected  to  affect  fer- 
tilization rates  (Birkhead  and  Moller,  1998)  and  these  can 
be  easily  studied  in  any  species  where  spawning  grounds 
are  accessible  to  researchers.  It  is  clear  that  new  manage- 
ment measures  must  be  developed  for  sex-changing  species 
that  consider  the  potential  for  sperm  limitation  because 
biomass  alone  may  miss  the  potential  for  rapid  population 
crashes.  One  purpose  of  theory  is  to  tell  us  what  we  need 
to  know  more  about  and  to  stimulate  further  research.  Our 
results  clearly  indicate  that  we  need  to  know  more  about 
sperm  production  and  fertilization  rates  when  managing 
protogynous  stocks. 


"P   V 


<=  s 

-  oi 

-^  <D 

CD  CD 


0.75 


0.5 


&      0.25 


i-  c\j 


Protogynous 


Dioecious 


B 


600,000 
500.000 


400.000 
300.000 


200,000 


100,000 


Protogynous 


Dioecious 


Figure  8 

The  effect  of  marine  reserves  on  protogynous  and  dioecious 
populations  when  fishing  effort  is  redistributed  (case  2).  iAi 
Percent  change  in  the  presence  of  fishing  iF=ll  in  the  produc- 
tion of  fertilized  eggs  compared  to  percent  change  in  the  absence 
of  fishing.  (Bl  Annual  biomass  removed  by  the  fisheries  varies 
with  marine  reserve  and  sex-change  pattern.  Numbers  shown 
are  for  10  mating  sites  when  F=l. 


Marine  reserves  and  size-selective  fishing  can  be  used 
to  manage  protogynous  stocks 

Marine  reserves  clearly  have  the  potential  to  decrease  the 
impact  of  fishing  on  populations.  Large  highly  fecund  or 
fertile  individuals  may  be  protected  from  size-selective 
fisheries.  However,  the  benefits  of  a  marine  reserve  will 
be  significantly  decreased  if  fishing  effort  is  simply  redis- 
tributed to  unprotected  sites  (Figs.  7  and  8;  Guenette  and 
Pitcher,  1999;  Apostolaki  et  al„  2002).  It  is  usually  rec- 
ognized that  the  larval  export  and  import  dynamics  will 
be  crucial  to  whether  reserves  increase  mean  population 
size.  We  predict  that  the  degree  to  which  stocks  respond 
to  no-take  reserves  will  also  depend  on  their  life-history 
pattern,  mating  system,  and  size-dependent  fecundity 
and  fertility.  The  protection  of  large  and  fecund  (or  fertile) 


12 


Fishery  Bulletin  102(1) 


fish  will  certainly  increase  reproduction  and  decrease  the 
impact  of  fishing  on  the  population.  However,  the  benefit  of 
marine  reserves  will  be  much  greater  in  populations  where 
larger  or  older  individuals  play  a  key  role  in  reproduction. 
Given  the  predicted  extreme  sensitivity  of  the  protogynous 
population  to  the  pattern  of  size-selective  fishing,  marine 
protected  areas  could  represent  a  precautionary  manage- 
ment strategy  to  ensure  that  some  males  are  not  subject 
to  fishing  mortality. 

A  comprehensive  approach  to  stock  dynamics 

Managing  fishing  on  stocks  of  sex-changing  fish  will  require 
considering  the  sex-change  pattern.  However,  one  must  also 
consider  the  sex  change  pattern  within  the  context  of  the 
mating  system.  Although  the  pattern  of  sex  determination 
does  affect  the  stock  dynamics,  simple  statements  regard- 
ing whether  dioecious  or  sex-changing  populations  are 
more  sensitive  to  fishing  are  not  possible.  The  differences 
among  dioecious  and  sex-changing  stocks  are  complex,  and 
the  management  of  these  stocks  will  depend  as  much  on 
their  mating  system,  the  type  of  fishing  strategies  used 
to  capture  them,  and  mating  aggregation  size  as  on  the 
sex  determination  pattern.  Classic  SPR  measures  cannot 
measure  sperm  limitation  and  reduced  fertilization  rates, 
and  thus  will  not  always  measure  or  predict  the  impact  of 
fishing  mortality  on  the  population.  Rather  than  relying 
on  measures  of  spawning  stock  biomass  per  recruit  alone, 
management  groups  should  also  monitor  protogynous  sex- 
changing  stocks  for  a  reduction  in  fertilization  rates 


Acknowledgments 

We  thank  Phil  Levin,  Alec  McCall,  Steve  Ralston,  and  Bob 
Warner  for  their  comments  on  an  earlier  version  of  this 
manuscript.  This  research  was  supported  by  National  Sci- 
ence Foundation  grant  IBN-01 10506  to  Suzanne  Alonzo 
and  the  Center  for  Stock  Assessment  Research  (CSTAR). 


Literature  cited 

Adams,  S.,  B.  D.  Mapstone,  G.  R.  Russ,  and  C.  R.  Davies. 

2000.  Geographic  variation  in  the  sex  ratio,  sex  specific  size, 
and  age  structure  of  Plectropomus  leopardus  (Serranidae) 
between  reefs  open  and  closed  to  fishing  on  the  Great  Bar- 
rier Reef.    Can.  J.  Fish.  Aquat.  Sci.  57: 1448- 1458. 

Apostolaki.  P.,  E.  J.  Milner-Gulland,  M.  K.  McAllister,  and 
G.  P.  Kirkwood. 

2002.     Modelling  the  effects  of  establishing  a  marine  reserve 
for  mobile  fish  species.     Can.  J.  Fish.  Aquat.  Sci.  59: 
405-415. 
Armsworth,  P.  R. 

2001.  Effects  of  fishing  on  a  protogynous  hermaphrodite. 
(  an  J.  Fish.  Aquat.  Sci.  58:568-578. 

Bannerot,  S.,  W.  F.  Fox,  and  J.  E.  Powers. 

1987.  Reproductive  strategies  and  the  management  of  snap- 
pers and  groupers  in  the  gulf  of  Mexico  and  Caribbean  In 
Tropical  snappers  and  groupers:  biology  and  fisheries  man- 
agement (J.  J.  Polovina  and  S.  Ralston,  eds.  I.  p.  561-603. 
Westview  Press,  Boulder,  CO. 


Barrowman,  N.  J.,  and  R.  A.  Myers. 

2000.     Still  more  spawner-recruitment  curves:  The  hockey 
stick  and  its  generalizations.     Can.  J.  Fish.  Aquat.  Sci.  57: 
665-676. 
Beets,  J.,  and  A.  Friedlander. 

1999.     Evaluation  of  a  conservation  strategy:  a  spawning 
aggregation  closure  for  red  hind.  Epinephelus  guttatus,  in 
the  U.S.  Virgin  Islands.     Environ.  Biol.  Fishes  55:91-98. 
Beverton,  R.  J.  H. 

1987.  Longevity  in  fish:  some  ecological  and  evolutionary 
considerations.  In  Evolution  of  longevity  in  animals.  (A.  D. 
Woodhead  and  K.  H.Thompson,  eds.  I  p.  161-185.  Plenum, 
New  York,  NY. 
1992.  Patterns  of  reproductive  strategy  parameters  in  some 
marine  teleost  fishes.  J.  Fish  Biol.  41  (suppl.  B):137-160. 
Birkhead,  T  R..  and  A.  P.  Meller. 

1998.  Sperm  competition  and  sexual  selection,  826  p.  Aca- 
demic Press,  San  Diego,  CA. 

Brule,  T,  C.  Deniel,  T.  Colas-Marrufo,  and  M.  Sanchez-Crespo. 

1999.  Red  grouper  reproduction  in  the  southern  Gulf  of 
Mexico.    Trans.  Am.  Fish.  Soc.  128:385-402. 

Charnov,  E.  L. 

1982.     The  theory  of  sex  allocation,  355  p.     Princeton  Univ. 
Press,  Princeton,  NJ. 
Coleman,  F.  C,  C.  C.  Koenig,  and  L.  A.  Collins. 

1996.  Reproductive  styles  of  shallow-water  groupers  ( Pisces: 
Serranidae)  in  the  eastern  Gulf  of  Mexico  and  the  conse- 
quences of  fishing  spawning  aggregations.  Environ.  Biol. 
Fishes  47:129-141. 

Cowen,  R.  K. 

1990.     Sex  change  and  life  history  patterns  of  the  labrid, 
Semicossyphus pulcher,  across  an  environmental  gradient. 
Copeia  1990:787-795. 
Cowen,  R.  K.,  K.  M.  M.  Lwiza,  S.  Sponaugle.  C.  B.  Paris,  and 
D  B.  Olson. 

2000.  Connectivity  of  marine  populations:  Open  or  closed? 
Science  287:857-859. 

Fisher,  R.  A. 

1930.    The  genetical  theory  of  natural  selection,  272  p.     Clar- 
endon Press,  Oxford. 
Fu.  C,  T  J.  Quinn,  II,  and  T.  C.  Shirley. 

2001.  The  role  of  sex  change,  growth  and  mortality  in  Panda- 
lus  population  dynamics  and  management.  ICES  J.  Mar. 
Sci.  58:607-621. 

Guenette,  S.,  and  T.  J.  Pitcher. 

1999.     An  age-structured  model  showing  the  benefits  of 

marine  reserves  in  controlling  overexploitation.     Fish. 

Res.  39:295-303. 
Gunderson,  D.  R. 

1997.  Trade-off  between  reproductive  effort  and  adult 
survival  in  oviparous  and  viviparous  fishes.  Can.  J.  Fish. 
Aquat.  Sci.  54:990-998. 

Haddon,  M. 

2001.     Modelling  and  quantitative  methods  in  fisheries, 
406  p.     Chapman  and  Hall.  Boca  Raton.  FL. 
Huntsman,  G.  R.,  and  W.  E.  Schaaf. 

1994.    Simulation  of  the  impact  of  fishing  on  reproduction 
of  a  protogynous  grouper,  the  graysby.     No.  Am.  J.  Fish. 
Manag.  14:41-52. 
Janctos,  A.  C. 

1980.     Strategies  of  female  mate  choice:  a  theoretical 
analysis.    Behav.  Ecol.  Sociobiol.  7:107-112. 
Jennings.  S.,  M.  J.  Kaiser,  and  J.  D.  Reynolds. 

2001.  Marine  fisheries  ecology,  417  p.  Blackwell  Science. 
Oxford. 


Alonzo  and  Mangel:  The  effects  of  size-selective  fisheries  on  the  stock  dynamics  of  and  sperm  limitation  in  sex-changing  fish       13 


Levin,  P.  R,  and  C.  B.  Grimes. 

2002.  Reef  fish  ecology  and  grouper  conservation  and 
management.  In  Coral  reef  fishes:  dynamics  and  diversity 
in  a  complex  ecosystem  (P.  F.  Sale,  ed.)  p.  377-390.  Aca- 
demic Press.  Amsterdam. 

Mangel,  M.,  and  C.  Tier. 

1993.  A  simple  direct  method  for  finding  persistence 
times  of  populations  and  application  to  conservation 
problems.     Proc.  Nat.  Acad.  Sci.  USA  90:1083-1086. 

1994.  Four  facts  every  conservation  biologist  should  know 
about  persistence.     Ecology  75:607-614. 

Parkes,  G. 

2000.  Understanding  SPR  and  its  use  in  U.S.  fishery  man- 
agement, 62  p.  Center  for  marine  conservation,  Washing- 
ton D.C. 

Petersen,  C.  W.,  R.  R.  Warner,  D.  Y.  Shapiro,  and  A.  Marconato. 

2001.  Components  of  fertilization  success  in  the  blue- 
head  wrasse,  Thalassemia  bifasciatum.  Behav.  Ecol.  12: 
237-245. 

Punt,  A.  E.,  P.  A.  Garratt,  and  A.  Govender. 

1993.     On  an  approach  for  applying  per-recruit  measures  to 
a  protogynous  hermaphrodite,  with  an  illustration  for  the 
slinger  Chrysoblephus puniceus  (Pisces:  Sparidae).     S.  Afr. 
J.  Sci.  13:109-119. 
Quinn,  T.  J.,  and  R.  B.  Deriso. 

1999.     Quantitative  fish  dynamics,  542  p.     Oxford  Univ. 
Press.  New  York,  NY, 
Sadovy,  Y 

2001.  The  threat  of  fishing  to  highly  fecund  fishes.  J.  Fish 
Biol.  59:90-108. 


Sale,  P.  F. 

2002.     The  science  we  need  to  develop  more  effective 
management.    In  Coral  reef  fishes:  dynamics  and  diversity 
in  a  complex  ecosystem  (P.  F.  Sale,  ed.)  p.  361-376.    Aca- 
demic Press.  Amsterdam. 
Schultz,  E.  T..  and  R.  R.  Warner. 

1991.     Phenotypic  plasticity  in  life-history  traits  of  female 
Thalassemia  bifasciatum  (Pisces:  Labridae):  2.  Correlation 
of  fecundity  and  growth  rate  in  comparative  studies.     En- 
viron. Biol.  Fishes  30:333-344. 
Shapiro.  D.  Y 

1987.     Reproduction  in  groupers.    In  Tropical  snappers  and 
groupers:  biology  and  fisheries  management  (J.  J.  Polvina 
and  S.  Ralston,  eds. )  p.  295-328.    Westview  Press,  Boulder, 
CO. 
Trivers,  R.  L. 

1972.     Parental   investment   and   sexual   selection.     In 
Sexual  selection  and  the  descent  of  man  (B.  Campbell,  ed.), 
p.  136-179.    Aldine-Atherton,  Chicago. 
Warner,  R.  R. 

1975.  The  reproductive  biology  of  the  protogynous  her- 
maphrodite Pimelometopon  pulchrum  (Pisces:  Labridae  I. 
Fish.  Bull.  73:262-283. 
Warner,  R.  R.,  D.  Y  Shapiro,  A.  Marcanato,  and  C.  W.  Petersen. 
1995.  Sexual  conflict:  Males  with  highest  mating  success 
convey  the  lowest  fertilization  benefits  to  females.  Proc. 
R.  Soc.  Lond.  B.  Biol.  Sci.  262:135-139. 


14 


Abstract— We  present  a  growth  analy- 
sis model  that  combines  large  amounts 
of  environmental  data  with  limited 
amounts  of  biological  data  and  apply  it 
to  Corbicula  japonica.  The  model  uses 
the  maximum-likelihood  method  with 
the  Akaike  information  criterion,  which 
provides  an  objective  criterion  for  model 
selection.  An  adequate  distribution  for 
describing  a  single  cohort  is  selected 
from  available  probability  density  func- 
tions, which  are  expressed  by  location 
and  scale  parameters.  Daily  relative 
increase  rates  of  the  location  parameter 
are  expressed  by  a  multivariate  logistic 
function  with  environmental  factors 
for  each  day  and  categorical  variables 
indicating  animal  ages  as  independent 
variables.  Daily  relative  increase  rates 
of  the  scale  parameter  are  expressed  by 
an  equation  describing  the  relationship 
with  the  daily  relative  increase  rate  of 
the  location  parameter.  Corbicula 
japonica  grows  to  a  modal  shell  length 
of  0.7  mm  during  the  first  year  in  Lake 
Abashiri.  Compared  with  the  attain- 
able maximum  size  of  about  30  mm, 
the  growth  of  juveniles  is  extremely 
slow  because  their  growth  is  less  sus- 
ceptible to  environmental  factors  until 
the  second  winter.  The  extremely  slow 
growth  in  Lake  Abashiri  could  be  a 
geographical  genetic  variation  within 
C.  japon  ica . 


An  environmentally  based  growth  model 

that  uses  finite  difference  calculus 

with  maximum  likelihood  method: 

its  application  to  the  brackish  water  bivalve 

Corbicula  japonica  in  Lake  Abashiri,  Japan 


Katsuhisa  Baba 

Hokkaido  Hakodate  Fisheries  Experiment  Station 

1-2-66,  Yunokawa,  Hakodate 

Hokkaido  042-0932,  Japan 

E-mail  address  babak@fjshexp  pref.hokkaido.jp 


Toshifumi  Kawajiri 

Nishiabashin  Fisheries  Cooperative  Association 
1-7-1,  Oomagan,  Abashiri 
Hokkaido  093-0045,  Japan 


Yasuhiro  Kuwahara 

Hokkaido  Abashiri  Fisheries  Experiment  Station 
31,  Masuura,  Abashiri 
Hokkaido  099-3119,  Japan. 

Shigeru  Nakao 

Graduate  School  of  Fisheries  Sciences 
Hokaido  University 
3-1-1,  Minato,  Hakodate 
Hokkaido  041-8611,  Japan 


Manuscript  approved  for  publication 
14  August  2003  by  Scientific  Editor 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:14-24  (2004). 


Extreme  fluctuations,  both  short-term 
and  seasonal,  in  food  availability  (e.g. 
phytoplankton  density )  make  it  difficult 
to  determine  relationships  between 
the  growth  of  filter-feeding  bivalves 
and  environmental  factors  (Bayne, 
19931.  However,  it  is  becoming  easier 
to  acquire  large  amounts  of  environ- 
mental data  through  the  use  of  data 
loggers,  submersible  fluorometers,  or 
remote-sensing  satellites,  which  enable 
environmental  monitoring  at  daily  or 
subdaily  intervals.  The  development 
of  these  devices  could  solve  difficulties 
in  data  collection.  However,  analytical 
methods  that  combine  large  amounts 
of  environmental  data  with  limited 
amounts  of  biological  data  (e.g.  shell 
length)  are  not  yet  well  developed. 
We  present  an  environmentally  based 
growth  model  that  combines  such 
unbalanced  data  sets.  This  model  is 
useful  in  elucidating  relationships 


between  environmental  factors  and 
growth  of  filter  feeders  from  field  data. 
Complex  box  models,  such  as  eco- 
physiological  models,  can  derive  the 
relationships  between  environmental 
factors  and  the  growth  of  filter-feeding 
bivalves  (Campbell  and  Newell,  1998; 
Grant  and  Bacher,  1998;  Scholten  and 
Smaal,  1998).  These  models  are  useful 
for  estimating  impacts  of  cultivated 
species  on  an  ecosystem  or  the  carrying 
capacity  of  a  species  (or  both)  (Dame, 
1993;  Heral,  1993;  Grant  et  al„  1993). 
They  are  suitable  for  animals  that  have 
been  widely  studied,  such  as  Mytilus 
edulis,  because  they  are  derived  by 
integrating  a  huge  amount  of  ecophysi- 
ological  knowledge  acquired  mainly 
from  laboratory  experiments.  However, 
extrapolation  of  such  knowledge  to 
natural  conditions  is  still  controver- 
sial (Jorgensen,  1996;  Bayne,  1998). 
Our  model  treats  complicated  eco- 


Baba  et  al.:  An  environmentally  based  growth  model  for  |uvenile  Corbicula  japonica 


15 


physiological  processes  as  a  black  box;  we  constructed  the 
model  directly  from  fluctuations  in  environmental  factors 
and  growth  rates.  Our  approach  is  reasonable  for  animals 
for  which  ecophysiological  knowledge  is  limited,  especially 
when  the  main  purpose  of  investigation  is  to  derive  the 
relationships  between  environment  and  growth. 

We  applied  the  model  to  a  single  cohort  of  Corbicula 
japonica  juveniles  spawned  in  August  1997.  We  did  not 
consider  any  bias  caused  by  adjacent  cohorts  because  C. 
japonica  failed  to  spawn  in  1995,  1996,  and  1998  in  Lake 
Abashiri  owing  to  low  water  temperatures  during  the 
spawning  season  (Baba  et  al.,  1999).  Such  investigations 
provide  important  basic  information,  such  as  the  shape  of 
the  distribution  of  a  single  cohort,  and  the  relationship 
between  growth  rate  and  expansion  rate  of  size  variation 
in  a  single  cohort. 

Corbicula  spp.  are  harvested  commercially  in  Japan.  The 
annual  catch  ranged  from  19,000  to  27,000  metric  tons  in 
1996  to  2000  ( Ministry  of  Agriculture,  Forestry  and  Fisher- 
ies1), of  which  C.  japonica  was  the  main  species.  Corbicula 
japonica  is  distributed  in  brackish  lakes  and  tidal  flats  of 
rivers  from  the  south  of  Japan  to  the  south  of  Sakhalin 
(Kafanov,  1991),  is  a  dominant  macrozoobenthos  in  these 
lakes,  and  has  important  roles  in  bioturbation  and  energy 
flow  (Nakamura  et  al.,  1988;  Yamamuro  and  Koike,  1993). 
Juvenile  C.  japonica  growth  is  fast  in  southern  habitats. 
Their  spats  collected  in  Lake  Shinji,  which  lies  in  the  south- 
ern part  of  its  range,  grow  to  a  mean  shell  length  of  around 
6.7  mm  in  natural  conditions  by  the  first  winter  ( Yamane  et 
al.2).  In  northern  habitats,  growth  is  also  believed  to  be  fast; 
Utoh  ( 1981 )  reported  that  mean  shell  length  at  the  first  an- 
nual mark  was  around  5.7  mm  in  Lake  Abashiri.  In  Utoh's 
study  differences  between  the  shell  lengths  at  the  first  an- 
nual marks  and  the  shell  lengths  of  individuals  aged  to  be 
one  year  were  also  reported.  The  purposes  of  the  present 
study  are  to  elucidate  juvenile  growth  and  its  relationship 
to  environmental  factors  in  Lake  Abashiri. 


Materials  and  methods 

Overview  of  the  model 

Our  model  expresses  relative  growth  rate  for  C.  japonica  by 
a  sigmoid  function  with  environmental  factors  and  animal 
ages  as  independent  variables.  Modeling  processes  in  gen- 
eral follow  five  steps:  1)  Shell  lengths  of  a  single  cohort  are 
summarized  by  an  adequate  probability  density  function, 
which  is  expressed  by  a  location  parameter  and  a  scale 
parameter;  2 )  Daily  relative  increase  rate  of  the  location 


1  Ministry  of  Agriculture,  Forestry  and  Fisheries.  1996- 
2002.  Statistics  on  fisheries  and  water  culture  production. 
Association  of  Agriculture  and  Forestry.  1-2-1  Kasumigaseki, 
Chiyoda,  Tokyo  100-0013.  Japan. 

2  Yamane,  K.,  M.  Nakamura,  T.  Kiyokawa,  H.  Fukui,  and  E. 
Shigemoto.  1999.  Experiment  on  the  artificial  spat  collec- 
tion. Bull.  Shimane  Pref.  Fish.  Exp.  Stn.,  p.  232-234.  Unpubl. 
rep.  Shimane  Prefectural  Fisheries  Experimental  Station, 
25-1  Setogashima,  Hamada,  Shimane  697-0051,  Japan.  |In 
Japanese.] 


parameter  (dRIRL)  is  approximated  by  a  sigmoid  function 
with  environmental  factors  and  animal  ages  as  indepen- 
dent variables;  3)  Daily  relative  increase  rate  of  the  scale 
parameter  is  approximated  by  a  simple  function  with  the 
dRIRL  as  an  independent  variable;  4)  The  model  is  opti- 
mized by  a  maximum  likelihood  method;  and  5)  The  best 
model  is  selected  by  Akaike  information  criterion  (AIC). 
The  AIC  is  an  information-theoretic  criterion  extended 
from  Fisher's  likelihood  theory  and  is  useful  for  simulta- 
neous comparison  of  models  (Akaike,  1973;  Burnham  and 
Anderson,  1998). 

Study  site  and  sampling  method 

To  collect  juveniles  of  C  japonica  spawned  in  August  1997, 
sediments  were  sampled  with  a  0.05-m2  Smith-Mclntyre 
grab  once  or  twice  a  month  from  September  1997  to  July 
1999  at  a  depth  of  3.5-4.0  m  in  Lake  Abashiri  (Fig.  1).  The 
habitat  of  C.  japonica  is  restricted  to  areas  shallower  than 
6-m  depth  because  the  deeper  area,  the  lower  stratum  of 
the  lake,  is  covered  by  anoxic  polyhaline  water.  We  assumed 
that  the  selectivity  of  the  sampling  gear  on  C.  japonica 
juveniles  was  negligible  because  the  gear  grabs  the  juve- 
niles with  the  sediment.  Because  the  magnitude  of  spawn- 
ing in  1997  was  relatively  small  (Baba  et  al.,  1999),  we 
selected  a  sampling  site  where  we  found  abundant  settled 
juveniles  in  our  preliminary  investigations.  Samples  could 
not  be  obtained  during  winter  because  of  ice  cover.  Sedi- 
ments were  washed  with  tap  water  on  2-  mm  and  0.125- 
mm  mesh  sieves  from  September  1997  to  October  1998, 
and  on  4.75-mm  and  0.125-mm  mesh  sieves  from  April 
to  July  1999.  To  separate  the  juveniles  from  the  retained 
sediments,  we  treated  the  sediments  with  zinc  chloride 
solution  as  described  by  Sellmer  (1956).  Then  we  sorted 
the  juveniles  under  a  binocular  microscope.  Identification 
of  the  cohort  spawned  in  1997  was  quite  easy  because  C. 
japonica  failed  to  spawn  in  1995,  1996,  and  1998  owing  to 
low  water  temperatures  during  the  spawning  season  ( Baba 
et  al.,  1999).  We  considered  all  the  individuals  that  passed 
through  the  larger-mesh  sieves  and  that  were  retained  on 
the  smaller-mesh  sieve  as  the  1997  cohort.  Shell  lengths 
were  measured  under  a  profile  projector  (V-12,  Nikon  Ltd., 
Chiyoda,  Tokyo)  at  50x  magnification  with  a  digital  caliper 
(Digimatic  caliper,  Mitsutoyo  Ltd.,  Kawasaki,  Kanagawa), 
which  has  a  0.02-mm  precision. 

Environmental  factors 

Values  for  water  temperature  (°C),  water  fluorescence 
(fluorescence  equivalent  to  uranin  density,  ug/L).  salinity 
(psu,  practical  salinity  unit),  and  turbidity  (equivalent  to 
kaolin  density,  ppm)  were  obtained  for  0.1-m  intervals 
from  unpublished  data  at  the  Abashiri  Local  Office  of 
the  Hokkaido  Development  Bureau.3  The  variables  were 
measured  by  a  submersible  fluorometer  (Memory  Chloro- 
tec,  ACL-1180-OK,  Alec  Electronics  Ltd.,  Kobe,  Hyogo)  at 
four  sites  in  Lake  Abashiri  at  intervals  of  about  one  week 
( Fig.  1 ).  The  average  values  of  each  variable  between  the 
depths  of  1  m  and  6  m  were  used  for  later  analyses.  Values 
between  the  measured  dates  were  interpolated  linearly 


16 


Fishery  Bulletin  102(1) 


E14CT  E142 


N44 


N42° 


Lake  Abashiri 

Figure  1 

Location  of  sampling  site  for  Corbicula  japonica  juveniles  in  Lake  Abashiri.  Japan  (#i.  Envi- 
ronmental factors — water  temperature,  water  fluorescence,  salinity,  and  turbidity — were 
measured  at  four  sites,  designated  by  ©■ 


for  subsequent  analysis  with  the  environmentally  based 
growth  model.  The  water  fluorescence  reflects  the  density 
of  phytoplankton. 

Model  structure 

Modeling  the  distribution  of  a  single  sample  Normal 
distribution  is  usually  used  to  describe  a  single  cohort  in 
fishes  and  aquatic  invertebrates  (e.g.  Pauly,  1987;  Founier 
and  Sibert,  1990;  Yamakawa  and  Matsumiya,  1997).  How- 
ever, an  adequate  function  to  describe  a  single  cohort  of 
each  animal  should  be  selected  to  avoid  biases  caused 
by  any  inadequacies  of  the  function.  Probability  density 
functions  of  many  distributions  are  applicable  for  that 
purpose,  and  the  appropriate  can  be  selected  among  easily 
calculable  functions  to  ensure  convergence  of  the  model. 
Characteristics  of  many  distributions  are  well  described 
by  Evans  et  al.  (1993).  We  used  two  distributions:  normal 
distribution  and  largest  extreme  value  distribution.  The 
normal  distribution  is  symmetric.  The  largest  extreme 
value  distribution  is  asymmetric  with  a  longer  tail  toward 
the  larger  side.  These  are  expressed  by  a  location  param- 
eter and  a  scale  parameter. 

To  use  all  the  information  inherent  in  data,  parameters 
of  the  distribution  functions  are  estimated  from  raw  data 
(e.g.  lengths  I,  not  from  summarized  data  such  as  length  fre- 
quency. This  estimation  method  is  described  by  Sakamoto 
et  al.  ( 1983 1.  The  most  adequate  distribution  is  selected  by 
AIC.  Log-likelihood  functions  of  the  distributions  take  the 
following  forms: 


Normal  distribution 


1 


logeLm)rmal(a,b)  =  £log(.  T=!=exp[-(/,-a)2/2&2]  ,    (1) 


(2) 


3  Abashiri  Local  Office  of  the  Hokkaido  Development  Bureau. 
2-6-1  Shinmachi,  Abashiri,  Hokkaido  093-0046,  Japan. 


Largest  extreme  value  distribution 

log,.  L,argt-Ja,b)  =  £log,.{<  1/  6)exp[-(/,  -a) lb] 

x  expj-  exp[-(  /,  -  a )  /  b]\\ , 

where  n  =  number  of  data; 

/,  =  length  of  (th  individual; 

a  =  location  parameter;  and 

b  =  scale  parameter. 


The  location  parameter  is  a  mean  in  the  normal  distri- 
bution. The  location  parameter  is  a  mode  in  the  largest 
extreme  distributions.  The  scale  parameter  is  a  standard 
deviation  in  the  normal  distribution. 
The  AIC  is  calculated  by 

AIC  =  -2  log  tma.ximum  likelihood)  +  2m.  (3) 

where  m  =  number  of  parameters  to  be  estimated. 

The  model  with  the  minimum  AIC  is  the  best  model.  A 
difference  of  more  than  1  or  2  is  regarded  as  significant  in 
terms  of  AIC  (Sakamoto  et  al.,  1983). 

Modeling  the  change  in  the  location  Values  of  the  location 
and  scale  parameters  usually  increase  with  the  growth  of 
an  animal.  The  relative  increase  rate  in  a  certain  time  step 
is  defined  as 


Baba  et  al.:  An  environmentally  based  growth  model  for  juvenile  Corbicula  /aponica 


17 


r,=(P, 


Pi-i)IPi=v 


(4) 


where  rt  =  relative  increase  rate  of  a  parameter  in  the  /th 
time  step;  and 
Pl  =  parameter  value  after  the  /th  time  step. 

Relationships  between  the  parameter  value  and  the  rela- 
tive increase  rate  of  the  parameter  can  be  expressed  by 


P,  =  P(,( !+/-,) 


P2=Pia+r2)  =  P0(l+r1Kl  +  r2) 
P3=P2(l+r3)  =  P0(l  +  r1)(l  +  7-2)(l  +  r!) 


(5) 


P.=P0fl(1+r')" 


where  P(l  =  parameter  value  at  the  first  sampling; 

P,    =  parameter  value  after  the  /'th  time  step;  and 
ri    =  relative  increase  rate  of  the  parameter  in  the 
/th  time  step. 

We  used  one  day  as  the  time  step  in  this  study  In  our  envi- 
ronmentally based  growth  model,  we  assumed  that  the 
daily  relative  increase  rate  of  location  parameter  (dRIRL) 
depends  on  the  age  of  the  animal  and  on  environmental 
factors  for  each  day.  Sigmoid  functions  that  take  values 
between  0  and  a  certain  maximum  are  empirically  appro- 
priate for  expressing  the  relationships  between  the  dRIRL 
and  independent  variables,  especially  for  measures  such  as 
shell  length  that  do  not  show  negative  growth.  Therefore, 
using  categorical  variables  indicating  animal  ages  and  envi- 
ronmental factors  for  each  day  as  independent  variables,  we 
express  the  dRIRL  by  the  multivariate  logistic  function 


related  because  the  dRIRS  is  larger  when  the  dRIRL  is 
larger.  Therefore,  we  estimated  the  dRIRS  from  an  equa- 
tion expressing  the  relationship  to  the  dRIRL.  We  tested 
two  functions, 


Yl  +  Y2Si       (7l+/2S,   >0) 

0  (yi  +  72s,  <0) 


and 


t,  = 


0 


s,  -  7i  >  0) 
(s,  -/,  <0) 


(7) 


(8) 


where  ti  =  dRIRS  on  the  /th  day  from  the  first  sampling; 
Yv  Yz    =  coefficients  of  the  equations;  and 

Sj  =  dRIRL  on  the  /th  day  from  the  first  sampling. 

Model  estimation 

Likelihood  function  The  location  and  scale  parameters 
at  the  first  sampling  (o0  and  fe0),  the  coefficients  of  Equa- 
tion 6  (smax,  a,  and  pk),  and  the  coefficients  of  Equations  7 
and  8  (y-j  and  y2)  are  estimated  as  values  that  maximize 
total  log-likelihood.  The  total  log-likelihood  is  evaluated 
by  the  adequate  probability  density  function  selected  in 
the  first  step.  The  log-likelihood  functions  take  the  follow- 
ing forms: 

Normal  distribution 


log,  L„ormal  (aQ,b„,  smax ,  a j ,  pk ,  yv  y2) 
=  X2>g*  -Arexp[-(Z<7i-a,)/242] 


2nb 


(9) 


Largest  extreme  value  distribution 


s,  =smax/    1  +  exp 


2>a+£ab* 


(61 


where      si    =   dRIRL  on  the  /th  day  from  the  first  sampling; 
smax   =   potential  maximum  dRIRL  of  the  animal; 
a.,  Pk   =  coefficients  of  each  independent  variable; 
A    =  categorical  variable  ( a  dummy  variable  indi- 
cating animal  ages )  that  takes  the  value  1 
orO; 
Ekl    =  the  kt\\  environmental  factor  on  the  /th  day 

from  the  first  sampling; 
nA   =   number  of  age  categories;  and 
nE   =   number  of  environmental  factors. 

The  categorical  variable  takes  the  value  of  1  when  the 
animal  is  the  category,  otherwise  it  takes  0.  The  multivari- 
ate logistic  function  with  smax  =  1  is  used  for  logistic  regres- 
sions (Sokal  and  Rohlf,  1995).  A  method  of  giving  a  value  to 
the  categorical  variable  is  described  by  Zar  ( 1999). 

Modeling  the  change  in  scale  The  daily  relative  increase 
rate  of  scale  parameter  (dRIRS)  and  dRIRL  must  be  cor- 


loge-L,argcs/o0,fe0,smax,a,,^„71,)'2) 

N       nq 

=XZ1°g«{(1/Vexp[-^-«,>/4] 

xexp{-exp[-(Z9i-<59)/feJU, 


(10) 


where  a0,  60  =  values  of  the  location  and  scale  param- 
eters, respectively,  at  the  first  sampling; 
smax>  aj>  Pk  =  coefficients  of  Equation  6; 

Yvy2  -  coefficients  of  Equations  7  and  8; 
N  =  number  of  samplings; 
nq  =  number  of  data  at  the  qth  sampling; 
aq  =  location  parameter  at  the  qth  sampling 

estimated  by  Equation  5  (r,=s, ); 
bq  =  scale  parameter  at  the  qth  sampling  esti- 
mated by  Equation  5  (r~^);  and 
/     =  length  of  the  /th  individual  at  the  <?th 
sampling. 

AIC  is  used  to  select  significant  environmental  factors, 
the  age  categorization,  and  the  equation  to  express  the 


18 


Fishery  Bulletin  102(1) 


relationship  between  dRIRL  and  dRIRS,  i.e.  Equation  7 
or  8. 

Confidence  intervals  To  evaluate  uncertainties  of  coef- 
ficient values  and  model  selection,  we  estimated  the  95r/f 
confidence  intervals  of  all  coefficients — i.e.  a0,  b0,  smax,  a-, 
jik,  yj,  and  y2 — based  on  profile  likelihood.  For  example,  the 


95%  confidence  interval  of  a,, — a, 


-was  estimated  as  an 


interval  that  suffices  in  the  following  equation: 


2\  max  log,.  L(a0,b0,  smax , a , , ft ,  y1 ,  y, ) 

max  log,.  U  d0 , 4,  smBX,  aJt ft, yu  y2 

K=a096)}<^(0.05), 


(11) 


where  .v^lO.OS)  =  value  of  a  chi-squared  distribution  at  an 
upper  probability  of  0.05  with  one  degree 
of  freedom,  i.e.  3.84. 

The  characteristics  of  the  interval  are  explained  by 
Burnham  and  Anderson  ( 1998). 

We  used  Microsoft  Excel  (Microsoft  Corp.,  Redmond,  WA) 
as  the  analysis  platform,  and  Solver  (Microsoft  Corp.,  Red- 
mond, WA)  as  the  nonlinear  optimization  tool. 

Model  selection 

We  used  three  procedures  for  model  selection  to  achieve 
the  best  model.  First,  we  constructed  an  a  priori  set  of  base 
models  based  on  biological  variables;  then  we  selected  the 
best  base  model.  Fixation  of  the  base  model  drastically 
decreases  possible  candidate  models  to  be  tested.  To  test  all 
possible  combinations  of  independent  variables  and  model 
forms  is  quite  impractical.  Second,  we  excluded  insignificant 
factors  from  the  best  base  model.  Third,  we  checked  the  sig- 
nificance of  environmental  factors  that  were  not  included  in 
the  base  models.  If  one  was  significant,  we  included  it  in  the 
best  base  model.  All  of  these  procedures  were  performed  by 
AIC.  The  construction  of  the  a  priori  set  of  candidate  models 
is  partially  subjective,  but  it  is  an  important  part  of  the 
model  construction  (Burnham  and  Anderson,  1998). 

Seasonal  growth  in  bivalves  is  influenced  by  water 
temperature  and  food  supply  (Bayne  and  Newell,  1983). 
The  growth  rate  of  Corbicula  fluminea  changes  with  age 
(McMahon,  1983).  Therefore,  we  constructed  base  models 
combining  water  temperature,  water  fluorescence,  and 
categorical  variables  indicating  age  for  the  independent 
variables  of  Equation  6.  We  tested  two  types  of  categoriza- 
tion of  age.  The  first  segregates  ages  based  on  real  age,  i.e. 
two  categories:  0+  or  1+.  The  second  segregates  ages  in  rela- 
tion to  winter,  i.e.  three  categories:  before  the  first  winter, 
from  the  first  to  the  second  winter,  and  after  the  second 
winter.  For  the  real-age  categorization,  age  was  segregated 
based  on  1  September,  because  the  spawning  season  was 
in  August  1997.  For  the  winter-base  age  categorization,  we 
segregated  ages  based  on  1  January.  No  biases  should  have 
occurred  because  of  the  segregation  date  of  the  winter  base 
categorization  and  because  the  growth  of  C.japonica  is  neg- 
ligible during  winter.  Four  base  models  were  constructed 


0.2  " 


g 
a 


0.1  -- 


0.0  -i 


Largest  extreme 
value  distribution 


Normal  distribution 


L. 


-+- 


0 


1  2 

Shell  length  (mm) 


Figure  2 

Two  distributions  fitted  by  the  maximum- 
likelihood  method  to  the  shell  lengths  of  Cor- 
bicula japonica  juveniles  spawned  in  1997  and 
sampled  on  22  April  1999.  Raw  data  are  shown 
by  +.  The  shell  length  composition  is  shown  by 
the  histogram. 


combining  the  two  types  of  age  categorization  and  two  types 
of  equations  expressing  the  relationship  between  the  dRIRL 
and  the  dRIRS,  i.e.  Equations  7  or  8.  We  selected  the  best 
base  model  by  AIC. 

To  check  the  significance  of  each  environmental  factor 
and  age  categorization,  we  removed  the  independent  vari- 
ables one  by  one  from  the  best  base  model  and  re-optimized 
the  model.  When  the  model  was  significantly  improved  by 
the  removal  in  terms  of  AIC,  the  effect  of  the  variable  was 
insignificant  on  the  model;  therefore  we  excluded  it. 

To  check  the  significance  of  salinity  and  turbidity,  which 
were  not  included  in  the  base  models,  we  included  them 
one  at  a  time  into  the  best  base  model  and  re-optimized  the 
model.  When  the  model  was  improved  by  the  inclusion,  the 
effect  of  the  variable  was  significant  on  the  model;  therefore 
we  included  it. 


Results 

Modeling  the  distribution  of  a  single  sample 

The  largest  extreme  value  distribution  was  the  best  in 
terms  of  AIC  except  for  data  sampled  on  13  May  1998 
(results  are  not  shown).  The  exception  is  due  probably 
to  the  small  sample  size  (rc=38)  on  that  date.  The  largest 
extreme  value  distribution  was  therefore  used  to  evaluate 
likelihood  in  later  analyses:  we  selected  Equation  10  from 
Equations  9  and  10.  The  result  of  fitting  the  two  distri- 
butions to  the  shell  lengths  sampled  on  22  April  1999  is 
shown  in  Figure  2  as  a  representative  example.  The  largest 
extreme  value  distribution  is  apparently  better  than  the 
normal  distribution  for  describing  the  single  cohort  of  C. 
japonica  spawned  in  1997. 


Baba  et  al.:  An  environmentally  based  growth  model  for  |uvenile  Corbicu/a  /aponica 


19 


Table  1 

Values  of  location  and  scale  parameters  at  the  first  sampling,  coefficients,  log-likelihood,  and  AIC  of  models  constructed  based  on 
the  largest  extreme  value  distribution.  The  best  AIC  among  four  base  models  ( models  1-4 )  is  enclosed  by  a  single  line.  The  best  AIC 
of  all  models  is  enclosed  by  a  double  line.  dRIRL  =  daily  relative  increase  rate  of  location  parameter,  dRIRS  =  daily  relative  increase 
rate  of  scale  parameter.  Temp.  =  water  temperature,  WF  =  water  fluorescence,  Sal.  =  salinity,  Turb.  =  turbidity,  CI  =  before  the  first 
winter,  C2  =  from  the  first  to  the  second  winter,  C3  =  after  the  second  winter. 

Model 
no. 

Parameters 

at  1st 

sampling 

Max. 
dRIRL 

Age  categorization 

Environmen 

tal  factors 

Expressing  relationship 

between  dRIRS 

and  dRIRL 

Log-L 

Al           A2 
a  j             a2 

A3 

«3 

Temp. 
ft 

WF 

ft 

Sal. 
ft 

Turb 
ft 

AIC 

ao 

^0 

smax 

)'i 

y2      Eq.  no 

1 

0.299 

0.040 

0.012 

0+             1  + 
-62.6       -23.7 

0.16 

2.61 

0.0000 

1.686 

7 

850.4 

-1682.9 

2 

3 

4 

0.297 

0.299 
0.299 

0.040 

0.042 
0.042 

0.011 
0.011 

0.011 

-56.1        -22.1 

0.20 

0.61 
0.65 

2.44 

0.41 
0.42 

0.0001 

-0.0076 
0.0034 

0.887 

2.902 
0.760 

8 

7 
8 

852.3 

950.4 
952.2 

-1686.5 

ci            C2 

C3 

-16.8       -16.7 
-17.5       -17.6 

-9.1 
-9.6 

-1880.9 

-1884.4 

4.1 
4.2 

0.299 
0.297 

0.042 
0.038 

0.011 
0.005 

-18.3'            -10.0 
127.9              -26.8' 

0.68 

0.34 

0.44 
4.15 

0.0034 
0.0000 

0.760 
0.895 

8 
8 

952.2 
735.0 

-1886.3 

-1451.9 

4.3 

0.295 

0.037 

0.008 

-47.3       -16.3 

-8.8 

1.47 

0.0033 

0.766 

8 

848.9 

-1679.9 

4.4 

0.299 

0.041 

0.013 

-4.9          -8.9 

-4.9 

0.40 

0.0020 

0.806 

8 

909.6 

-1801.1 

4.5 

0.299 

0.042 

0.011 

-16.7' 

-9.1 

0.62 

0.42 

-0.25 

0.0033 

0.762 

8 

952.4 

-1884.8 

4.6 

0.299 

0.042 

0.011 

-18.5' 

-10.2 

0.68 

0.44 

0.007 

0.0034 

0.760 

8 

952.2 

-1884.4 

1  One  common  coefficient  was  used  for  the  two  categorical 

/ariables. 

Model  selection  and  application 

Model  4  was  the  best  in  terms  of  AIC  among  four  base 
models  (Table  1,  models  1-4);  ages  were  categorized  in 
relation  to  winter;  and  the  relationship  between  dRIRL 
and  dRIRS  was  expressed  by  Equation  8. 

Four  models  were  made  by  removing  each  independent 
variable  from  model  4  (Table  1,  models  4.1  to  4.4).  The  effect 
of  one  age  categorization — segregation  of  ages  between  the 
first  and  second  winters — was  insignificant  on  the  model, 
because  the  model  was  significantly  improved  by  its  re- 
moval in  terms  of  AIC.  The  effects  of  the  other  independent 
variables  were  significant  on  the  model,  because  the  model 
was  significantly  worse  by  their  removal  in  terms  of  AIC. 
The  effects  of  salinity  and  turbidity  were  insignificant  on 
the  model,  because  adding  each  variable  made  the  model 
significantly  worse  in  terms  of  AIC  (Table  1,  models  4.5  and 
4.6).  Consequently,  model  4. 1  was  the  best  model  to  describe 
the  relationships  among  environmental  factors,  ages,  and 
growth  of  C.  japonica  juveniles  spawned  in  1997. 

The  coefficient  value  for  age  categorization  of  before  the 
second  winter  (-18.3)  is  much  smaller  than  that  of  after  the 
second  winter  (-10.0)  (Table  1).  This  difference  suggests 
that  the  growth  response  of  C.  japonica  juveniles  is  much 
less  susceptible  to  environmental  factors  before  the  second 
winter  than  after. 

Peaks  of  the  dRIRL  corresponded  with  peaks  of  water 
fluorescence,  when  the  water  temperature  was  warmer 
than  about  10°C,  especially  before  the  second  winter  (Fig.  3, 


B  and  C).  Therefore,  food  supply  is  the  most  influential  fac- 
tor when  the  water  temperature  is  above  about  10°C.  The 
slow  growth  or  no  growth  during  winter  is  due  to  the  low 
water  temperatures.  The  dRIRL  reached  a  plateau  after  30 
May  1999.  This  was  due  to  two  factors:  water  fluorescence 
was  relatively  intense  after  30  May  1999  (Fig.  3B);  and  the 
growth  response  of  C.  japonica  to  the  environmental  factors 
was  more  susceptible  after  the  second  winter  than  before. 

The  confidence  limits  of  all  the  coefficients  seem  to  be 
reasonably  estimated  by  the  profile  likelihood  method 
(Table  2).  These  results  also  guarantee  the  convergence  of 
the  model  because  the  model  was  frequently  optimized  to 
seek  each  confidence  limit  with  different  starting  values. 
We  repeated  the  optimization  at  least  20  times  to  seek  each 
confidence  limit.  On  other  models,  we  also  confirmed  the 
convergences  as  well. 

The  largest  extreme  value  distributions  estimated  by 
model  4.1  fitted  the  shell  lengths  of  C.  japonica  juveniles 
very  well  (Fig.  4). 


Discussion 

Model  formulation  and  application 

Largest  extreme  value  distribution  is  apparently  better 
than  normal  distribution  to  describe  the  single  cohort  of 
C.  japonica  that  spawned  in  1997.  This  distribution  has  a 
mode  and  a  longer  tail  toward  the  larger  side.  If  the  shell 


20 


Fishery  Bulletin  102(1) 


±    0.02 


£     0.01 


rr 


D 


5 

4 
3  " 
2  " 

1   " 

0 
fc- 


^         > 


fiO  ; 

M 

4^%, 

-^—  Turbidity 
-•—  Salinity 

-■ 

40  - 

1/    \ 

20-                     l"'"l       I 

■ 1 1 

1 

~wj 

Mode  (estimated  by  model  4.1) 

90%  confidence  interval  (estimated  by 

model  4.1) 
°    Mode  (sample) 


Date 

Figure  3 

Environmental  fluctuations  and  prediction  of  the  growth  oiCorbiculajapon- 
ica  juveniles  spawned  in  1997  in  Lake  Abashiri  by  the  best  model  (Model  4.1 
in  Tablel).  (Al  Insignificant  environmental  factors  (factors  excluded  in  the 
model  selection),  turbidity  (equivalent  to  kaolin  density,  ppmi  and  salinity 
(psu,  practical  salinity  unitl.  (Bl  Significant  environmental  factors  (factors 
included  in  the  model  selection  I,  temperature  (°C)  and  water  fluorescence 
(equivalent  to  uranin  density,  /'g/L>.  (Cl  Daily  relative  increase  rate  of  loca- 
tion parameter  (dRIRLl  and  daily  relative  increase  rate  of  scale  parameter 
(dRIRS)  estimated  by  the  model.  (Di  Growth  of  Corbicula  japonica;  verti- 
cal bars  represent  90%  confidence  intervals  for  the  shell  lengths  of  the 
samples. 


length  distribution  becomes  asymmetric  during  growth, 
skcwness  of  the  distribution  would  increase  according 
to  growth.  However,  there  is  no  correlation  between  the 
skewness  and  the  means  of  the  shell  lengths.  Therefore,  we 
thought  that  the  shell  length  distribution  of  the  cohort  was 
already  asymmetric  just  after  settlement.  Such  a  distribu- 


tion might  be  influenced  by  fluctuations  in  larval  settle- 
ment during  the  spawning  season;  and  larval  settlement 
would  be  influenced  by  fluctuations  in  larval  supply  from 
the  water  column.  During  the  spawning  season  of  1997. 
the  average  planktonic  larval  density  gradually  increased 
from  26  ind/m3  on  25  July  to  a  maximum  of  603  ind/m3  on 


Baba  et  al.:  An  environmentally  based  growth  model  for  |uvenile  Corbicula  japonica 


21 


Table  2 

95%  confidence  limits  of  location  and  scale  parameters  at  the  first  sampling  and  coefficients  of  the  best  model  constructed  based 
on  the  largest  extreme  value  distribution  (models  4.1  in  Table  1)  estimated  by  profile  likelihood  method.  dRIRL  =  daily  relative 
increase  rate  of  location  parameter.  dRIRS  =  daily  relative  increase  rate  of  scale  parameter.  Temp.  =  water  temperature,  WF  = 
water  fluorescence,  Sal.  =  salinity,  Turb.  =  turbidity. 


Parameters 

at  1st 

sampling 


Max. 
dRIRL 


Age  categorization 


Environmental  factors 


Expressing 

relationship 

between  dRIRS 

and  dRIRL 


A, 
a, 


a. 


Temp. 
ft 


WF 
ft 


Sal. 

ft 


Turb. 

ft 


Lower  95  % 
Upper  95  % 


0.294 
0.304 


0.039 
0.045 


0.010  -26.6' 

0.013         -11.5' 


-14.6 
-6.4 


0.41 
1.00 


0.27 
0.64 


0.0027 
0.0039 


0.734 
0.793 


1  One  common  coefficient  for  the  two  categorical  variables. 


13  August.  Then  it  sharply  decreased  to  3  ind/m3  on  19 
August  (Baba  et  al.,  1999).  Such  a  pattern  of  larval-density 
fluctuation  might  have  caused  the  asymmetric  distribution 
of  shell  lengths  of  the  settled  juveniles.  Another  possible 
factor  that  influenced  the  shapes  of  the  shell  length  distri- 
butions and  the  relationship  between  dRIRL  and  dRIRS 
is  size-dependent  mortality,  e.g.  predations  and  fisheries. 
Size-dependent  mortality  has  been  reported  in  several 
marine  bivalves  (e.g.  Nakaoka,  1996).  Potential  predators 
of  C.japonica  are  fishes,  such  as  Japanese  dace  (Tribolo- 
don  hakonensis)  (also  known  as  big-scaled  Pacific  redfin, 
FAO),  Pacific  redfin  (Tribolodon  brandtii),  common  carp 
(Cyprinus  carpio),  and  the  So-iny  mullet  (Liza  haemato- 
cheila  )  (Kawasaki4).  In  our  study,  the  size-dependent  mor- 
tality was  negligible  because  the  range  of  the  shell  lengths 
observed  in  this  study  was  very  narrow. 

The  shape  of  the  distribution  to  describe  a  single  cohort 
should  be  determined  from  the  data.  In  contrast,  single 
cohorts  are  usually  separated  from  multicohort  data  by  as- 
suming a  normal  distribution  of  lengths  in  a  single  cohort 
(e.g.  Fournier  and  Sibert,  1990).  Therefore,  it  is  possible 
that  multicohort  analysis  done  without  selection  of  an 
adequate  distribution  to  describe  a  single  cohort  causes 
substantial  bias  in  estimations  of  various  stock  features 
of  animal  populations,  such  as  age  composition,  growth, 
mortality,  and  recruitment.  In  our  preliminary  analyses, 
we  also  tested  smallest  extreme  value  distribution,  inverse 
Gaussian  distribution,  and  lognormal  distribution.  The  in- 
verse Gaussian  distribution  was  the  best  for  two  samples; 
the  lognormal  distribution,  was  the  best  for  two  samples; 
the  largest  extreme  value  distribution  was  the  best  for  ten 
samples.  Therefore,  it  is  reasonable  to  select  the  largest  ex- 
treme value  distribution.  We  selected  a  single  distribution 


for  our  analyses,  otherwise  a  discontinuous  point  would 
have  appeared  in  the  growth  curve. 

Relatively  large  confidence  intervals  were  obtained  in 
the  coefficients  of  the  linear  component  of  Equation  6,  i.e. 
a ,  and  /3;,  (Table  2).  The  relatively  large  confidence  inter- 
vals may  indicate  that  the  number  of  estimated  coefficients 
is  somewhat  larger  than  the  number  of  samplings.  There- 
fore, to  estimate  these  coefficients  more  precisely,  we  may 
need  to  investigate  more  cohorts  spawned  in  other  years 
in  future  investigations. 

Growth  of  C.  japonica 

We  identified  extremely  slow  growth  in  C.  japonica  juve- 
niles, which  grew  to  a  modal  shell  length  of  0.7  mm  during 
the  first  year  in  Lake  Abashiri,  which  lies  at  43.7°N.  Spats 
of  C.  japonica  collected  from  1992  to  1997  in  Lake  Shinji, 
which  lies  at  35.5°N,  grew  to  a  mean  shell  length  of  6.7 
mm  in  natural  conditions  by  the  first  winter  (Yamane  et 
al.2).  Using  environmental  factors  measured  in  Lake  Shinji 
from  1990  to  1998  at  monthly  intervals  (Seike5),  we  simu- 
lated the  growth  of  C.  japonica  with  model  4.1.  Corbicula 
japonica  grew  to  a  mean  shell  length  of  1.4  mm  (standard 
error,  0.37 )  by  the  first  winter  in  the  simulations.  Therefore, 
the  large  difference  in  juvenile  growth  between  the  two 
habitats  cannot  be  explained  by  environmental  differences 
because  the  results  of  the  simulation  were  apparently  an 
underestimate.  We  think  that  the  extremely  slow  growth 
of  the  juveniles  (prolonged  phase  of  meiobenthic  develop- 
ment )  in  Lake  Abashiri  is  probably  a  geographical  varia- 
tion, which  is  genetically  determined,  within  C.  japonica. 
However,  there  remains  a  possibility  that  the  juvenile 
growth  differences  depend  on  other  environmental  factors 
not  measured  in  this  study.  Therefore,  the  geographical 


4  Kawasaki,  K.  1997.  Lagoon  structure  and  fish  produc- 
tion in  Ogawara-ko  Lagoon.  /;;  Final  reports  on  fisheries  in 
Ogawara-ko  Lagoon  (Tohoku  Construction  Corporation  ed.), 
p.  4-33.  Unpubl.  rep.  Construction  Office  for  Takasegawa 
General  Development  of  Tohoku  Regional  Construction  Bureau, 
3  Ishido,  Hachinohe,  Aomori  039-1165,  Japan. 


6  Seike,  Y.  1990-98.  Gobiusu:  monthly  report  of  water  quality 
in  Lake  Shinji  and  Lake  Nakaumi.  Unpubl.  rep.  Faculty  of 
Science  and  Engineering.  Shimane  University,  1060  Nishi- 
kawatsu,  Matsue,  Shimane  690-0S23,  Japan. 


22 


Fishery  Bulletin  102(1) 


10  Sep  1997;  mode:  0.30mm, 
scale:  0.04,  n=341 


13  May  1998;  mode:  0.41mm, 
scale:  0.06,  n=38 


11  Jun  1998;  mode:  0.51  mm, 
scale:  0.12,  n=292 


10  Jul  1998;  mode:  0.57mm. 
scale:  0.10,  n=610 


13  Aug  1998;  mode:  0.64mm, 
scale:  0.12.  n=456 


1 1  Sep  1998;  mode:  0.70mm. 
scale:  0.17,  n=202 


14  Oct  1998;  mode:  0.76mm. 
scale:  0.17.  n=162 


0.0 


22  Apr  1999;  mode:  0.74mm. 
scale:  0.15,  n=265 


+  + 

13  May  1999;  mode:  0.81mm, 
scale:  0.20.  n=241 


0.2  t 


t 


H 


0.1 


00 


28  Jul  1999;  mode:  2.14mm, 
scale:  1.06,  n=63 


^#T>fffi^^ 


3456  0123456 

Shell  length  (mm) 

Figure  4 

Shell-length  compositions  of  a  single  cohort  of  Corbicula  japonica  spawned  in  1997.  The  raw  data 
(shell  lengths)  are  shown  by  +.  The  largest  extreme  value  distribution  estimated  by  the  best  model 
( model  4. 1  in  Table  1 1  is  shown  by  a  solid  line.  The  largest  extreme  value  distribution  independently 
fitted  by  the  maximum  likelihood  method  is  shown  by  a  dashed  line.  The  sampling  date  and  values 
of  location  parameter  I  mode)  and  scale  parameter  independently  fitted  by  the  maximum  likelihood 
method  are  shown  in  each  panel. 


variation  should  be  validated  by  reciprocal  transplanta- 
tions or  laboratory  experiments  (or  both)  in  future  inves- 
tigations. Prolonged  phases  of  meiobenthic  development 
have  been  reported  in  some  marine  bivalves  (Nakaoka, 
1992;  Harvey  and  Gage,  1995).  However,  a  prolonged  phase 
of  meiobenthic  development  as  a  geographical  variation  is 
rarely  reported. 


In  many  species  of  bivalve,  populations  from  higher  lati- 
tudes have  a  slower  initial  growth  rate;  but  longevity  and  ul- 
timate size  in  these  populations  are  frequently  greater  than 
at  lower  latitudes  (Newell,  1964;  Seed,  1980).  The  extremely 
slow  growth  of  C.  japonica  juveniles  in  Lake  Abashiri  may  be 
an  extreme  example  of  this  phenomenon.  In  Lake  Abashiri, 
C.  japonica  failed  to  spawn  in  ten  out  of  21  years  for  which 


Baba  et  al 


An  environmentally  based  growth  model  for  juvenile  Corbicu/a  japonica 


23 


data  were  available  because  of  low  water  temperatures  dur- 
ing the  summer  spawning  season  (Baba  et  al.,  1999).  This 
means  that  a  long  life  span  is  essential  to  sustain  popula- 
tions of  C.  japonica  in  northern  habitats.  We  think  that  a 
long  life  span  is  the  ultimate  factor  for  the  extremely  slow 
growth  rate  of  C.  japonica  juveniles  in  Lake  Abashiri. 

The  growth  response  of  C.  japonica  juveniles  is  much  less 
susceptible  to  environmental  factors  before  the  second  win- 
ter than  after  and  is  the  proximate  factor  for  an  extremely 
slow  growth  rate.  Nuculoma  tenuis,  a  detritus  feeder,  de- 
velops its  palp  proboscides,  its  feeding  apparatus,  during 
the  prolonged  phase  of  meiobenthic  development  (Harvey 
and  Gage,  1995).  The  change  of  growth  susceptibility  to  en- 
vironmental factors  in  young  ages  may  suggest  that  some 
functional  morphological  changes  occur  in  C.  japonica,  also 
a  filter  feeder.  In  our  preliminary  analyses,  we  could  not 
find  a  better  model  when  we  used  different  values  of  smax 
in  Equation  6  between  ages  instead  of  categorical  variables 
indicating  ages.  Therefore,  we  conclude  that  the  difference 
in  growth  rates  between  ages  is  not  due  to  a  difference  in 
potential  maximum  growth  rate,  at  least  in  the  range  of  the 
shell  length  observed  in  our  study.  When  our  model  is  ap- 
plied to  a  wider  range  of  the  shell  lengths  or  other  species, 
it  is  best  to  examine  the  age  dependence  of  smax. 


Acknowledgments 

We  express  our  thanks  to  T  Kato,  Vice-Head  of  the  River 
Improvement  Section  in  the  Abashiri  Local  Office  of  the 
Hokkaido  Development  Bureau,  for  providing  environmen- 
tal data  on  Lake  Abashiri.  We  also  thank  the  reviewers  of 
Fishery  Bulletin  for  providing  helpful  suggestions  on  our 
manuscript. 


Literature  cited 

Akaike,  H. 

1973.     Information  theory  and  an  extension  of  the  maximum 
likelihood  principle.     In  Second  international  symposium 
on  information  theory  (B.  N.  Petrov  and  F.  Csaki,  eds. ), 
p.  267-281.    Akademiai  Kiado,  Budapest. 
Baba.  K..  T.  Kawajiri.  and  Y.  Kuwahara. 

1999.     Effects  of  water  temperature  and  salinity  on  spawning 
of  the  brackish  water  bivalve  Corbicula  japonica  in  Lake 
Abashiri,  Hokkaido,  Japan.     Mar.  Ecol.  Prog.  Ser.  180: 
213-221. 
Bayne,  B.  L. 

1993.  Feeding  physiology  of  bivalves:  time  dependence  and 
compensation  for  changes  in  food  availability.  In  Bivalve 
filter  feeders  and  marine  ecosystem  processes  (R.  F.  Dame 
ed. ).  p.  1-24.     Springer  Verlag.  New  York,  NY. 

1998.     The  physiology  of  suspension  feeding  by  bivalve 
molluscs:  an  introduction  to  the  Plymouth  "TROPHEE" 
workshop.    J.  Exp.  Mar.  Biol.  Ecol.  219:1-19. 
Bayne,  B.  L.,  and  R.  C.  Newell. 

1983.     Physiological  energetics  of  marine  molluscs.     In  The 
Mollusca,  4(1)  (A.  S.  M.  Saleuddin  and  K.  M.  Wilbur  eds.), 
p.  407-515.    Academic  Press,  New  York,  NY. 
Burnham.  K.  P.,  and  D.  R.  Anderson. 

1998.  Model  selection  and  inference,  349  p.  Springer,  New 
York,  NY. 


Campbell.  D.  E.,  and  C.  R.  Newell. 

1998.     MUSMOD,  a  production  model  for  bottom  culture  of 
the  blue  mussel,  Mytilus  edulis  L.     J.  Exp.  Mar.  Biol.  Ecol. 
219:171-203. 
Dame,  R.  F. 

1993.    The  role  of  bivalve  filter  feeder  material  fluxes  in  estua- 
rine  ecosystems.    In  Bivalve  filter  feeders  in  estuarine  and 
coastal  ecosystem  processes:  NATO  ASI  Series  (R.  F.  Dame 
ed. ).  p.  245-269.     Springer- Verlag,  Berlin,  Heidelberg. 
Evans,  M.,  N.  Hastings,  and  B.  Peacock. 

1993.     Statistical  distribution,  170  p.    John  Wiley  &  Sons, 
New  York,  NY. 
Fournier,  D.  A.,  and  J.  R.  Sibert. 

1990.  MULTIFAN  a  likelihood-based  method  for  estimat- 
ing growth  parameters  and  age  composition  from  multiple 
length  frequency  data  sets  illustrated  using  data  for  south- 
ern bluefin  tuna  iThunnus  maccoyii).  Can.  J.  Fish.  Aquat. 
Sci.  47:301-317. 

Grant,  J.,  and  C.  Bacher. 

1998.  Comparative  models  of  mussel  bioenergetics  and  their 
validation  at  field  culture  sites.  J.  Exp.  Mar.  Biol.  Ecol. 
219:21-44. 

Grant,  J.,  M.  Dowd,  K.  Thompson,  C.  Emerson,  and  A.  Hatcher. 
1993.  Perspectives  on  field  studies  and  related  biological 
models  of  bivalve  growth  and  carrying  capacity.  In  Bivalve 
filter  feeders  in  estuarine  and  coastal  ecosystem  processes: 
NATO  ASI  Series  (R.  F.  Dame  ed.  I,  p.  371-420.  Springer- 
Verlag,  Berlin,  Heidelberg. 

Harvey,  R.,  and  J.  D.  Gage. 

1995.  Reproduction  and  recruitment  of  Nuculoma  Tenuis 
(Bivalvia:  Nuculoida*  from  Loch  Etive.  Scotland.  J.  Moll. 
Stud.  61:409-419. 

Heral.  M. 

1993.  Why  carrying  capacity  models  are  useful  tools  for 
management  of  bivalve  molluscs  culture.  In  Bivalve  filter 
feeders  in  estuarine  and  coastal  ecosystem  processes:  NATO 
ASI  Series  (R.  F.  Dame  ed. ),  p.  455-477.  Springer- Verlag, 
Berlin,  Heidelberg. 

Jorgensen,  C.  B. 

1996.  Bivalve  filter  feeding  revisited.  Mar.  Ecol.  Prog.  Ser. 
142:287-302. 

Kafanov.  A.  I. 

1991.  Bivalves  on  continental  shelf  and  continental  slope 
of  Northern  Pacific,  p.  81-82.  Science  Acad.  USSR, 
Vladivostok. 

MacMahon,  R.  F. 

198.3.     Ecology  of  an  invasive  pest  bivalve,  Corbicula.     In 
The  Mollusca,  61 12)  (W.  D.  Russell-Hunter,  ed.),  p.  505-561. 
Academic  Press,  Orlando,  FL. 
Nakamura,  M.,  M.  Yamamuro,  M.  Ishikawa,  and  H.  Nishimura. 
1988.     Role  of  the  bivalve  Corbicula  japonica  in  the  nitrogen 
cycle  in  a  mesohaline  lagoon.     Mar.  Biol.  99:  369-374. 
Nakaoka,  M. 

1992.  Age  determination  and  growth  analysis  based  on 
external  shell  rings  of  the  protobranch  bivalve  Yoldia 
notabilis  Yokoyama  in  Otsuchi  Bay.  Northeastern  Japan. 
Benthos  Res.  43:53-66. 

1996.     Size-dependent  survivorship  of  the  bivalve  Yoldia 
notabilis  (Yokoyama,  1920):  the  effect  of  crab  predation.     J. 
Shellfish  Res.  15:355-362. 
Newell,  G.  E. 

1964.     Physiological  aspects  of  the  ecology  of  intertidal 
molluscs.    In  Physiology  of  Mollusca  (K.  M.  Wilbur  and  C. 
M.  Yonge  eds. ),  p.  59-81.    Academic  Press,  London. 
Pauly,  D. 

1987.    A  review  of  the  ELEFAN  system  for  analysis  of  length 


24 


Fishery  Bulletin  102(1) 


frequency  data  in  fish  and  aquatic  invertebrates.  In 
Length-based  methods  in  fisheries  research  (D.  Pauly  and 
G.  R.  Morgan  eds.),  p.  7-34.  ICLARM  conference  pro- 
ceedings 13.  688  International  Center  for  Living  Aquatic 
Resource  Management.  Manila,  Philippines  and  Kuwait 
Institute  for  Scientific  Research.  Safet.  Kuwait. 

Sakamoto,  Y,  M.  Ishiguro,  and  G.  Kitagawa. 

1983.  Statistics  based  on  information  amount  (Jouhour- 
you  Toukei  Gaku),  236  p.  Kyoritu  Shuppan,  Tokyo.  [In 
Japanese.] 

Scholten,  H.,  and  A.  C.  Smaal. 

1998.  Responses  of  Mytilus  edulis  L.  to  varying  food  concen- 
trations: testing  EMMY,  an  ecophysiological  model.  J.  Exp. 
Mar.  Biol.  Ecol.  219:217-39. 

Seed.  R. 

1980.  Shell  growth  and  form  in  the  Bivalvia.  //;  Skeletal 
growth  of  aquatic  organisms  (D.  G.  Rhoads  ed.),  p.  23-67. 
Plenum  Press,  New  York,  NY. 

Sellmer,  G.  P. 

1956.  A  method  for  the  separation  of  small  bivalve  molluscs 
from  sediments.     Ecology  37:206. 


Sokal,  R.  R.  and  F.  J.  Rohlf. 

1995.     Biometry,  3rd  ed.,  887  p.    W.  H.  Freeman  and  Com- 
pany, New  York,  NY. 
Utoh,  H. 

1981.     Growth  of  the  brackish  water  bivalve,  Corbicula 
japomca  Prime,  in  Lake  Abashiri.     Sci.  Rep.  Hokkaido 
Fish.  Exp.  Stn.  23:65-81.     [In  Japanese.] 
Yamakawa,  T..  and  Y.  Matsumiya. 

1997.     Simultaneous  analysis  of  multiple  length  frequency 
data  sets  when  the  growth  rates  fluctuate  between  years. 
Fish.  Sci.  63:708-714. 
Yamamuro,  M.,  and  I.  Koike. 

1993.     Nitrogen  metabolism  of  the  filter-feeding  bivalve 
Corbicula  japonica  and  its  significance  in  primary  produc- 
tion of  a  brackish  lake  in  Japan.     Limnol.  Oceanogr.  38: 
997-1007. 
Zar,  J.  H. 

1999.     Biostatistical  analysis.  663  p.     Prentice  Hall,  Upper 
Saddle  River,  NJ. 


25 


Abstract— Information  is  summarized 
on  juvenile  salmonid  distribution,  size, 
condition,  growth,  stock  origin,  and 
species  and  environmental  associations 
from  June  and  August  2000  GLOBEC 
cruises  with  particular  emphasis  on 
differences  related  to  the  regions  north 
and  south  of  Cape  Blanco  off  Southern 
Oregon.  Juvenile  salmon  were  more 
abundant  during  the  August  cruise  as 
compared  to  the  June  cruise  and  were 
mainly  distributed  northward  from 
Cape  Blanco.  There  were  distinct  differ- 
ences in  distribution  patterns  between 
salmon  species:  chinook  salmon  were 
found  close  inshore  in  cooler  water  all 
along  the  coast  and  coho  salmon  were 
rarely  found  south  of  Cape  Blanco.  Dis- 
tance offshore  and  temperature  were 
the  dominant  explanatory  variables 
related  to  coho  and  chinook  salmon 
distribution.  The  nekton  assemblages 
differed  significantly  between  cruises. 
The  June  cruise  was  dominated  by  juve- 
nile rockfishes,  rex  sole,  and  sablefish, 
which  were  almost  completely  absent 
in  August.  The  forage  fish  community 
during  June  comprised  Pacific  herring 
and  whitebait  smelt  north  of  Cape 
Blanco  and  surf  smelt  south  of  Cape 
Blanco.  The  fish  community  in  August 
was  dominated  by  Pacific  sardines  and 
highly  migratory  pelagic  species.  Esti- 
mated growth  rates  of  juvenile  coho 
salmon  were  higher  in  the  GLOBEC 
study  area  than  in  areas  farther  north. 
An  unusually  high  percentage  of  coho 
salmon  in  the  study  area  were  preco- 
cious males.  Significant  differences  in 
growth  and  condition  of  juvenile  coho 
salmon  indicated  different  oceano- 
graphic  environments  north  and  south 
of  Cape  Blanco.  The  condition  index 
was  higher  in  juvenile  coho  salmon  to 
the  north  but  no  significant  differences 
were  found  for  yearling  chinook  salmon. 
Genetic  mixed  stock  analysis  indicated 
that  during  June,  most  of  the  chinook 
salmon  in  our  sample  originated  from 
rivers  along  the  central  coast  of  Oregon. 
In  August,  chinook  salmon  sampled 
south  of  Cape  Blanco  were  largely  from 
southern  Oregon  and  northern  Cali- 
fornia; whereas  most  chinook  salmon 
north  of  Cape  Blanco  were  from  the 
Central  Valley  in  California. 


Manuscript  approved  for  publication 
30  June  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull  102:25-46  (2004). 


Juvenile  salmonid  distribution,  growth,  condition, 
origin,  and  environmental  and  species  associations 
in  the  Northern  California  Current* 


Rick  D.  Brodeur 

Northwest  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
2030  S.  Marine  Science  Drive 
Newport,  Oregon  97365 
E-mail  address:  Rick-Brodeuriffinoaa-gov 

Joseph  P.  Fisher 

College  of  Ocean  and  Atmospheric  Sciences 
Oregon  State  University 
Corvallis,  Oregon  97331 

David  J.  Teel 

Northwest  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
Seattle,  Washington  98112 


Robert  L.  Emmett 

Northwest  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
2030  S  Marine  Science  Drive 
Newport,  Oregon  97365 

Edmundo  Casillas 

Northwest  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
Seattle,  Washington  98112 


Todd  W.  Miller 

Cooperative  Institute  for  Marine  Resources 

Studies 
Oregon  State  University 
Newport,  Oregon  97365 


The  need  to  understand  the  direct 
and  indirect  linkages  between  oceano- 
graphic  conditions  and  salmon  sur- 
vival in  the  marine  environment  has 
increased  with  the  listing  of  many 
West  Coast  salmon  stocks  as  threat- 
ened or  endangered.  Recent  studies 
have  shown  that  long-term  changes 
in  climate  affect  oceanic  structure  and 
produce  abrupt  differences  in  salmon 
marine  survival  and  returns  (Francis 
and  Hare,  1994:  Mantua  et  al.,  19971.  A 
major  regime  shift  in  the  subarctic  and 
California  Current  ecosystems  during 
the  late  1970s  may  have  been  a  factor 
in  reducing  ocean  survival  of  salmon  in 
the  Pacific  Northwest  and  in  increas- 
ing marine  survival  in  Alaska  ( Hare  et 
al.,  1999).  Fluctuations  in  mortality  of 
salmon  in  the  freshwater  and  marine 
environments  have  been  shown  to  be 
almost  equally  significant  sources  of 
annual  salmonid  recruitment  variability 
( Bradford,  1995 ).  Unlike  in  the  freshwa- 
ter environment,  the  physical  and  bio- 
logical mechanisms  and  factors  in  the 
marine  environment  that  cause  mor- 
tality of  salmon  are  largely  unknown. 
Predation,  inter-  and  intraspecific 
competition,  food  availability,  smolt 
quality  and  health,  and  physical  ocean 
conditions  likely  influence  survival  of 
salmon  in  the  marine  environment. 


Thus,  increasing  our  understanding  of 
nearshore  ocean  environments,  their 
linkages  to  oceanographic  conditions, 
and  the  role  they  play  in  salmonid 
survival,  could  provide  management 
options  for  increasing  adult  returns. 
Characterization  of  the  space-time  vari- 
ability of  the  environmental  conditions 
that  smolts  encounter  when  they  enter 
the  nearshore  ocean,  and  the  eventual 
survival  of  these  smolts  will  allow  us  to 
identify  which  biotic  and  abiotic  ocean 
conditions  are  correlated  with  various 
ocean  survival  levels. 

Many  anadromous  salmonid  popula- 
tions along  the  west  coast  of  the  United 
States  have  declined  over  the  last  few 
decades  (Nehlsen  et  al.,  1991),  and  most 
stocks  show  a  regional  north-south  pat- 
tern in  degree  of  extinction  risk  (Kope 
and  Wainwright,  1998).  This  pattern 
suggests  that  both  marine  habitat  con- 
ditions and  mesoscale  climate  patterns 
affect  salmonid  population  status  (e.g. 
Lawson,  1993).  A  dramatic  example  is 
the  population  trend  of  coho  salmon 
(Oncorhynchus  kisutch)  along  the  Or- 
egon coast.  Populations  along  the  coast 
north  of  Cape  Blanco  (43°N)  have  exhib- 


;  Contribution  number  364  of  the  U.S. 
GLOBEC  program.  NEP  Office,  Oregon 
State  University,  Corvallis.  OR. 


26 


Fishery  Bulletin  102(1) 


ited  a  strong  decline  in  size  and  survival  in  the  mid-1990s; 
whereas  populations  south  of  Cape  Blanco  have  not  shown 
this  trend  (Lewis1).  This  finding  suggests  that  these  two 
populations  have  experienced  different  ocean  conditions. 

The  quality  of  the  marine  habitat  (in  terms  of  habitat 
complexity,  prey  density,  and  temperature)  undoubt- 
edly influences  fish  growth  and  condition.  Growth  and 
indices  of  condition  can  be  used  as  measures  of  habitat 
quality  for  juvenile  salmon  and  to  identify  essential  links 
between  oceanographic  conditions  and  survival  of  salmon 
populations  during  the  critical  juvenile  life  history  phase. 
Measures  such  as  growth  (growth  rate,  size  variation,  and 
allometric  relationships)  (Lorenzen,  1996;  McGurk,  1996) 
and  accumulation  of  energetic  reserves  used  in  growth  and 
sustenance  during  the  low-productivity  winter  periods 
have  been  used  previously  to  characterize  habitat  quality 
and  to  describe  how  it  ultimately  affects  the  individual  and 
the  population  (Perry  etal.,  1996;  Paul  and  Willette,  1997). 
Environmental  factors  are  known  to  affect  growth,  repro- 
duction, survival,  and  ultimately  population  recruitment 
(Hinch  et  al.,  1995;  Marschall  and  Crowder,  1995;  Fried- 
land  and  Haas,  1996).  As  such,  fish  condition,  growth  rate, 
and  size  in  the  pre-adult  stages  are  parameters  that  can  be 
used  to  identify  the  influence  of  natural  and  anthropogenic 
ocean  conditions  on  marine  survival. 

Much  of  our  current  knowledge  of  the  dominant  nekton 
of  the  pelagic  ecosystem  off  the  coasts  of  Oregon  and  Wash- 
ington is  derived  from  a  series  of  17  cruises  conducted  by 
Oregon  State  University  (OSU)  from  1979  to  1985.  These 
collections,  consisting  of  >900  quantitative  purse  seine  sets 
in  the  northern  California  Current,  were  made  to  examine 
geographic  distributions  and  temporal  trends  of  the  domi- 
nant nekton  and  how  these  relate  to  physical  and  biotic 
conditions  at  the  time  of  capture.  The  primary  purpose 
of  these  cruises  was  to  collect  data  for  assessment  of  the 
abundance,  distribution,  growth,  migration,  and  ecology  of 
juvenile  salmon  in  coastal  waters.  Data  on  the  distribution, 
migration  and  growth  of  juvenile  salmon  from  these  cruises 
have  been  summarized  in  Fisher  and  Pearcy  (1988;  1995). 
Pearcy  and  Fisher  ( 1988,  1990),  and  Pearcy  ( 1992).  Analy- 
sis of  the  nonsalmonid  data  includes  studies  on  their  abun- 
dance and  distribution  (Brodeur  and  Pearcy,  1986;  Emmett 
and  Brodeur,  2000),  feeding  habits  (Brodeur  et  al.,  1987) 
and  interannual  variability  in  relation  to  oceanographic 
conditions  (Brodeur  and  Pearcy,  1992).  In  addition,  the 
distribution  of  juvenile  salmon  (mainly  coho  and  chinook 
salmon  [O.  tshawytscha})  has  been  studied  more  recently 
as  a  component  of  a  multiyear  Columbia  River  Plume  study 
(Emmett  and  Brodeur,  2000;  Teel  et  al.,  2003;  Brodeur  et 
al.,  2003).  However,  all  these  cruises  extended  only  as  far 
south  as  Cape  Blanco,  with  the  exception  of  one  cruise  (July 
1984),  which  extended  as  far  south  as  Eureka,  California, 
but  included  only  a  few  collections  south  of  Cape  Blanco 
(Pearcy  and  Fisher,  1990).  Thus,  the  region  south  of  Cape 
Blanco  is  almost  completely  unknown  in  terms  of  juvenile 


1  Lewis,  M.  A.  2002.  Stock  assessment  of  anadromous  salmo- 
nids  2001.  Monitoring  program  report  OPSW-ODFW-2002-04, 
57  p.     Oregon  Dept.  Fish  Wildlife,  Portland.  OR  97207. 


salmon  distribution,  pelagic  nekton,  and  biological  ocean- 
ography in  general,  despite  being  an  area  of  very  strong 
upwelling  and  high  productivity.  Also,  the  fine-scale  dis- 
tribution of  juvenile  salmon  in  relation  to  environmental 
variables  has  not  been  studied  in  any  detail. 

The  California  Current  is  not  homogeneous  but  rather 
can  be  divided  into  distinct  subunits  or  regions,  each  with 
its  own  physical  and  biological  characteristics  (U.S.  GLO- 
BEC,  1994).  A  break  between  the  northernmost  two  regions 
occurs  at  Cape  Blanco,  where  the  equatorward  upwelling 
jet  veers  sharply  off  the  shelf  and  into  the  California  Cur- 
rent (Barth  et  al.,  2000).  The  upwelling  zone  north  of  the 
cape  is  narrow,  extending  out  about  30  km,  whereas  south 
of  Cape  Blanco,  it  can  extend  up  to  100  km  offshore.  This 
area  also  appears  to  represent  a  faunal  break  for  some  zoo- 
plankton  communities  (McGowan  et  al.,  1999;  Peterson  and 
Keister,  2002)  and  is  a  break  point  for  alternative  salmon 
migration  strategies  (Weitkamp  et  al.,  1995;  Weitkamp  and 
Neely,  2002). 

During  the  summer  of  2000,  we  conducted  broad-scale 
sampling  and  fine-scale  process  studies  from  central  Or- 
egon to  northern  California  to  examine  the  distribution 
of  juvenile  salmon  and  associated  species  in  relation  to 
environmental  conditions.  This  was  one  component  of  a 
multidisciplinary  U.S.  Global  Ocean  Ecosystem  Dynamics 
(GLOBEC)  Northeast  Pacific  study  examining  the  north- 
ern California  Current  ranging  in  scope  from  the  physics 
up  to  the  top  trophic  levels  (Batchelder  et  al.,  2002).  We 
were  interested  in  examining  the  distribution  of  juvenile 
salmon  north  and  south  of  Cape  Blanco,  the  origin  of  these 
fish,  and  any  regional  differences  in  growth  and  condition 
of  salmon  across  the  range  of  sampling.  Evidence  exists 
that  the  physical  conditions  and  the  associated  biota  are 
different  within  this  geographical  scale.  Thus,  analyses  of 
the  relationship  between  oceanographic  conditions  and  the 
response  of  resident  biota  can  provide  insights  into  the 
linkages  associated  with  physical  and  biological  processes 
that  shape  the  biological  community,  and  in  particular, 
those  associated  with  salmon  recruitment. 


Methods 

Field  surveys 

Surveys  were  conducted  over  two  time  periods — early 
summer  (29  May-18  June,  2000)  and  late  summer  (28 
July-15  August,  2000).  Each  survey  consisted  of  a  meso- 
scale  grid  along  designated  GLOBEC  transects  that  had 
been  monitored  for  several  years  and  by  fine-scale  pro- 
cess sampling  at  stations  of  interest  based  on  features 
observed  in  the  physical  environment  (fronts  or  eddies) 
or  by  acoustic  sampling  conducted  by  two  accompanying 
oceanographic  vessels  (RV  Wecoma  and  RV  New  Horizon). 
Further  details  on  the  physical  and  biological  conditions 
occurring  at  the  time  of  our  sampling  have  been  reported 
by  Batchelder  et  al.  (2002). 

For  the  mesoscale  survey,  stations  were  established  at 
1,  5,  10,  15,  20,  25  and  30  nautical  miles  from  shore  on 
each  of  five  transects.  Inclement  weather,  particularly 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  |uvenile  salmonids 


27 


during  the  first  cruise,  prevented  us  from  sampling  all 
the  stations  along  each  transect.  At  each  station,  a  Nordic 
264  rope  trawl  built  by  Nor'Eastern  Trawl  Systems,  Inc. 
(Bainbridge  Island,  WA)  was  towed  in  surface  waters  by 
a  chartered  fishing  vessel  (FV  Sea  Eagle)  at  a  speed  of  6 
km/h.  This  rope  trawl  has  a  maximum  mouth  opening  of 
approximately  30  m  x  18  m.  Mesh  sizes  ranged  from  162.6 
cm  in  the  throat  of  the  trawl  near  the  jib  lines  to  8.9  cm  in 
the  codend.  To  maintain  catches  of  small  fish  and  squid,  a 
6.1-m  long,  0.8-cm  mesh  knotless  liner  was  sewn  into  the 
codend.  All  tows  were  30  minutes  in  duration.  All  fish  and 
squid  caught  were  counted  and  measured  at  sea.  After  fork 
length  (FL)  was  measured  to  the  nearest  mm,  all  juvenile 
salmon  were  immediately  frozen  for  later  determinations 
of  growth,  condition,  food  habits,  genetic  analysis,  and  as- 
sessment of  pathological  condition. 

The  physical  and  biological  environment  was  monitored 
and  sampled  at  each  station  immediately  prior  to  setting 
the  trawl.  A  CTD  (conductivity,  temperature,  and  depth) 
cast  was  made  with  a  Sea-Bird  SBE  19  Seacat  profiler  to 
100  m  at  deep  stations  or  within  10  m  of  the  bottom  at 
shallow  stations.  Chlorophyll  and  nutrient  samples  were 
collected  from  3  m  depth  with  a  Niskin  water  sampler.  A 
neuston  tow  with  a  1-m2  mouth  containing  333-,(im  mesh 
net  was  towed  for  5  minutes  out  of  the  wake  of  the  vessel 
at  each  station.  General  Oceanics  flow  meters  were  placed 
inside  the  net  to  measure  the  amount  of  water  sampled. 
Additional  details  on  the  analysis  of  these  neuston  trawls 
are  available  in  Reese  et  al.2 

Condition  and  growth  analysis 

Each  salmonid  was  remeasured  (FL  to  the  nearest  mm) 
and  weighed  (to  the  nearest  0.1  g)  in  the  laboratory.  A  por- 
tion of  hepatic  and  muscle  tissue  was  excised,  placed  in 
individual  capsules,  frozen  in  liquid  nitrogen,  and  stored 
at  -80°C  until  analyzed.  The  bioenergetic  health  of  juve- 
nile salmon  was  evaluated  by  assessing  changes  in  water 
content  (as  a  surrogate  measure  of  fat  accumulation)  of 
liver  and  muscle  to  estimate  dry  tissue  weight.  The  water 
content  was  determined  by  drying  tissue  samples  to  a  con- 
stant weight  at  105°C.  The  accumulation  of  energy  reserves 
during  the  growth  season  ( energy  reserves  of  salmon  in 
August  in  relation  to  salmon  collected  in  June)  that  would 
enhance  survival  of  juveniles  during  the  winter  when  food 
availability  is  lower  was  also  measured.  The  condition  of 
juvenile  salmon  was  assessed  by  examining  weight  residu- 
als (by  using  either  the  wet  weight  or  dry  weight)  derived 
from  the  allometric  relationship  between  length  and  weight 
of  individual  juvenile  salmon  after  logarithmic  transforma- 
tion (Jakob  et  al.,  1996)  of  salmon  captured  in  June  and 
August.  Wet-weight  residuals  are  representative  of  the 
traditional  condition  index  of  animals  and  are  a  reflection 


2  Reese,  D.C.,  T.W.Miller,  and  R.D.  Brodeur.  2003.  Community 
structure  of  neustonic  zooplankton  in  the  northern  California 
Current  in  relation  to  oceanographic  conditions.  22  p.  Unpubl. 
manuscript.  Northwest  Fisheries  Science  Center,  NMFS.  2030 
S.  Marine  Science  Drive,  Newport,  OR  97365. 


of  somatic  tissue  growth.  Dry-weight  residuals  are  respon- 
sive to  accumulation  of  fat  stores  and  are  a  reflection  of  the 
bioenergetic  health  of  the  individual  animal  (Sutton  et  al., 
2000;  Post  and  Parkinson,  2001). 

To  contrast  growth  characteristics  during  2000  in  differ- 
ent latitudinal  ranges  of  the  California  Current,  we  com- 
pared ocean  growth  rates  of  juvenile  coho  salmon  south 
and  north  of  Cape  Blanco  in  the  GLOBEC  study  area, 
and  in  the  area  from  Newport,  Oregon,  north  to  northern 
Washington.  The  physical  and  biological  characteristics  of 
these  three  regions  of  the  coastal  ocean  differ  greatly  (U.S. 
GLOBEC,  1994),  and  these  differences  may  impact  the  dis- 
tribution and  abundance  of  prey  of  juvenile  salmonids  and 
therefore  may  also  affect  salmonid  growth.  Data  north  of 
Newport,  Oregon,  were  collected  during  a  separate  study  of 
the  Columbia  River  plume  and  the  adjacent  coastal  ocean 
(hereafter  called  the  "plume  study")  using  the  same  trawl 
and  a  similar  sampling  strategy  as  in  the  GLOBEC  study 
(see  Emmett  and  Brodeur  [2000]  and  Teel  et  al.  [2003]  for 
details). 

Scales  were  examined  from  45  juvenile  coho  salmon 
caught  during  the  June  and  August  2000  GLOBEC 
cruises  and  252  juvenile  coho  salmon  caught  during  the 
2000  plume  cruises.  The  scales  were  mounted  on  gummed 
cards  from  which  acetate  impressions  were  made.  Using 
a  video  camera  attached  to  a  compound  microscope  and 
Optimas®  imaging  software  (vers.  5.1,  Optimas  Inc.,  Se- 
attle, WA)  we  measured  the  distance  (scale  radius)  along 
the  anterior-posterior  axis  of  each  scale  from  the  focus 
(F)  to  the  ocean  entry  mark  (OE)  and  to  the  scale  margin 
(Fig.  1).  The  fork-length  of  each  fish  at  the  time  of  ocean 
entry  (FL0E)  was  estimated  from  the  scale  radius  (SR0E) 
at  ocean  entry  using  the  Fraser  and  Lee  back-calculation 
method  (Ricker,  1992): 


FL„ 


(FL- 36.07) 
SR 


xSRof.  +36.07, 


where  FL  =  length  at  capture; 

SR  =  scale  radius  at  capture;  and 
36.07  =  the  intercept  from  a  regression  of  SR  on  FL 
for  juvenile  coho  salmon  caught  in  the  ocean 
(Fig.  2A). 


In  an  analogous  fashion,  fish  weight  at  time  of  ocean  entry 
(Wr0£)  was  back-calculated  f 
length  at  ocean  entry  (FL0E): 


(Wt0E)  was  back-calculated  from  the  estimated  fish  fork 


\ni  Wt0E)  = 


(ln(Wr  1  +  12.633) 
ln(FL) 


xln(FLr,F  1-12.633, 


where  Wt  =  weight  at  capture;  and 
-12.633  =  the  intercept  from  a  linear  regression  of 
ln(Wr)  on  ln(FL)  for  juvenile  coho  salmon 
caught  in  the  ocean  (Fig.  2B). 

The  growth  rate  in  FL, 

(FL-FL0E)lAd, 


28 


Fishery  Bulletin  102(1) 


Figure  1 

Scale  from  a  352-mm  FL  male  juvenile  coho  salmon 
(Oncorhynchus  kisutch)  caught  during  the  August  2000 
GLOBEC  cruise  showing  the  axis  of  measurement  (black 
line),  the  focus  (F),  the  mark  of  ocean  entry  (OE),  and  the 
scale  margin  (SM). 


and  the  instantaneous  growth  rate  in  weight: 

G  =  (MWt)-MWt0E))/M, 

where  Ad  =  estimated  days  between  ocean  entry  and  cap- 
ture, were  estimated  for  each  salmon. 

The  meaning  of  the  instantaneous  growth  rate  G  can  be 
stated  as  follows:  if  salmon  growth  is  exponential  between 
ocean  entry  and  capture,  then 

Wt 


Wt„ 


and  at  any  instant  the  fish's  weight  increases  at  the  rate  of 
G  of  its  body  weight  per  day.  G  can  be  multiplied  by  100  to 
give  the  instantaneous  growth  rate  in  terms  of  percentage 
of  body  weight  per  day. 

Although  the  dates  of  ocean  entry  of  individual  lish 
were  unknown,  seaward  migration  of  coho  salmon  smolts 
in  California,  Oregon,  and  Washington  rivers  occurs  mainly 
between  mid-April  and  mid-June,  and  there  is  no  consis- 
tent latitudinal  trend  in  timing  of  the  migration  ( Weitkamp 
et  al.,  1995).  Peak  downstream  migration  of  coho  salmon 
smolts  was  between  mid-May  and  very  early  June  in  the 
Columbia  River  estuary,  1978-83  (Dawley  et  al.,  1985), 
and  in  the  lower  Trinity  River,  California,  1997-2000  (US- 


A   FL  (mm)  vs  scale  radius  (mm) 


GM  Regression:  FL  =  152.22  SR  +36.07 
r2  =  0.94,  n=370 


1  2 

Scales  radius  (mm) 


B   Wt(g)vs  FL(mm) 

In(WI)  =  3.2273'ln(FL)  -  12  6329 
or  Wt(g)  =  3.263x1  (T5  FL(mm)32273 
n=1018V  =  0.99 


s 

—         5 


In  (FL) 

Figure  2 

(A)  Regression  of  fork  length  (FL)  on  scale  radius  and.  'Bi 
regression  of  ln(WY)  on  ln(FL)  for  juvenile  coho  salmon  {On- 
corhynchus  kisutch)  caught  during  the  May  1998-September 
2000  Columbia  River  plume  study. 


FWS3).  In  2000,  peak  downstream  migration  of  mainly 
nonhatchery  coho  salmon  smolts  at  13  monitoring  sites  in 
coastal  Oregon  rivers  north  of  Cape  Blanco  occurred  from 
April  2  to  May  20;  median  peak  migration  occurred  26  April 
( Solazzi  et  al.4)  From  the  information  available  on  timing  of 
seaward  migration  of  coho  salmon  smolts.  we  used  an  ocean 
entry  date  of  15  May  when  calculating  Ad  and  estimating 
ocean  growth  rates  of  unmarked  coho  salmon  from  scales. 
In  addition  to  estimating  growth  rates  of  juvenile 
coho  salmon  from  scales,  we  also  estimated  instantaneous 
growth  rates  in  weight  between  hatchery  release  and  cap- 
ture in  the  ocean  of  28  coded-wire-tagged  (CWT)  juvenile 
coho  salmon: 


USFWS  (U.S.  Fish  and  Wildlife  Service).  2001.  Juvenile  sal- 
monid  monitoring  on  the  mainstem  Klamath  River  at  Big  Bar 
and  mainstem  Trinity  River  at  Willow  Creek,  1997-2000,  106  p. 
Annual  report  of  the  Klamath  River  Fisheries  Assessment  Pro- 
gram. Areata  Fish  and  Wildlife  Office,  Areata,  CA  9552 1 . 
Solazzi,  M.F.,  S.L.Johnson,  B.Miller,  and  T.Dalton.  2002.  Sal- 
monid  life-cvele  monitoring  project  2001.  Monitoring  program 
report  OPSW-ODFW-2002-2,  25  p.  Oregon  Dept.  Fish  and 
Wildlife,  Portland.  OR  97207. 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  juvenile  salmonids 


29 


G  =  (MWt)-MWtR))/M, 

where  Wt  =  weight  of  the  CWT  fish  at  capture; 

WtR  =  the  average  weight  of  fish  in  the  CWT  group 

at  time  of  release;  and 
Ad  =  days  between  hatchery  release  and  capture 
in  the  ocean. 

Estimated  growth  rates  of  these  CWT  fish,  of  known  release 
date  and  known  average  release  weight  were  used  to  vali- 
date the  growth  rates  estimated  from  scale  analysis 

Our  analysis  of  the  growth  of  chinook  salmon  based  on 
scale  characteristics  is  not  far  enough  advanced  to  report 
in  this  article.  We  plan  to  present  these  data  in  a  later 
article. 

Contribution  of  hatchery  coho  salmon  to  catches 

The  total  numbers,  percentages  of  marked  fish  ( any  exter- 
nal fin  clips  or  internal  tags)  and  grand  average  weights 
of  hatchery  coho  salmon  smolts  released  in  2000  are  sum- 
marized for  different  release  regions  in  Appendix  Table  1. 
These  data  were  used  to  compare  the  estimated  average 
weights  of  fish  at  time  of  ocean  entry  (from  scale  analy- 
sis )  with  the  average  weights  of  hatchery  fish  at  time  of 
release,  and  also  to  estimate  the  proportions  of  hatchery 
coho  salmon  in  our  catches.  We  calculated  the  expected 
percentage  (E%)  of  marked  fish  in  each  catch  if  100%  of 
the  fish  were  hatchery  fish: 


E% 


X*.*4, 


where  i?,  =  the  proportional  contribution  of  region  i  to  the 
catch  (this  paper  for  the  GLOBEC  catches, 
and  from  Teel  et  al.,  2003  for  the  plume  study 
catches);  and 
A,  =  the  percentage  of  hatchery  fish  that  were 
marked  in  region  i  ( from  Appendix  Table  1 ). 

The  percentage  of  hatchery  fish  in  each  catch  sample  (H%) 
was  then  estimated  as 

0% 

H%  =  —  xlOO, 
E% 

where  OcA  =  observed  percentage  of  marked  fish. 

Genetic  analysis 

The  freshwater  origins  of  juvenile  chinook  and  coho  salmon 
and  steelhead  (O.  my  kiss)  were  studied  by  using  standard 
methods  of  genetic  mixed  stock  analysis  (Milner  et  al., 
1985;  Pella  and  Milner,  1987).  According  to  the  methods 
described  by  Aebersold  et  al.  (1987),  samples  of  eye,  liver, 
heart,  and  skeletal  muscle  were  extracted  from  frozen  whole 
juvenile  salmon  and  analyzed  with  horizontal  starch-gel 
protein  electrophoresis.  Data  from  previous  studies  char- 
acterizing genetic  (allozyme)  differences  among  spawning 
populations  in  California  and  the  Pacific  Northwest  were 
then  used  as  baseline  data  to  estimate  the  stock  composi- 
tions of  our  ocean  caught  mixed-stock  samples.  Baselines 


consisted  of  32  gene  loci  and  116  populations  for  chinook 
salmon  (Teel  et  al.5),  58  loci  and  49  populations  for  coho 
salmon  (Teel  et  al.,  2003),  and  55  loci  and  57  populations 
for  steelhead  (Busby  et  al.,  1996).  Estimates  of  stock  com- 
positions were  made  by  using  the  maximum  likelihood 
procedures  described  by  Pella  and  Milner  (1987)  and  the 
Statistical  Package  for  Analyzing  Mixtures  (Debevec  et  al., 
2000).  Estimates  of  individual  baseline  populations  were 
then  summed  to  estimate  contributions  of  regional  stock 
groups.  Precision  of  the  stock  composition  estimates  was 
estimated  by  bootstrapping  the  estimates  100  times  with 
resampling  of  the  baseline  and  mixture  genetic  data  as 
described  in  Pella  and  Milner  (1987). 

Habitat  and  assemblage  analysis 

The  raw  numbers  offish  and  squid  caught  from  each  trawl 
were  converted  to  densities  based  on  the  volume  filtered 
per  trawl  to  standardize  for  differences  in  effort  between 
tows.  Density  contours  of  juvenile  salmon  and  other  nekton 
were  produced  using  specialized  graphics  programs.  We 
then  tested  whether  the  habitat  associations  of  the  domi- 
nant salmonids  were  significantly  different  from  the  total 
habitat  sampled  by  following  the  methods  outlined  in  Perry 
and  Smith  ( 1994).  This  procedure  involved  comparing  the 
cumulative  distributions  of  salmon  catch  with  observed 
environmental  conditions  (temperature,  salinity,  chloro- 
phyll-a  at  one  meter,  water  depth,  and  neuston  displace- 
ment volume).  We  performed  5000  randomizations  of  the 
data  and  used  the  Cramer-von  Mises  test  statistic  recom- 
mended by  Syrjala  ( 1996)  as  being  robust  to  the  effects  of 
inordinately  large  catches. 

To  explore  the  relationship  between  juvenile  salmon  and 
other  fish  species  and  environmental  variables,  we  used 
several  types  of  multivariate  analyses  (McCune  and  Grace, 
2002 ).  Original  data  from  each  of  the  two  cruises  formed 
complimentary  species  and  environmental  matrices.  The 
June  and  August  cruises  were  analyzed  individually  to 
look  at  spatial  patterns  of  species  composition  in  relation  to 
environmental  gradients  (Gauch,  1982).  To  avoid  spurious 
effects  of  rare  species,  we  excluded  species  from  the  data 
matrix  that  had  a  frequency  of  occurrence  of  less  than  10% 
of  the  possible  occurrences  for  each  cruise  (McCune  and 
Grace,  2002).  To  minimize  the  effect  of  very  large  catches, 
the  data  were  log  transformed.  Stations  with  no  species 
present  were  eliminated  from  the  data  set  to  allow  for  anal- 
ysis of  sample  units  in  species  space.  Data  transformations 
and  their  effects  on  the  summary  statistics  were  examined 
prior  to  analysis.  Analyses  of  data  were  performed  by  using 
PC-ORD  version  4.28  (McCune  and  Mefford,  1999). 

Agglomerative  hierarchical  cluster  analysis  (AHCA) 
using  the  Bray-Curtis  dissimilarity  measure  and  Wards 


Teel,  D.  J„  P.  A.  Crane.  C.  M.  Guthrie,  III,  A.  R.  Marshall.  D. 
M.  Van  Doornik,  W.  D.  Templin,  N.  V.  Varnavskaya,  and  L.  W. 
Seeb.  1999.  Comprehensive  allozyme  database  discriminates 
chinook  salmon  from  around  the  Pacific  Rim.  (NPAFC  docu- 
ment 440),  25  p.  Alaska  Department  of  Fish  and  Game,  Divi- 
sion of  Commercial  Fisheries,  333  Raspberry  Road,  Ancorage,  AK 
99518. 


30 


Fishery  Bulletin  102(1) 


linkage  function  was  applied  to  arrange  the  nekton  spe- 
cies assemblages  and  stations  into  cluster  groups.  The 
cutoff  level  to  form  optimal  groups  within  the  species 
and  station  dendrograms  was  based  on  several  criteria:  1) 
biological  meaning;  2)  significance  tests  of  groups  using 
a  multi-response  permutation  procedure  (MRPP);  and  3i 
comparison  of  cutoff  level  MRPP  results  with  those  groups 
obtained  from  one  cutoff  level  below  and  above  the  level  of 
interest.  A  nonparametric  procedure,  MRPP  compares  the 
a  priori  groupings  from  AHCA  and  tests  the  hypothesis 
of  no  difference  between  the  groups.  For  cluster  analysis 
of  stations,  indicator  species  analysis  (ISA)  was  used  to 
determine  nekton  species  strongly  associated  with  indi- 
vidual groups.  ISA  assigns  indicator  values  to  each  spe- 
cies according  to  relative  abundance  and  frequency,  then 
tests  the  significance  (Monte-Carlo  permutation  test)  of 
the  highest  species-specific  indicator  value  assigned  to  a 
particular  group. 

Nonmetric  multidimensional  scaling  (NMS;  Kruskal, 
1964)  was  used  to  ordinate  sample  units  in  species  space 
and  to  compare  station  cluster  groups  to  environmental 
gradients.  NMS  was  chosen  for  this  analysis  because  it  is 
robust  to  data  that  are  non-normal  and  that  have  high 
numbers  of  zeros.  Initial  runs  of  NMS  from  both  cruise  da- 
tasets  resulted  in  three-dimensional  solutions.  Subsequent 
reapplication  of  NMS  using  a  three-dimensional  solution 
(Sorensen  distance,  400  maximum  iterations,  and  40  runs 
with  real  data)  was  applied  for  the  final  ordinations.  To 
examine  the  environmental  or  station  factors  associated 
with  each  NMS  axis  that  may  have  affected  the  distribu- 
tion of  the  dominant  taxa,  we  correlated  the  NMS  station 
and  species  scores  to  a  suite  of  environmental  variables 
including  water  depth,  distance  offshore,  latitude,  surface 
temperature,  surface  salinity,  chlorophyll-a  concentration, 
and  neuston  zooplankton  settled  volumes.  Pearson  and 
Kendall  correlations  with  each  ordination  axis  were  used 
to  measure  strength  and  direction  of  individual  species  and 
environmental  parameters. 


Results 

Distribution  of  juvenile  salmon  and  other  species 

We  collected  a  total  of  18,852  nekton  individuals:  two  ceph- 
alopod,  one  agnathan,  two  elasmobranch,  and  57  fish  taxa 
from  163  surface  trawls  (see  Table  1  for  scientific  names 
of  all  species).  With  the  exception  of  market  squid  in  June 
and  blue  shark  in  August,  most  of  the  nonteleost  nekton 
occurred  in  only  a  few  collections.  Substantially  fewer  fish 
were  caught  in  the  June  cruise  than  in  the  August  cruise, 
but  the  diversity  was  much  higher  in  the  June  cruise.  The 
catch  in  June  was  dominated  by  forage  fishes  such  as 
Pacific  herring,  surf  and  whitebait  smelt,  and  juvenile  rock- 
fishes,  sablefish,  and  flatfishes.  Salmonids,  mainly  juvenile 
chinook  and  coho  salmon  and  steelhead,  comprised  a  rela- 
tively minor  proportion  of  the  catches  (only  114  juvenile 
salmonids;  1.9  %  of  the  total). 

The  August  cruise  was  dominated  by  several  large 
catches  of  Pacific  sardine  (Table  1 ).  Jack  mackerel  was  the 


most  common  nonsalmonid  caught.  Many  of  the  juvenile 
fish  taxa  caught  during  the  June  cruise  were  absent  during 
the  August  cruise;  those  that  did  occur  ( sablefish.  rex  sole) 
were  much  lower  in  abundance.  Mesopelagic  fishes  of  the 
family  Bathylagidae  and  Myctophidae  were  collected  only 
during  the  August  cruise,  mainly  because  of  the  inclusion 
of  more  offshore  stations  and  occasional  collections  during 
nondaylight  hours.  As  in  the  earlier  cruise,  salmonids  com- 
prised a  relatively  minor  percentage  of  the  catch  (3.19f )  but 
were  more  common  and  abundant  during  this  survey. 

Juvenile  chinook  salmon  were  broadly  distributed  lati- 
tudinally  during  both  cruises,  but  their  distribution  was 
mainly  restricted  to  nearshore  stations  within  the  100-m 
isobath  (Fig.  3).  Coho  salmon  juveniles  were  more  common 
north  of  Cape  Blanco  during  both  cruises  and  were  found 
generally  farther  offshore  than  chinook  salmon  juveniles 
(Fig.  3).  In  contrast,  steelhead  juveniles  were  found  mainly 
south  of  Cape  Blanco,  especially  in  June,  but  their  zonal 
distribution  overlapped  that  of  coho  salmon  juveniles. 

Size  and  condition  of  juvenile  salmon 

Fork  length  of  yearling  chinook  salmon  averaged  227  ±42 
mm  FL  in  June  and  230  ±30  mm  FL  in  August  and  aver- 
aged 135  ±12  mm  FL  for  subyearling  chinook  salmon  in 
August,  whereas  juvenile  coho  salmon  averaged  162  ±32 
mm  FL  in  June  and  286  ±46  mm  FL  in  August  ( Table  2 ).  No 
significant  differences  in  fork  length  of  juvenile  chinook  or 
coho  salmon  north  or  south  of  Cape  Blanco  were  evident. 

Juvenile  coho  salmon  weighed  significantly  more  on  a 
wet-weight  basis  for  a  given  fork  length  in  the  region  north 
of  Cape  Blanco  compared  to  juveniles  captured  south  of 
Cape  Blanco  (Fig.  4).  This  pattern  was  also  similar  and 
significant  when  evaluated  on  a  dry-weight  basis  (bioen- 
ergetic  growth).  Although  the  stock  composition  in  the  two 
regions  could  account  for  some  of  these  differences,  the 
growth  responses  likely  reflect  habitat-specific  features  in 
the  region  north  of  Cape  Blanco  that  benefit  coho  salmon. 
No  difference  in  condition  of  yearling  chinook  salmon  cap- 
tured north  or  south  of  Cape  Blanco,  on  either  a  wet-  or  dry- 
weight  basis,  was  evident  (Fig.  4).  Information  regarding 
size  and  condition  of  subyearling  chinook  salmon  are  not 
presented  because  few  subyearling  chinook  salmon  were 
caught  in  June  and  all  but  one  subyearling  chinook  salmon 
in  August  were  caught  in  the  region  south  of  Cape  Blanco, 
OR.  Insufficient  subyearling  chinook  salmon  were  avail- 
able for  an  analysis  comparable  to  that  done  for  yearling 
chinook  and  coho  salmon. 

Proportions  of  wild  and  hatchery  coho  salmon 

Most  of  the  juvenile  coho  salmon  caught  during  the  plume 
study  north  of  Newport,  Oregon,  originated  in  hatcher- 
ies (Table  3).  In  June  and  September  2000  we  estimated 
that  wild  fish  comprised  only  W9i  and  25r< .  respectively, 
of  the  catch.  Wild  fish,  however,  comprised  a  proportion- 
ally much  higher  percentage  of  the  catch  of  coho  salmon 
in  the  GLOBEC  study  area  in  June  north  of  Cape  Blanco 
(67$  I,  and  in  August  south  of  Cape  Blanco  (619!  I,  than  in 
the  plume  study  area  farther  to  the  north.  Most  jacks  and 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  |uvenile  salmonids 


31 


Table  1 

Phylogenetic  listing  of  nekton  catch  in  numerical  composition,  frequency  of  occurrence 

(F.O.)  and  size 

range  cau 

ght  for  each  cruise. 

(j)  indicates  juvenile  stage;  (a)  adult.  ML  = 

mantle  length,  TL  =  total  length. 

FL  =  fork  length,  SL  =  standard  length  ( 

in  mm). 

Class  and  Family 

Common  name 

June  (84  stations) 

August  (79  stations) 

Scientific  name 

dumber 

F.O. 

Size  range 

Number 

F.O. 

Size  range 

Cephalopoda 

Onychoteuthidae 

Pacific  clubhook 
squid 

Onychoteuthis 
borealijaponicus 

19 

6 

21-80  ML 

302 

6 

21-227  ML 

Loliginidae 

Market  squid 

Loligo  opalescens 

301 

14 

33-122  ML 

1 

1 

35  ML 

Agnatha 

Petromyzontidae 

Pacific  lamprey 

Lampetra  tridentata 

1 

1 

625  TL 

Chondrichthyes 

Alopiidae 

Thresher  shark 

Alopias  vulpinus 

1 

1 

36-576  TL 

Carcharhinidae 

Blue  shark 

Prionace  glauca 

18 

10 

1300-1660  TL 

Osteichthyes 

Xenocongridae 

Eel  leptocephalus 

Thalassenchelys  coheni 

3 

1 

214-243  TL 

2 

2 

260-305  TL 

Clupeidae 

Pacific  herring 

Clupea  pallasi 

1022 

9 

127-195  FL 

Pacific  sardine 

Sardinops  sagax 

7 

2 

237-260  FL 

10,327 

15 

178-290  FL 

Engraulididae 

Northern  anchovy 

Engraulis  mordax 

49 

12 

148-165  FL 

Salmonidae 

Chinook  salmon  (j,a) 

Oncorhynchus 
tshawytscha 

56 

18 

121-780  FL 

252 

26 

109-910  FL 

Coho  salmon  (j,a) 

Oncorhynchus  kisutch 

35 

15 

122-580  FL 

111 

25 

210-736  FL 

Cutthroat  trout  (j,a) 

Oncorhynchus  clarki 

1 

1 

186  FL 

3 

3 

258-341  FL 

Steelhead  trout  (j,a) 

Oncorhynchus  mykiss 

22 

8 

176-284  FL 

36 

13 

261-430  FL 

Osmeridae 

Smelt  (j) 

Osmeridae 

14 

4 

37-52  SL 

74 

5 

31-50  SL 

Surf  smelt 

Hypomesus  pretiosus 

846 

8 

128-184  FL 

351 

7 

140-187  FL 

Whitebait  smelt 

Allosmerus  elongatus 

946 

6 

60-114  FL 

79 

3 

76-132  FL 

Bathylagidae 

Popeye  blacksmelt 

Bathylagus  ochotensis 

1 

1 

76  SL 

Paralepidae 

Slender  barracudina 

Lestidium  ringens 

3 

1 

72-76  SL 

Myctophidae 

Northern  lampfish 

Stenobrachius  leucopsarus 

96 

4 

14-70  SL 

Bigfin  lanterfish 

Symbolophorus  californiensis 

61 

4 

89-102  SL 

Blue  laternfish 

Tarletonbeama  crenularis 

10 

3 

33-87  SL 

Gadidae 

Gadid(j) 

Gadidae 

10 

3 

42-58  SL 

13 

3 

53-57  SL 

Pacific  cod  1  j ) 

Gadus  macrocephalus 

23 

1 

38-60  SL 

Pacific  tomcod  ( j ) 

Microgadus  proximus 

6 

4 

35-55  SL 

8 

2 

49-80  SL 

Scomberesocidae 

Pacific  saury 

Cololabis  saira 

26 

1 

182-229  FL 

66 

6 

131-194  FL 

Atherinidae 

Jacksmelt 

Atherinopsis  californiensis 

1 

1 

302  FL 

Trachipteridae 

King-of-the-salmon  (j ) 

Trachipterus  altivelis 

2 

2 

71-270  SL 

12 

2 

40-83  SL 

Gasterosteidae 

Threespine  stickleback 

Gasterosteus  aculeatus 

1 

1 

60  SL 

Scorpaenidae 

Pacific  ocean  perch  (j ) 

Sebastes  alutus 

1 

1 

33  SL 

Darkblotched  rockfish  (j 

Sebastes  crameri 

154 

14 

29-54  SL 

1 

1 

53  SL 

Yellowtail  rockfish  (j) 

Sebastes  flavidus 

1350 

24 

20-63  SL 

1 

1 

18  SL 

Shortbelly  rockfish  (j ) 

Sebastes  jordani 

1 

1 

37  SL 

Black  rockfish  (j,a) 

Sebastes  melanops 

1 

1 

30  SL 

1 

1 

335  FL 

Bocaccio  (j ) 

Sebastes  paucispinis 

20 

5 

21-36  SL 

Canary  rockfish  (j ) 

Sebastes  pinniger 

27 

5 

22-39  SL 

Bank  rockfish  (j ) 

Sebastes  rufus 

8 

1 

16-28  SL 

Stripetail  rockfish  (j) 

Sebastes  saxicola 

13 

3 

32-37  SL 

Hexagrammidae 

Lingcod  (j) 

Ophiodon  elongatus 

20 

9 

76-81  FL 

Anoplopomatidae 

Sablefish  (j ) 

Anoplopoma  fimbria 

182 

14 

55-136  FL 

4 

2 

173-241  FL 
continued 

32 


Fishery  Bulletin  102(1) 


Table  1  (continued) 

Class  and  Family 

Common  name 

Scientific  name 

June  (84  stations) 

August  179  stations) 

Number 

F.O. 

Size  range 

Number 

F.O. 

Size  range 

Cottidae 

Irish  lord  Ij) 

Hemilepidotus  spp. 

2 

1 

38-40  FL 

Cabezon  (j ) 

Scorpeanichthys 
marmoratus 

12 

7 

33-38  SL 

Pacific  staghorn 
sculpin 

Leptocottus  armatus 

1 

1 

180  TL 

Agonidae 

Sturgeon  poacher  (j) 

Podothecus  acipenserinus 

1 

1 

80  TL 

Cyclopteridae 

Pacific  spiny 
lumpsucker 

Eumierotremus  orbis 

1 

1 

253  TL 

Carangidae 

Jack  mackerel 

Trachurus  symmetricus 

111 

3 

364-583  FL 

839 

20 

227-589  FL 

Bramidae 

Pacific  pomfret 

Brama  japonica 

5 

2 

387-434  FL 

Anarhichadidae 

Wolf-eel  (j) 

Anarrhichthys  ocellatus 

15 

13 

215-555  TL 

8 

7 

442-582  TL 

Ammodytidae 

Pacific  sandlance 

Ammodytes  hexapterus 

4 

4 

45-82  SL 

Zaprodidae 

Prowfish  (j) 

Zaprora  silenus 

1 

1 

68  SL 

Scombridae 

Chub  mackerel 

Scomber  japonicus 

74 

6 

266-421  FL 

Centrolophidae 

Medusafish 

Icichthys  lockingtoni 

3 

3 

37-50  SL 

8 

6 

87-129  FL 

Bothidae 

Sanddabs  (j) 

Citharichthys  spp. 

23 

13 

35-43  SL 

3 

2 

269-288  TL 

Pacific  sanddab  (j ) 

Citharichthys  sordidus 

32 

4 

32^4  SL 

Speckled  sanddab  (j ) 

Citharichthys  stigmaeus 

60 

10 

30-43  SL 

Pleuronectidae 

Dover  sole  (j) 

Microstomas  pacificus 

2 

2 

40-50  SL 

3 

1 

27-34  SL 

Sand  sole  (j) 

Psettichthys  melanostictus 

3 

3 

22-39  SL 

Slender  sole  (j) 

Eopsetta  exilis 

1 

1 

66  SL 

Starry  flounder 

Platichthys  stellatus 

2 

1 

349-399  TL 

Curlfin  sole  (j) 

Pleuronichthys  decurrens 

5 

3 

25-31  SL 

English  sole 

Parophrys  vetulus 

1 

1 

303  TL 

Rex  sole  (j ) 

Errex  zachirus 

581 

12 

34-79  SL 

48 

11 

44-70  SL 

Molidae 

Ocean  sunfish 

Mola  mola 

1 

1 

620  TL 

Total 

5974 

12,878 

about  one  half  of  the  nonjacks  caught  north  of  Cape  Blanco 
in  August  were  hatchery  fish. 

Two  factors,  however,  may  have  lead  to  inaccuracies  in 
estimation  of  hatchery-wild  ratios  of  coho  salmon  in  the 
GLOBEC  study  area.  First,  because  of  low  sample  sizes, 
the  data  were  pooled  from  both  June  and  August  catches 
for  the  genetic  stock  analysis;  therefore  we  do  not  know  the 
proportional  contributions  of  the  different  release  areas 
to  the  catches  in  either  month  alone.  Second,  all  the  fish 
released  from  Klamath  River  and  Trinity  River  hatcheries 
had  been  clipped  on  the  maxillary.  We  were  unaware  that 
the  maxillary  clip  was  being  used,  did  not  look  for  it,  and 
consequently  may  have  classified  fish  with  this  mark  as 
unmarked.  Therefore,  the  proportion  of  hatchery  fish  in 
the  catch  of  coho  salmon  during  GLOBEC  may  have  been 
higher  than  is  shown  in  Table  3. 

Age  and  growth  of  juvenile  coho  salmon 

Forty-three  percent  (24  of  56)  of  the  juvenile  coho  salmon 
caught  during  the  August  GLOBEC  cruise  were  preco- 


cious males  ("jacks")  according  to  the  testes-weight  to 
body-weight  criteria  of  Pearcy  and  Fisher  ( 1988).  This  is  a 
much  higher  percentage  of  jacks  than  found  among  juve- 
nile fish  caught  in  September  2000  in  the  plume  study  off 
Oregon  and  Washington,  where  only  4.5%  offish  (6  of  132) 
were  precocious  males  or  females  according  to  the  same 
criteria.  Because  the  jacks  were  considerably  larger  than 
the  nonjacks,  average  growth  rates  of  the  two  groups  were 
reported  separately. 

Estimated  average  growth  rates  in  FL  between  ocean 
entry  and  capture  were  higher  for  fish  caught  in  the 
August  2000  GLOBEC  cruises  (1.56-2.22  mm/d)  than 
for  fish  caught  in  any  other  cruises  (Table  3).  The  fish 
caught  in  August  2000  were  also  larger  when  they  entered 
the  ocean  (average  170- 178  mm  FL)  than  fish  caught  in 
other  cruises  (averagel54-160  mm  FL).  Average  growth 
rate  of  jacks  from  north  of  Cape  Blanco  (2.22  mm/d),  was 
significantly  higher  (/-test,  P<0.05)  than  growth  rates  of 
nonjacks  (1.56-1.67  mm/d).  Growth  rates  of  nonjacks  north 
and  south  of  Cape  Blanco  were  not  significantly  different  la- 
test, P<0.05).  The  combination  of  large  size  at  ocean  entry 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  juvenile  salmonids 


33 


45.0 


44.5 


44.0 


43.5 


43.0 


42.5 


42.0 


41.5 


Newport 


Chinook 

> 

0 

1  to 

5 

0 

6  to 

150 

Coho 

A 

1     10 

5 

A 

6  to 

150 
J 

Oregon 
r     California 


45.0 


44.5 


44.0 


43.5 


43.0 


42.5 


42.0 


41.5 


125.5 


125.0 


124.5 


124,0 


123.5    125.5 
Longitude  (W) 


125.0 


124.5 


124.0 


123.5 


Figure  3 

Catch  distribution  for  juvenile  coho  (Oncorhynchus  kisutch)  and  chinook  salmon  (O.  tshawytscha) 
for  the  (A)  June  and  (B)  August  cruise  overlaid  on  surface  temperature  contours.  Plus  signs  are 
stations  sampled  where  no  salmon  were  caught. 


and  favorable  conditions  for  growth  in  the  ocean  probably 
contributed  to  the  very  high  percentage  of  jack  coho  salmon 
in  August  2000  in  the  GLOBEC  study  area. 

Estimated  average  growth  rates  between  ocean  entry  and 
capture  of  juvenile  coho  salmon  were  higher  in  the  GLOBEC 
area  than  in  the  plume  study  area  U-tests,  P<0.05).  For  fish 
caught  in  June,  average  growth  rate  was  1.06  mm/d  and  0.63 
mm/d  in  the  GLOBEC  and  plume  study  areas,  respectively. 
For  fish  caught  in  August  or  September,  average  growth 
rate  was  1.57-2.22  mm/d  in  the  GLOBEC  study  area  and 
1.17  mm/d  plume  in  the  study  area  (Table  3).  The  higher 
growth  rates  of  coho  salmon  caught  in  the  GLOBEC  study 
area  suggests  that  in  2000  conditions  for  growth  were  bet- 
ter there  than  those  in  the  plume  study  area  farther  north 
off  Oregon  and  Washington.  Average  instantaneous  growth 
rates  in  weight  were  also  higher  (/-tests,  P<0.05)  for  the 
fish  caught  in  the  June  and  August  2000  GLOBEC  cruises 
(2.0  and  2.1-2.8%  body  wt/d,  respectively)  than  for  the  fish 
caught  in  the  June  and  September  2000  plume  study  cruises 
(1.2  and  1.7  %  body  wt/d,  respectively;  Table  4A). 

In  addition,  the  average  condition  index  (CI)  of  juve- 
nile coho  salmon  in  June  was  significantly  higher  (/-test, 


P=0.03)  in  the  GLOBEC  study  area  (1.12,  n=32,  SD=0.087) 
than  in  the  plume  study  area  (1.07,  n=245,  SD=0.117). 
Similarly,  the  average  CI  of  nonjack  juvenile  coho  salmon 
was  higher  (/-test,  P=0.002)  in  August  in  the  GLOBEC 
study  area  (1.24,  n=32,  SD=0.096)  than  in  September  in 
the  plume  study  area  (1.18,  n=132,  SD=0.100).  Both  the 
high  instantaneous  growth  rates  in  weight  and  the  high 
CI  of  juvenile  coho  salmon  caught  in  the  GLOBEC  study 
area  suggest  that  conditions  for  growth  of  coho  salmon  in 
this  area  were  very  good  in  2000.  Growth  rates  estimated 
from  the  few  CWT  fish  caught  during  these  cruises  (Table 
4B)  were  similar  to,  and  help  validate,  the  growth  rates 
estimated  from  scales  (Table  4A). 

Average  weights  at  time  of  ocean  entry  back-calculated 
from  scales  for  coho  salmon  caught  in  June  in  the  GLOBEC 
area  and  in  all  months  in  the  plume  study  area  (Table  4A) 
were  slightly  higher  than  the  average  weights  of  hatchery 
coho  salmon  at  time  of  release  (Appendix  Table  1).  For  ex- 
ample, in  the  plume  study  area,  average  back  calculated 
weights  at  ocean  entry  ranged  from  37.5  g  to  42.4  g  (Table 
4A) — slightly  higher  than  the  expected  average  weights 
at  release  of  about  32-33  g  based  on  the  stock  composi- 


34 


Fishery  Bulletin  102(1) 


Table  2 

Summary  of  mean,  standard  deviation,  and  range  of  FL  measured  in  the  field,  weight  measured  in  the  laboratory,  and  condition 
index  (CI)  of  subyearling  (age  0.0)  and  yearling  (age  1.0)  chinook  salmon  and  yearling  (age  1.0)  coho  salmon  caught  during  the  June 
and  August  cruises  north  (N)  and  south  (S)  of  Cape  Blanco  (latitude  42.837°).  Precocious  coho  salmon  are  indicated  with  a  "J". 

Field  FL  (mm) 

Laboratory 

weight  (g) 

C.I. 
(wtx  105/  FL3  ) 

n 

Mean 

SD 

Range 

Mean 

SD 

Range 

Mean 

SD 

Chinook  (age  0.0) 

June (N) 

1 

121 

— 

— 

18 

— 

— 

1.04 

— 

August  (N) 

1 

172 

— 

— 

70 

— 

— 

1.37 

— 

August  (S) 

125 

134 

12 

109-175 

28 

9 

12-70 

1.10 

0.08 

Chinook  (age  1.0) 

June  (N) 

27 

229 

42 

144-280 

178 

91 

33-306 

1.32 

0.10 

June(S) 

1 

174 

— 

— 

67 

— 

— 

1.28 

— 

August  (N) 

54 

229 

26 

187-318 

164 

72 

80-468 

1.32 

0.09 

August  (S) 

35 

231 

35 

190-349 

176 

94 

80-535 

1.32 

0.07 

Coho  (age  1.0) 

June  (N) 

30 

161 

33 

122-276 

56 

51 

19-292 

1.13 

0.08 

June (S) 

2 

172 

0 

172-172 

49 

1 

48-49 

0.95 

0.01 

August  (N-J) 

24 

365 

31 

310-415 

690 

209 

375-1198 

1.38 

0.12 

August  (N) 

24 

285 

51 

210-385 

326 

188 

97-766 

1.26 

0.10 

August  (S) 

8 

293 

33 

239-334 

308 

103 

157-433 

1.19 

0.05 

Table  3 

Catch,  percentage  of  the  catch  that  was  marked,  estimated  percentage  of  hatchery  origin,  size  of  scale  sample,  FL  at  ocean  entry 
(OE)  back  calculated  from  scales,  FL  at  capture,  and  estimated  growth  rate  in  FL  while  in  the  ocean  for  juvenile  coho  salmon 
caught  during  the  2000  GLOBEC  and  Columbia  River  plume  studies.  All  length  data  are  from  the  scale  sample  only.  An  ocean  entry 
date  of  15  May  was  used  when  calculating  growth  rate  in  FL. 

Cruise 

Catch 

(n) 

Marked 

Estimated  %        Scale  sample 
hatchery  origin              (n) 

Back- 
calculated  FL 
at  OE  (mm) 
mean  (SD) 

FL  at  capture 
(mm) 

Growth  rate 

(mm/d) 

mean  (SD) 

mean(SD) 

GLOBEC 

June  2000 

32 

32% 

33%                      11 

155  (29.0) 

177(42.3) 

1.06(1.01) 

Aug  2000 

North  of  C.Blanco 

Jacks 

24 

71% 

74%                       19 

170(22.8) 

370(28.1) 

2.22  (0.35) 

Nonjacks 

24 

46% 

48%                        9 

178(21.6) 

309(46.1) 

1.67  (0.51) 

South  of  C.  Blanco 

Nonjacks 

8 

38% 

39%                        6 

178(13.0) 

303  (29.3) 

1.56  (0.22) 

Plume  study 

May  2000 

165 

68% 

76-80%;               79 

157(16.5) 

166(17.7) 

0.97(1.15) 

Jun  2000 

245 

76% 

90%                      97 

160(14.5) 

185(23.4) 

0.63  (0.53) 

Sep  2000 

132 

65% 

75%                      76 

154(19.0) 

305  (24.9) 

1.17(0.23) 

'  No  genetic  stock  analysis  was  available.  The  higher  estimate  assumes  the  same  stock  composition  as  in  June, 
hatchery  fish  were  from  the  Columbia  River. 

the  lower  estimate 

assumes  that  all 

Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  |uvenile  salmonids 


35 


A 

0.004 

0002 

0 

-0.002 

-0.004 

□  Wet  Wt  (Somatic  Growth) 

to 

as     -0.006 

"D 

■  Dry  Wl  (Energetic  Growth) 

CO 

1)     0.02  -, 
o 

B 

0.01  - 

—L— H^H 

-0.01  - 

-0.02  - 

-0.03  - 
-0.04  - 
-0.05  - 

l 

-0.06  - 

-0.07  - 

Cape  Blanco                                       Cape  Blanco 

North                                                     South 

Figure  4 

Wet  and  dry  weight  residuals  (  +  1  standard  error)  for  (A)  yearling  chinook  (On- 

corhynchus    tshawytscha)  and  (B)  juvenile  coho  salmon  (O.  kisutch)  collected 

North  and  South  of  Cape  Blanco.  Weight  residuals  are  derived  from  the  linear 

relationship  between  fork  length  and  wet  or  dry  weight  (log-transformed  data) 

of  juvenile  salmon  captured  in  June  and  August. 

tion  of  these  catches  (Teel  et  al.,  2003)  and  the  release 
weights  (Appendix  Table  1).  Similarly,  the  back-calculated 
weight  at  ocean  entry  in  June  in  the  GLOBEC  area  (45.5  g) 
was  slightly  higher  than  the  expected  average  weight  at 
hatchery  release  (about  41  gl  based  on  the  stock  compo- 
sition (Table  5)  and  the  average  release  weights.  These 
fairly  small  differences  between  back-calculated  size  at 
ocean  entry  and  average  size  at  release  could  be  due  to 
growth  during  downstream  migration,  selectively  higher 


mortality  of  small  smolts,  or  a  bias  in  the  back-calculation 
procedure. 

However,  the  average  back-calculated  weights  at  time  of 
ocean  entry  offish  caught  in  August  in  the  GLOBEC  study 
area  (60-69  g)  were  over  two  standard  deviations  above  the 
average  weights  of  hatchery  fish  released  from  the  Oregon 
coast  or  northern  California — the  main  contributors  to  this 
catch  (Appendix  Table  1).  These  were  obviously  atypical 
coho  salmon,  and  the  very  high  proportion  of  jacks  (preco- 


36 


Fishery  Bulletin  102(1) 


Table  4 

(A)  Weights  at  ocean  entry 

I OE )  back-calculated  from  scales,  weights  at  capture 

and  estimated  instantaneous  rates 

of  growth  while 

in  the  ocean  iGl  for  juvenile  coho  salmon 

caught  during  the  2000  GLOBEC  and  Col 

umbia  River  plume  studies. 

An  ocean  entry 

date  of  15  May  was  used  when  calculating  growth  rate.  (B)  Similar  data  for  CWT  fish. 

Growth  rates  of  the  CWT  coho  salmon  were 

estimated  for  the  periods 

between  hatchery  release  and  capture  in  the  ocean. 

A    Cruise 

;; 

Back-calc.  Wt.  at  OE  (g) 

Weight  at  capture  (g) 

G 

mean (SD) 

mean(SD) 

mean  (SD) 

GLOBEC 

June  2000 

11 

45.5  (26.8) 

78.0(76.4) 

0.020(0.015) 

Aug  2000 

North  of  C. 

Blanco 

Jacks 

19 

68.9(27.2) 

719.7(200.0) 

0.028  (0.005) 

Nonjacks 

9 

59.5  (26.3) 

419.2(177.2) 

0.023  (0.006) 

South  of  C. 

Blanco 

Nonjacks 

6 

60.3(12.8) 

336.2  (96.2) 

0.021  (0.002) 

Plume  study 

May  2000 

79 

39.4  (10.8) 

47.9(14.6) 

0.020(0.024) 

Jun  2000 

97 

42.4(12.5) 

71.9(33.3) 

0.012(0.009) 

Sep  2000 

75 

37.5(13.7) 

347.2(158.3) 

0.017(0.003) 

B     Cruise 

n 

Wt.  at  release  (g) 

Wt.  at  capture  (g) 

G 

mean  (SD) 

mean  (SD) 

mean (SD) 

GLOBEC 

Jun  2000 

4 

44.4(1.3) 

86.6  (30.9) 

0.018(0.005) 

Aug  2000 

3 

35.6  (9.8) 

395.7(215.0) 

0.024(0.003) 

Plume  study 

Jun  2000 

11 

28.3(4.5) 

66.1(32.3) 

0.012(0.005) 

Sep  2000 

10 

33.4(10.91 

392.4(283.3) 

0.018(0.002) 

cious,  sexually  developed  males)  among  the  fish  was  prob- 
ably a  consequence  of  their  very  large  size  at  ocean  entry 
and  their  high  rates  of  growth  in  the  ocean. 

Freshwater  origins  of  juvenile  salmonids 

Allozyme  data  were  collected  from  samples  of  247  chinook 
salmon,  88  coho  salmon,  and  58  steelhead.  Genetic  mixed 
stock  analyses  indicated  that  chinook  salmon  in  June  were 
predominately  (54%,  SD=0.18)  from  rivers  and  hatcheries 
along  the  mid  Oregon  coast,  an  area  immediately  north  of 
Cape  Blanco  (Table  5,  Fig.  5).  In  August,  chinook  salmon 
were  largely  from  rivers  that  enter  the  sea  south  of  Cape 
Blanco.  Fish  from  the  Sacramento  and  San  Joaquin  rivers 
in  northern  California  were  estimated  to  comprise  90% 
(SD=0.07)  of  the  chinook  salmon  sampled  in  August  north 
of  Cape  Blanco.  The  largest  concentration  of  chinook 
salmon  we  sampled  was  south  of  Cape  Blanco  in  August, 
and  these  fish  were  mostly  from  rivers  in  southern  Oregon 
(539(,  SD=0.10)  and  the  Sacramento  and  San  Joaquin 
rivers  (20%,  SD=0.05).  Chinook  salmon  from  the  Colum- 
bia River  Basin  were  also  present,  but  were  estimated 


to  comprise  only  18%  (SD=0.15)  of  the  June  sample  and 
8%  (SD=0.05)  of  the  August  sample  north  of  Cape  Blanco. 
Recoveries  of  hatchery  chinook  salmon  bearing  coded-wire 
tags  (CWT)  provided  direct  evidence  of  stock  origins  for 
ten  fish,  all  taken  in  trawls  north  of  Cape  Blanco  (Table 
5).  These  data  reveal  that  hatchery  fish  released  from 
the  Umpqua  River  on  the  central  Oregon  coast  (;;=6), 
Columbia  River  Basin  («=3)  and  Sacramento  River  («  =  1) 
contributed  to  our  sample  of  chinook  salmon.  The  propor- 
tion of  CWT  fish  from  the  Umpqua  River  in  our  August 
catch  north  of  Cape  Blanco  (8%)  indicated  that  the  con- 
tribution of  mid  Oregon  coastal  fish  was  underestimated 
in  the  genetic  analysis  likely  because  of  the  small  size  of 
the  mixture  sample. 

Genetic  estimates  of  coho  salmon  indicated  that  most 
fish  originated  from  coastal  Oregon  rivers  north  of  Cape 
Blanco  (479S ,  SD=0.10)  and  from  the  Columbia  River  (13%, 
SD=0.08 )  (Table  5 ).  However,  a  substantial  proportion  (40r/i , 
SD=0.09)  of  coho  salmon  were  from  coastal  rivers  south  of 
Cape  Blanco,  a  region  that  includes  spawning  populations 
in  the  Rogue  and  Klamath  rivers.  Eight  coho  salmon  in 
our  sample  contained  CWTs  and  showed  that  fish  from 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  juvenile  salmonids 


37 


Table  5 

Estimated  percentage  stock  compositions,  samples  sizes,  and  recoveries 

of  coded  wire  tags  (CWTs)  for  chinook  and  coho  salmon  and 

steelhead  sampled  in  trawl  surveys  along  the  Oregon  and  California  coasts  in  2000 

Some  of  tht 

major  baseline  stocks  are  given  for 

coastal  stock  groups.  None  of  the  steelhead  sampled  contained  coded  wire  tags. 

June  (rc=35) 

August  (?!=157)            August  (n=55) 

Entire 

South  of 

North  of 

Study  Area 

Cape  Blanco               Cape  Blanco 

Chinook  salmon  stock  group                                                                    Est. 

SD       CWT       Est. 

SD      CWT       Est. 

SD     CWT 

Columbia  and  Snake  Rivers                                                                       18 

0.15 

2             3 

0.03                       8 

0.05        1 

North  Oregon  coast  (Nehalem,  Trask,  Alsea,  and  Siuslaw  Rivers)        0 

0.00 

0 

0.00                       0 

0.00 

Mid  Oregon  coast  (Umpqua,  Coquille,  Sixes,  and  Elk  Rivers)             54 

0.18 

3             3 

0.03                        1 

0.02        3 

South  Oregon  coast  (Rogue.  Chetco,  and  Winchuck  Rivers)                 26 

0.16 

53 

0.10                       0 

0.00 

Klamath  and  Trinity  Rivers                                                                         0 

0.00 

14 

0.07                       0 

0.00 

North  California  Coast  (Mad,  Eel,  and  Mattole  Rivers)                          2 

0.05 

7 

0.07                         1 

0.04 

Sacramento  and  San  Joaquin  Rivers                                                         0 

0.00 

20 

0.05                      90 

0.07        1 

June  and  August  (rc=88) 

Coho  salmon  stock  group 

Entire  study  area 

Est. 

SD      CWT 

Columbia  River 

13 

0.08        2 

North  and  Mid  Oregon  coast  (Nehalem,  Siletz,  Alsea,  Umpqua,  and  Coos  Rivers) 

47 

0.10         5 

Rogue  and  Klamath  Rivers 

40 

0.09         1 

North  California  Coast  (Mad,  Russian,  Little,  and  Scott  Rivers) 

0 

0.00 

June  and  August  (n=58) 

Steelhead  trout  stock  group 

Entire  study  area 

Est. 

SD 

Columbia  and  Snake  Rivers 

0 

0.00 

North  and  Mid  Oregon  coast  (Nehalem,  Siletz,  Alsea,  Umpqua,  Coos, 

and  Coquille  Rivers) 

1 

0.03 

South  Oregon  coast  (Elk,  Rogue,  Chetco,  and  Winchuck  Rivers) 

53 

0.08 

Smith,  Klamath,  and  Trinity  Rivers 

0 

0.00 

North  California  Coast  (Mad,  Eel,  and  Ten  Mile  Rivers) 

10 

0.05 

Sacramento  and  San  Joaquin  Rivers 

14 

0.05 

Central  and  South  California  Coast  (San  Lorenzo  River  and  Scott,  Pauma, 

and  Gaviota  Creeks 

3 

0.02 

Unknown 

19 

— 

hatcheries  in  the  Umpqua  River  (n=5),  Rogue  River  (n=l), 
and  Columbia  River  (n=2)  were  in  our  study  area. 

Genetic  analysis  of  steelhead  samples  showed  that  a 
large  proportion  were  from  the  Rogue  River  and  nearby 
coastal  streams  (53%,  SD=0.08).  Steelhead  from  the  Sacra- 
mento and  San  Joaquin  rivers  (14%,  SD=0.05)  and  north- 
ern California  coastal  rivers  (10%,  SD=0.05)  were  also 
present.  Estimates  for  steelhead  originating  from  rivers 
north  of  Cape  Blanco  and  from  south  of  the  San  Francisco 
Bay  were  near  zero.  Approximately  19%  of  the  steelhead 
mixture  was  not  allocated  to  any  source  population,  sug- 
gesting that  our  baseline  data  for  the  species  is  incomplete. 
No  steelhead  in  our  collections  contained  CWTs. 


Species  associations  of  juvenile  salmonids  and  other 
species 

From  cluster  analysis  of  species  based  on  station  assem- 
blages (Fig.  6),  MRPP  of  both  sample  periods  showed  strong 
within-group  agreement  (P<0.0001)  at  the  first  level  (two 
groups);  all  subsequent  groups  had  sequentially  higher 
levels  of  within-group  agreement.  As  a  result,  the  cutoff 
level  was  determined  by  balancing  a  lower  percent  infor- 
mation remaining  (<30%)  in  the  model  while  retaining  bio- 
logically meaningful  groups.  For  June  this  cutoff  resided  at 
the  second  level  (three  groups)  and  for  August,  at  the  third 
level  (four  groups ).  For  the  June  cruise,  all  salmonids  includ- 


38 


Fishery  Bulletin  102(1) 


1 
A 

I 

127° 

i 

122" 

1 

117"W 

—  50°N 

Vancouver    "~-~~ 
Island      ^fc--- 

B.C. 

- 

Pacific 

Ocean 

Olympic 
Peninsula 

Puge! 
^  Sound  ,  r"  r£*\    •/ 

-  46" 

Columbia  R 

-5sk^  "X Columbia  R 

L  wA 

J  Snake  R     _ 

—  42" 

N 

-38"               ^ 

Newport 

Cape  Blanco    /       ,-.-, 

/   •  Yj 

Crescent  City    vC7 

Eel  R.   /\\ 
I 

s"  Umpqua 
Rogue  R 

CU 

/    o>  r    /, 
3  )  A 

3  (  L- 

o 

\    ;o  V 

i 

R.                                     1 

OR 

V 

ID 

CA 
3arvJu 



1                  1                   1 

0       1 00     200  km 
I 

B 

1                1 

127-                    122- 

1 

117"W 

June 

-so-N    entire 

study  area 

.^^ 

- 

© 

f|°oo 

~46'  August 
north  of 
Cape  Blanco 

°  oo 

7\ 

-4-'  W 

i 

August 
south  of 

#•'-. 

i 

Cape  Blanco 

i 

# 

• 

0      100     200  km 

1                 1 

•• 

1             1 

—  50*  N 


c 


— r 

132" 


127° 


122° 


'46'June  and  August 
entire  study  area 


—  42" 


O       • 


N 


J_ 


_L 


J. 


Figure  5 

(A)  Map  of  study  area  and  location  of  GLOBEC  sampling  (hatching).  (B)  Stock  compositions  of  chinook  salmon  (Oncorhynchus 
tshawytscha).  Stock  groups  are  North  of  Columbia  River  (grey),  Columbia  River  Basin  (green),  north  Oregon  coast  (pink),  mid  Oregon 
coast  (yellow),  south  Oregon  coast  (dark  blue),  Klamath  River  Basin  (black),  north  California  coast  (light  blue),  and  Central  Valley 
(red).  (C)  Stock  compositions  of  coho  salmon  (O.  kisutch).  Stock  groups  are  Columbia  River  (green),  mid  and  north  Oregon  coast  (dark 
pink),  Rogue  and  Klamath  rivers  (blue),  and  north  California  coast  (orange). 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  juvenile  salmonids 


39 


June 


100 


Information  remaining  (%) 
75                                 50                                25 
H 1 1 1 H 


Coho 

Chinook  (a) 

Wolfeel 

Chinook  (j) 

Lmgcod 

Steelhead 

Sablefish 

Market  squid 

Whitebait  smelt 

Pacific  herring 

Surf  smelt 


Darkblotched  rockfish  — , 
Yellowtail  rockfish  — T~ 
Rex  sole  — 
Speckled  sanddab  — 


i 


August 


100 

r- 


Information  remaining  (%) 
75  50  25 

H 1 1 1 h 


i 


Coho  (a) 

Coho  (j) 

Chinook  (a) 

Chinook  (j) 

Surf  smelt 

Steelhead 

Medusafish 

Pacific  saury 

Wolfeel 

Osmeriid  (j) 

Blue  shark 

Northern  anchovy 

Rex  sole 

Chub  mackerel 

Pacific  sardine 

Jack  mackerel 

Figure  6 

Cluster  species  groupings  by  cruise.  The  dashed  lines  indicate  the  cutoff  levels  for  each 
cluster  group.  See  Table  1  for  scientific  names. 


i> 


ing  steelhead  were  classified  within  the  same  grouping  that 
included  several  pelagic  juvenile  taxa,  including  wolf-eel, 
lingcod,  and  sablefish  (Fig.  61.  Two  other  clusters  that  were 
not  associated  with  juvenile  salmon  included  a  southern 
inshore  group  consisting  of  market  squid.  Pacific  herring, 
and  two  species  of  smelt  and  an  offshore  northern  group 
consisting  primarily  of  juvenile  rockfish  and  rex  sole.  For 
the  August  cruise,  all  salmonid  juveniles  and  adults  again 
clustered  together  in  one  large  group  with  surf  smelt  and 
medusafish  ( Fig.  6 ).  The  remaining  three  groups  were  much 
smaller  and  consisted  primarily  of  offshore  pelagic  species. 
Cluster  analysis  of  stations  based  on  species  assem- 
blages, and  subsequent  examination  of  the  cutoff  level  us- 
ing MRPP,  resulted  in  three  groupings  from  both  sample 
periods  (Fig.  7).  MRPP  revealed  strong  within-group 
agreement  for  all  levels  (P<0.0001);  however,  delineation 
at  three  groups  was  based  on  maintaining  lower  percent  in- 
formation remaining  (<30%)  and  still  having  a  meaningful 


level  of  resolution.  There  was  some  measure  of  geographic 
separation  among  the  three  groups  (Fig.  7).  In  June,  group 
A  was  predominantly  inshore  and  mostly  in  the  southern 
half  of  the  sampling  area,  group  B  was  found  mainly  in 
the  middle  shelf  region  and  was  more  northern,  and  group 
C  was  found  predominantly  offshore.  In  August,  group  A 
consisted  of  only  three  stations,  all  south  of  Cape  Blanco, 
whereas  groups  B  and  C  both  spanned  the  entire  shelf  and 
offshore  region  and  had  no  particular  north-south  affin- 
ity (Fig.  7).  ISA  of  the  groups  from  both  sampling  periods 
showed  that  only  groups  A  and  C  had  indicator  species 
(Tables  6  and  7),  whereas  the  intermediate  groups  had 
none. 

Ordination  analyses  and  environmental  correlates 

NMS  ordination  of  the  June  sampling  period  (Fig.  8A) 
revealed  most  of  the  variance  in  the  data:  axes  1  and 


40 


Fishery  Bulletin  102(1) 


June 

2000        AAAAAn]            a 

44.5- 

□  A          * 

rw 

44.0- 

cPa    A    a 

n      nrriAAA 

d 

43.5- 

n  nriAA'fY/ 

• 

43.0- 

A  \ 

D 

AA      \    ( 

□□      ao/ 

duster  Groupings 

42.5- 

'  Group  A     O  , 
Group  B     A 

aA 

Group  C    [ 

42.0- 

OnA 

,  ,  ,  i  i  ,  ,  ,  , 

1 

42.0- 


August  2000  rr    D|    &A 


A 


125.5  125.0  124.5  124.0  1235      125.5  125.0  124.5  124.0  123.5 

Longitude  (W) 

Figure  7 

Map  showing  locations  of  cluster  station  groupings  by  cruise. 


Table  6 

Indicator  species  analysis  showing  indicator  values  for  dominant  pelagic  nekton  captu 
mean,  standard  deviations  (SD),  and  P- values  for  each  cluster  grouping.  Cluster  Group 
mined  to  be  indicators  of  that  group. 

•ed  in  pelagic  trawls  during  June  2000  and 
B  did  not  have  any  species  that  were  deter- 

Group 

Species 

Observed  indicator 
value  (IV) 

Indicator  value  IV  from  randomized 

groups 

P-value 

Mean 

SD 

A 

chinook  (age  0.0 1 

61.0 

15.7 

6.54 

<0.001 

A 

lingcod 

26.1 

12.6 

5.67 

0.024 

A 

Pacific  herring 

71.7 

12.8 

5.88 

<0.001 

A 

surf  smelt 

86.5 

11.8 

5.59 

<0.001 

A 

whitebait  smelt 

31.5 

10.4 

5.55 

0.007 

A 

market  squid 

50.8 

15.0 

6.20 

<0.001 

C 

darkblotched  rockfish 

66.8 

1 5  8 

6.31 

<0.001 

C 

rex  sole 

46.0 

15.0 

6.24 

0.002 

C 

sablefish 

31.1 

16.2 

6.32 

0.035 

C 

speckled  sanddab 

52.5 

13.4 

5.94 

0.001 

C 

yellowtail  rockfish 

98.8 

19.0 

6.30 

<0.001 

3  represented  31%  and  237f,  respectively  (stress=16.3). 
Temperature,  depth  and  salinity  best  explained  the  ordi- 
n;it  ion  of  stations,  representing  a  cross  shelf  gradient  from 
nearshore  high  levels  of  salinity  to  increasing  temperature 
and  depth  offshore.  Ordination  of  August  stations  (Fig. 
8B)  represented  42'  i  of  the  variance  in  the  data,  and  23% 


of  the  variance  was  loaded  on  axis  2  and  19%  on  axis  3 
(stress=19.4).  As  with  June,  salinity  increased  toward  the 
coast  and  temperature  and  depth  increased  off  the  shelf. 
The  groups  derived  from  the  cluster  analysis  tended  to 
group  together  in  multivariate  space,  with  the  exception 
of  group  B  in  the  June  cruise  (triangles  in  Fig.  8A). 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  |uvenile  salmonids 


41 


Table  7 

Indicator  Species  Analysis  showing  indicator  values  for  dominant  pelagic  nekton  captured  in  pelagic  trawls  during  August  2000 
and  mean,  standard  deviations  (SD),  and  P-values  for  each  cluster  grouping.  Cluster  Group  B  did  not  have  any  species  that  were 
determined  to  be  indicators  of  that  group. 

Group 

Species 

Observed  indicator 
value  (IV) 

Indicator  value  IV  from  randomized  groups 

P-value 

Mean 

SD 

A 

chinook  (age  1.0) 

76.5 

21.3 

11.18 

0.004 

A 

A 

chinook  (age  0.0) 
surf  smelt 

80.4 
97.9 

22.1 
12.4 

11.62 
8.21 

0.003 
<0.001 

C 

chub  mackerel 

33.3 

12.8 

8.88 

0.021 

C 

jack  mackerel 

73.7 

23.0 

11.86 

0.006 

Table  8 

Results  of  statistical  tests  for  habitat  associations  between  juvenile  salmon  and  environmental  or  station  variables  from  each 
cruise  in  2000.  Fish  marked  by  zeros  indicate  subyearlings  and  those  marked  with  one  indicate  yearlings.  Shown  are  the  P-levels 
for  5000  randomizations  of  the  cumulative  frequency  of  the  habitat  variable  and  the  proportion  of  the  standardized  salmon  catch 
associated  with  each  habitat  observation.  Results  are  based  on  the  Cramer  von-Mises  test  statistic  determined  from  binned  data 
for  depth  and  neuston  biomass.  Significance  values  <0.05  are  shown  in  boldface. 

Cruise 

Jun 


Aug 


Taxon  and  age 

Surface  temp. 

Surface  salinity 

1-m  chlorophyll 

Bottom  depth 

Neuston  biomass 

chinook  (age  1.0) 

0.30 

0.60 

0.13 

0.18 

0.13 

coho  (age  1.0) 

0.33 

0.48 

0.21 

0.17 

0.31 

chinook  (age  0.0) 

0.36 

0.25 

0.13 

0.35 

0.42 

chinook  (age  1.0) 

0.04 

<0.01 

<0.01 

0.02 

0.29 

coho  (age  1.0) 

0.68 

0.04 

0.07 

0.02 

0.45 

There  were  few  instances  where  the  habitat  associations 
of  juvenile  salmon  were  significantly  different  from  the 
distribution  of  environmental  variables  sampled  (Table  8). 
None  of  the  variables  were  significant  for  yearling  chinook 
and  coho  salmon  in  the  June  sampling  (no  subyearling 
salmon  were  caught  during  that  cruise).  In  August,  all 
the  variables  except  neuston  biomass  were  significant  for 
yearling  chinook  salmon.  These  fish  were  collected  at  cooler 
temperatures,  higher  salinities,  higher  chlorophyll-o  con- 
centrations, and  at  shallower  depths  than  have  been  typi- 
cally recorded  (Fig.  9).  Coho  salmonjuveniles  were  found  in 
higher  salinities  and  shallower  depths  than  at  the  sampled 
habitat  (Fig.  9).  These  results  correlated  with  the  capture 
of  juvenile  chinook  salmon  and  to  a  lesser  with  extent  coho 
salmon  at  nearshore  stations  in  the  upwelling  zone. 


Discussion 

Understanding  the  mechanisms  underlying  the  dynamics 
of  multispecies  communities  is  one  of  the  biggest  challenges 
in  ecology.  Most  communities  contain  many  interacting  spe- 
cies, each  of  which  is  likely  to  be  affected  by  multiple  biotic 
and  abiotic  factors.  In  order  to  effectively  characterize  a 


system,  we  need  to  differentiate  variability  resulting  from 
both  temporal  and  spatial  factors.  Our  observations  took 
place  during  two  time  periods  of  about  20  days  each  and 
thus  were  not  synoptic  "snapshots"  of  the  system.  Indeed, 
during  our  June  sampling,  conditions  changed  markedly 
from  the  beginning  to  the  end  of  the  cruise  because  of  the 
arrival  of  an  anomalous  major  southwest  storm  ( Batch- 
elder  et  al.,  2002),  which  likely  completely  altered  the 
hydrography  and  biology  of  the  system.  Thus,  short-term 
temporal  variability  may  obscure  patterns  observed  over 
the  spatial  scale  of  our  sampling. 

The  pelagic  nekton  community  sampled  during  these 
cruises  was  not  that  different  from  what  had  previously 
been  shown  for  purse  seine  and  trawling  collections  off 
the  coast  of  Oregon  and  Washington  ( Brodeur  and  Pearcy, 
1986;  Emmett  and  Brodeur,  2000;  Brodeur  et  al.,  2003). 
The  early  summer  nekton  community  was  dominated  by 
coastal  forage  fishes  such  as  smelts  and  Pacific  herring, 
but  also  comprised  juveniles  of  many  rockfish,  sculpin, 
and  flatfish  species.  These  winter-spring  spawning  species 
eventually  settle  out  to  demersal  habitats  sometime  in 
summer  (Shenker,  1988;  Doyle,  1992),  which  may  in  part 
explain  the  paucity  of  these  taxa  in  the  August  cruise.  In 
contrast,  the  August  nekton  community  consisted  of  large, 


42 


Fishery  Bulletin  102(1) 


highly  migratory  species  such  as  Pacific  sardines,  jack 
mackerel,  and  chub  mackerel.  Pacific  sardine,  which  was 
almost  completely  absent  from  the  system  in  the  1980s,  has 
undergone  a  substantial  resurgence  and  is  now  one  of  the 
most  abundant  species  off  the  coast  in  summer  (Brodeur 
et  al.,  2000;  Emmett  and  Brodeur,  2000;  McFarlane  and 
Beamish,  2001).  It  should  be  noted,  however,  that  some  of 
the  differences  between  cruises  could  be  accounted  for  by 
the  inclusion  of  substantially  more  offshore  stations  during 


A; 

aa 

REXS 

SPSD 

cr 

A 

A 

A 

3 

A 

U 

STHD                  t 

A 

MASO                 o 
WBSM 

DBRF 

°-,YTRF 
D 

cP 

SABF 

A 

Temperature 
Depth 

,           Salinityn  n 

LGCD 

-         <% 

PHER 

»* 

D 

d1 

CHIN1 
COHO 

A 
A 

* 

1 

Axisl  (r2=0.31) 


CO 
CO 
X 

< 


B 

A 

A 

*  A 

* 

REXS 

A 

D 

STHD 

A 

COHOA 

A 

D 

a 

coftoj 

D 

A 

a       D 

-Depth 

a    Salinity 

D 

Temperature 

CHINO 

0 

BLSH 

OSMJ    ° 

CO 

CD 

PSAR 

CHIN1 

O 

o 

a 

o    ^OEL 

a 

°0 

CO 

o 

Axis  2  (r2=0.23) 

Figure  8 

Nonmetric  multidimensional  scaling  (NMS)  ordination 
plot  of  stations  and  nekton  species  with  environmental 
parameters  from  June  (A)  and  August  (B)  2000  GLOBEC 
cruises.  Station  symbols  denote:  onshore  tO>.  mid-shelf 
!▲).  and  slope  (D)  groupings;  Species  abbreviations  denote 
the  following  taxa:  CHIN  0  (chinook,  age  0),  CHIN  1 
(chinook,  age  al.ll,  STHD  (steelhead  trout).  SUSM  (surf 
smelt),  PSAU  (Pacific  saury),  WOEL  (wolf-eel  juvenile), 
OSM  J  (osmerid  juvenile),  REXS  (rex  sole,  larval  i,  MEDF 
(medusafish ),  PSAR  (Pacific  sardine),  .JAMA  (jack  mack- 
erel), CHMA  (chub  mackerel),  NANC  (northern  anchovy). 
BLSH  (blue  shark). 


the  second  cruise.  Our  results  from  the  community  analy- 
ses suggest  that  juvenile  salmon  tend  to  co-occur  with  each 
other  and  with  a  variety  of  other  pelagic  nekton,  including 
adult  salmon,  and  that  this  spatial  overlap  varies  tempo- 
rally. Brodeur  et  al.  (2003),  in  analyzing  community  struc- 
ture based  on  previous  pelagic  sampling  data  off  Oregon 
and  Washington,  arrived  at  similar  results.  In  both  studies, 
the  geographic  boundaries  of  the  pelagic  assemblages  often 
overlap  and  are  not  as  distinct  as  demersal  assemblages. 
However,  the  pelagic  environment  is  much  more  spatially 
and  temporally  heterogeneous  than  the  demersal  environ- 
ment, and  many  of  the  species  examined  in  our  study  are 
highly  mobile  and  are  likely  to  respond  quickly  to  changing 
conditions.  Research  is  presently  underway  to  examine  the 
trophic  interactions  among  salmonids  and  with  other  sym- 
patric  nekton  species  in  order  to  determine  what  ecological 
relationships  (e.g.  predation,  competition),  if  any,  are  occur- 
ring in  this  system. 

From  the  results  of  our  sampling,  we  concluded  that  ju- 
venile salmonids,  with  the  possible  exception  of  steelhead, 
occupy  the  cool,  high  salinity,  inshore  upwelling  regions  off 
the  southern  Oregon  coast.  However,  particularly  for  the 
June  cruise,  many  of  the  coho  and  chinook  salmon  juveniles 
collected  may  have  recently  entered  the  ocean  with  little 
time  to  disperse  offshore,  so  that  the  capture  location  may 
not  reflect  true  habitat  preferences.  Moreover,  the  vertical 
dimensions  of  our  trawl  also  precluded  us  from  sampling 
the  nearshore,  subtidal  regions  where  some  subyearling 
chinook  may  reside  shortly  after  entering  the  ocean. 

Salmon  and  steelhead  differed  considerably  in  stock  com- 
position. The  pattern  for  coho  salmon  was  similar  to  that 
of  chinook  salmon  in  that  fish  from  sources  both  north  and 
south  of  Cape  Blanco  contributed  to  our  catches.  However, 
steelhead  from  rivers  north  of  Cape  Blanco  were  absent, 
presumably  having  migrated  offshore  and  north  shortly 
after  entering  the  sea,  as  shown  by  Pearcy  et  al.  (1990). 
Although  our  stock  composition  estimates  for  steelhead 
should  be  viewed  with  caution  because  of  an  incomplete  ge- 
netic baseline  and  a  relatively  small  number  of  samples,  our 
findings  support  Pearcy  et  al.'s  suggestion  that  steelhead 
from  rivers  south  of  Cape  Blanco  have  a  unique  marine 
distribution  and  reside  throughout  the  summer  in  the  up- 
welling  zone  off  northern  California  and  southern  Oregon. 

Our  study  revealed  seasonal  shifts  in  the  abundance  and 
stock  composition  of  juvenile  salmonids.  Although  salmo- 
nids comprised  small  portions  of  the  vertebrate  catches  of 
both  the  June  and  August  cruises,  juvenile  chinook  salmon, 
coho  salmon,  and  steelhead  were  much  more  abundant 
later  in  the  summer,  likely  indicating  that  fish  moving 
into  our  study  area  are  from  shoreline  or  riverine  habitats. 
The  greater  abundance  of  chinook  salmon  in  late  summer 
can  be  explained  in  part  by  the  northern  migration  offish 
that  originated  in  rivers  south  of  our  study  area.  Chinook 
salmon  from  the  Sacramento  and  San  Joaquin  rivers  in 
California's  Central  Valley  comprised  substantial  propor- 
tions in  the  August  catches  both  south  (20%)  and  in  nth 
i 90'  i  )  of  Cape  Blanco.  In  contrast,  the  few  chinook  salmon 
caught  in  June  were  mostly  (549r )  from  streams  that  en- 
ter the  sea  immediately  north  of  Cape  Blanco  such  as  the 
Umpqua,  Coquille,  Sixes,  and  Elk  rivers.  Chinook  salmon 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  juvenile  salmonids 


43 


E 

o 


•Chinook  1  0 
Coho  1 .0 
-Habitat 


12  14 

Water  temperature  (C) 


j  ■*" 

09  - 

~~, r. ..---' 

OR  - 

,  * 

0 

07  - 

/                 ■" 

.    .. 

06  - 

r- ' 

05  - 

r' 

04  - 

03  ■ 

-    -    -Chinook  1  0 

02  - 

f         J" 

Coho  1  0 

n  1  - 

1    V 

Habitat 

n  - 

10  15 

Chla  concentration 


1 

09 
08 
07 
06 
0.5 
04 
03 
0.2 
0  1 


0 
31.50 


J     i 

-   -   -Chinook  1  0 

X   ' 

Coho  1.0 

Habitat 

Y                                     1 

^     } 

#  > 

,' 

4 

ja      _    _      _      J 

3250  3300 

Salinity  (PSU) 


•Chinook  1  0 
-Coho  1.0 
-Habitat 


100  150  200 

Water  depth  (m) 


Figure  9 

Cumulative  distribution  curves  for  salmon  and  environmental  or  station  variables.  Only  the  August  variables  that  showed  at  least 
one  significant  difference  are  included.  See  Table  8  for  results  of  the  statistical  tests. 


from  these  rivers  are  known  to  primarily  migrate  north 
of  our  study  area  along  the  coast  (Nicholas  and  Hankin, 
1988).  By  August,  fish  from  these  stocks  were  nearly  absent 
from  our  samples.  Oregon  rivers  south  of  Cape  Blanco,  an 
area  that  includes  the  Rogue,  Chetco,  and  Winchuck  riv- 
ers, produce  chinook  salmon  with  a  more  southerly  pattern 
of  ocean  migration  (Nicholas  and  Hankin,  1988;  Myers  et 
al.,  1998).  Chinook  salmon  from  these  rivers  were  found 
throughout  the  summer  and  contributed  53%  to  our  largest 
catches  of  chinook  salmon  along  transects  south  of  Cape 
Blanco  in  August. 

Results  from  our  2000  GLOBEC  cruises  identified  Cape 
Blanco  as  an  important  breakpoint  in  salmonid  life-his- 
tory variation.  Stock  distributions  of  both  juvenile  salmon 
and  steelhead  indicated  that  different  migration  patterns 
of  fish  originating  from  southern  and  northern  rivers  are 
evident  during  their  early  marine  phase.  Our  August  sur- 
vey also  revealed  sharp  contrasts  in  life-history  type  and 
freshwater  origin  between  the  juvenile  chinook  salmon 
population  in  the  marine  area  north  of  Cape  Blanco  and 


that  to  the  south.  Chinook  salmon  captured  north  of  Cape 
Blanco  were  nearly  all  yearlings  and  largely  from  the  Sac- 
ramento River  drainage.  Subyearlings  predominated  in  our 
catches  south  of  Cape  Blanco  and  included  a  much  larger 
proportion  offish  from  coastal  streams  in  southern  Oregon 
and  northern  California. 

Comparisons  of  our  results  with  similar  studies  conduct- 
ed further  north  show  differences  in  salmonid  migrations 
on  a  somewhat  broader  geographic  scale.  In  several  years  of 
sampling  during  the  summers  of  1981  through  1985  off  the 
central  Oregon  to  northern  Washington  coast,  most  juvenile 
chinook  salmon  bearing  CWTs  were  from  Columbia  River 
hatcheries  (Pearcy  and  Fisher,  1990;  Fisher  and  Pearcy, 
1995).  Only  one  tagged  chinook  salmon  from  a  river  south 
of  Cape  Blanco  (Klamath  River)  was  captured.  Pearcy  and 
Fisher  also  found  that  juvenile  coho  salmon  were  largely 
from  the  Columbia  River  and  that  smaller  contributions 
were  from  coastal  rivers  north  of  Cape  Blanco.  Their  find- 
ings have  been  corroborated  by  more  recent  surveys  in  the 
same  region  (Emmett  and  Brodeur,  2000)  using  genetic 


44 


Fishery  Bulletin  102(1) 


data  (Teel  et  al.,  2003).  Samples  from  subsequent  cruises 
will  be  used  to  examine  the  persistence  of  such  fine-  and 
broad-scale  geographic  structure  in  the  juvenile  migrations 
of  salmonid  stocks. 

A  major  source  of  error  in  our  estimates  of  growth  rates 
of  juvenile  coho  salmon  back-calculated  from  scales  was 
uncertainty  of  when  individual  fish  entered  the  ocean.  We 
used  a  single  date  of  ocean  entry  for  all  fish  (15  May),  but 
individual  fish,  of  course,  entered  the  ocean  at  different 
times  over  the  course  of  a  month  or  more.  Consequently, 
coefficients  of  variation  were  relatively  large  (84—119%  and 
75-120%  of  mean  growth  rate  in  FL  and  weight,  respec- 
tively) for  fish  caught  in  May  and  June,  when  errors  in  es- 
timated growth  periods  likely  were  large  in  relation  to  the 
actual  growth  periods.  Conversely,  coefficients  of  variation 
were  relatively  small  ( 14-30%  and  10-26%  of  growth  rate 
in  FL  and  weight,  respectively)  for  fish  caught  in  August  or 
September,  when  errors  in  estimated  growth  periods  likely 
were  small  in  relation  to  the  actual  growth  periods.  (Note 
the  decrease  in  standard  deviation  of  mean  growth  rates 
with  month  of  capture  in  Tables  3  and  4A).  Growth  rates 
of  CWT  coho  salmon  between  hatchery  release  and  capture 
in  the  ocean  (Table  4B)  were  very  similar  to  the  growth 
rates  of  unmarked  salmon  estimated  from  scales  for  the 
same  months  and  areas.  In  addition,  the  growth  rates  of 
the  former  group  ( CWT  coho  salmon )  helped  to  validate  the 
growth  rates  of  the  latter  group  (Table  4A). 

Significant  differences  in  growth  and  condition  of  ju- 
venile coho  salmon  indicate  that  different  oceanographic 
environments  exist  north  and  south  of  Cape  Blanco.  The 
length  of  the  fish  indicated  that  substantial  growth  oc- 
curred in  juvenile  coho  salmon  during  the  study  period.  As- 
sessment of  other  growth  features  (condition)  revealed  that 
juvenile  coho  salmon  grew  better  north  of  Cape  Blanco. 
Because  we  included  measurement  of  condition  in  both  the 
June  and  August  period  in  the  evaluation,  changes  in  stock 
composition,  described  earlier,  may  be  partly  responsible 
for  this  observation.  Although  genetic  stock  composition 
was  different  between  months,  month  of  sampling  was  not 
a  significant  factor,  suggesting  that  stock  composition  is 
not  likely  a  significant  factor  affecting  the  difference  in 
condition  (a  performance  metric)  of  juvenile  salmon  north 
and  south  of  Cape  Blanco. 

Several  lines  of  evidence  further  support  the  hypothesis 
that  areas  north  of  Cape  Blanco  benefit  juvenile  yearling 
chinook  and  coho  salmon.  There  were  greater  numbers  of 
juvenile  yearling  chinook  and  coho  salmon  to  the  north  of 
Cape  Blanco.  Although  our  overall  sampling  effort  was 
greater  north  of  Cape  Blanco,  in  the  mesoscale  portion  of 
our  survey  designed  to  assess  general  distribution  patterns, 
more  yearling  chinook  and  coho  salmon  were  captured 
north  of  Cape  Blanco.  Secondly,  when  we  evaluated  the 
growth  rate  of  juvenile  coho  salmon  in  the  GLOBEC  region 
compared  to  juveniles  captured  off  northern  Oregon  and 
Washington,  juveniles  from  the  GLOBEC  region  grew  much 
better.  The  similar  tracking  of  resource  (distribution  and 
abundance)  and  performance  (measured  in  terms  of  either 
somatic  and  energetic  growth  or  growth  rate)  metrics  for 
juvenile  yearling  chinook  salmon  and  coho  salmon  ninth 
of  Cape  Blanco  suggests  that  habitat  quality  in  this  region 


was  better.  The  results  of  this  study  help  define  the  biogeo- 
graphical  zones  for  salmon  growth  and  establish  regional- 
based  management  strategies  for  depleted  salmon  stocks. 


Acknowledgments 

We  thank  the  captain  and  crew  of  the  FV  Sea  Eagle  for  their 
expert  help  in  conducting  the  trawling  operations  under 
sometimes  adverse  weather  conditions.  We  are  grateful 
to  Jackie  Popp-Noskov,  Paul  Bentley,  Marcia  House,  and 
Becky  Baldwin  for  assistance  in  field  sampling.  Donald 
Van  Doornik  and  David  Kuligowski  collected  the  genetic 
data.  We  thank  Anne  Marshall  for  the  use  of  unpublished 
chinook  salmon  allele  frequency  data.  Stephen  Smith 
and  Alex  De  Robertis  helped  with  the  statistical  analy- 
sis. Earlier  versions  of  this  manuscript  were  improved 
by  the  helpful  comments  of  two  anonymous  journal 
reviewers.  Research  was  conducted  as  part  of  the 
U.S.  GLOBEC  program  and  was  jointly  funded  by  the 
National  Science  Foundation  (Grant  no.  OCE-0002855) 
and  the  National  Oceanic  and  Atmospheric  Administra- 
tion (NOAA).  We  also  acknowledge  the  Bonneville  Power 
Administration  for  funding  the  plume  study. 


Literature  cited 

Aebersold,  P.  B.,  G.  A.  Winans,  D.  J.  Teel.  G.  B.  Milner.  and 
F.  M.  Utter. 

1987.     Manual  for  starch  gel  electrophoresis:  a  method  for 
the  detection  of  genetic  variation.     NOAA  Tech.  Rep.  NMFS 
61,  19  p. 
Batchelder,  H.  P.,  J.  A.Barth,  M.  P.  Kosro,  P.  T.  Strub, 
R.  D.  Brodeur,  W.  T.  Peterson,  C.  T.  Tynan.  M.  D.  Ohman, 
L.  W.  Botsford,  T.  M.  Powell,  F.  B.  Schwing,  D.  G.  Ainley. 
D.  L.  Mackas,  B.  M.  Hickey,  and  S.  R.  Ramp. 

2002.  The  GLOBEC  Northeast  Pacific  California  Current 
System  program.     Oceanogr.  15:36-47. 

Barth,  J.  A.,  S.  D.  Pierce,  and  R.  L.  Smith. 

2000.    A  separating  coastal  upwelling  jet  at  Cape  Blanco. 
Oregon  and  its  connection  to  the  California  Current  System. 
Deep-Sea  Res.  47(part  II):783-810. 
Bradford,  M.  J. 

1995.     Comparative  review  of  Pacific  salmon  survival  rates. 
Can.  J.  Fish.  Aquat.  Sci.  52:1327-1338. 
Brodeur,  R.  D..  P.  J.  Bentley,  R.  L.  Emmett.  T.  Miller,  and 
W.  T  Peterson. 

2000.     Sardines  in  the  ecosystem  of  the  Pacific  Northwest. 
Proc.  sardine  2000  symposium.     Pacific  States  Marine  Fish. 
Comm.,  p.  90-102.     [Available  from  Pacific  States  Marine 
Fisheries  Commission,  Portland,  OR. I 
Brodeur,  R.  D.,  J  .P.  Fisher.  Y.  Ueno,  K.  Nagasawa,  and 
W.  G.  Pearcy. 

2003.  An  east-west  comparison  of  the  Transition  Zone 
coastal  pelagic  nekton  of  the  North  Pacific  Ocean.  J. 
Oceanog.  59(41:415-434. 

Brodeur,  R.  D.,  H.  V.  Lorz,  and  W.  G.  Pearcy. 

1987.     Food  habits  and  dietary  variability  of  pelagic  nekton 
off  Oregon  and  Washington.  1979-1984.     NOAA  Tech.  Rep. 
NMFS  57,  32  p. 
Brodeur.  R.  D.,  and  W.  G.  Pearcy. 

1986.     Distribution  and  relative  abundance  of  pelagic  non- 


Brodeur  et  al.:  Distribution,  growth,  condition,  origin,  and  associations  of  juvenile  salmonids 


45 


salmonid  nekton  off  Oregon  and  Washington,  1979-1984. 
NOAA  Tech.  Rep.  NMFS  46,  85  p. 
1992.     Effects  of  environmental  variability  on  trophic  inter- 
actions and  food  web  structure  in  a  pelagic  upwelling 
ecosystem.    Mar.  Ecol.  Prog.  Sen  84:101-119. 
Busby,  P.  J.,  T.  C.  Wainwright,  G.  J.  Bryant,  L.  Lierheimer, 
R.  S.  Waples,  F.  W.  Waknitz,  and  I.  V.  Lagomarsino. 

1996.     Status  review  of  west  coast  steelhead  from  Washing- 
ton, Idaho,  Oregon,  and  California.     NOAA  Tech.  Memo. 
NMFS-NWFSC-27,  261  p. 
Dawley,  E.  M.,  R.  D.  Ledgerwood,  and  A.  Jensen. 

1985.     Beach  and  purse  seine  sampling  of  juvenile  salmonids 

in  the  Columbia  River  estuary  and  ocean  plume,  1977-1983. 

Vol.  1,  Procedures,  sampling  effort,  and  catch  data.     NOAA 

Tech.  Memo.  NMFS  F/NWC-74,  260  p. 

Debevec,  E.  M.,  R.  B.  Gates,  M.  Masuda.  J.  Pella,  J.  Reynolds,  and 

L.  W.  Seeb. 

2000.     SPAM  (Version  3.2):  Statistics  program  for  analyzing 
mixtures.    J.  Hered.  91:509-511. 
Doyle,  M.  J. 

1992.  Neustonic  ichthyoplankton  in  the  northern  region  of 
the  California  Current  ecosystem.  Calif.  Coop.  Oceanic 
Fish.  Invest.  Rep.  33:141-161. 

Emmett,  R.  L.,  and  R.  D.  Brodeur. 

2000.  The  relationship  between  recent  changes  in  the 
pelagic  nekton  community  off  Oregon  and  Washington 
and  physical  oceanographic  conditions.  Bull.  North  Pac. 
Anadr.  Fish.  Comm.  2:11-20. 

Fisher,  J.  P.,  and  W.  G.  Pearcy. 

1988.  Growth  of  juvenile  coho  salmon  (Oncorhynchus 
kisutch)  off  Oregon  and  Washington,  USA  in  years  of  dif- 
fering coastal  upwelling.  Can.  J.  Fish.  Aquat.  Sci.  45: 
1036-1044. 

1995.  Distribution,  migration,  and  growth  of  juvenile  Chi- 
nook salmon,  Oncohychus  tshawytscha,  off  Oregon  and 
Washington.     Fish.  Bull.  93:274-289. 

Francis,  R.  C,  and  S.  R.  Hare. 

1994.  Decadal-scale  regime  shifts  in  large  marine  ecosys- 
tems of  the  Northeast  Pacific:  a  case  for  historical  science. 
Fish.  Oceanogr.  3:279-291. 

Friedland,  K.  D.,  and  R.  E.  Haas. 

1996.  Marine  post-smolt  growth  and  age  at  maturity  of 
Atlantic  salmon.     J.  Fish  Biol.  48:1-15. 

Gauch,  H.  G.Jr. 

1982.     Multivariate  analysis  in  community  ecology.     Cam- 
bridge Univ.  Press,  N.Y.  298  p. 
Hare,  S.  R.,  N.  J.  Mantua,  and  R.  C.  Francis. 

1999.     Inverse  production  regimes:  Alaska  and  West  Coast 
salmon.     Fisheries  24:6-14. 
Hinch,  S.  G,  M.  C.  Healey,  R.  E.  Diewert,  K.  A.Thomson, 
R.  Hourston,  M.  A.  Henderson,  and  F.  Juanes. 

1995.  Potential  effects  of  climate  change  on  marine  growth 
and  survival  of  Fraser  River  sockeye  salmon.  Can.  J.  Fish. 
Aquat.  Sci.  52:2651-2659. 

Jakob,  E.  M.,  S.  D.  Marshall,  and  G.  W.  Uetz. 

1996.  Estimating  fitness:  a  comparison  of  body  condition 
indices.     Oikos  77:  61-67. 

Kope,  R..  and  T.  Wainwright. 

1998.    Trends  in  the  status  of  Pacific  salmon  populations  in 
Washington,  Oregon,  California,  and  Idaho.     North  Pac. 
Anadr.  Fish.  Comm.  Bull.  1:1-12. 
Kruskal,  J.  B. 

1964.     Nonmetric  multidimensional  scaling:  a  numerical 
method.     Psychometricka  29:115-129. 
Lawson,  P.  W. 

1993.  Cycles  in  ocean  productivity,  trends  in  habitat  quality, 


and  the  restoration  of  salmon  runs  in  Oregon.     Fisheries 
18:6-10. 
Lorenzen,  K. 

1996.  The  relationship  between  body  weight  and  natural 
mortality  in  juvenile  and  adult  fish:  A  comparison  of  natural 
ecosystems  and  aquaculture.     J.  Fish  Biol.  49:  627-647. 

Mantua,  N.  J.,  S.  R.  Hare,  Y.  Zhang,  J.  M.  Wallace,  and 
R.  C.  Francis. 

1997.  A  Pacific  interdecadal  oscillation  with  impacts 
on  salmon  production.  Bull.  Am.  Meteorol.  Soc.  78: 
1069-1079. 

Marschall,  E.  A.,  and  L.  B.  Crowder. 

1995.  Density-dependent  survival  as  a  function  of  size  in 
juvenile  salmonids  in  streams.  Can.  J.  Fish.  Aquat.  Sci. 
52:136-140. 

McCune,  B„  and  J.  B.  Grace. 

2002.    Analysis  of  ecological  communities.  300  p.     MjM  Soft- 
ware Design,  Gleneden  Beach,  OR. 
McCune,  B.,  and  M.  J.  Mefford. 

1999.     PC-ORD,  multivariate  analysis  of  ecological  data, 
users  guide,  237  p.     MjM  Software  Design,  Gleneden 
Beach,  OR. 
McFarlane,  G  A.,  and  R.  J.  Beamish. 

2001.     The  re-occurrence  of  sardines  off  British  Columbia 
characterises  the  dynamic  nature  of  regimes.     Prog. 
Oceanogr.  49:151-165. 
McGowan,  J.  A..  D.  B.  Chelton,  and  A.  Conversi. 

1999.  Plankton  patterns,  climate,  and  change  in  the  Cali- 
fornia Current.  In  Large  marine  ecosystems  of  the  Pacific 
Rim:  assessment,  sustainability.  and  management  (K. 
Sherman  and  Q.  Tang,  eds.),  p.  63-105.  Blackwell  Sci- 
ence, Maiden,  MA. 
McGurk,  M.  D. 

1996.  Allometry  of  marine  mortality  of  Pacifc  salmon.  Fish. 
Bull.  94:77-88. 

Milner,  G  B.,  D.  J.  Teel,  F.  M.  Utter,  and  G.  A.  Winans. 

1985.    A  genetic  method  of  stock  identification  in  mixed 
populations  of  Pacific  salmon,  Oncorhynchus  spp.     Mar. 
Fish.  Rev.  47:1-8. 
Myers,  J.  M.,  R.  G.  Kope.  G  J.  Bryant,  D.  J.  Teel,  L.  J.  Lierheimer, 
T  C.  Wainwright,  W  S.  Grant,  F.  W.  Waknitz,  K.  Neely, 
S.  T  Lindley,  and  R.  S.  Waples. 

1998.  Status  review  of  chinook  salmon  in  Washington, 
Idaho,  Oregon,  and  California.  NOAA  Tech.  Memo.  NMFS- 
NWFCS-35,  443  p. 

Nehlsen,  W,  J.  E.  Williams,  and  J.  A.  Lichatowich. 

1991.  Pacific  salmon  at  the  crossroads:  stocks  at  risk  from 
California.  Oregon.  Idaho,  and  Washington.  Fisheries  16: 
4-21. 

Nicholas,  J.  W.,  and  D.  G.  Hankin. 

1988.  Chinook  salmon  populations  in  Oregon  coastal  river 
basins:  Description  of  life  histories  and  assessment  of  recent 
trends  in  run  strengths.  Oregon  Department  of  Fish  and 
Wildlife,  Fish  Division  Information  Report  88-1,  359  p. 

Paul.  A.  J.,  and  T.  M.  Willette. 

1997.  Geographical  variation  in  somatic  energy  content 
of  migrating  pink  salmon  fry  from  PWS:  a  tool  to  mea- 
sure nutritional  status.  In  Proceedings  of  the  interna- 
tional symposium  on  the  role  of  forage  fishes  in  marine 
ecosystems.  Alaska  Sea  Grant  College  Program,  Report 
97-01,  p.  707-720.     Univ.  Alaska,  Fairbanks,  AK. 

Pearcy.  W.  G 

1992.  Ocean  ecology  of  North  Pacific  salmonids,  179 
p.     Univ.  Washington  Press,  Seattle,  WA. 

Pearcy,  W  G,  R.  D.  Brodeur,  and  J.  P.  Fisher. 

1990.     Distribution  and  biology  of  juvenile  cutthroat  trout 


46 


Fishery  Bulletin  102(1) 


(Oncorhynchus  clarki  clarki)  and  steelhead  10.  mykiss)  in 
coastal  waters  off  Oregon  and  Washington.     Fish.  Bull.  88: 
697-711. 
Pearcy,  W.  G..  and  J.  P.  Fisher. 

1988.  Migrations  of  coho  salmon,  Oncorhynchus  kisutch, 
during  their  first  summer  in  the  ocean.  Fish.  Bull.  86: 
173-195. 
1990.  Distribution  and  abundance  of  juvenile  salmonids  off 
Oregon  and  Washington,  1981-1985.  NOAA  Tech.  Rep. 
NMFS  93:1-83. 
Pella,  J.,  and  G.  B.  Milner. 

1987.     Use  of  genetic  marks  in  stock  composition  analyses. 
In  Population  genetics  and  fishery  management,  (N.  Ryman 
and  F.  Utter,  eds.),  p.  247-276.     Washington  Sea  Grant, 
Univ.  Washington  Press,  Seattle,  WA. 
Perry,  R.  I..  B.  Hargreaves,  B.  J.  Waddell,  and  D.  L.  Mackas. 

1996.     Spatial  variations  in  feeding  and  condition  of  juve- 
nile pink  and  chum  salmon  off  Vancouver  Island,  British 
Columbia.     Fish.  Oceanogr.  5:73-88. 
Perry,  R.  I.,  and  S.  J.  Smith. 

1994.     Identifying  habitat  associations  of  marine  fishes  using 
survey  data:  an  application  to  the  Northwest  Atlantic.     Can. 
J.  Fish.  Aquat.  Sci.  51:589-602. 
Peterson,  W.  T.,  and  J.  E.  Keister. 

2002.     The  effect  of  a  large  cape  on  distribution  patterns 
of  coastal  and  oceanic  copepods  off  Oregon  and  northern 
California  during  the  1998-1999  El  Nino-La  Nina.     Prog. 
Oceanogr.  53:389-411. 
Post.  J.  R.,  and  E.  A.  Parkinson. 

2001.     Energy  allocation  strategy  in  young  fish:  allometry 
and  survival.     Ecology  82:1040-1051. 


Ricker.W  E. 

1992.     Back-calculation  of  fish  lengths  based  on  proportion- 
ality between  scale  and  length  increments.     Can.  J.  Fish. 
Aquat.  Sci.  49:1018-1026. 
Shenker,  J.  M. 

1988.     Oceanographic  associations  of  neustonic  larval  and 
juvenile  fishes  and  Dungeness  crab  megalopae  off  Oregon. 
Fish.  Bull.  86:299-317. 
Sutton,  S.  G,  T.  P.  Bult,  and  R.  L.  Haedrich. 

2000.     Relationship  among  fat  weight,  body  weight,  water 
weight,  and  condition  factors  in  wild  Atlantic  salmon  parr. 
Trans.  Am.  Fish.  Soc.  129:527-538. 
Syrjala,  S.  E. 

1996.    A  statistical  test  for  a  difference  between  the  spatial 
distributions  of  two  populations.     Ecology  77:75-80. 
Teel,  D.  J.,  D.  M.  Van  Doornik,  D.  R.  Kuligowski,  and  W.  S.  Grant. 
2003.     Genetic  analysis  of  juvenile  coho  salmon  [Oncorhyn- 
chus kisutch)  off  Oregon  and  Washington  reveals  few 
Columbia  River  wild  fish.     Fish.  Bull.  101:640-652. 
U.S.  GLOBEC. 

1994.  A  science  plan  for  the  California  Current.  U.S. 
GLOBEC  Rep.  11.  134  p.    Univ.  California,  Berkeley,  CA. 

Weitkamp,  L.,  and  K.  Neely. 

2002.     Coho  salmon  [Oncorhynchus  kisutch  I  ocean  migration 
patterns:  insight  from  marine  coded-wire  tag  recoveries. 
Can.  J.  Fish.  Aquat.  Sci.  59: 1 100- 1 1 1 5. 
Weitkamp,  L.  A.,  T.  C.  Wainright,  G.  J.  Bryant,  G.  B.  Milner, 
D.  J.  Teel,  T  G.  Kope,  and  R.  S.  Waples. 

1995.  Status  review  of  coho  salmon  from  Washington. 
Oregon,  and  California.  NOAA  Tech.  Memo.  NMFS- 
NWFSC-24,  258  p. 


Appendix 

Table  1 

Summary  of  releases  of  coho  salmon  smolts  in  2000  by  region.  This  summary  of  releases  of  all  hatchery  coho  salmon  smolts 
by  region  was  calculated  from  data  in  the  Pacific  States  Marine  Fisheries  Commission  Regional  Mark  Information  System 
(http://www.rmis.org/  [accessed  5  April  2003])  and  in  USFWS  2001  (see  Footnote  2  in  the  general  text). 

No.  of  release 
groups 

ToHl  fish 

Release 

weight  (gl 

released 

Marked 

mean  I SD ) 

All  British  Columbia 

250 

13,612,715 

71.4', 

19.6(5.7) 

Washington:  St.  Juan  de  Fuca,  Puget  Sound,  Skagit  River, 
Nooksack  River,  etc. 

83 

15,316,299 

86  r, 

29.1  (19.7) 

Washington: 

North  of  Columbia  River  to  Cape  Flattery 

63 

7,630,257 

76  7', 

31.6(5.3) 

Columbia  River 

140 

29,879,137 

89.09i 

32.0^  1 

Oregon  Coast  north  of  Cape  Blanco 

14 

809,962 

95.69! 

41.6(7.41 

Southern  Oregon  and  Northern  California:  Rogue,  Klamath, 
and  Trinity  Rivers 

5 

745.060 

99.8^' 

42.1  (4.4) 

'  100%  of  the  fish  released  from  Klamath  and  Trinity  Rivers  were  clipped  on  the  maxillary. 

47 


Abstract— Between  June  1995  and  May 
1996  seven  rookeries  in  the  Gulf  of  Cali- 
fornia were  visited  four  times  in  order 
to  collect  scat  samples  for  studying  spa- 
tial and  seasonal  variability  California 
sea  lion  prey.  The  rookeries  studied 
were  San  Pedro  Martir,  San  Esteban. 
El  Rasito,  Los  Machos,  Los  Cantiles. 
Isla  Granito,  and  Isla  Lobos.  The  1273 
scat  samples  collected  yielded  4995 
otoliths  (95.3%)  and  247  (4.7%)  cepha- 
lopod  beaks.  Fish  were  found  in  97.4% 
of  scat  samples  collected,  cephalopods 
in  11.2%,  and  crustaceans  in  12.7%.  We 
identified  92  prey  taxa  to  the  species 
level,  11  to  genus  level,  and  10  to  family 
level,  of  which  the  most  important  were 
Pacific  cutlassfish  (Trichiuruslepturus), 
Pacific  sardine  (Sardinops  caeruleus), 
plainfin  midshipman  (Porichthys  spp. ), 
myctophid  no.  1,  northern  anchovy 
(Engraulis  mordax).  Pacific  mackerel 
(Scomber- japonicus),  anchoveta  (Ceten- 
graulis  mysticetus),  and  jack  mackerel 
(Trachurus  symmetricus).  Significant 
differences  were  found  among  rooker- 
ies in  the  occurrence  of  all  main  prey 
(P<0.04),  except  for  myctophid  no.  1 
(P>0.05).  Temporally,  significant  dif- 
ferences were  found  in  the  occurrence 
of  Pacific  cutlassfish,  Pacific  sardine, 
plainfin  midshipman,  northern  an- 
chovy, and  Pacific  mackerel  (P<0.05). 
but  not  in  jack  mackerel  lx2=2.94,  df=3, 
P=0.40 1,  myctophid  no.  l(;r=  1.67,  df=  3, 
P=0.64 ),  or  lanternfishes  ( x2=2.08,  df=3, 
P=0.56).  Differences  were  observed  in 
the  diet  and  in  trophic  diversity  among 
seasons  and  rookeries.  More  evident 
was  the  variation  in  diet  in  relation  to 
availability  of  Pacific  sardine. 


Spatial  and  temporal  variation  in  the  diet 

of  the  California  sea  lion  (Zalophus  californianus) 

in  the  Gulf  of  California,  Mexico 

Francisco  J.  Garcia-Rodriguez 

David  Aurioles-Gamboa 

Centra  Interdisciplinary  de  Ciencias  Mannas-lnstituto  Politecnico  Nacional 

Departamento  de  Biologia  Manna  y  Pesquerias 

Apdo.  Postal  592 

La  Paz,  Ba|a  California  Sur,  Mexico 

E-mail  address  (for  F  J.  Garcia-Rodriguez)  fjgrodriifflcibnor.mx 


Manuscript  approved  for  publication 
9  October  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:47-62  (2004). 


The  population  of  the  California  sea 
lion  (Zalophus  californianus),  in  the 
Gulf  of  California  numbers  approxi- 
mately 23,000  individuals,  82%  of 
which  inhabit  the  northern  region  of 
the  gulf  above  latitude  28°  (Aurioles- 
Gamboa  and  Zavala-Gonzalez,  1994). 
In  this  region  are  found  the  most 
important  reproductive  areas  and  the 
highest  pup  production  of  the  Gulf. 
Aurioles-Gamboa  and  Zavala-Gonzalez 
(1994)  suggested  that  the  high  con- 
centration of  animals  in  this  region  is 
related  to  high  abundance  of  pelagic 
fish  such  as  Pacific  sardine  (Sardinops 
caeruleus)  (also  known  as  South  Ameri- 
can pilchard,  FAO),  Pacific  mackerel 
(Scomber  japonicus).  Pacific  thread 
herring  (Opisthonema  libertate),  and 
anchoveta  (Cetengraulis  mysticetus) 
(Cisneros-Mata  et  al.,  19871;  Cisneros- 
Mata  et  al.,  19912;  Cisneros-Mata  et  al., 
19973). 

Despite  the  importance  of  the  north- 
ern gulf  region,  feeding  studies  of  the 
California  sea  lion  at  Gulf  of  California 
rookeries  have  been  few  and  have  been 
conducted  at  different  time  periods. 
Researchers  have  studied  sea  lion  diet 
in  Los  Islotes  (Aurioles-Gamboa  et  al., 
1984;  Garcia-Rodriguez,  1995),  Los 
Cantiles  (Isla  Angel  de  la  Guarda),  Isla 
Granito  (Sanchez-Arias,  1992;  Bautista- 
Vega,  2000),  and  Isla  Racito  (Orta-Davi- 
la,  1988).  These  studies  have  shown  that 
sea  lions  consume  a  variety  of  prey  and 
that  differences  exist  between  the  diet 
of  sea  lions  found  at  different  rookeries 
within  the  Gulf  of  California.  At  Los 
Islotes,  the  most  important  prey  were 
cusk  eel  (Aulopus  bajacali),  bigeye  bass 


(Pronotogrammus  eos),  threadfin  bass 
(Pronotogrammus  multifasciatus),  and 
splitail  bass  (Hemanthias  sp.)  (Aurioles- 
Gamboa  et  al,  1984;  Garcia-Rodriguez. 
1995).  At  Los  Cantiles  and  Isla  Granito 
important  prey  were  lanternfish  (Dia- 
phus  sp.),  northern  anchovy  (Engraulis 
mordax).  Pacific  cutlassfish  (Trichiurus 
nitens),  shoulderspot  (Caelorinchus 
scaphopsis),  and  Pacific  whiting  (Mer- 
luccius  productus)  (Sanchez-Arias, 
1992;  Bautista-Vega,  2000),  whereas  at 
Isla  Racito,  important  prey  were  Pacific 
sardine  (Sardinops  caeruleus).  Pacific 
mackerel  (Scomber  japonicus),  grunt 
(Haemulopsis  spp.),  rockfish  (Sebastes 


1  Cisneros-Mata,  M.  A..  J.  P.  Santos-Molina, 
J.  A.  DeAnda  M.,A.  Sanchez-Palafox,  and  J. 
J.  Estrada.  1987.  Pesqueria  de  sardina 
en  el  noroeste  de  Mexico  ( 1985/86 ).  Informe 
Tecnico,  79  p.  Centro  Regional  de  Inves- 
tigaciones  Pesqueras  de  Guaymas.  INP. 
SEPESCA.  Calle  20  No.  605  Sur  Col.  La 
Cantera.  Guaymas,  Son.  CP.  85400. 

2  Cisneros-Mata,  M.  A.,  M.  O.  Nevarez- 
Martinez,  G.  Montemayor-Lopez,  J. 
P.  Santos-Molina,  and  R.  Morales- 
Azpeitia.  1991.  Pesqueria  de  sardina  en 
el  Golfo  de  California  de  1988/89-1989/90. 
Informe  Tecnico.  80  p.  Centro  Regional  de 
Investigaciones  Pesqueras  de  Guaymas. 
INP.  SEPESCA.  Calle  20  No.  605  Sur  Col. 
La  Cantera.  Guaymas,  Son.  CP.  85400. 

3  Cisneros-Mata,  M.  A.,  M.  O.  Nevarez- 
Martinez,  M.  A.  Martinez-Zavala,  M.  L. 
Anguiano-Carranza,  J.  P.  Santos-Molina, 
A.  R.  Godinez-Cota,  and  G.  Montemayor- 
Lopez.  1997.  Diagnosis  de  la  pesqueria 
de  pelagicos  menores  del  Golfo  de  Califor- 
nia de  1991/92  a  1995/96.  Informe  Tecnico, 
59  p.  Centro  Regional  de  Investigaciones 
Pesqueras  de  Guaymas.  INP.  SEMARNAP. 
Calle  20  No.  605  Sur  Col.  La  Cantera. 
Guavmas,  Son.  CP.  85400. 


48 


Fishery  Bulletin  102(1) 


spp. ),  and  Pacific  whiting  (Merluccius  spp. ) 
(Orta-Davila,  1988). 

Some  California  sea  lion  prey  are  important 
fisheries  resources  in  Mexico.  The  Pacific  sar- 
dine, for  example,  is  the  target  of  a  fishery  be- 
gun in  1967  and  which,  along  with  the  northern 
anchovy,  contributed  to  most  of  the  volume  of 
the  catch  (200,870  t  during  the  1995-96  season) 
obtained  in  the  Gulf  (Cisneros-Mata  et  al.3). 
The  central  and  northern  regions  of  the  Gulf 
of  California  harbor  the  greatest  abundance  of 
sea  lions  and  schooling  fishes,  such  as  the  sar- 
dine and  anchovy.  Because  of  this,  knowledge  of 
sea  lion  feeding  habits  and  their  temporal  and 
spatial  variability  is  relevant  to  understanding 
the  potential  interaction  between  sea  lions  and 
fisheries  in  the  area  (Orta-Davila,  1988;  San- 
chez-Arias, 1992;  Bautista-Vega,  2000). 

In  this  article,  we  present  the  results  of 
concurrent  diet  studies  conducted  at  various 
rookeries  and  haulout  areas  of  sea  lions  in  the 
northern  rookeries  of  the  Gulf  of  California  to 
examine  the  prey  consumed,  and  spatial  and 
temporal  variability  in  their  diet. 


Materials  and  methods 


32° 


28° 


24° 


20° 


16° 


12° 


Scat  samples  of  California  sea  lions  were 
collected  at  Isla  San  Pedro  Martir  (SPM, 
28°24'00"N,  112°25'3"W),  Isla  San  Esteban 
(EST,  28°42'00"N,  112°36'00"W),  Isla  Rasito 
(RAS,  28°49'30"N,  112°59'30"W),  Isla  Granito 
(GRA,  29°34'30"N,  113°32'15"W),  Isla  Lobos 
(LOB,  30°02'30"N,  114°.  28'30"W),  and  at  two 
colonies  of  Isla  Angel  de  la  Guarda  known  as 
Los  Machos  (MAC,  29°20'00"N,  113°30'00"W), 
and  Los  Cantiles  (CAN,  29°32'00"N,  113°29'00"W,  Fig.  1). 
The  total  number  of  California  sea  lions  in  these  seven 
rookeries  was  approximately  15,000  animals  (that  were 
hauled  out)  of  which  about  12.2%  inhabit  San  Pedro  Martir. 
34.7%  San  Esteban,  2.8%  El  Rasito,  10.0%  Los  Machos, 
8.7%.  Los  Cantiles,  11.0%  Isla  Granito,  and  20.6%  Isla 
Lobos  (Aurioles-Gamboa  and  Zavala-Gonzalez,  1994).  All 
the  animals  were  spread  out  along  the  shoreline  of  each 
island,  except  at  Isla  Angel  de  la  Guarda,  where  they  were 
clustered  within  two  areas:  Los  Cantiles,  on  the  eastern 
shoreline  and  Los  Machos  on  the  western  shoreline. 

Scat  samples  were  obtained  at  reproductive  and  non- 
reproductive  haulout  areas  between  June  1995  and  May 
1996.  At  El  Rasito,  sampling  was  done  only  at  one  reproduc- 
tive area;  fresh  and  dried  samples  were  collected  (Fig.  2). 
If  for  any  reason  a  scat  was  not  collected  (because  it  was 
found  in  pieces  or  in  poor  condition),  it  was  destroyed  and 
the  site  was  cleared  to  avoid  collection  during  subsequent 
trips.  All  fresh  and  dried  samples  collected  were  pooled  to 
represent  each  sampling  period.  We  assumed  that  the  diet 
information  corresponded  to  a  time  period  close  to  the  col- 
lection trip,  but  some  dried  scats  could  have  been  deposited 
shortly  after  the  last  collection. 


Pacific 
Ocean 


122° 


118° 


114° 


110° 


106° 


Figure  1 

Map  of  Baja  California  showing  location  of  California  sea  lion  rook- 
eries that  were  studied  in  the  Gulf  of  California. 


Scats  were  stored  in  plastic  bottles  and  then  dried 
shortly  thereafter  to  prevent  decomposition  offish  otoliths 
and  other  hard  parts  (which  were  used  in  subsequent 
prey  identification)  until  the  scats  could  be  processed  at 
a  later  date.  The  samples  were  processed  by  soaking  in 
a  weak  biodegradable  detergent  solution  for  1  to  7  days 
before  being  sifted  through  nested  sieves  of  2. 00-,  1.18-. 
and  0.5-mm  mesh  size.  Fish  bones  and  scales,  eye  lenses  of 
fish  and  squid,  otoliths,  cephalopod  beaks,  and  crustacean 
fragments  were  extracted  from  the  samples.  Cephalopod 
beaks  were  stored  in  70%  ethanol,  and  the  other  items  were 
dried  and  stored  in  vials.  Sagittal  otoliths  and  cephalopod 
beaks  were  used  to  identify  teleost  fish  and  cephalopods,  re- 
spectively. Identifications  were  made  by  using  photographs 
and  diagrams  from  Clarke  (1962),  Fitch  ( 1966),  Fitch  and 
Brownell  (1968),  and  Wolff  (1984),  as  well  as  voucher 
specimen  material  from  the  1)  Center  Interdiseiplinario 
de  Marinas  Ciencias  (CICIMAR),  2)  Instituto  Tecnologico 
y  de  Estudios  Superiores  de  Monterrey,  Guaymas,  3)  Los 
Angeles  County  Museum  of  Natural  History,  California, 
and  4)  Centro  de  Investigacion  Cientifica  y  de  Educacion 
Superior  de  Ensenada  (CICESE).  Baja  California,  Mexico. 
Prey  species  identifed  to  family  level  were  coded  by  using 


Garcia-Rodriquez  and  Aurioles-Gamboa:  Spatial  and  temporal  variation  in  the  diet  of  Zalophus  californianus 


49 


San  Pedro  Martir  (SPM) 


28°  24'- 


HA 


San  Esteban  (EST) 

112°40'     112=38'    1 12=36'    112=34'    112=32' 
J L 


28=44- 


28=42' 


El  Rasito  (RAS) 


Angel  de  la  Guarda 


28=49' 


113=40'     113=30'     113=20'     113=10'     113=00' 
I                I                i                i                i 

29=30'- 

r^\  «—  Los  Cantiles  (CAN) 
/                           (RAyHA) 

29°  20'- 

A\    ^-. 

29°  10- 

Los  Machos  (MAC)\            ^ 
(RA  y  HA)          ^v         \ 

M 

Isla  Granito  (GRA) 


Isla  Lobos  (LOB) 


113' 

34' 

113°  33' 

i 

29°  35'- 

s              RA 
RA 

ha   *\y 

29=34'- 

30=03'. 


114=29' 

1 

114=  28 
I 

|      RA 

K 

HA  - 

Figure  2 

Location  of  sites  where  samples  of  California  sea  lion  scats  were  collected  at  each  island. 
RA  =  reproductive  area;  HA  =  haulout  area. 


the  family  name  plus  a  sequential  number.  Otoliths  from 
prey  species  that  were  not  identified  to  species,  genus,  or 
family  level  were  coded  with  "fish  species"  plus  a  number. 
Three  indices  were  used  to  describe  the  diet  of  sea  lions. 
Percent  number  (PN)  represents  the  percentage  of  the 
number  of  individuals  for  each  prey  taxon  over  the  total 
number  of  individuals  found  in  all  scat  samples.  Percent 
of  occurrence  (PO)  represents  the  percentage  of  scats  hav- 
ing a  given  prey  taxon  and  indicates  the  percentage  of  the 
population  that  is  consuming  a  particular  prey  species.  The 
third  index,  index  of  importance  (IIMP)  incorporates  PN 
and  PO  and  is  defined  as 


IIMP, 


'T  ^ 


u 


X 


(1) 


where  xt   =  number  of  individuals  of  taxon  z'  in  scatj; 

X  =  total  number  of  individuals  from  all  taxa  found 

in  scat  J;  and 
U  =  total  number  of  samples  with  prey. 

The  IIMP,  developed  for  scat  analysis  (Garcfa-Rodriguez, 
1999),  was  used  to  determine  the  importance  of  prey 
species,  their  spatial  and  temporal  variation  in  the  diet. 


50 


Fishery  Bulletin  102(1) 


diversity  of  prey  estimates,  and  measures  of  similarity 
among  rookeries.  Crustaceans  were  not  incorporated  into 
the  IIMP  index  because  it  was  not  possible  to  quantify  the 
number  of  individuals  in  the  samples. 

We  used  the  IIMP  Index  because  it  is  less  sensitive  to 
changes  in  the  number  of  prey  in  an  individual  scat  com- 
pared to  PN.  For  example,  if  a  scat  contains  a  single  prey 
taxon,  the  IIMP  does  not  change  regardless  of  the  number 
of  individuals  of  that  taxon,  in  that  scat.  However,  as  one 
increases  the  number  of  individuals  of  a  given  prey  taxon 
in  the  scat,  the  PN  will  also  increase  for  that  prey.  The 
IIMP  allows  each  scat  to  contribute  an  equal  amount  of 
information,  whereas  PN  can  be  dominated  by  a  few  scats 
with  a  large  number  of  a  single  prey-taxon  otoliths.  In  this 
manner  the  IIMP  is  similar  to  the  split-sample  frequency 
of  ocurrence  (SSFO)  index,  developed  by  Olesiuk  (1993), 
where  each  scat  is  treated  as  a  sampling  unit  and  does 
not  assume,  as  does  PN,  that  the  distribution  of  prey  hard 
parts  between  scats  is  uniform.  However,  with  the  SSFO 
index,  each  prey  taxon  in  a  given  scat  is  given  an  equal 
weight  for  that  scat.  If  only  one  species  occurs  in  a  sample, 
its  occurrence  is  scored  as  1,  if  two  species  occur,  each  oc- 
currence is  scored  as  0.5,  and  so  forth  (Olesiuk,  1993).  The 
IIMP  index  incorporates  more  information  than  the  SSFO 
index,  regardless  of  the  number  of  individuals  of  each  taxon 
in  the  scat.4 

Percent  number  (PN)  was  used  only  to  show  differences 
between  broad  prey  groups  (fishes  and  cephalopods)  and 
PO  was  used  to  review  the  temporal  and  spatial  changes 
from  each  main  prey  (those  with  average  IIMP  of  at  least 
10%  at  any  rookery).  For  all  estimations,  a  single  otolith 
(right  or  left)  or  single  cephalopod  beak  (upper  or  lower) 
represented  one  individual  prey.  We  tested  the  hypothesis 
that  the  occurrence  of  the  main  prey  was  independent  of 
the  rookery  and  of  the  date  collection  using  contingency 
tables  and  an  estimator  of  chi-square  (x~)  (Cortes,  1997). 

Total  length  of  the  otoliths  (mm)  and  the  linear 
equation  obtained  by  Alvarado-Castillo5  were  used  to 
estimate  the  length  of  the  Pacific  sardine  (total  length 
mm=7. 41+147. 24xotolith  length  mm);  r=0.85,  n=2740). 
Insufficient  data  did  not  allow  estimating  the  size  of  speci- 
mens from  May.  All  estimated  lengths  were  transformed  us- 
ing loglO,  followed  by  an  ANOVA  among  sampling  periods. 
The  size  of  Pacific  sardine  consumed  by  California  sea  lion 
was  compared  to  those  caught  in  the  commercial  fishery. 
We  chose  to  estimate  the  size  of  Pacific  sardines  because  of 
the  broad  information  available  concerning  age  and  growth 
and  other  aspects  about  the  fishery  and  because  we  found 
many  sardine  otoliths  in  good  condition. 

Spatial  and  temporal  correlation  in  composition  of  diet 
was  compared  by  using  the  Spearman  rank  correlation  co- 


4  Garcfa-Rodriguez,  F.  J.,  and  J.  De  la  Cruz-Agiiero.  In  prep.  An 
index  to  measure  the  specie  prey  importance  of  California  sea 
lion  ^Zalophus  californianus)  based  on  scat  samples. 

'Alvarado-Castillo,  R.  Unpubl.  data.  Departamento  de 
Biologia  y  Pesquerias,  Centro  Interdisciplinary  de  Ciencias 
Marinas.  Avenida  IPN  S/N  Col.  Palo  Playa  de  Santa  Rita,  La 
Paz,  Baja  California  Sur,  Mexico  23070. 


efficient  (Rs)  (Fritz.  1974).  Pairs  of  IIMP  values  were  used 
for  each  prey  taxon  in  a  pair  of  sampling  events  (rookeries 
or  sampling  dates)  to  examine  the  correlation  among  them. 
This  analysis  was  limited  to  prey  that  had  an  IIMP  value 
of  10%  or  more.  Cluster  analysis  of  average  IIMP  data  for 
the  seven  rookeries  was  used  to  assess  the  similarity  of 
the  diet  among  rookeries.  The  dendrogram  for  the  cluster 
analysis  was  based  on  relative  Euclidean  distances  and 
unweighted  pair-grouping  methods  (UPGMA)  strategy 
(Ludwig  and  Reynolds,  1988).  We  included  only  prey  that, 
for  at  least  one  occasion,  had  IIMP  values  >10%. 

Trophic  diversity  was  evaluated  by  using  diversity  curves 
(Hurtubia,  1973)  developed  from  IIMP  index  data.  For  each 
date  and  colony,  the  cumulative  diversity  was  calculated  for 
increasing  numbers  of  sequentially  numbered  (but  we  as- 
sumed randomly  deposited  and  collected)  scat  samples.  The 
diversity  curves  also  allowed  us  to  evaluate  sample  size 
(Hurtubia,  1973;  Hoffman.  1978;  Magurran,  1988,  Cortes, 
1997)  by  identifying  the  point  at  which  the  curve  flattens. 
The  diversity  was  estimated  by  using  the  Shannon  index: 


H' 


-^P,\nPr 


(2) 


where  H'  =  trophic  diversity; 

S    =  total  number  of  prey  taxa;  and 

Pl  =  IIMPr  and  represents  the  relative  abundance 
of  taxon  i  obtained  from  each  scat  and  pooled 
from  scat  1  up  to  the  total  number  of  scats 
collected. 

The  values  of  trophic  diversity  were  then  plotted  against 
the  number  of  pooled  scats. 


Results 

Identification  of  prey 

The  1273  scat  samples  collected  during  June  1995  through 
May  1996  (Table  1)  yielded  fish  remains  in  97.4%  of  the 
samples,  cephalopod  remains  in  11.2%,  and  crustacean 
remains  in  12.7%.  Fish  and  cephalopods  represented 
95.39;  and  4.7%,  respectively,  of  the  5242  individual  prey 
(excluding  crustaceans).  The  occurrence  and  number 
of  these  prey  groups  changed  temporally  and  spatially 
(Fig.  3).  We  identified  92  prey  taxa  to  the  species  level,  11 
to  the  genus  level,  and  10  to  family  level  from  851  scats 
(Table  2).  Remaining  scats  had  damaged  prey  structures 
in  a  condition  that  prevented  us  from  identifying  species 
to  the  genus  or  family  level. 

We  found  nine  main  prey  with  IIMP  average  values  a  10% 
(Table  3):  the  Pacific  cutlassfish  {THchiurus  lepturus),  the 
Pacific  sardine  (Sardinops  eaeruleus),  the  plainfin  mid- 
shipman (Porichthys  spp.),  myctophid  no.  1,  the  northern 
anchovy  iEngraulis  mordax),  Pacific  mackerel  {Scomber 
japonicus),  the  anchoveta  (Cetengraulis  mysticetus),  jack 
mackerel  iTrachurus  symmetricus),  and  the  lanternfish 
(unid.  myctophid). 


Garcia-Rodriquez  and  Aurioles-Gamboa:  Spatial  and  temporal  variation  in  the  diet  of  Zalophus  californianus 


51 


Table  1 

Number  of  scats  collected  at  each  rookery  for  each  sampling  period. 


June  1995 


San  Pedro  Martir  (SPM) 

SanEsteban(EST) 

ElRasito(RAS) 

Los  Cantiles  (CAN) 

IslaGranito(GRAl 

Los  Machos  (MAC) 

IslaLobos(LOB) 

Total 


22 
50 
11 
20 
24 
39 
72 
238 


September  1995 


January  1996 


33 
66 
56 
58 
20 
32 
139 
404 


91 
58 
47 
41 
36 
72 
433 


Mav  1996 


29 
67 
25 
35 
19 
0 
23 
198 


Total 


172 
274 
150 
160 
104 
107 
306 
1273 


Spatial  and  temporal  variability  of  the  main  prey 

The  importance  (IIMP)  of  the  Pacific  cutlassfish  was 
greater  in  Los  Cantiles  (28.4%),  Isla  Lobos  (20.8%),  and 
Isla  Granito  (48.5%)  than  at  other  sites  (Fig.  4).  The  Pacific 
sardine  was  the  dominant  prey  at  Los  Machos  and  to  the 
south.  There  was  a  significant  correlation  across  the  sea- 
sons between  Los  Machos  and  El  Rasito  (r=0.998.  P=0.04), 
but  not  between  Los  Machos  and  Isla  Granito  U-0.602, 
P=0.59).  The  IIMP  of  sardine  was  also  correlated  between 
El  Rasito  and  San  Esteban  (r=0.95,  P=0.04).  The  plainfin 
midshipman  did  not  show  a  clear  pattern,  but  its  presence 
in  the  diet  increased  northeastward  from  Isla  Angel  de  la 
Guarda.  Lanternfishes,  especially  myctophid  no.  1,  were 
the  dominant  prey  at  San  Pedro  Martir,  San  Esteban,  and 
El  Rasito.  The  presence  of  Pacific  mackerel  was  positively 
correlated  with  the  presence  of  the  Pacific  sardine.  The 
anchoveta  was  only  found  at  Isla  Lobos,  and  jack  mackerel 
at  El  Rasito,  San  Pedro  Martir,  and  Isla  Granito. 

The  changes  in  the  PO  of  the  main  prey  coincided  with 
the  variations  of  the  IIMP.  The  occurrence  of  Pacific  cut- 
lassfish. Pacific  sardine,  plainfin  midshipman,  northern 
anchovy,  Pacific  mackerel,  and  jack  mackerel  was  signifi- 
cantly different  (P<0.04)  among  rookeries.  Myctophid  no. 
1  showed  no  significant  difference  in  ocurrence  <x2=11.04, 
df=6,  P=0.09);  but  when  all  lanternfishes  were  pooled, 
their  occurrence  among  rookeries  was  significantly  differ- 
ent (x2=H.13,df=6,P=0.04).  We  found  significant  temporal 
differences  in  the  occurrence  of  Pacific  cutlassfish.  Pacific 
sardine,  plainfin  midshipman,  northern  anchovy,  and  Pa- 
cific mackerel  (P<0.05),  but  no  significant  differences  were 
found  among  seasons  in  the  occurrence  of  jack  mackerel 
(*2=2.94,  df=3,  P=0.40),  myctophid  no.  1  <x2=1.67,  df=3, 
P=0.6428),  or  lanternfish  <x2=2.08,  df=3,P=0.5562). 

Size  of  Pacific  sardine  consumed  by  sea  lions 

The  estimated  size  of  the  Pacific  sardine  found  in  scat 
was  between  101.8  mm  and  179.7  mm  (mean  length  of 
150.4  mm  ±13.7  mm).  Significant  differences  were  found 
among  sampling  periods  (P=4. 79,  df=2,  P=0. 01 ),  specifically 
between  June  and  January  (Newman-Keuls  test;  P=0.04) 
and  between  September  and  January  (Newman-Keuls  test; 


P=0.01).  The  average  size  was  147.4  mm  (±16.1  mm)  in 
June,  151.7  mm  (±13.0  mm)  in  September,  and  136.5  mm 
( ±13.7  mm )  in  January  ( Fig.  5 ).  A  similar  pattern  was  found 
in  Los  Cantiles,  Los  Machos,  and  Isla  Granito. 

Spatial  and  temporal  correlation  in  diet 

We  identified  25  prey  taxa  that  had  an  IIMP  index  value 
of  >10%  (Table  3)  for  a  given  collection.  The  Spearman 
rank  correlation  coefficient  of  IIMP  between  any  pair  of 
rookeries  during  June,  September,  January,  and  May  was 
not  significant  (P>0.08).  There  was  no  positive  correla- 
tion among  any  pair  of  sampling  periods  for  any  rookery 
(P>0.14),  except  between  January  and  May  at  San  Pedro 
Martir  (Ps=0.64,  P<0.05)  and  El  Rasito  (Ps=0.66,  P<0.05) 
and  between  January  and  June  as  well  as  between  Janu- 
ary and  May  at  Isla  Lobos  (Rs=0.56,  P=0.05;  and  Ps=0.59, 
P=0.05.  respectively). 

The  similarity  in  diet  was  related  to  the  distance  between 
rookeries.  A  clustering  for  the  seven  rookeries  was  obtained 
from  these  25  prey  taxa  (Fig  6).  We  arbitrarily  used  a  "cut- 
off" distance  of  0.3  and  0.4  on  the  dendrogram  as  reference 
points  for  identifying  clusters.  The  group  obtained  by  us- 
ing the  first  distance  (0.3)  showed  four  feeding  areas:  one 
located  in  the  south  ( area  I;  San  Pedro  Martir,  San  Esteban, 
and  El  Rasito),  another  in  Canal  de  Ballenas  (area  II:  Los 
Machos)  and  two  in  the  north  (area  III:  Los  Cantiles  and 
Isla  Lobos;  and  area  IV:  Isla  Granito).  When  the  second 
distance  (0.4)  was  used,  the  seven  rookeries  grouped  into 
two  clusters:  1)  the  southern  region  and  Canal  de  Ballenas, 
and  2)  the  region  north  of  Angel  de  la  Guarda. 

Spatial  and  temporal  variability  in  trophic  diversity 

Temporal  variability  in  trophic  diversity  was  evident 
between  the  rookeries  (Fig.  7).  In  general,  two  patterns 
could  be  differentiated:  one  in  which  the  diversity  varied 
little  throughout  the  year  and  the  other  in  which  diversity 
was  high  in  January  and  low  in  September.  The  rookeries 
San  Pedro  Martir  and  Isla  Lobos  showed  the  first  pattern 
and  Los  Machos  and  Isla  Granito  (and  to  a  lesser  extent, 
San  Esteban  and  El  Rasito)  showed  the  second  pattern.  In 
September,  when  diversity  was  low,  the  dominant  prey  at 


52 


Fishery  Bulletin  102(1) 


100 

80 
60 
40 
20 
0 


100  T 

80 
60 
40 
20 
0.- 


Percent  number 

D  Fishes         ■  Cephalopods 
JUNE  1995- MAY  1996 


SPM        EST       RAS        MAC        CAN       GRA         LOB 

□  Fishes  ■  Cephalopods 

JUNE  1995 


100 
80 
60 
40 
20 
0 


100 
80 
60 
40 
20 
0 


100 
80 
60 

40  ■ 

20 

0 


SPM        EST        RAS        MAC       CAN      GRA        LOB 

■  Fishes  ■   Cephalopods 

SEPTEMBER  1995 


SPM        EST       RAS        MAC       CAN      GRA        LOB 


□  Fishes         ■  Cephalopods 
JANUARY  1996 


SPM        EST       RAS        MAC       CAN       GRA         LOB 

□  Fishes  ■  Cephalopods 

MAY  1996 


SPM        EST        RAS        MAC       CAN      GRA        LOB 


100 
80 1 
60 
40 1 

20 
0 


100' 
80 
60 
40' 
20' 
0. 


Percent  occurrence 

Q  Fishes      ■  Cephalopods         □  Crustaceans 
JUNE  1995- MAY  1996 


M 


XI 


Jl 


SPM        EST       RAS        MAC       CAN      GRA         LOB 

□  Fishes      ■  Cephalopods         □  Crustaceans 
JUNE  1995 


n 

n 

^3*. 

SPM        EST       RAS        MAC        CAN      GRA         LOB 

□  Fishes      ■  Cephalopods         D  Crustaceans 
SEPTEMBER  1995 


n  Jl    n 


100, 
80 
60 1 
40 1 

20 

0 


SPM        EST        RAS        MAC        CAN      GRA         LOB 


O  Fishes      H  Cephalopods         D  Crustaceans 
JANUARY  1996 


SPM   EST   RAS   MAC   CAN   GRA    LOB 

D  Fishes      I  Cephalopods         D  Crustaceans 
MAY  1996 


n^Q 


SPM        EST       RAS        MAC       CAN      GRA         LOB 


Figure  3 

Percent  number  (PNi  and  percent  occurrence  (POl  index  values  for  fishes,  cephalopods,  and  crustaceans  found  in 
samples  of  California  sea  lion  scats  collected  at  seven  rookeries  in  the  Gulf  of  California,  Mexico,  for  all  sampling 
periods  combined  and  for  each  sampling  period. 


San  Esteban,  El  Rasito,  and  Los  Machos  was  Pacific  sar- 
dine, whereas  at  Isla  Granito,  it  was  Pacific  cutlassfish 
(Fig.  4 1.  The  curves  obtained  for  Los  Cantiles  showed  a 


clear  pattern  of  diversity  only  in  September,  although  the 
trend  in  the  January  curve  would  suggest  a  higher  diver- 
sity in  January  than  in  September. 


Garcia-Rodriquez  and  Aurioles-Gamboa:  Spatial  and  temporal  variation  in  the  diet  of  Zalophus  califomianus 


53 


Table  2 

Prey  of  California  sea  lion 

identified  from  scat  samples 

collected  at  Isla  San  Pedro  Martir, 

Isla  San  Esteban,  Isla  El  Rasito,  Los 

Cantiles,  Isla  Granito,  Los  Machos  and  Isla  Lobos  from  June  1995  through  May  1996.  n  ind.  = 

=  number  of  individuals  in 

the  sample; 

PN  =  percent  number;  n  occurr  =  number  of  occurrences 

PO  =  percentage 

of  occurrence;  IIMP  =  index  of 

importance. 

Scientific  name 

Common  name 

n  Ind. 

PN 

n  Occurr. 

PO 

IIMP 

Trichiurus  lepturus 

Pacific  cutlassfish 

306 

5.837 

128 

15.041 

16.392 

Sardinops  caeruleus 

Pacific  sardine 

358 

6.829 

88 

10.341 

10.020 

Porichthys  spp. 

midshipman 

456 

8.699 

95 

11.163 

9.297 

Myctophid  no.  1 

lanternfish 

714 

13.621 

119 

13.984 

7.901 

Engraulis  mordax 

northern  anchovy 

430 

8.203 

43 

5.053 

5.199 

Scomber  japonicus 

Pacific  mackerel 

103 

1.965 

42 

4.935 

3.403 

Cetengraulis  mysticetus 

anchoveta 

410 

7.821 

15 

1.763 

2.404 

Loliolopsis  diomedeae 

squid 

77 

1.469 

35 

4.113 

2.399 

Trachurus  symmetricus 

jack  mackerel 

111 

2.118 

41 

4.818 

2.273 

Merluccius  spp. 

Pacific  whiting 

55 

1.049 

25 

2.938 

2.206 

Pontinus  spp. 

scorpionfish 

61 

1.164 

26 

3.055 

1.842 

Enoploteuthid  no.  1 

squid 

95 

1.812 

23 

2.703 

1.754 

Caelorinchus  scaphopsis 

shoulderspot 

65 

1.240 

25 

2.938 

1.728 

Octopus  sp.  no.  1 

octopus 

24 

0.458 

17 

1.998 

1.614 

Sebastes  macdonaldi 

Mexican  rockfish 

42 

0.801 

18 

2.115 

1.496 

Citharichthys  sp  no.  1 

sanddab 

120 

2.289 

23 

2.703 

1.220 

Fish  species  no.  1 

— 

49 

0.935 

25 

2.938 

1.153 

Haemulopsis  leuciscus 

white  grunt 

176 

3.357 

21 

2.468 

1.093 

Peprilus  snyderi 

salema  butterfish 

163 

3.110 

33 

3.878 

1.025 

Prionotus  spp. 

searobin 

12 

0.229 

9 

1.058 

0.855 

Prionotus  stephanophrys 

lumptail  searobin 

53 

1.011 

14 

1.645 

0.814 

Argentina  sialis 

Pacific  argentine 

19 

0.362 

13 

1.528 

0.754 

Fish  species  no.  2 

— 

55 

1.049 

27 

3.173 

0.737 

Hemanthias  peruanus 

splittail  bass 

60 

1.145 

22 

2.585 

0.602 

Fish  species  no.  3 

— 

9 

0.172 

6 

0.705 

0.592 

Micropogomas  ectenes 

slender  croaker 

13 

0.248 

9 

1.058 

0.547 

Lepophidium  spp. 

cusk-eel 

9 

0.172 

3 

0.353 

0.532 

Fish  species  no.  4 

— 

10 

0.191 

3 

0.353 

0.511 

Sebastes  exsul 

buccanner  rockfish 

15 

0.286 

10 

1.175 

0.505 

Cranchiid  no.  2 

Squid 

20 

0.382 

12 

1.410 

0.501 

Haemulon  flaviguttatum 

yellowspotted  grunt 

11 

0.210 

3 

0.353 

0.468 

Sela r  cru men oph  th aim  us 

bigeye  scad 

24 

0.458 

12 

1.410 

0.431 

Fish  species  no.  5 

— 

33 

0.630 

19 

2.233 

0.384 

Paralabrax  sp.  no.  1 

sea  bass 

9 

0.172 

5 

0.588 

0.373 

Synodus  sp.  no.  3 

lizardfish 

10 

0.191 

3 

0.353 

0.341 

Lepophidium  prorates 

prowspine  cusk-eel 

5 

0.095 

4 

0.470 

0.335 

Fish  species  no.  6 

— 

9 

0.172 

5 

0.588 

0.324 

Synodus  sp.  no.  1 

lizardfish 

25 

0.477 

10 

1.175 

0.324 

Octopus  sp,  no.  2 

octopus 

8 

0.153 

7 

0.823 

0.308 

Gonatus  berryi 

squid 

5 

0.095 

5 

0.588 

0.274 

Mugil  cephalus 

striped  mullet 

1 

0.019 

1 

0.118 

0.265 

Paranthias  colonus 

Pacific  creole-fish 

1 

0.019 

1 

0.118 

0.265 

Batistes  polylepis 

finescale  triggerfish 

13 

0.248 

4 

0.470 

0.245 

Physiculus  nematopus 

charcoal  mora 

30 

0.572 

12 

1.410 

0.244 

Hemanthias  spp. 

sea  bass 

9 

0.172 

6 

0.705 

0.234 

Fish  species  no.  7 

— 

10 

0.191 

8 

0.940 

0.233 

Uroconger  varidens 

conger  eel 

8 

0.153 

5 

0.588 

0.189 

Larimus  spp. 

drum 

8 

0.153 

6 

0.705 

0.174 

Apogon  retrosella 

barspot  cardinalfish 

5 

0.095 

4 

0.470 

0.173 

Dosidicus  gigas 

squid 

8 

0.153 

5 

0.588 

0.171 
continued 

54 


Fishery  Bulletin  102(1) 


Table  2  (continued) 

Scientific  name 

Common  name 

n  Ind. 

PN 

n  Occurr. 

PO 

IIMP 

Merluccius  productus 

Pacific  whiting 

1 

0.019 

1 

0.118 

0.167 

Fish  species  no.  8 

— 

2 

0.038 

2 

0.235 

0.159 

Synodus  sp.  no.  2 

lizardfish 

12 

0.229 

5 

0.588 

0.132 

Scorpaena  sonorae 

Sonora  scorpionfish 

2 

0.038 

1 

0.118 

0.130 

Eucinostomus  spp. 

mojarra 

13 

0.248 

5 

0.588 

0.129 

Fish  species  no.  9 

— 

3 

0.057 

3 

0.353 

0.127 

Cynoscion  reticulatus 

striped  weakfish 

23 

0.439 

7 

0.823 

0.124 

Fish  species  no.  10 

— 

10 

0.191 

1 

0.118 

0.122 

Caulolatilus  affinis 

bighead  tilefish 

4 

0.076 

3 

0.353 

0.114 

Paralabrax  auroguttatus 

goldspotted  sand  bass 

18 

0.343 

4 

0.470 

0.110 

Fish  species  no.  11 

— 

3 

0.057 

2 

0.235 

0.102 

Cranchiid  no.  5 

squid 

1 

0.019 

1 

0.118 

0.097 

Bodianus  diplotaenia 

mexican  hogfish 

1 

0.019 

1 

0.118 

0.087 

Prionotus  sp.  no.  1 

searonbin 

2 

0.038 

2 

0.235 

0.087 

Strongylura  exilis 

California  needlefish 

1 

0.019 

1 

0.118 

0.083 

Synodus  spp. 

lizardfish 

6 

0114 

5 

0.588 

0.146 

Fish  species  no.  12 

— 

3 

0.057 

3 

0.353 

0.074 

Fish  species  no.  13 

— 

2 

0.038 

1 

0.118 

0.065 

Fish  species  no.  14 

— 

3 

0.057 

1 

0.118 

0.060 

Fish  species  no.  15 

— 

2 

0.038 

1 

0.118 

0.058 

Fish  species  no.  16 

2 

0.038 

2 

0.235 

0.056 

Porichthys  sp.  1 

midshipman 

1 

0.019 

1 

0.118 

0.052 

Fish  species  no.  17 

— 

5 

0.095 

3 

0.353 

0.049 

Calamus  brachysomus 

Pacific  porgy 

5 

0.095 

2 

0.235 

0.043 

Fish  species  no.  18 

— 

1 

0.019 

1 

0.118 

0.042 

Fish  species  no.  19 

— 

5 

0.095 

2 

0.235 

0.041 

Ophididae  no.  1 

— 

1 

0.019 

1 

0.118 

0.040 

Fish  species  no.  20 

— 

5 

0.095 

3 

0.353 

0.039 

Sebastes  sinesis 

blackmouth  rockfish 

2 

0.038 

1 

0.118 

0.039 

Symphurus  spp. 

tonguefish 

3 

0.057 

1 

0.118 

0.038 

Fish  species  no.  21 

— 

2 

0.038 

1 

0.118 

0.036 

Pronotogrammus  multifasciatus 

threadfin  bass 

8 

0.153 

2 

0.235 

0.029 

Fish  species  no.  22 

— 

2 

0.038 

2 

0.235 

0.027 

Fish  species  no.  23 

— 

2 

0.038 

1 

0.118 

0.021 

Orthopristis  reddingi 

Bronze-striped  grunt 

16 

0.305 

1 

0.118 

0.020 

Fish  species  no.  24 

— 

2 

0.038 

1 

0.118 

0.020 

Fish  species  no.  25 

— 

1 

0.019 

1 

0.118 

0.016 

Cranchiidae  no.  4 

squid 

2 

0.038 

2 

0.235 

0.014 

Fish  species  no.  26 

— 

2 

0.038 

2 

0.235 

0.014 

Histioteuthis  heteropsis 

squid 

0.019 

1 

0.118 

0.014 

Scorpaenidae  no.  1 

— 

0.019 

1 

0.118 

0.011 

Fish  species  no.  27 

— 

0.057 

2 

0.235 

0.011 

Fish  species  no.  28 

— 

0.019 

1 

0.118 

0.010 

Fish  species  no.  29 

— 

0.019 

1 

0.118 

0.008 

Cranchiidae  no.  3 

squid 

0.019 

1 

0.118 

0.006 

Bollmannia  spp. 

goby 

0.019 

1 

0.118 

0.006 

Fish  species  no.  30 

— 

0.019 

1 

0.118 

0.005 

Cranchiidae  no.  1 

squid 

0.019 

1 

0.118 

0.004 

Paralabrax  maculatofasciatus 

spotted  sand  bass 

0.019 

1 

0.118 

0.003 

Ophidian  scrippsae 

basketweave  cusk-eel 

0.019 

1 

0.118 

0.003 

Physiculus  spp. 

cod.  codling,  mora 

2 

0.038 

1 

0.118 

0.003 

Ophididae  no.  2 

— 

4 

0.076 

1 

0.118 

0.002 

Unid.  Carangidae 

jacks 

8 

0.153 

3 

0.353 

0.141 
continued 

Garcia-Rodriquez  and  Aurioles-Gamboa:  Spatial  and  temporal  variation  in  the  diet  of  Za/ophus  californianus 


55 


Table  2  (continued) 

Scientific  name 

Common  name 

n  Ind. 

PN 

n  Occurr. 

PO 

IIMP 

Unid.  Engraulidae 

anchovies 

1 

0.019 

1 

0.118 

0.248 

Unid.  Haemulidae 

grunts 

13 

0.248 

11 

1.293 

0.509 

Unid.  Labridae 

wrasses 

1 

0.019 

1 

0.118 

0.005 

Unid.  Mycthophidae 

lanternifishes 

216 

4.121 

71 

8.343 

4.895 

Unid.  Ophididae 

cusk-eel 

2 

0.038 

1 

0.118 

0.098 

Unid.  Scianidae 

drums 

13 

0.248 

9 

1.058 

0.643 

Unid.  Scorpaenidae 

scorpionfishes 

30 

0.572 

18 

2.115 

1.078 

Unid.  Serranidae 

sea  bass 

13 

0.248 

6 

0.705 

0.176 

Unid.  Triglidae 

searobins 

1 

0.019 

1 

0.118 

0.002 

Unid.  fishes 

39 

0.744 

16 

1.880 

1.819 

Unid.  cephalopod 

4 

0.076 

4 

0.470 

0.373 

Unid.  fishes  (very 

eroded ) 

381 

7.268 

231 

27.145 

Remains  of  cephalopods 

14 

1.645 

Remains  of  crustaceans 

162 

19.036 

Discussion 

Stomach  acids  attack  otoliths,  affecting  their  size  and 
number  and  consequently  the  estimate  of  prey  occurrence 
and  importance.  Erosion  of  otoliths  during  digestion  has 
been  analyzed  in  studies  of  pinnipeds  in  captivity.  Bowen 
(2000)  reviewed  nine  studies  that  estimated  the  propor- 
tion of  otoliths  recovered  in  scat  samples  to  obtain  a 
prey-number  correction  factor  (NCF).  He  found  that  NCF 
is  greater  than  1.0  because  many  prey  species  are  not 
recovered  in  the  scat  samples.  Additionally,  the  erosion 
level  can  be  significantly  different  among  prey  species 
(Bowen,  2000)  because  of  differences  in  the  shape  and 
microstructure  of  otoliths.  Therefore,  estimates  of  biomass 
based  on  scat  analysis  should  be  carefully  interpreted 
because  the  consumption  of  some  prey  species  can  be 
under-  or  overestimated.  Correction  factors  are  needed 
to  compensate  for  differential  erosion  for  the  prey  species 
of  each  pinniped. 

In  this  study  the  most  important  prey  of  California  sea 
lions  were  pelagic  fish  with  small,  thin,  and  fragile  otoliths 
(Nolf,  1993).  The  lanternfish  also  have  small  otoliths — 
perhaps  smaller  than  those  of  any  other  prey  taxa  found 
in  the  scats.  Their  true  importance  in  California  sea  lion 
feeding  may  be  underestimated  because  of  erosion  caused 
by  stomach  acids  (Da  Silva  and  Neilson,  1985;  Murie  and 
Lavigne,  1985;  Jobling  and  Breiby,  1986;  Jobling,  1987;  Toll- 
it  et  al.,  1997).  Similarly,  the  presence  of  cephalopods  may 
have  been  underestimated  because  their  jaws  are  composed 
of  chitin,  which  is  harder  to  digest,  and  frequently  are  re- 
gurgitated (Pitcher,  1980;  Hawes,  1983).  However,  the  high 
resistence  to  digestion  of  cephalopod  beaks  allows  recovery 
of  them  in  good  shape.  Thus  they  are  a  good  choice  to  use  in 
such  diet  analyses  (Lowry  and  Carretta,  1999). 

A  numerical  index  of  prey  species  importance  may  over- 
or  underestimate  the  dominance  of  prey  species  in  the  diet 
because  it  does  not  consider  the  weight  of  the  prey.  For 
IIMP,  a  numerical  index  that  assumes  a  similar  weight  for 


all  prey  species,  the  true  importance  of  the  individual  large 
prey  in  the  diet  can  be  underestimated  and  the  importance 
of  individual  small  prey  can  be  overestimated.  This  prob- 
lem is  also  present  when  the  PO,  PN,  and  the  SSFO  index 
are  used  because  these  are  all  based  only  upon  the  number 
and  occurrence  of  otoliths  and  cephalopods  beaks.  As  when 
using  PN.  and  the  SSFO,  the  IIMP  does  not  work  for  species 
that  cannot  be  enumerated,  such  as  crustaceans. 

Given  the  tendencies  of  the  trophic  diversity  curves,  the 
sample  size  was  suitable  in  almost  all  cases.  However,  at 
San  Pedro  Martir  a  few  more  samples  would  have  been 
useful  to  fully  depict  the  diet.  At  Los  Cantiles,  except 
during  September  1995,  the  samplings  should  have  been 
more  intense  because  the  flattened  portion  of  the  diversity 
curves  are  not  clear.  The  information,  therefore,  that  comes 
from  those  samples  could  be  biased.  However,  the  number 
of  scats  that  we  analyzed  contained  a  high  percentage  of 
the  consumed  species,  especially  the  main  prey. 

The  results  of  this  study  indicate  that  the  California 
sea  lion  consumed  mainly  fish  and  some  crustaceans  and 
cephalopods.  According  to  the  PN  index,  fish  were  more 
important  than  cephalopods  in  the  diet  of  sea  lions.  In  ad- 
dition, fish  were  more  frequent  (PO)  than  crustacean  and 
cephalopods. 

Crustaceans  were  represented  in  a  similar  manner  in 
scats  from  all  rookeries.  Cephalopods,  however,  were  more 
important  at  San  Pedro  Martir  and  San  Esteban,  prob- 
ably because  they  are  more  common  towards  the  southern 
gulf.  Species  of  the  suborder  Oegopsida,  which  includes 
oceanic  species  (Roper  and  Young,  1975),  were  most  com- 
monly found  in  scats  from  these  rookeries.  Orta-Davila 
(1988)  and  Sanchez-Arias  (1992)  have  also  noted  the  low 
consumption  of  cephalopods  at  the  northern  rookeries. 
Fish  were  the  most  diverse  and  commonly  eaten  prey.  In 
contrast  to  cephalopods,  fish  were  slightly  less  important 
in  the  southern  region. 

The  availability  and  abundance  of  the  various  prey 
resources  influenced  the  diet  of  the  sea  lions  in  the  Gulf 


56 


Fishery  Bulletin  102(1) 


Table  3 

Prey  of  California  sea  lions  having  IIMP  index  values  alO^  in  at  leas 

t  one  sampling 

period  for  seven  rookeries  in  the  Gulf  of  Cali- 

fornia,  Mexico 

Blank  indicate  that  species  were 

not  recorded  in  diet; ' 

— "  means  unavailable  data. 

Prey  species 

June  1995         September  1995 

January  1996 

May  1996 

Average 

San  Pedro 

Engraulis  mordax 

29.7 

2.1 

0.5 

8.1 

Marti  r 

myctophid  no.  1 

29.0 

10.5 

9.0 

20.5 

17.3 

Porichthys  spp. 

11.2 

2.0 

6.8 

15.5 

8.9 

Prionotus  stephanophrys 

0.6 

3.3 

3.3 

10.9 

4.5 

enopleoteuthid  no.l 

27.3 

0.8 

7.0 

Sebastes  macdonaldi 

10.4 

2.6 

Haeumulopsis  leuciscus 

16.7 

6.0 

5.7 

San  Esteban 

Trichiurus  lepturus 

24.9 

3.4 

3.0 

7.8 

Sardinops  caeruleus 

10.0 

34.1 

4.2 

12.1 

unid.  Myctophidae 

13.79 

3.4 

4.3 

10.9 

8.1 

myctophid  no.  1 

2.8 

11.8 

8.9 

18.8 

10.6 

enopleoteuthid  no.  1 

16.9 

4.2 

Sebastes  macdonaldi 

2.1 

9.7 

1.4 

3.3 

fish  species  no.  1 

1.7 

11.0 

3.2 

El  Rasito 

Porichthys  spp. 

26.2 

4.0 

2.3 

8.1 

unid.  Myctophidae 

16.4 

1.5 

8.1 

16.4 

10.6 

Scomber  japonicus 

13.8 

3.2 

3.7 

2.5 

5.8 

Pontinus  spp. 

11.5 

5.1 

4.1 

10.9 

7.9 

Octopus  sp.  no.  1 

11.5 

2.9 

7.7 

5.5 

myctophid  no.  1 

6.6 

5.1 

21.4 

6.8 

10.0 

Sardinops  caeruleus 

1.6 

40.1 

0.9 

7.3 

12.5 

Trachurus  symmetricus 

22.0 

5.0 

23.4 

12.6 

Caelorinchus  scaphopsis 

3.6 

13.5 

10.5 

6.9 

Los  Machos 

Sardinops  caeruleus 

21.0 

64.1 

16.8 

— 

34.0 

Scomber  japonicus 

19.0 

10.9 

— 

10.0 

Merluccius  spp. 

15.4 

8.2 

— 

7.9 

Trichiurus  lepturus 

11.7 

5.4 

— 

5.7 

Sebastes  macdonaldi 

1.8 

11.3 

— 

4.4 

Los  Cantiles 

Porichthys  spp. 

66.7 

15.5 

20.6 

Trichiurus  lepturus 

22.2 

38.2 

53.1 

28.4 

Engraulis  mordax 

3.7 

0.4 

14.3 

4.6 

myctophid  no.  1 

17.6 

4.8 

5.6 

Sardinops  caeruleus 

6.8 

19.0 

6.5 

fish  species  no.  3 

0.9 

14.3 

3.8 

unid.  fishes 

0.9 

19.0 

5.0 

unid.  Scianidae 

14.3 

3.6 

Lepophidium  spp. 

14. 

3.5 

Lo/iolopsis  diomedcav 

21.1 

5.3 

Isla  Granito 

Engraulis  mordax 

49.3 

7.8 

14.3 

Trichiurus  lepturus 

22.0 

70.1 

2.0 

100.0 

48.5 

unid.  myctophidae 

1.7 

1.1 

12.6 

3.9 

Sardinops  caeruleus 

0.9 

18.7 

4.9 

Porichthys  spp. 

0.5 

18.2 

4.6 

5.8 

Citharichthys  sp.  no.  1 

21.7 

5.4 

Isla  Lobos 

Cetengraulis  mysticetus 

32.7 

0.1 

6.8 

27.8 

16.9 

Trichiurus  lepturus 

25.2 

27.7 

15.8 

14.3 

20.8 

Porichthys  spp. 

9.0 

10.3 

23.2 

35.5 

19.5 

Loliolopsis  diomedeae 

4.9 

2.2 

11.6 

3.5 

5.6 

Peprilus  snyderi 

23.5 

5.2 

7.2 

Garcia-Rodriquez  and  Aurioles-Gamboa:  Spatial  and  temporal  variation  in  the  diet  of  Zalophus  califomianus  57 


100 
80 
60 

SPM         EST         RAS         MAC         CAN       GRA         LOB 

11111         Ml          H 

l                             l                            i                             1                      ^_                      ^B 

Trichiurus  lepturus 

20 

«                            hJI  L^l  h^^^m. 

I  a  5  fe'i  a  i  fell  a  i  fell  Si  ?  fell  &  i  fell  &  5  fe'i  &  i  | 

n     10     *     SI  3     $     t     SH     to     n     Sl=5     to     t     Sin     (0     t     Sin     to     n     5  1  =>     to     ->     S 

100 
80 

1                         1                        1                        1                        1                         1 

40 
20 

\_m    i  ■    J  -  _«  :    _  i 

Sardinops  caeruleus 

■7      Q.      ;r      >-  1   Z      Q_      Z      >-  1  2     Q.     z      >-  1  z     0-      Z      >-'Z      0-      Z      >;!   Z      0-      Z      >r'Z      0-      z      ^ 
3     W     *     S't     W     3     5  '  =>     co     =5     5 '  =^     w>     ^     5I=3     «     ^     5l=i     W     ^     5' =5     w     ->     5 

100, 
80 

1                             I                            1                             1                            1                             1 

1                             I                            1                             1                            I                             1 

60 
40 

I                    !         - 

Porichthys  spp 

20 

- |            ■                    P-      '   -      ' — ^ 

iMI;ilil;iill;ili  l;l  ft  i  £;5  M  1;5  8l  1 

100 
80 
60 
40 
20 

Myctophidae  no.  1 

Z      Cl      Z      >'Z      0-      Z      >"'z      Q-      Z      V'z      Q-      Z      i  '  Z      CL      Z      >'   Z      0-      Z      b'z      CL      Z      £ 

D      111      <      <iD      UJ      <      <iD      HI      <      <|D      UJ      <      <|D      HI      <      <|D      UJ      <      <|D      Hi      <      * 
=i     CO     3     S1   =5     (0     ^     5't     0)     ^     SI=»W     ^     S  '  =?     CO     -»     S'->     (0     -»     S     ->     CO     ->5 

CD 
C_> 

c 
a 

o 

Q. 

1 

CD 
CD 

ro 
c 

100 
80 
60 
40 
20 

Myctophidae 

z    cl    z    >-'z    n    z    v'zcl    z    v'z    cl    z    >'z    0-    z    b'z    cl    z    i'z    cl    z    > 

i  8  1  1;1  8  1  i;l  8  1  S;i  8  1  I;3  8  1  3;3  8  1  I;r  8  1  i 

100 
80 
60 

Scomber  japonicus 

o 

20 

__)■ 

Q_ 

z    o.    z    v'z    0-    z    >-'z    Q.    z    bz    Q-    z    i'z    Q.    z    >: '  z    0.    z    >: '  z    Q-    z    >; 

TWn5TU)nS=50)-)S-iV)->S->WnS->W-)S->W-iS 

100 
80 

60 
40 
20 

■                                                            ■_ 

Engraulis  mordax 

Z     0.     Z     >-'z     0-     Z     >-'z     D-     z     >-'z     0.     Z     i'z     Q-     z     £  '  Z     0-     Z     £:     Z     0.     Z     >: 

t      CO      n      S      =)      CO      *      S  '  =j      0)      n      So     CO      n      S      =3      CO      n      S      =5      CO      n      S      =i      CO      =3      5 

100 
80 

40 
20 

■  -m 

Cetengraulis  mysticetus 

i  &  i  fell  a  s  s!i  a  ?  si?  &  s  fell  a  i  fell  a  i  fell  a  ?  | 

T     CO     n     5     n     CO     *     S     n     CO     n     St     »     n     S     n     (0     n     St     0)     n     5  (  =>     to     -,     S 

100 
80 

40 
20 

^_    _L    ^_    _L    _^    -L 

Trachurus  symmetricus 

Z      0-      Z      >".Z      tL      Z      >,Z     D-     Z      >-.Z      0-      Z      >ZtZ      0-      Z      ^,Z      0-      Z      £,Z      Q-      Z      5j 

D      S      tf       -t'D      QJ      <      <It      111      rf      <>D      Ul      <      <b      11)      <      <'D      Ul      <      <'3      111      <       < 
n     B     n     S,   =>     to     *     S|=i     CO     t     S,=5     to     n     S,n     to     t     S,=i     to     n     Srn     W     n     S 

spm  :  est    :  ras   :  mac  :  can   :  GRA    :  LOB 

Figure  4 

Index  of  importance  (IIMP)  for  nine  prey  species  identified  from  samples  of  California 
at  seven  rookeries  in  the  Gulf  of  California,  Mexico,  during  June  and  September  1995 

sea  lions  scats  collected 
.  and  January  and  May 

1996. 

of  California.  The  distribution  pattern  of  Pacific  sardine 
closely  agrees  with  its  importance  in  the  sea  lions  diet. 
The  Pacific  sardine  occurred  in  high  concentrations  around 


Angel  de  la  Guarda  and  Isla  Tiburon  during  the  summer 
and  along  the  coast  of  southern  Sonora  during  the  winter, 
where  spawning  occurs  (Cisneros-Mata  et  al.3).  Sardines 


58 


Fishery  Bulletin  102(1] 


were  consumed  in  the  Canal  de  Ballenas  region  during 
the  summer  (September),  when  they  are  very  abundant. 
Larger  size  Pacific  sardines  were  consumed  by  sea  lions 
most  frequently  during  the  summer  when  adult  sardines 
occur  more  frequently  in  the  Canal  de  Ballenas.  As  adult 
sardine  left  Canal  de  Ballenas  ( Cisneros-Mata  et  al.,  1997 ), 
the  proportion  of  young  individuals  in  the  diet  of  sea  lions 
increased.  The  fish  eaten  by  sea  lions  were  apparently 
smaller  than  those  captured  by  the  commercial  fisher- 
ies. The  average  estimated  size  of  the  sardines  consumed 
was  150.4  mm,  whereas  the  average  size  of  commercially 
caught  fish  during  the  1995-96  season  was  162.4  mm  (Cis- 
neros-Mata et  al.3).  This  7%  difference  in  size  may  have 
been  caused  by  an  underestimation  of  otolith  size  because 
of  digestive  erosion  ( Jobling  and  Breiby,  1986).  If  this  is  so, 
then  the  size  of  Pacific  sardines  consumed  by  sea  lions  is 
similar  to  the  size  of  those  captured  by  the  fishery. 

Isla  Lobos  was  the  only  rookery  where  Pacific  sardine  was 
not  consumed.  This  finding  differs  from  those  of  Cisneros- 
Mata  et  al.3  which  show  the  Pacific  sardines  present  as  far 
north  as  Isla  Lobos.  However,  their  study  period  was  during 
the  1991-92  El  Nino  episode,  whereas  our  study  occurred 
during  normal  oceanographic  conditios  in  1995-96. 

Less  is  known  about  the  spatial 
and  temporal  availability  of  other 
important  prey.  As  with  commercial 
captures  (Arvizu-Martinez,  1987), 
Pacific  mackerel  occurred  together 
with  Pacific  sardine.  Similar  varia- 
tions in  occurrence  for  both  species 
have  been  noticed  from  stomach 
content  analyses  of  the  giant  squid 
(Dosidicus  gigas)  (Ehrhardt,  1991). 
Lanternfishes  were  abundant  north 
of  Isla  Angel  de  la  Guarda  (Robison, 
1972);  however  they  were  not  im- 
portant in  the  diet  of  the  California 
sea  lion  in  this  region.  Their  greater 
importance  in  the  diet  at  southern 
rookeries  was  probably  due  to  the 
absence  of  more  preferred  prey  such 
as  Pacific  sardine,  Pacific  cutlass- 
fish,  or  anchoveta.  The  consump- 
tion of  northern  anchovy  tended  to 
be  less  important  towards  Canal 
de  Ballenas,  where  Pacific  sardine 

reached  its  maximum  importance.  The  low  spatial  overlap 
of  these  two  species  has  also  been  noted  in  other  studies. 
The  anchoveta  was  present  only  at  Isla  Lobos.  This  is 
an  estuarine-lagoon  species,  typical  of  coastal  lagoons  of 
northern  Sinaloa  and  Sonora  (Castro-Aguirre  et  al.,  1995). 
The  presence  of  this  prey  in  Isla  Lobos  is  possibly  due  to 
the  sandy  coast  (Walker,  1960),  which  is  similar  to  that  of 
the  Sinaloa-Sonora  coast. 

The  diet  of  California  sea  lions  differed  among  rooker- 
ies, probably  due  to  differences  in  feeding  sites  and  prey 
availability.  Antonelis  et  al.  (1990)  studied  the  foraging 
characteristics  of  the  northern  fur  seal  (Callorhinus  ur- 
sinus)  and  the  California  sea  lion  at  San  Miguel  Island 
and  found  differences  between  foraging  areas  among 


0.15 


200-i 
180- 

160- 

140- 

|      120- 

•£_     100- 

£       80- 

_J 

60- 

40- 

20- 

0 

n=121 

JUN95                SEP95                JAN95 

Figure  5 

Size  of  Pacific  sardine  iSardinops  caeruleus)  estimated 
from  otoliths  found  in  California  sea  lions  scats  collected 
in  Isla  San  Esteban,  El  Rasito,  Granito,  Los  Cantiles,  and 
Los  Machos.  One  standard  deviation  is  indicated  from  each 
mean. 

0.2 


0.25 


0.3 


0.35 


0.4 


0.45 


Figure  6 

Dendrogram  of  cluster  analysis  of  seven  rookeries  determined  with  Euclidean  dis- 
tance (computed  from  the  IIMP  of  the  25  prey  that  had  on  at  least  one  occasion  a 
value  >10%)  and  the  UPGMA  (unweighted  pair-grouping  methods)  strategy.  The 
vertical  lines  represent  the  points  of  references  to  delimit  the  groups. 


species.  The  northern  fur  seal  was  found  most  frequently 
foraging  in  oceanic  water  within  72.4  km  from  the  island, 
whereas  Califorinia  sea  lions  forgaged  more  often  in  the 
shallower  neritic  zone,  within  54.2  km  from  the  island. 
Different  foraging  distances  in  California  sea  lions  from 
San  Miguel  Island  were  found  by  Melin  and  DeLong 
( 1999).  During  the  nonbreeding  season  a  higher  percent- 
age of  foraging  locations  occurred  at  distances  less  than 
100  km,  whereas  during  the  breeding  season  most  of  the 
foraging  locations  occurred  at  distances  greater  than 
100  km.  These  differences  are  probably  due  to  the  in- 
creased California  sea  lion  population  in  San  Miguel; 
this  increase  in  population  forces  sea  lions  to  exploit  new 
areas  as  a  density-dependent  response  to  population 


Garcia-Rodriquez  and  Aurioles-Gamboa:  Spatial  and  temporal  variation  in  the  diet  of  Zalophus  californianus 


59 


SPM 


■Jun95 
»Sep96 


■4    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    t    I 


0     2     4     6     8     10    G    M    16    18   2022242628303234  36  384042'! 


EST 


0      2     4     6     8     10    12 


2022  24  26283032  34  36  384042444648  50 


RAS 


0     2     4     6     8     10   12    14   16   18  202224262830323436384042444648  50 


3  50  ■ 

MAC 

3  00    . 

*     ■     "     * 

2  50   . 

t 

„     / 

2  00   . 

1  50  . 

*/        i 

- 

- 

- 

Jan 

t  00  . 

0  50    . 

1      t      1 

1      1      1      1      1      1      1 

*   I   I   l   l   l   l 

I   I   l 

■t-f 

H- 

■h-i 

0     2     A     6     8     10    12    14    16    18  20  22  24  26  28  30  32  34  36  38  40  42  44  46  48  50 

Sample  size 


CAN 


I'l   I   I   I   i   I   i   i   i   I   I   I   I   I   I   I   I   I   i   I   I   I   I   l   l 

0  2  4  6  8  10  12  14  16  18  20  22  24  26  28  30  32  34  36  38  40  42  44  46  48  50 


GRA 


I'l  i   i  i  i  i  i  i  i i  i  i   i  i  i  i  i  i  i  i 

0246610121416    18  20  22  24  26  28  30  32  34  36  38  40  42  44  46  48  50 


3  50   - 

LOB 

3  00    . 

2  50    . 

_, 

_.  -  -J—v 

2  00    . 

r  v  J 

C^y 

1  50  . 

100    . 

.  A 
/• 

- 

- 

- 

Jan96 

0.50   ■ 

•_i/v 

- 

- 

May9 

/,  i 

-t-t- 

Mill 

i  i  i  i   i  i  i 

0246810121416    18  20  22  24  26  28  30  32  34  36  38  40  42  44  46  40  50 

Sample  size 


Figure  7 

Trophic  diversity  curves  for  California  sea  lions  determined  from  scat  samples  collected  at  seven  rookeries  in  the  Gulf 
of  California,  Mexico.  SPM  =  San  Pedro  Martir;  EST  =  San  Esteban;  RAS  =  Isla  Rasito;  MAC  =  Los  Machos;  CAN  = 
Los  Cantiles;  GRA  =  Isla  Granito;  LOB  =  Isla  Lobos. 


60 


Fishery  Bulletin  102(1) 


growth.  Although,  these  differences  could  also  be  due  to 
variability  in  the  distribution  of  prey  (Melin  and  DeLong, 
1999),  as  suggested  by  Antonelis  and  Fiscus  (1980),  forag- 
ing areas  might  change  with  season  and  annual  variations 
in  prey  availability  and  abundance. 

Foraging  areas  in  the  Gulf  of  California  could  lie  closer 
to  rookeries  than  those  recorded  for  San  Miguel  Island  sea 
lions  because  the  diet  was  different  among  rookeries  in 
spite  of  the  shorter  distance  between  them  (54.2  km).  At 
Los  Islotes,  Baja  California  Sur,  adult  females  fed  within 
20  km  of  the  colony  (Duran-Lizarraga.  1998).  Kooyman  and 
Trillmich  (1986a,  1986b)  reported  similar  data  in  sea  lion 
colonies  of  the  Galapagos  Islands.  In  the  northern  region  of 
the  Gulf  of  California,  feeding  range  could  be  shorter  than 
that  at  Los  Islotes  because  of  the  higher  concentration  of 
food  at  high  nutrient  concentrations  (phosphate,  nitrate, 
nitrite,  and  silicate)  in  Canal  de  Ballenas  that  is  associated 
with  strong  tidal  mixing  (Alvarez-Borrego,  1983). 

Four  foraging  zones  were  discerned  from  dietary  differ- 
ences in  sea  lions  from  the  seven  rookeries  studied.  Zone 
I,  which  included  San  Pedro  Martir.  San  Esteban.  and  El 
Rasito,  was  characterized  by  the  consumption  of  lantern- 
fish;  zone  II,  which  included  Los  Machos  was  characterized 
by  the  consumption  of  Pacific  sardine  and  Pacific  mackerel; 
zone  III,  which  included  Isla  Granito,  by  the  consumption 
of  Pacific  cutlassfish  and  the  northern  anchovy;  and  zone 
IV,  Los  Cantiles  and  Isla  Lobos,  was  characterized  by  the 
consumption  of  the  plainfin  midshipman  and  the  Pacific 
cutlassfish.  These  four  zones  may  indicate  differences  in 
habits  used  by  sea  lions  or  may  indicate  different  oceano- 
graphic  conditions  exploited  by  sea  lions.  The  eastern 
coast  of  the  Gulf  of  California  displays  high  photosyn- 
thetic  pigment  concentrations,  associated  with  upwelling 
induced  by  winds  from  the  northwest  in  the  winter.  These 
conditions  may  make  Canal  de  Ballenas  one  of  the  most 
important  for  the  distribution  of  Pacific  sardine  during 
the  summer. 

Trophic  diversity  varied  spatially  and  temporally.  San 
Pedro  Martir  and  Isla  lobos  sea  lions  seem  to  depend  on  a 
more  stable  feeding  areas  compared  to  sea  lions  at  rook- 
eries on  Isla  Granito  and  Los  Machos,  where  changes  in 
diversity  of  consumed  species  indicated  that  sea  lions  feed 
on  fewer  species  during  certain  times  of  the  year.  Similar 
results  in  relation  to  the  changes  in  diversity  were  also 
noticed  in  the  rookeries  of  the  Channel  Islands  and  Faral- 
lon  Islands,  California  (Bailey  and  Ainley,  1982;  Antonelis 
et  al.,  1984;  Lowry  et  al.,  1990;  Lowry  et  al.,  1991 ).  Perhaps 
the  tendency  to  have  the  highest  values  of  diversity  and 
little  seasonal  variation  at  San  Pedro  Martir  is  the  result 
of  this  rookery  being  located  in  a  zone  of  transition  between 
two  biogeographical  areas.  This  geographical  position  con- 
fers greater  environmental  heterogeneity  and  greater 
ecological  diversity  (Walker,  1960). 

California  sea  lions  in  the  upper  region  of  the  Gulf  of 
California  obtain  the  main  portion  of  their  diet  from  a 
relatively  small  number  of  species.  The  decrease  in  abun- 
dance of  any  of  these  food  resources  can  seriously  affect  the 
population,  particularly  at  Isla  Granito  and  Los  Machos 
because  sea  lions  from  these  rookeries  depend  on  a  few 
species. 


Acknowledgments 

We  wish  to  thank  Secretaria  de  Marina,  Armada  de  Mexico, 
for  its  great  support  during  the  field  activities,  and  the 
Consejo  Nacional  de  Ciencia  y  Tecnologia  (CONACYT) 
for  funding  this  study  under  grant  number  26430-N.  The 
Secretaria  de  Medio  Ambiente,  Recursos  Naturales  y  Pesca 
(SEMARNAP)  provided  permits  for  field  work  (DOO.-700- 
(2)01104  and  DOO.-700(2).-1917).  We  would  like  to  thank 
Robert  Lavenberg  and  Jeff  Siegel  for  allowing  us  the  use 
of  otoliths  from  the  collection  at  the  Natural  Museum  His- 
tory of  Los  Angeles  County  and  also  Lawrence  Barnes  for 
his  logistical  support  during  the  stay  of  first  author  at  Los 
Angeles;  we  also  thank  Manuel  Nava  for  allowing  us  the  use 
of  otoliths  from  the  collection  in  Tecnologico  de  Monterrey, 
Campus  Guaymas.  We  are  also  grateful  to  Unai  Markaida 
for  his  assistance  in  prey  identification  based  on  the 
examination  of  cephalopods  beaks.  We  thank  Mark  Lowry 
for  commenting  on  an  earlier  draft  of  the  paper,  Norman 
Silverberg  for  reviewing  the  manuscript  in  English,  and 
two  anonymous  reviewers  for  their  valuable  suggestions 
and  criticism.  The  first  author  would  like  to  thank  Centro 
Interdisciplinario  de  Ciencias  Marinas-IPN  for  a  scholar- 
ship (PIFI,  Programa  Institucional  para  la  Formacion  de 
Investigadores)  assigned  for  postgraduate  studies. 


Literature  cited 

Alvarez-Borrego,  S. 

1983.  Gulf  of  California.  In  Estuaries  and  enclosed  seas  (B. 
H.  Ketchum.  ed.i,  p  427-449.  Elsevier  Scientific  Publish- 
ing Co..  Amsterdam. 

Antonelis,  G.  A,  and  C.  H.  Fiscus. 

1980.     The  pinnipeds  of  the  California  Current.     Calif.  Coop. 
Oceanic  Fish.  Invest.  Rep.  21:68-78. 
Antonelis,  G  A..  C.  H.  Fiscus.  and  R.  L.  DeLong. 

1984.  Spring  and  summer  prey  of  California  sea  lions. 
Zalophus  californianus,  at  San  Miguel  Island,  California 
1978-79.     Fish.  Bull.  82:67-75. 

Antonelis,  G.  A.,  B.  S.  Stewart,  and  W.  F.  Perryman. 

1990.     Foraging  characteristics  of  females  northern  fur  seals 
(Callorhinus  ursinus)  and  California  sea  lion  \Zalophus 
californianus).     Can  J.  Zool.  68:150-158. 
Arvizu-Martinez,  J. 

1987.     Fisheries  activities  in  the  Gulf  of  California.  Mexico. 
Calif.  Coop.  Oceanic  Fish.  Invest.  Rep.  28:32-36. 
Aurioles-Gamboa,  D.,  C.  Fox,  F.  Sinsel,  and  G.  Tanos. 

1984.     Prey  of  sea  lions  iZalophus  californianus)  in  the  bay 
of  La  Paz,  Baja  California  Sur.  Mexico.     J.  Mammal.  65(3): 
519-21. 
Aurioles-Gamboa,  D.,  and  A.  Zavala-Gonzalez. 

1994.    Algunos  factores  ecolbgicos  que  determinan  la  distri- 
bucion  y  abundancia  del  lobo  marino  de  California,  Zalo- 
phus californianus.  en  el  Golfo  de  California.     Ciencias 
Marinas  20(4):535-553. 
Bailey,  K.  M.,  and  D.  G.  Ainley. 

1982.    The  dynamics  of  California  sea  lion  predation  on 
Pacific  hake.     Fish.  Res.  1:163-176. 
Bautista-Vega,  A.  A. 

2000.  Vartiacion  estacional  en  la  dieta  del  lobo  marino  comun. 
Zalophus  californianus.  en  las  Islas  Angel  de  la  Guarda  y 
Granito.  Golfo  de  California,  Mexico.    Tesis  de  Licenciatura. 


Garcfa-Rodrfquez  and  Aurioles-Gamboa:  Spatial  and  temporal  variation  in  the  diet  of  Zalophus  californianus 


61 


1 13  p.    Universidad  Nacional  Autonoma  de  Mexico.  Mexico, 
D.F. 

Bowen,  W.  D. 

2000.  Reconstruction  of  pinniped  diets:  accounting  for  com- 
plete digestion  of  otoliths  and  cephalopod  beaks.  Can.  J. 
Fish.  Aquat.  Sci.  57:898-905. 

Castro-Aguirre.  J.  L„  E.  F.  Balart,  and  J.  Arvizu-Martinez. 

1995.  Contribucibn  al  conocimiento  del  origen  y  distribution 
de  la  ictiofauna  del  Golfo  de  California,  Mexico.  Hidro- 
biolbgica  5(1-21:57-78. 

Clarke,  M.  R. 

1962.  The  identification  of  cephalopod  "beaks"  and  the  rela- 
tionship between  beak  size  and  total  body  weight.  Bull. 
British  Museum  (Natural  History)  Zoology  8(  101:419-480. 

Cortes,  E. 

1997.  A  critical  review  of  methods  of  studying  fish  feeding 
based  on  analysis  of  stomach  contents:  application  to  elas- 
mobranch  fishes.     Can.  J.  Fish.  Aquat.  Sci.  54:726-738. 

Da  Silva,  J.,  and  J.  D.  Neilson. 

1985.     Limitations  of  using  otoliths  recovered  in  scats  to 

estimate  prey  consumptions  in  seals.     Can.  J.  Fish.  Aquat. 

Sci.  42:1439-1442. 
Duran-Lizarraga.  M.  E. 

1998.  Caracterizacibn  de  los  patrones  de  buceos  de  alimen- 
tation de  lobo  marino  Zalophus  californianus  y  su  relacibn 
con  variables  ambientales  en  la  Bahia  de  La  Paz,  B.C.S. 
Tesis  de  Maestria,  82  p.    CICIMAR,  La  Paz,  B.C.S.,  Mexico. 

Ehrhardt,  N.  M. 

1991.     Potential  impact  of  a  seasonal  migratory  jumbo  squid 
(Dosidicus  gigas)  stock  on  a  Gulf  of  California  sardine 
(Sardinops  sagax  caerulea)  population.     Bull.  Mar.  Sci. 
49(1-21:325-332. 
Fitch,  J.  E. 

1966.    Additional  fish  remains,  mostly  otoliths,  from  a  Pleis- 
tocene deposit  at  Playa  del  Rey,  California.     Los  Angeles 
County  Mus.,  Cont.  in  Sci.  (1191:1-16. 
Fitch,  J.  E.,  and  R.  L.  Brownell. 

1968.     Fish  otoliths  in  Cetacean  stomachs  and  their  impor- 
tance in  interpreting  feeding  habits.     J.  Fish.  Res.  Board 
Canada.  25(121:2561-2574. 
Fritz,  E.  S. 

1974.    Total  diet  comparison  in  fishes  by  Spearman  rank  cor- 
relation coefficients.     Copeia  (19741:210-214. 
Garcia-Rodriguez,  F.  J. 

1995.  Ecologia  alimentaria  del  lobo  marino  de  California, 
Zalophus  californianus  californianus  en  Los  Islotes, 
B.C.S.,  Mexico.  Tesis  de  Licenciatura,  106  p.  Univer- 
sidad Autonoma  de  Baja  California  Sur.  La  Paz,  B.C.S. , 
Mexico. 
1999.  Cambios  espaciales  y  estacionales  en  la  estructura  tro- 
fica  del  lobo  marino  de  California, Zalophus  caliornianus,  en 
la  region  de  la  grande  islas,  Golfo  de  California.  Tesis  de 
Maestria,  73  p.  CICIMAR.  La  Paz,  B.C.S,  Mexico. 
Hawes,  D. 

1983.    An  evaluation  of  California  sea  lion  scat  samples  as 
indicators  of  prey  importance.     Master's  thesis,  50  p.     San 
Francisco  State  Univ.,  San  Francisco,  CA. 
Hoffman,  M. 

1978.  The  use  of  Pielou's  method  to  determine  sample  size 
in  food  studies.  In  Fish  food  habits  studies:  proceedings 
of  the  second  Pacific  Northwest  technical  workshop  (S.  J. 
Lipovsky  and  C.  A.  Simenstad,  eds.),  p  56-61.  Washington 
Sea  Grant  Publication,  Seattle,  WA. 
Hurtubia,  J. 

1973.    Trophic  diversity  measurement  in  sympatric  preda- 
tory species.     Ecology  54(41:885-890. 


Jobling,  M. 

1987.     Marine  mammal  faeces  samples  as  indicators  of  prey 
importance — a  source  of  error  in  bioenergetics  studies. 
Sarsia  72:255-260. 
Jobling,  M.,  and  A.  Breiby. 

1986.    The  use  and  abuse  of  fish  otoliths  in  studies  of  feeding 
habits  of  marine  piscivores.     Sarsia  71:265-274. 
Kooyman,  G  L.,  and  F.  Trillmich. 

1986a.  Diving  behavior  of  Galapagos  fur  seals.  In  Fur 
seals:  maternal  strategies  on  land  and  at  sea  (R.  L.  Gentry 
and  G.  L.  Kooyman,  eds),  p.  186-195.  Princeton  Univ. 
Press,  New  Jersey,  NJ. 
1986b.  Diving  behavior  of  Galapagos  sea  lions.  In  Fur 
seals — maternal  strategies  on  land  and  at  sea  (R.  L. 
Gentry  and  G.  L.  Kooyman,  eds.  1,  p.  209-220.  Princeton 
LIniv.  Press,  Princeton.  NJ. 
Lowry,  M.  S.,  and  J.  V.  Carretta. 

1999.     Market  squid  (Loligo  opalescens)  in  the  diet  of 
California  sea  lions  (Zalophus  californianus)  in  southern 
California  ( 1981-1995).    Calif.  Coop.  Oceanic  Fish.  Invest. 
Rep.  40:196-207. 
Lowry,  M.  S„  C.  W.  Oliver,  C.  Macky,  and  J.  B.  Wexler. 

1990.  Food  habits  of  California  sea  lions  Zalophus  califor- 
nianus at  San  Clemente  Island,  California,  1981-86.  Fish. 
Bull.  88:509-521. 

Lowry,  M.  S.,  B.  S.  Stewart,  C.  B.  Heath,  P.  K.  Yochem. 
and  J.  M.  Francis. 

1991.  Seasonal  and  annual  variability  in  the  diet  of  Cali- 
fornia sea  lions  Zalophus  californianus  at  San  Nicolas 
Island,  California,  1981-86.    Fish.  Bull.  89:331-336. 

Ludwig,  J.  A.,  and  J.  F  Reynolds. 

1988.     Statistical  ecology:  a  primer  on  methods  and  comput- 
ing. 338  p.    John  Wiley  and  Sons,  New  York,  NY. 
Magurran,  A.  E. 

1988.     Ecological  diversity  and  its  measurement,  178  p. 
Princeton  Univ.  Press,  Princeton,  NJ. 
Melin,  S.  R.,  and  R.  L.  DeLong. 

1999.  At-sea  distribution  and  diving  behavior  of  California 
sea  lion  females  from  San  Miguel  Island,  California.  In 
Proceeding  of  the  fifth  California  islands  symposium  (D.  R. 
Browne,  K.  L.  Mitchell,  and  H.  W.  Chaney,  eds.),  p.  407-402. 
Santa  Barbara  Museum  of  Natural  History,  Santa  Barbara, 
CA. 
Murie,  D.  J.,  and  D.  M.  Lavigne. 

1985.    A  technique  to  recovery  of  otoliths  from  stomach 
contents  of  piscivorous  pinnipeds.     J.  Wildl.  Manag.  49: 
910-912. 
Nolf,  D. 

1993.    A  survey  of  perciform  otoliths  and  their  interest  for 
phylogenetic  analysis,  with  an  iconographic  synopsis  of  the 
Percoidei.     Bull.  Mar.  Sci.  52(1  ):220-239. 
Olesiuk,  P.  F. 

1993.     Annual  prey  consumption  by  harbor  seals  (Phoca 
vitulina )  in  the  Strait  of  Georgia,  British  Columbia.     Fish. 
Bull.  91:491-515. 
Orta-Davila,  F. 

1988.  Habitos  alimentarios  y  censos  globales  del  lobo  marino 
(Zalophus  californianus)  en  el  Islote  El  Racito,  Bahia  de 
las  Animas,  Baja  California.  Mexico  durante  octubre 
1986-1987.  Tesis  de  Licenciatura,  59  p.  Universidad 
Autonoma  de  Baja  California.  Ensenada,  B.C. 
Pitcher,  K.  W. 

1980.  Stomach  contents  and  feces  as  indicators  of  harbour 
seal,  Phoca  vitulina,  foods  in  the  Gulf  of  Alaska.  Fish. 
Bull.  78:797-798. 


62 


Fishery  Bulletin  102(1 


Robison.  B.  H. 

1972.     Distribution  of  the  midwater  fishes  of  the  Gulf  of 
California.     Copeia  (19721:449-61. 
Roper.  C.  F.  E.,  and  R.  E.  Young. 

1975.    Vertical  distribution  of  pelagic  cephalopods.     Smith- 
sonian Contribution  to  Zoology  209(51 1:31. 
Sanchez-Arias,  M. 

1992.  Contribucion  al  conocimiento  de  los  habitos  alimen- 
tarios  del  lobo  marino  de  California  Zalophus  califomianus 
en  las  Islas  Angel  de  la  Guarda  y  Granito,  Golfo  de  Cali- 
fornia, Mexico.  Tesis  de  Licenciatura,  63  p.  Universidad 
Nacional  Autonoma  de  Mexico.  Mexico,  D.F. 


Tollit,  D.  J.,  M.  J.  Steward,  P.  M.  Thompson.  G.  J.  Pierce, 
M.  B.  Santos,  and  S.  Hughes. 

1997.     Species  and  size  differences  in  the  digestion  of  oto- 
liths and  beaks:  implications  for  estimates  of  pinniped  diet 
composition.     Can.  J.  Fish.  Aquat.  Sci.  54:105-119. 
Walker,  B.  W. 

1960.    The  distribution  and  affinities  of  the  marine  fish  fauna 
of  the  Gulf  of  California.     System.  Zool.  9(3):123-133. 
Wolff,  G  A. 

1984.  Identification  and  estimation  of  size  from  the  beaks  of 
18  species  of  cephalopods  from  the  Pacific  Ocean.  NOAA 
Tech.  Rep.  NMFS  17,  49  p. 


63 


Abstract— Recruitment  of  bay  anchovy 
{Anchoa  mitchilli)  in  Chesapeake  is 
related  to  variability  in  hydrologi- 
cal  conditions  and  to  abundance  and 
spatial  distribution  of  spawning  stock 
biomass  (SSB  I.  Midwater-trawl  surveys 
conducted  for  six  years,  over  the  entire 
320-km  length  of  the  bay,  provided 
information  on  anchovy  SSB,  annual 
spatial  patterns  of  recruitment,  and 
their  relationships  to  variability  in 
the  estuarine  environment.  SSB  of 
anchovy  varied  sixfold  in  1995-2000; 
it  alone  explained  little  variability  in 
young-of-the-year  (YOY)  recruitment 
level  in  October,  which  varied  ninefold. 
Recruitments  were  low  in  1995  and 
1996  (47  and  31xl09)  but  higher  in 
1997-2000  (100  to  265 xlO9).  During 
the  recruitment  process  the  YOY  popu- 
lation migrated  upbay  before  a  subse- 
quent fall-winter  downbay  migration. 
The  extent  of  the  downbay  migration 
by  maturing  recruits  was  greatest  in 
years  of  high  freshwater  input  to  the 
bay.  Mean  dissolved  oxygen  (DO)  was 
more  important  than  freshwater  input 
in  controlling  distribution  of  SSB  and 
shifts  in  SSB  location  between  April- 
May  (prespawningl  and  June-August 
(spawning)  periods.  Recruitments  of 
bay  anchovy  were  higher  when  mean 
DO  was  lowest  in  the  downbay  region 
during  the  spawning  season.  It  is 
hypothesized  that  anchovy  recruit- 
ment level  is  inversely  related  to  mean 
DO  concentration  because  low  DO  is 
associated  with  high  plankton  produc- 
tivity in  Chesapeake  Bay.  Additionally, 
low  DO  conditions  may  confine  most 
bay  anchovy  spawners  to  the  downbay 
region,  where  production  of  larvae  and 
juveniles  is  enhanced.  A  modified  Ricker 
stock-recruitment  model  indicated  den- 
sity-compensatory recruitment  with 
respect  to  SSB  and  demonstrated  the 
importance  of  spring-summer  DO  levels 
and  spatial  distribution  of  SSB  as  con- 
trollers of  bay  anchovy  recruitment. 


Recruitment  and  spawning-stock  biomass 
distribution  of  bay  anchovy  (Anchoa  mitchilli) 
in  Chesapeake  Bay* 

Sukgeun  Jung 
Edward  D.  Houde 

University  of  Maryland  Center  for  Environmental  Science 

Chesapeake  Biological  Laboratory 

1  Williams  St.,  P.O.  Box  38 

Solomons,  Maryland  20688 

E-mail  address  (for  S  Jung):  iung@cbl.umces.edu 


Manuscript  approved  for  publication 
30  September  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:63-77  (20041. 


Recruitment  for  marine  fishes  is  vari- 
able and  is  regulated  or  controlled  by  a 
combination  of  density-dependent  and 
density-independent  processes.  It  has 
been  hypothesized  that  density-inde- 
pendent processes  dominate  from  the 
egg  to  larval  stages  whereas  density- 
dependent  control  by  predation  may  be 
more  important  in  the  juvenile  stage 
(Sissenwine,  1984;  Houde,  1987).  Den- 
sity-dependent processes  may  be  stock 
dependent,  regulated  by  adult  abun- 
dances, or  dependent  on  abundances  of 
the  early-life  stages  (Ricker,  1975).  In 
estuarine  systems,  where  hydrological 
conditions  (e.g.  dissolved  oxygen,  tem- 
perature, and  circulation)  vary  widely, 
the  roles  of  density-independent  physi- 
cal factors  on  fish  recruitments  may 
be  dominant,  making  it  difficult,  but 
still  important,  to  partition  density- 
dependent  and  density-independent 
processes,  particularly  for  short-lived 
small  pelagic  fishes  such  as  anchovies 
and  sardines. 

Bay  anchovy  {Anchoa  mitchilli)  (En- 
graulidae)  is  a  coastal  species  distrib- 
uted broadly  in  the  western  Atlantic 
from  Maine  to  Mexico.  This  small  fish  is 
the  most  abundant  and  ubiquitous  fish 
in  Chesapeake  Bay,  the  largest  estu- 
ary on  the  east  coast  of  North  America 
(Houde  and  Zastrow,  1991;  Able  and 
Fahay,  1998).  It  is  not  fished,  yet  there 
is  evidence  that  recruitment  is  variable 
(Newberger  and  Houde,  1995).  It  feeds 
on  zooplankton — primarily  copepods  and 
other  small  Crustacea — and  is  a  major 
prey  of  piscivores,  including  several  eco- 
nomically important  fishes  (Baird  and 
Ulanowicz,  1989;  Luo  and  Brandt,  1993; 


Hartman  and  Brandt,  1995).  Male  and 
female  bay  anchovy  in  Chesapeake  Bay 
mature  at  40^15  mm  fork  length  (44-50 
mm  total  length)  at  about  10  months 
of  age,  and  peak  spawning  occurs  in 
July  (Zastrow  et  al..  1991).  Most  eggs 
are  produced  by  age-1  individuals  (Luo 
and  Musick,  1991;  Zastrow  et  al.,  1991). 
Bay  anchovy  may  survive  to  age  3+  and 
reach  approximately  100  mm  length  and 
5  g  wet  weight  ( Newberger  and  Houde, 
1995;  Wang  and  Houde,  1995). 

Newberger  and  Houde  (1995)  noted 
large  differences  in  annual  survey 
abundances  of  bay  anchovy  that  appar- 
ently resulted  from  variability  in  an- 
nual recruitments.  In  Chesapeake  Bay, 
abundance,  growth,  and  mortality  rates 
of  bay  anchovy  eggs  and  larvae  vary 
temporally  and  spatially  (Dorsey  et 
al,  1996;  MacGregor  and  Houde,  1996; 
Rilling  and  Houde,  1999a,  1999b).  Indi- 
vidual-based models  were  developed  to 
test  the  hypothesis  that  recruitment  of 
bay  anchovy  is  determined  by  variable 
growth  and  mortality  during  early-life 
stages  that  are  regulated  by  density-de- 
pendent processes  (Wang  et  al.,  1997; 
Cowan  et  al.,  1999;  Rose  et  al„  1999). 
In  previous  research,  there  was  little 
knowledge  of  levels  of  spawning  stock 
biomass  or  density-independent  envi- 
ronmental factors  that  may  control  re- 
cruitment through  their  effects  on  spa- 
tial and  temporal  variability  in  growth 
and  mortality  of  prerecruit  anchovy. 


*  Contribution  3696  of  the  University  of 
Maryland  Center  for  Environmental  Sci- 
ence, Chesapeake  Biological  Laboratory, 
Solomons,  MD  20688. 


64 


Fishery  Bulletin  102(1) 


39°N 


38°N 


^vt^w 


N      37°N 


gi 

quehanna 

Upper 

— 

Middle 

Lower 

Atlantic 
Ocean 

77°W 


76°W 


Figure  1 

Chesapeake  Bay  and  stations  sampled  by  the  midwater  trawl  in  the  1995-2000  surveys. 
Horizontal  lines  indicate  boundaries  of  three  designated  regions. 


We  evaluated  environmental  factors,  spatial  distribution 
of  spawning  stock  biomass  (SSB),  and  possible  ontogenetic 
migrations  of  prerecruits  (Dovel,  1971;  Loos  and  Perry. 
1991;  Wang  and  Houde,  1995;  Kimura  et  al,  2000)  with 
respect  to  bay  anchovy  recruitment  variability.  Our  objec- 
tives were  1)  to  estimate  annual  and  regional  variability 
in  bay  anchovy  recruitment.  2)  to  evaluate  effects  of  hy- 
drological  conditions  (mainly,  freshwater  input,  and  dis- 
solved oxygen  concentration)  on  stage-specific  distribution, 
ontogenetic  migration,  and  recruitment,  and  3)  to  identify 
mechanisms  and  describe  patterns  or  trends  in  the  bay 
anchovy  recruitment  process.  Data  were  obtained  in  a 
six-year,  multidisciplinary  research  program  conducted 
throughout  Chesapeake  Bay. 


Materials  and  methods 

Study  area 

Chesapeake  Bay  is  a  coastal  plain  estuary  of  partially  mixed 
fresh  water  and  sea  water.  Its  320-km  mainstem  varies  in 
width  from  about  6  to  50  km  (Fig.  1 ).  The  Bay  is  shallow; 
less  than  10' r  of  its  area  is  >18  m  deep  and  approximately 
50'  i  is  <6  m  deep.  More  than  809&  of  the  freshwater  entering 
the  bay  is  from  tributaries  on  its  northern  and  western  sides 


(Chesapeake  Bay  Program1 ).  Salinity  grades  from  near-full 
seawater  at  the  mouth  of  the  bay  to  freshwater  near  its 
head.  Water  temperatures  reach  28-30°C  in  mid  summer, 
and  fall  to  1^°C  in  late  winter  (Murdy  et  al,  1997 ).  Despite 
shallow  depth,  the  bay  usually  has  a  strongly  developed 
pycnocline,  and  has  seasonally  strong  vertical  gradients  in 
temperature,  salinity,  and  dissolved  oxygen. 

Surveys 

Trawl  surveys  were  conducted  three  times  annually  over 
the  entire  bay  (April-May,  June-August,  and  October). 
1995-2000  (Table  l.Fig.  1).  Midwater-trawl  (MWT)  fish  col- 
lections2 were  made  on  transects  in  three  regions:  the  lower 
bay  (37°05'N-37°55'N),  middle  bay  (37°55'N-38°45'N  I,  and 
upper  bay  (38°45'N-39°25'N).  As  defined,  the  lower  bay 
contains  51%  of  water  volume,  the  middle  bay  32^  .and  the 
upper  bay  17^  (Fig.  1).  The  number  of  midwater  trawl  sta- 


Chesapeake  Bay  Program.  2000.  Chesapeake  Bay:  Introduc- 
tion to  an  ecosystem.  U.S.  Environmental  Protection  Agency, 
publ.  EPA  903-R-00-001.  30  p.  EPA.  410  Severn  Ave,  Suite  109, 
Annapolis.  MD  21403. 

Trophic  interactions  in  estuarine  systems,  midwater  trawl  sur- 
vey. University  of  Maryland  Center  for  Environmental  Sci- 
ence, Chesapeake  Biological  Laboratory,  http://www.ch.esa 
peake.org/  ties/mwt  laccessed  15  October  20031. 


Jung  and  Houde:  Recruitment  and  spawning-stock  biomass  distribution  of  Anchoa  mitchilli  65 


Table  1 

Cruise  dates,  mean 
standard  errors  for 

temperatures  (°C) 
individual  cruises. 

salinities  (psul,  and  dissolved  oxygen  (mg/L).  ir 
years,  seasons,  and  regions  of  Chesapeake  Bay, 

tegrated  from  surface  to  bottom,  and  pooled 
1995-2000.  CV  =  coefficient  of  variation  for 

annual  means. 

Temperature 

SE                       Salinity 

SE 

Oxygen                       SE 

Cruise  date  ( depart 
28  Apr  95 

ure) 

13.88 

0.11                         15.01 

0.42 

8.53                       0.1.3 

23  Jul   95 

28.13 

0.12                         15.48 

0.44 

6.50                       0.14 

28  Oct  95 

17.26 

0.12                         17.39 

0.45 

7.59                        0.14 

28  Apr  96 

13.87 

0.10                        10.84 

0.36 

10.21                        0.11 

17  Jul   96 

24.66 

0.11                        11.80 

0.41 

7.43                       0.13 

22  Oct  96 

16.10 

0.10                        11.26 

0.36 

8.55                        0.11 

20  Apr  97 

10.93 

0.13                        11.41 

0.50 

10.01                        0.16 

11  Jul   97 

25.28 

0.13                        13.59 

0.51 

7.10                        0.16 

29  Oct  97 

14.64 

0.13                        18.19 

0.51 

8.01                       0.16 

11  Apr  98 

12.26 

0.12                         8.90 

0.44 

9.95                       0.14 

04  Aug  98 

26.15 

0.12                        12.89 

0.46 

7.01                        0.15 

19  Oct  98 

18.60 

0.13                         16.64 

0.49 

8.64                       0.15 

19  Apr  99 

11.97 

0.13                        13.51 

0.49 

10.04                      0.16 

26  Jun  99 

23.52 

0.15                        16.02 

0.56 

5.75                        0.18 

23  Oct  99 

16.30 

0.14                        17.38 

0.53 

8.87                       0.17 

29  Apr  00 

12.95 

0.17                        12.51 

0.64 

8.98                       0.20 

25  Jul   00 

24.26 

0.14                        14.06 

0.53 

5.17                        0.17 

17  Oct  00 

17.89 

0.15                        16.73 

0.56 

7.63                        0.18 

Year 

1995 

19.76 

0.07                        15.96 

0.25 

7.54                         0.08 

1996 

18.21 

0.06                        11.30 

0.22 

8.73                         0.07 

1997 

16.95 

0.08                        14.40 

0.29 

8.37                         0.09 

1998 

19.00 

0.07                        12.81 

0.27 

8.53                       0.08 

1999 

17.26 

0.08                        15.64 

0.30 

8.22                        0.10 

2000 

18.36 

0.09                        14.43 

0.33 

7.26                        0.11 

CV 

5.8% 

12.5% 

1.2% 

Season 

April-May 

12.64 

0.05                         12.03 

0.20 

9.62                       0.06 

June-August 

25.33 

0.05                        13.97 

0.20 

6.49                       0.06 

October 

16.80 

0.05                         16.27 

0.20 

8.22                       0.06 

Region  of  bay 
Lower 

18.40 

0.04                         21.19 

0.16 

8.15                       0.05 

Middle 

18.33 

0.05                         14.06 

0.19 

8.33                        0.06 

Upper 

18.04 

0.06                          7.02 

0.23 

7.85                        0.07 

tions  per  survey  ranged  from  24  to  52  (six-year  total=597). 
Additional  baywide  surveys  (August  1997  and  September 

1998)  and  partial  surveys  (June  1997,  July  1998,  and  July 

1999)  also  provided  data  (total  stations  =146). 

An  18-m2  mouth-opening  midwater  trawl  (MWT),  with 
3-mm  codend  mesh  was  deployed  from  the  stern  of  the 
37-m  research  vessel  Cape  Henlopen.  All  trawling  was 
conducted  at  night.  Standardized  tows  of  20-min  dura- 
tion were  conducted  and  the  trawl  was  deployed  at  graded 
depth  intervals  from  surface  to  bottom  ( 2  minutes  at  each 
depth  interval )  in  order  to  provide  a  sample  of  fish  from 


the  entire  water  column.  Fish  catches  (or  subsamples)  were 
counted,  measured  (to  the  nearest  1.0  mm),  and  weighed 
on  deck  immediately  after  a  tow. 

Abundance  and  biomass  of  bay  anchovy  recruits  and 
spawners 

We  separated  bay  anchovy  catches  into  YOY  and  spawn- 
ers based  on  total  length  (TL).  The  minimum  length  of  bay 
anchovy  retained  by  the  MWT  was  21  mm  TL,  which  we 
also  defined  as  the  minimum  TL  for  recruited  YOY  bay 


66 


Fishery  Bulletin  102(1) 


anchovy.  Modal  lengths  of  young-of-the-year  (YOY)  bay 
anchovy  cohorts  were  determined  from  length-frequency 
distributions  in  MWT  catches  and  a  modal  analysis  (Bhat- 
tacharya,  1967;  King,  1995).  Based  on  the  modal  analysis 
of  summer  and  fall  survey  data,  the  maximum  TL  of  YOY 
bay  anchovy  and,  therefore,  the  minimum  TL  of  spawners, 
were  estimated  (Table  2). 

Length-dependent  gear  selectivity  for  bay  anchovy  was 
adjusted  by  comparing  catches  of  the  MWT  and  a  2-m2 
Tucker  trawl  with  catches  from  707-iim  meshes  at  the 
same  stations  during  a  September  1998  baywide  survey. 
The  length-specific  MWT:Tucker-trawl  catch  ratios  (N^^j/ 
iVj^catch  per  unit  of  effort  MWT  4-  catch  per  volume  of 
water  Tucker  trawl)  for  anchovies  21-70  mm  TL  indicated 
that  both  gears  fished  with  a  consistent  selectivity  for  bay 
anchovy  of  30-48  mm  TL,  and  with  a  slight  decrease  in  NTT 
for  48-70  mm  TL.  However,  the  values  ofNMWTINTT  were 
lower  by  factors  of  1-7  for  21-30  mm  TL  fish,  indicating 
that  small  anchovies  were  collected  less  efficiently  by  the 
MWT.  We  concluded  that  length  classes  of  anchovies  >30 
mm  TL  were  equally  vulnerable  to  the  MWT  and  those  >48 
mm  TL  were  less  vulnerable  to  the  Tucker  trawl.  Accord- 
ingly, we  adjusted  MWT  catches  of  ^30  mm  TL  anchovy 
by  multiplying  them  by  a  weighting  factor  estimated  from 
the  regression  of  values  of  iVMH,T/./V.r7.  for  21-30  mm  TL 
bay  anchovy. 

( Weighting  factor)  =  -0.59  TL  +  19.08,  (r2=0.96) 

where  TL  =  total  length. 

The  weighting  factor  equals  1.0  for  anchovy  >30  mm  TL 
because  MWT  selectivity  is  constant  for  anchovy  >30  mm 
TL.  To  estimate  water  sampled  in  a  20-min  MWT  tow, 


and 


where  D« 


dn  =  nmwt/  vmwt  =  ( 1/s '  x  Nt/Vtt 


MWT  —  ^  x  ^-^  MWT       TT    x      TT  ♦ 


bay 


N, 


MWT 


N 


77' 


the  concentration  of  31-48  mm  TL 
anchovy  at  a  station  (i.e.  number/m3); 
the  number  of  31-48  mm  TL  bay  anchovy 
collected  per  20-min  MWT  tow  at  a  station; 
VMWT  =  the  effective  water  volume  sampled 
by  a  20-min  MWT  tow  (m:!); 
the  number  of  3 1-48  mm  TL  bay  anchovy  col- 
lected by  the  2-m-  Tucker  trawl  at  the  same 
station; 

vulnerability  to  the  Tucker  trawl  (s=l  if  all 
bay  anchovies  in  water  volume,  V^,  are  col- 
lected); and  VTT  is  the  volume  filtered  by  the 
Tucker  trawl  (m3)  estimated  from  a  flowme- 
ter in  its  mouth. 


The  mean  of  AfWHT/./V7T  for  30-48  mm  TL  bay  anchovy 
during  the  September  1998  survey  indicated  that  V'WUT  = 
4961  m\  if  30-48  mm  TL  bay  anchovy  did  not  significantly 
avoid  the  mouth  of  the  2-m2  Tucker  trawl  (i.e.  s=l).  Assum- 
ing s=l  (i.e.  VMVVT=4961  m3),  we  estimated  "relative"  bay- 


Table  2 

Estimated 

maximum 

total  lengths 

of  young-of-the-year 

bay  anchov 

y  (mm 

)  from  Chesapeake  Bay, 

based  on  analy- 

sis  of  length-frequency 

distributions. 

Year 

Date 

Length  (mm) 

1995 

23  Jul 

28  Oct 

52 
69 

1996 

17  Jul 
22  Oct 

57 
68 

1997 

11  Jul 

2  Aug 
29  Oct 

30 
56 
66 

1998 

4  Aug 

7  Sep 

19  Oct 

50 
62 
69 

1999 

26  Jun 
23  Oct 

30 
65 

2000 

25  Jul 
17  Oct 

52 

67 

wide  abundance  and  biomass  of  YOY  and  spawners  for  the 
18  surveys  from  1995  to  2000. 

To  coarsely  estimate  a  typical  value  of  s.  "absolute"  bay- 
wide  spawner  biomasses  in  June— August  were  estimated 
for  1995-2000  according  to  an  egg  production  method 
(Parker,  1985;  Rilling  and  Houde,  1999a).  Bay  anchovy 
eggs  had  been  collected  in  a  1-m2  Tucker  trawl  during  the 
same  surveys  and  provided  estimates  of  egg  abundance. 
The  coverage  of  stations  and  sampling  design  for  the 
Tucker  trawl  was  comparable  to  that  of  the  MWT,  but  the 
Tucker  trawl  was  deployed  during  both  day  and  night.  We 
presumed  that  all  eggs  collected  between  00:00  and  20:00 
h  had  been  spawned  near  a  midnight  peak  1 00:00  h)  (Za- 
strow  et  al.,  1991)  and  decreased  in  abundance  at  a  mean 
instantaneous  mortality  (reported  for  bay  anchovy  eggs 
in  Chesapeake  Bay  as  M  =  0.066/h;  Dorsey  et  al.,  1996). 
Based  on  the  estimated  number  of  eggs  spawned  at  00:00 
h  for  each  station,  the  regional  mean  weight  of  individual 
spawners  (defined  by  the  minimum  TL  in  Table  2)  in  MWT 
catches,  and  the  reported  fecundity-weight  relationship  for 
females  (Zastrow  et  al.,  1991),  we  were  able  to  coarsely 
estimate  "absolute"  baywide  spawner  biomass.  We  as- 
sumed that  the  spawning  fraction  of  adult  females  per  day- 
was  essentially  1.0  (i.e.  all  adult  females  participated  in 
spawning,  Zastrow  et  al.,  1991)  and  the  fecundity-weight 
relationship  was  constant  over  years. 

Comparison  of  the  baywide  estimates  of  spawner  bio- 
mass in  June-August  based  on  the  egg  production  method 
("absolute"  biomass)  with  estimates  based  on  the  MWT 
catch-per-unit-of-effort  ("relative"  biomass)  indicated  that. 
on  average,  for  1995  to  2000,  s  is  equal  to  0.20.  Therefore, 
the  mean  effective  water  volume  fished  by  a  20-min  MWT 
tow  was  4961x0.20  =  989  m3. 


Jung  and  Houde:  Recruitment  and  spawning-stock  biomass  distribution  of  Anchoa  mitchilli 


67 


Because  NmvT  of  bay  anchovy  was  highly  variable,  even 
at  stations  on  the  same  sampling  transect,  and  a  mixed 
model  (SAS  version  6.12,  SAS  Inst.  Inc.,  Cary,  NC)  includ- 
ing spatial  covariance  ( variogram )  did  not  significantly  im- 
prove precision  in  annual,  seasonal,  and  regional  means  or 
differences  of  NMWT,  a  stratified  sampling  design  ( Steel  and 
Torrie,  1980),  i.e.  stratum  =  region,  was  adopted.  Based  on 
the  mean  effective  water  volume  (=sxVMWJ, ),  we  estimated 
regional  "absolute"  abundance  and  biomass  (number  and 
wet  weight)  and  related  standard  errors  of  the  linear  com- 
bination by  regional  subvolumes  (Samuels,  1989)  of  bay 
anchovy  >21  mm  TL  for  all  MWT  surveys  from  1995  to 
2000  by  multiplying  regional  mean  MWT  catch  by  Vr/989, 
where  Vr  represents  the  water  volume  (m3)  in  each  bay 
region  (Cronin,  1971): 


N!olal=(N^Vl+Nn 


V,„  +  N. 


Vj/(sxVMWT)xVlotal 


SEN=ScNjVr/n,+V*/nn 


+v:?/n„ 


where  Nlotal 


v„  vm,  vu 


SEX 


ScN  = 


baywide  absolute  abundance; 

mean  values  of  NmvT  for  the  lower  (1), 

middle  (m),  and  upper  (u)  bay; 

bay  subvolumes  for  the  lower  (1),  middle 

(m),  and  upper  (u)  bay  (from  Cronin,  1971), 

V,  =  26.7  x  109  m3,  Vm=  16.8  x  109  m3,  V„  = 

8.7  x  109  m3,  V„„„,  =  V,  +  Vm  +  V„  =52.1  x 

109  m3; 

standard  error  of  Nlolal; 

number  of  midwater  trawl  stations  for  the 

lower  (1),  middle  (m),  and  upper  (u)  bay; 

pooled  standard  deviation  of  NMWT  = 

square   root   of  mean   squares   within 

groups  in  analysis  of  variance  table  = 

t/< SS,  +  SSm  +  SS„ ) / ( n,lMl  -3i,  where  SS,,  SSm, 

SStl  =  sum  of  squares  of  NMWT  for  the 

lower  (1),  middle  (m),  and  upper  (u)  bay, 

and  "total  =  nl  +  nm  +  nu- 

Environmental  factors 

Depth  profiles  of  temperature,  salinity,  and  dissolved 
oxygen  ( DO )  concentration  were  determined  from  conduc- 
tivity-temperature-depth ( CTD )  casts  at  sampling  stations. 
DO  data  were  adjusted  by  calibrating  against  Winkler 
titration  data  from  water  samples  collected  in  Niskin  bot- 
tles deployed  with  the  CTD  cast.  However,  DO  data  from 
the  CTD  could  not  be  adjusted  for  the  1999  summer  and  all 
calendar  year  2000  cruises  because  Winkler  titrations  were 
not  conducted.  To  estimate  regional  means  for  the  water 
column,  we  averaged  temperature,  salinity,  and  DO  values 
by  integrating  the  observed  values  with  respect  to  depth, 
after  dividing  the  water  column  into  "above  pycnocline"  and 
"subpycnocline"  layers. 

Ontogenetic  migration 

We  analyzed  length-frequency  distributions  along  the 
south-north  axis  of  the  bay  (i.e.  by  latitude)  to  delineate 


possible  ontogenetic  migrations  of  YOY  and  adult  bay 
anchovy.  To  parameterize  the  distribution  of  YOY  and 
adult  abundance  and  biomass,  we  estimated  the  biomass- 
weighted  mean  latitudes  of  occurrence  for  each  length  class 
(3-mm  interval). 


lb.i  =  2_,BkjLk/2jBtl, 


where  LB ,  =  biomass-weighted  mean  latitude  of  a  length 
class,  /; 
Lk  =  latitude  of  the  station,  k;  and 
B   =  biomass  (g,  wet  weight)  per  20-min  tow. 

We  devised  a  metric  to  parameterize  the  location  of  bay 
anchovy  SSB.  We  assumed  that  the  baseline  boundary  for 
SSB  distribution  during  the  spring  was  at  the  mouth  of 
the  bay  (37°00'N).  Then,  the  upbay  difference  between 
biomass-weighted  mean  latitude  of  SSB  (in  decimal  units) 
in  Jun-August  and  the  baseline  for  SSB  during  the  spring 
lAL  i  was  calculated: 


SL 


biomass-weighted  mean  latitude  of 
SSB  in  June  -  August 


-37.00. 


Recruitment  model 

As  an  exploratory  step,  a  correlation  analysis  was  under- 
taken to  examine  the  relationships  between  bay  anchovy 
SSB,  migration  patterns,  and  recruitment  levels  with 
respect  to  regional  and  depth-layer-specific  mean  tempera- 
ture, mean  salinity,  mean  DO,  their  gradients,  and  monthly 
mean  freshwater  flow  from  the  Susquehanna  River.  Cross- 
correlations  revealed  that  SSB  migration  pattern  {AD, 
regional  mean  DO  concentrations,  and  October  YOY 
recruitment  level  were  closely  correlated.  Regional  mean 
DO  concentration  provided  the  best  fit  to  YOY  recruitment 
level  in  October  when  baywide  SSB  also  was  included  as 
an  explanatory  variable  in  multiple  regressions.  However, 
because  there  is  uncertainty  in  the  uncalibrated  DO 
measurements  in  1999  and  2000.  we  did  not  use  regional 
mean  DO  in  our  recruitment  model.  Instead,  we  developed 
a  modified  Ricker-type  stock-recruitment  model  (Ricker, 
1975)  that  included  AL  as  an  explanatory  variable: 

Rx  =  a  S  exp  (-/3j  S  -  /i,  AL)  +  e      (modified  Ricker  model ) 


where 


R, 


recruitment  level  =  October  YOY  abun- 
dance in  each  year  ( 1995-2000); 
y;  a,  l\  and  p.-,  =  regression  coefficients; 

S  =  estimated  baywide  SSB  (male-i- female)  in 

metric  tons  for  April-May;  and 
£   =  the  error  term. 

In  this  model,  if  AL  is  held  constant,  Rs.  is  maximum  at  S  = 
l//3j.  Although  no  abiotic  factor  was  included  explicitly  in 
the  model,  AL  is  strongly  correlated  with  regional  mean  DO 
and  serves  as  a  proxy  for  it.  For  the  modified  Ricker  model, 
collinearity,  and  jackknife  influence  diagnostic  tools  were 


68 


Fishery  Bulletin  102(1) 


Table  3 

Seasonal  mean  freshwater  flow  entering  Chesapeake 
chesbay/RIMP/adaps.html. 

Bay  ft' 

Dm  the  Susqu 

ehanna  River  ( m3/s ).  Data  source 

:  http://va. water. 

usgs.gov/ 

Period 

1995 

1996 

1997 

1998 

1999 

2000 

Jan-Mar 

1289 

2495 

1474 

2563 

1325 

1379 

Apr-Jun 

728 

1702 

920 

1625 

791 

1627 

Jul-Sep 

238 

768 

239 

334 

294 

393 

Oct-Dec 

923 

2230 

746 

194 

642 

504 

Annual  mean 

795 

1799 

845 

1179 

763 

976 

applied  to  evaluate  reliability  of  the  regression  model 
(Belsley  et  al„  1980;  SAS,  1989). 


Results 

Environmental  factors 

Stream  flows  from  the  Susquehanna  River  (Table  3) 
varied  annually  and  seasonally.  Freshwater  stream 
flows  were  higher  in  1996  and  1998  than  in  other 
years.  Baywide  mean  values  of  water  temperature, 
salinity,  and  DO  concentration,  averaged  from  surface 
to  bottom,  varied  annually,  seasonally,  and  regionally 
(Table  1 ).  Annually,  mean  temperature  was  highest  in 
1995  and  lowest  in  1997.  Mean  salinity  was  highest 
in  1995  and  lowest  in  1996.  Mean  DO  concentration 
was  highest  in  1996  and  lowest  in  2000.  Regionally, 
salinity  was  more  variable  than  temperature  and 
DO  concentration.  Seasonally,  temperature  and  DO 
concentration  were  more  variable  than  salinity.  Tem- 
perature was  highest  in  the  June-August  period,  the 
spawning  season  of  bay  anchovy.  Seasonally,  salinity 
increased  progressively  from  April-May  to  October. 
Mean  DO  concentration  was  consistently  lowest  in 
June-August. 

Trends  in  abundance  and  recruitment 

Estimates  of  bay  anchovy  abundance  reported  in  our 
study  are  for  the  entire  mainstem  of  Chesapeake  Bay. 
The  estimated  recruitment  levels  (baywide  abundance 
of  YOY  bay  anchovy  >30  mm  TL  in  October)  varied 
ninefold  and  were  low  in  1995  and  1996  (47.5  ±16.6 
and  30.6  ±8.6xl09  individuals)  but  much  higher  in 
1997-2000  (99.6  ±12.4  to  264.8  ±32.6xl09).  Baywide 
estimates  of  bay  anchovy  biomass  for  individuals  >30 
mm  TL  increased  from  April  to  October  in  each  year 
(Table  4).  October  baywide  biomass  varied  sevenfold 
from  27.1  ±5.5  x  103  to  192.9  ±20.4  x  103  tons  and  was 
highest  in  1998  and  lowest  in  1996. 

Estimated  spawning  stock  biomass  (SSB)  in 
April-May  was  lowest  in  1995  (3.3  ±1.1  x  103  tons), 
and  highest  in  1997  (20.1  ±5.3  x  103  tons),  indicating 
sixfold  variability.  SSB  in  June-August  was  lowest 


in  1996  (2.4  ±0.2 x  103  tons),  and  highest  in  1997  (21.1 
±2.3 x  103  tons).  The  SSBs  in  April-May  and  June-August 
did  not  show  any  obvious  relationship  to  YOY  abundance 
(recruitment)  in  October. 

Ontogenetic  migration 

The  length-specific  mean  locations  (latitudes  of  occur- 
rence )  of  bay  anchovy  revealed  an  apparent  ontogenetic 
migration.  Small  juveniles  of  bay  anchovy  tended  to  move 
upbay  and  were  located  primarily  upbay  until  they  were 
approximately  45  mm  TL,  after  which  they  began  to  move 
downbay  (Fig.  2).  In  April-May,  age-1  bay  anchovy  <60  mm 
TL,  consisting  of  individuals  recruited  from  the  previous 
year,  varied  annually  in  their  mean  latitude  of  occurrence, 
whereas  large  (sage  1,  a60  mm  TL)  bay  anchovy  had 
relatively  stable  locations  near  the  boundary  between  the 
lower  and  middle  bay  regions,  centered  at  latitude  37°40'N 
(Fig.  2A).  Compared  to  April-May,  age-l+  bay  anchovy  in 
June-August  were  more  variable  in  their  annual  mean 
locations,  but  both  YOY  and  adult  bay  anchovy  tended  to 
occur  upbay  of  latitude  38°00'N,  except  in  year  2000  (Fig. 
2B).  In  1997  and  1999,  when  annual  mean  temperatures 
were  lowest  (Table  1),  YOY  bay  anchovy  were  too  small 
to  be  sampled  by  the  MWT  in  June-August  and  are  not 
represented  in  Figure  2B.  In  October,  mean  latitudes  of 
occurrence  (Fig.  2C)  indicated  a  consistent  distribution 
pattern  and  an  apparent  ontogenetic  migration  by  YOY 
anchovy.  The  most  probable  explanation  for  the  observed 
latitudinal  distributions  was  that  small  YOY  bay  anchovy 
tended  to  move  upbay  initially,  but  then  downbay  at  about 
45  mm  TL.  Distribution  of  age-l-t-  individuals  in  October 
was  variable. 

The  SSB  of  bay  anchovy  (excludes  YOY)  from  1995  to 
2000  was  centered  near  38°00'N  in  April-August  except 
in  June-August  of  1995  and  1996,  when  the  SSB  was 
centered  farther  upbay  (Fig.  3A).  In  2000,  the  migration 
pattern  differed  from  other  years.  Spawning  bay  anchovy 
in  2000  were  located  farther  downbay  in  July  than  in  April 
(Fig.  3A).  The  April-May  location  of  prespawning  SSB  was 
mostly  explained  by  the  mean  flow  of  the  Susquehanna 
River  from  June  of  the  previous  year  to  February  of  the 
current  year  (r2=0.94,  P=0.0012;  Fig.  3B ).  But,  in  June-Au- 
gust, the  mean  location  of  spawning  fish  was  more  strongly 
and  significantly  related  to  the  subpycnocline-layer  mean 


Jung  and  Houde:  Recruitment  and  spawning-stock  biomass  distribution  of  Anchoa  mitchilli 


69 


Table  4 

Baywide  abundance  and  biomass 

estimates  for  bay  anchovy 

>30  mm  TL  (young-of-the-year 

+  adult).  SE  = 

=  standard 

error. 

Year 

Period 

Abundance  I 

xlO9) 

Biomass 

xlO3  metric  tons) 

Estimate 

SE 

Estimate 

SE 

1995 

April-May 

2.1 

0.7 

3.3 

1.1 

June-August 

57.8 

28.1 

32.6 

17.5 

October 

47.5 

16.6 

51.9 

21.0 

1996 

April-May 

4.9 

1.1 

8.9 

2.0 

June-August 

5.3 

1.6 

3.7 

1.3 

October 

30.6 

8.6 

27.1 

5.5 

1997 

April-May 

11.8 

3.3 

20.1 

5.3 

June-August 

9.4 

2.3 

21.1 

5.0 

October 

99.6 

12.4 

85.6 

10.8 

1998 

April-May 

3.5 

0.7 

6.1 

1.3 

June-August 

14.4 

4.5 

17.0 

7.9 

October 

264.8 

32.6 

192.9 

20.4 

1999 

April-May 

6.9 

1.4 

10.6 

2.2 

June-August 

5.5 

1.2 

10.6 

2.4 

October 

124.5 

28.3 

115.3 

25.0 

2000 

April-May 

6.2 

4.1 

13.0 

6.6 

June-August 

144.6 

51.2 

56.0 

17.0 

October 

169.1 

43.7 

152.9 

40.0 

DO  during  that  same  period  in  the  middle  bay  (/•-=(). 75, 
P=0.02;Fig.  3C). 

Correlations 

Correlation  analyses  suggested  that  regional  mean  DO 
concentrations  are  the  most  important  environmental 
correlate  associated  with  spatial  distribution  of  SSB  and 
recruitment  processes  of  bay  anchovy.  The  mean  locations 
(latitudes  of  occurrence),  abundances,  and  biomasses  for 
YOY  and  adult  bay  anchovy  were  analyzed  with  respect 
to  environmental  variables  (Table  5).  Recruitment  levels 
(YOY  abundance)  in  October  were  consistently  inversely 
correlated  with  DO  concentrations  in  the  lower  and 
middle  bay  in  June-August  (/-=-0.13  to  -0.89).  Biomass- 
weighted  mean  latitude  of  SSB  (age  1+)  in  April-May  was 
consistently  and  positively  correlated  with  regional  salini- 
ties in  April-May  (r=0.30  to  0.88).  On  the  other  hand,  in 
June-August,  surface-layer  mean  salinity  in  the  lower  Bay 
and  subpycnocline-layer  mean  DO  in  the  lower  and  middle 
bay  were  significantly  and  positively  correlated  with  mean 
latitude  of  SSB  or  AL  (r=0.82  to  0.91).  Baywide  SSB  in 
April-May  and  June-August  tended  to  be  negatively  cor- 
related with  water  temperature  in  April-May  (r=-0.45  to 
-0.90). 

Recruitment  model 

Although  SSB  alone  did  not  correlate  significantly  with 
recruitment  level,  mean  DO  in  June-August  was  signifi- 


cantly related  to  the  mean  latitude  of  SSB  in  June-August 
(or  AL)  and  bay  anchovy  recruitment  level  in  October  (Figs. 
3C  and  4).  AL  was  selected  as  the  explanatory  variable, 
rather  than  DO,  because  DO  data  were  uncalibrated  in 

1999  and  2000.  The  correlation  observed  between  AL  and 
DO  ( Fig  3C )  suggested  that  AL  can  serve  as  a  proxy  for  DO 
in  the  stock-recruitment  model.  Including  AL  and  SSB  for 
April-May  in  a  modified  Ricker  model  provided  a  good  fit 
to  bay  anchovy  recruitment  levels  observed  from  1995  to 

2000  (Fig.  5).  The  model  is 

Rv  =  365  S  exp  (-0.19  S  1.35  AL)    (modified  Ricker  model). 

In  the  model,  if  AL  is  held  constant,  predicted  recruitment 
level  of  bay  anchovy  is  maximum  when  baywide  SSB  in 
April-May  is  approximately  5.3  x  103  tons.  Collinearity  and 
influence  diagnostic  statistics  did  not  indicate  collinearity 
between  the  two  independent  variables  (S  and  AL),  or  that 
an  observation  in  any  year  had  a  dominating  influence  on 
parameter  estimates. 


Discussion 

Complex  environmental  processes  and  biological  interac- 
tions control  bay  anchovy  recruitment  in  Chesapeake  Bay. 
Dissolved  oxygen  (DO),  freshwater  flow,  salinity,  and  tem- 
perature acting  on  prerecruits  and  adults  are  important 
factors  affecting  bay  anchovy  distribution  and  levels  of 
recruitment.  Spawning  stock  size  also  is  related  to  recruit- 


70 


Fishery  Bulletin  102(1) 


ment  level.  Our  results  have  demonstrated  that  there  is 
a  strong  spatial  component  in  the  recruitment  dynamics 
of  bay  anchovy.  Although  fish  recruitment  processes  his- 
torically have  been  difficult  to  understand,  our  six-year, 
spatially  extensive  research  has  provided  new  insights  into 
processes  that  control  bay  anchovy  recruitment. 

Ontogenetic  migration  pattern 

It  is  apparent  that  ontogenetic  migration  plays  a  role  in 
the  spatial  and  temporal  patterns  in  abundance,  biomass, 
and  production  of  bay  anchovy.  There  are  several  lines  of 
evidence.  Rilling  and  Houde  (1999a),  in  a  baywide  analy- 
sis, reported  that  mean  density  of  eggs  and  larvae  in  June 
and  July  1993  was  very  high  in  the  lower  Chesapeake  Bay 
compared  to  more  upbay  sites.  Dovel  (1971)  and  Loos  and 
Perry  (1991)  reported  possible  upbay  or  upriver  migra- 
tion of  bay  anchovy  larvae  and  juveniles  in  the  mainstem 
and  tributaries  of  the  Bay.  Recent  otolith  microchemical 
analyses  have  strongly  supported  the  hypothesis  that 
an  upbay  ontogenetic  migration  by  small  YOY  anchovy 
(>25  mm,  late  larvae  and  small  juveniles)  occurs  (Kimura 
et  al.,  2000).  In  the  middle  Hudson  River  estuary  (Schultz 


April-May 


39°00' 


£    38°00 


37°00 


39°00 


38°00 


37°00' 

30  40  50  60  70  80  90         100 

TL  (mm) 

1995   1996   1997   1998   1999   2000 

Figure  2 

Abundance-weighted  mean  latitude  of  occurrence  of  bay  anchovy 
(Am  hoa  mitchilli)  in  Chesapeake  Bay,  1995-2000. 


et  al.,  2000)  and  Chesapeake  Bay  (North  and  Houde,  in 
press),  selective  tidal-stream  transport  was  suggested  as 
a  mechanism  for  up-estuary  movements  of  bay  anchovy 
larvae.  Our  conceptual  model  of  the  bay  anchovy  life  cycle 
includes  migration  patterns  in  the  bay  based  on  available 
knowledge  and  evidence  (Fig.  6). 

It  is  uncertain  what  benefits  YOY  bay  anchovy  derives 
from  upbay  migration  in  summer  and  whether  the  migra- 
tion is  passive  or  active  before  a  subsequent  reverse  migra- 
tion in  the  fall.  To  explain  upbay  movements  of  estuarine 
fishes,  Dovel  ( 1971 )  proposed  that  there  is  a  "critical  zone" 
of  low  salinity  and  high  prey  production  in  the  upper  bay, 
which  is  important  as  a  nursery  for  bay  anchovy  and 
other  fish  species.  In  late  spring  and  early  summer,  age-1 
and  age-l+  bay  anchovy  mature  and  move  upbay  while 
spawning,  although  the  year  2000,  when  mean  freshwater 
streamflow  during  the  previous  fall-winter  was  lowest,  was 
an  exception.  Recruited  YOY  bay  anchovy  apparently  over- 
winter primarily,  but  not  entirely,  downbay  until  spring. 

There  remains  a  possibility  of  significant  immigration 
to  the  bay  by  adult  bay  anchovy  in  some  years  from  the 
coastal  ocean  or  tidal  tributaries  of  the  bay.  Without  such 
immigration,  baywide  adult  abundance  would  decrease 
continuously  during  the  April-October  period  through 
natural  mortality  However,  in  two  years  of  our  six-year 
study,  1995  and  1998,  estimated  adult  abundance  in- 
creased substantially  from  April  to  July,  and  in  1999 
adult  abundance  increased  from  June  to  October, 
implying  significant  immigration  to  the  bay  in  those 
years  (Jung,  2002). 

Recruitment  control  and  regulation 

The  modified  Ricker  recruitment  model  that  included 
SSB  and  AL  as  explanatory  variables  provided  a  good 
fit  to  bay  anchovy  recruitments.  Although  the  model 
fitted  well,  there  were  only  six  years  of  data,  and 
the  underlying  mechanisms  explaining  relationships 
between  the  distribution  and  level  of  SSB,  hydro- 
logical  conditions,  and  density-dependent  regulatory 
processes  in  recruitment  of  bay  anchovy  are  not  yet 
clear.  Nevertheless,  correlations  and  the  recruitment 
model  clearly  indicated  a  density-dependent  effect  of 
SSB  level  and  also  implicated  environmental  factors 
(at  the  mesoscale)  that  are  related  to  mean  DO  concen- 
tration, latitudinal  distribution  of  SSB  (AL),  and  the 
recruitment  level  of  bay  anchovy  (Fig.  4). 

The  modified  Ricker  model  for  bay  anchovy  <  Fig.  5) 
indicates  a  density-compensatory  stock-recruitment 
relationship  (Ricker,  1975).  although  we  do  not  know 
at  what  life  stages  density-dependent  processes  are 
most  important.  Without  accounting  for  the  control- 
ling effect  of  AL  and  mean  DO  on  a  regional  scale, 
the  density-dependence  might  have  gone  undetected 
(Fig.  4 1.  Recent  individual-based  models  suggest  that 
density-dependent  processes  during  early-life  stages 
could  stabilize  bay  anchovy  recruitments  (Wang  et 
al.,  1997;  Cowan  et  al.,  1999;  Rose  et  al,  1999).  At  the 
small  scales  of  several  meters  modeled  by  Wang  et  al. 
(1997)  and  Cowan  et  al.  (1999),  larval-stage  feeding 


Jung  and  Houde:  Recruitment  and  spawning-stock  biomass  distribution  of  Anchoa  mitchilli 


71 


Zl      39°00' 


38°00' 


CO 
CO 
CO 


37°00' 


April-May 
June-August 


1995 

1996 

1997    1998 

1999    2000 

2000 

38°00' 

1999^ 

1=38.30  -  0.00087  X 
r-=0.94(/)=0.0012) 

37045' 

1995 

--4?^- 1998 

B 

1996_ 

37°30' 

i         , 

300 

< 

39°00' 

c 

C 

CO 

CO 

c 

38°00' 

400  500  600  700 

Mean  river  flow  from  June  to  Feb  (m3/sec) 


1995 
1996^ 


1999 


1998 


37°00' 


2000 


1997 

Y=  35.78  +  0.53  A' 
r2=0.75(p=0.02) 


3.0  3.5  4.0  4.5  5.0 

Dissolved  oxygen  (mg/L) 


5.5 


Figure  3 

Mean  location  (latitude)  of  adult  bay  anchovy  {Anchoa  mitchilli)  spawn- 
ing stock  biomass  (SSB)  in  Chesapeake  Bay.  (A)  Mean  latitude  and 
standard  deviation  in  April-May  and  in  June-August.  The  upper  verti- 
cal bar  represents  mean  +  standard  deviation  for  June-August,  and  the 
lower  vertical  bar  represents  mean-standard  deviation  for  April-May, 
I B  l  Mean  latitude  in  April-May  and  mean  Susquehanna  River  flow  from 
June  of  the  previous  year  to  February  of  the  current  year.  (C)  Mean  lati- 
tude in  June-August  and  mean  dissolved  oxygen  in  the  subpycnocline 
layer  of  the  middle  bay  in  June-August. 


processes  were  important  and  high  adult  SSB  could  pro- 
duce abundant  first-feeding  larvae  with  subsequent  den- 
sity-dependent food  competition.  In  Tampa  Bay,  Florida, 
Peebles  et  al.  ( 1996)  hypothesized  that  bay  anchovy's  size- 
specific  fecundity  is  directly  related  to  prey  availability 
for  adults.  Modeled  results  of  Rose  et  al.  (1999)  suggested 
that  density-dependent  growth  of  bay  anchovy  larvae  and 
juveniles  in  Chesapeake  Bay  would  lead  to  density-depen- 
dent survival  of  these  stages.  Hunter  and  Kimbrell  (1980) 
and  Alheit  (1987)  proposed  that  cannibalism  by  adults  on 


eggs  and  larvae  provides  a  degree  of  density-dependent 
regulation  in  anchovies  of  the  genus  Engraulis.  Analyses 
of  feeding  by  adult  bay  anchovy  did  not  indicate  that  pe- 
lagic fish  eggs  were  a  significant  part  of  bay  anchovy  diet 
(Vazquez-Rojas,  1989;  Klebasko,  1991),  although  no  specific 
study  of  cannibalism  has  been  undertaken. 

We  propose  three  hypotheses  that  may  explain  the  rela- 
tionships among  regional  DO  concentration,  the  latitudi- 
nal shift  in  SSB  distribution  during  the  spawning  season 
(AL),  and  recruitment  levels  of  bay  anchovy  in  October.  The 


72 


Fishery  Bulletin  102(1) 


hypotheses  are  the  following:  1)  averaged  DO  concentra- 
tion is  inversely  related  to  levels  of  plankton  productivity 
in  a  region  and  high  plankton  productivity  favors  high  re- 


cruitments of  planktivorous  bay  anchovy;  2 )  low  dissolved 
oxygen  concentrations  can  restrict  spatial  distribution  of 
bay  anchovy  SSB  to  the  lower  bay  insuring  high  egg  and 


Table  5 

Cross-correlation  coefficients  for  bay  anchovy  distribution  and  abundance  with  respect  to  region-  and  layer-specific  means  of  tem- 
perature, salinity,  and  dissolved  oxygen  from  1995  to  2000.  Mean  latitude  is  biomass-weighted  mean  latitude  of  occurrence  of  bay 
anchovy.  Abundance  and  biomass  are  baywide  total  estimates.  AL  =  (mean  latitude  in  June-August)  -37.00.  Abbreviations  are 
as  follows:  SAL  =  salinity,  TEM  =  water  temperature,  OXY  =  dissolved  oxygen;  the  fourth  and  fifth  digits:  04  =  April-May,  07  = 
June-August;  the  sixth  character:  L  =  lower  bay,  M  =  middle  bay,  U  =  upper  bay;  The  last  character:  S  =  layer  above  the  pycnocline. 
B  =  layer  below  the  pycnocline.  *  =  significant  at  a  =  0.05. 

Young-of-the-year 

Adult 

Mean  latitude 

Abundance 

Mean  latitude 

Biomass 

April-May 

June-August 
(orAL) 

October 

October 

April-May 

June-August 

SAL04LS 

0.29 

-0.43 

0.74 

0.26 

-0.17 

-0.52 

SAL04MS 

0.45 

-0.63 

0.30 

0.71 

-0.41 

-0.22 

SAL04US 

0.27 

-0.60 

0.42 

0.53 

-0.18 

-0.02 

SAL04LB 

-0.24 

0.01 

0.88* 

-0.16 

-0.14 

-0.31 

SAL04MB 

0.08 

-0.17 

0.59 

0.33 

-0.39 

-0.05 

SAL04UB 

0.29 

-0.61 

0.45 

0.46 

-0.03 

0.05 

SAL07LS 

0.83* 

-0.75 

0.91* 

-0.46 

SAL07MS 

-0.12 

0.06 

0.14 

0.31 

SAL07US 

0.06 

-0.03 

-0.04 

-0.33 

SAL07LB 

0.70 

-0.75 

0.64 

-0.11 

SAL07MB 

-0.41 

0.60 

-0.31 

0.19 

SAL07UB 

0.15 

-0.20 

0.01 

-0.42 

TEM04LS 

0.16 

-0.25 

-0.03 

0.65 

-0.90* 

-0.48 

TEM04MS 

0.50 

-0.46 

0.14 

0.65 

-0.71 

-0.85* 

TEM04US 

0.53 

-0.32 

-0.36 

0.52 

-0.56 

-0.85* 

TEM04LB 

0.29 

-0.49 

0.19 

0.71 

-0.72 

-0.45 

TEM04MB 

0.22 

-0.42 

0.39 

0.47 

-0.55 

-0.62 

TEM04UB 

0.40 

-0.26 

-0.39 

0.48 

-0.60 

-0.77 

TEM07LS 

-0.49 

-0.04 

0.11 

0.45 

TEM07MS 

-0.16 

-0.21 

0.47 

0.14 

TEM07US 

-0.29 

-0.08 

0.39 

0.38 

TEM07LB 

-0.68 

0.24 

-0.11 

0.38 

TEM07MB 

-0.24 

-0.10 

0.37 

-0.04 

TEM07UB 

-0.45 

0.16 

0.2] 

0.46 

OXY04LS 

0.63 

-0.22 

-0.80 

0.39 

-0.10 

-0.30 

OXY04MS 

-0.27 

0.56 

0.23 

-0.81 

0.55 

-0.04 

OXY04US 

-0.43 

0.41 

-0.30 

-0.30 

0.30 

0.88* 

OXY04LB 

0.93** 

-0.68 

-0.59 

0.63 

0.04 

-0.38 

OXY04MB 

0.47 

-0.35 

-0.31 

-0.09 

0.70 

-0.12 

OXY04UB 

-0.57 

0.65 

-0.32 

-0.46 

0.21 

0.78 

OXY07LS 

0.18 

-0.30 

0.29 

0.32 

OXY07MS 

0.01 

-0.13 

0.29 

0.56 

OXY07US 

0.23 

-0.32 

0.50 

0.10 

OXY07LB 

0.67 

-0.48 

0.82* 

-0.28 

OXY07MB 

0.72 

(l,SH 

0.87* 

-0.04 

OXY07TJB 

0.01 

0.16 

0.21 

0.37 

Jung  and  Houde:  Recruitment  and  spawning-stock  biomass  distribution  of  Anchoa  mitchilli 


73 


larval  production  there;  and  3)  density-depensatory 
predator  satiation  occurs  when  concentrations  of  bay 
anchovy  larvae  and  juveniles  at  the  mesoscale  ( 10-100 
km )  are  high  in  relation  to  satiation  potential  of  preda- 
tors, which  favors  larval  production  and  high  anchovy 
recruitments. 

First,  averaged  DO  level  in  the  bay  or  its  regions 
may  be  an  indicator  of  ecosystem  metabolism  and  sec- 
ondary production.  DO  level  in  the  subeuphotic  layer 
is  an  indicator  of  respiration  and  secondary  produc- 
tion by  planktonic  and  benthic  communities  (Kemp 
and  Boynton,  1980;  Kemp  et  al.,  1992).  Recruitment 
levels  of  bay  anchovy  increased  substantially  in  1997 
and  in  subsequent  years.  We  speculate  that  enhanced 
detrital  production  potentially  increased  zooplankton 
prey  abundances  in  the  subsequent  year  and  that  asso- 
ciated elevated  levels  of  respiration  by  detrital  micro- 
organisms and  zooplankton  contributed  to  low  mean 
DO.  Increased  zooplankton  prey  abundances,  in  turn, 
may  have  promoted  production  of  larval  and  juvenile 
bay  anchovy  in  1997  and  1998.  Thus,  increased  prey 
availability,  associated  with  low  mean  DO  concentra- 
tion, could  have  enhanced  recruitment  (Fig.  4). 

The  second  hypothesis  proposes  that  spatial  restric- 
tion of  SSB  by  low  DO  is  a  factor  controlling  bay  anchovy 
recruitment.  Based  on  our  results,  hypoxic  conditions  in 
the  bay  appear  to  define  the  distribution  and  potential  for 
upbay  migration  of  bay  anchovy  SSB  (Fig.  3C).  In  years 


300 

1998 

7= -88  .V+ 5 10 

C     200 

o      1 00 
cr 

r-=0.79P=0.01S 
2000 

^~"-\1999 

^^19,97 

~"\J995 

1W(, 

0 

3.0               3.5               4.0               4.5                5.0 

Dissolved  oxygen  (mg/L) 

Figure  4 

Relationship  between  mean  dissolved  oxygen  below  the  pycno- 

cline  in  the  middle  Chesapeake  Bay  during  the  June-August 
period  and  recruitment  level  of  bay  anchovy  in  October,  r2  = 

coefficient  of  determination. 

when  the  baywide  subpycnocline  mean  DO  level  was  low, 
spawning  bay  anchovy  tended  to  be  most  concentrated  in 
the  lower  bay  (Table  5,  Fig.  3,  A  and  C),  possibly  because 
hypoxia  in  deeper  waters  of  the  mid-bay  region  discouraged 
upbay  migration.  The  region  selected  by  adult  anchovy  as 
the  predominant  spawning  area  and  its  variability  played 


R  =  365  Sexpf-O.l0-  S  -  1.354Z.) 

r2-- 


2.0  0 


Figure  5 

Stock-recruitment  model  (modified  Ricker  model).  R  =  baywide  number  of  recruits  in 
October  (xlO9).  AL  =  location  of  bay  anchovy  iA?iclioa  mitchilli)  spawning  stock  biomass 
in  June-August  in  relation  to  the  baseline  latitude  at  the  mouth  of  the  bay,  37°00'N.  S  = 
baywide  spawning  stock  biomass  (SSB  xlO3  metric  tons  for  April-May  1.  Balloon  symbols 
are  observed  data  from  1995  to  2000. 


74 


Fishery  Bulletin  102(1) 


a  strong  role  in  controlling  YOY  recruitment  levels.  The 
four  highest  recruitment  years  in  our  series  had  the  lowest 
mean  subpycnocline  DO  levels  and  had  distribution  pat- 
terns of  SSB  that  differed  little  between  the  prespawning 
April-May  and  spawning  June-August  periods  (Fig.  4).  Al- 
though we  do  not  fully  understand  how  DO,  and  possibly 
hypoxic  conditions,  affect  migratory  behavior  and  distribu- 
tion patterns  of  bay  anchovy,  hypoxia  in  Chesapeake  Bay 
has  been  demonstrated  in  other  research  to  affect  spatial 
and  temporal  patterns  of  fish  abundance,  including  bay 
anchovy  (Breitburg,  1992;  Keister  et  al.,  2000). 

Our  third  hypothesis  proposes  that  predation  is  an  im- 
portant regulator  of  fish  recruitment  in  early-life  stages 
(Sissenwine,  1984;  Bailey  and  Houde,  1989).  We  hypoth- 
esize that  abundant  and  spatially  concentrated  larval 
or  juvenile  anchovy,  as  observed  in  the  lower  bay,  could 
promote  early-life  survival  by  satiating  predators,  even  if 
some  predators  migrate  to  areas  where  larval  and  juvenile 
anchovy  are  abundant.  At  mesoscale  distances  of  10-100 
km,  distribution  of  predators  (e.g.  YOY  and  age-1  weakfish 
[Cynoscion  regalis] )  may  be  important.  If  the  maximum 
number  of  prey  that  can  be  eaten  by  predators  is  reason- 
ably constant,  the  effect  of  predation  can  be  density-depen- 
satory  (Hilborn  and  Walters,  1992),  i.e.  predation  mortality 
rate  decreases  as  prey  density  increases. 

In  support  of  the  third  hypothesis,  a  correspondence 
analysis  on  fish  species  assemblages  by  year,  season,  re- 
gion, and  life  stage  (Jung  and  Houde,  2003)  indicated  that 
distributions  and  abundances  of  YOY  weakfish,  a  major 
predator  of  bay  anchovy  in  Chesapeake  Bay  (Hartman 
and  Brandt,  1995),  and  YOY  bay  anchovy  were  closely  as- 
sociated spatially,  seasonally,  and  annually  in  our  six-year 
study.  The  major  spawning  area  of  bay  anchovy  is  spatially 
restricted.  If  predator  migration  to  the  area  is  limited,  then 
as  the  supply  of  larvae  and  juveniles  increases,  it  may  satu- 
rate predator  demand,  the  condition  necessary  for  depensa- 
tion  to  be  important. 

It  may  seem  contradictory  to  propose  that  density-com- 
pensation with  respect  to  SSB  (the  negative  sign  of  j\) 
and  density-depensation  with  respect  to  AL  (the  second  or 
third  hypothesis )  can  act  simultaneously  during  larval  and 
juvenile  stages.  Under  this  circumstance,  the  number  of 
surviving  postlarval  anchovies  is  hypothesized  to  decrease 
because  of  food  limitation  when  larval  abundance  is  high, 
reducing  subsequent  predation-related  mortality  rate  on 
postlarvae  and  small  juveniles.  Low  abundance  of  anchovy 
early-life  stages  will  lead  to  the  opposite  effect  (Fig.  7).  The 
proposed  opposing  responses  of  the  early-larval  and  late- 
larval-juvenile  stages  are  explained  by  differences  in  the 
spatial  scales  of  distribution  and  densities  of  life  stages  of 
bay  anchovy  (Fig.  7).  The  spatial  scale  of  processes  that 
affect  distributions  of  late-stage  larvae  and  juveniles  is 
large  compared  to  that  for  early-stage  larvae  because  of 
the  increased  dispersal  and  swimming  ability  of  juveniles. 
Comparing  early-larval  and  late-larval-juvenile  stages  of 
bay  anchovy,  we  propose  that  effects  of  prey  concentration 
(the  first  hypothesis)  and  SSB  level  (density-compensa- 
tion) act  primarily  on  the  dynamics  of  early-larval  stages, 
whereas  predation  mortality  and  the  inhibitory  effects  of 
low  DO  (density-depensation;  the  second  and  third  hy- 


Nursery 
Ground 


(3)  Fall 


YOY  recruits, 
adults 


Late-stage  larvae, 
juveniles,  some  adults 
Eggs  and  larvae 


Overwintering 

Recruited 

anchovy 


Adult 

Immigration  from 
tributaries'? 


Major 

Spawning        Mature  adults. 

/  eggs,  larvae 

ground 


(1)  Spring 


Adult 

Immigration  from 
ocean? 


Figure  6 

Conceptual  model  representing  bay  anchovy  (Anchoa 
mitchilli)  life  cycle  and  ontogenetic  migration  within 
Chesapeake  Bay,  and  possible  immigration  of  adults 
from  tributaries  and  coastal  ocean. 


potheses)  are  more  important  regulators  and  controllers, 
respectively,  during  late-larval  and  juvenile  stages. 

The  three  hypotheses  that  relate  DO,  SSB  distribution, 
and  recruitment  of  bay  anchovy  are  not  mutually  exclusive. 
If  low  mean  DO  level  is  an  indicator  of  enhanced  prey  pro- 
duction and  availability  to  larvae  and  juveniles,  increased 
prey  productivity  in  the  lower  bay  could  enhance  bay 
anchovy  recruitment  potential  by  supplying  enough  zoo- 
plankton  prey  to  spawning  adults,  larvae,  and  juveniles.  At 
the  same  time,  low  mean  DO  in  the  mid-Bay  could  confine 
most  spawning  bay  anchovy  to  the  lower  bay.  thus  increas- 
ing spawning  and  larval  production  there,  and  possibly 
enhancing  survival  of  juveniles  by  predator  satiation.  Ul- 
timately, other  hypotheses  may  provide  better  explanations 
of  the  relationships  between  regional  mean  DO.  latitudinal 
shifts  in  distribution  of  spawners,  abundances  of  spawners. 
and  recruitment  of  bay  anchovy.  For  example,  abundant 
gelatinous  organisms,  such  as  the  scyphomedusa  (Chn'sa- 
ora  quinquecirra)  and  the  lobate  ctenophore  \Mnemiopsis 
leidyi),  can  be  important  predators  on  early-stage  anchovy 
and  competitors  with  juveniles  and  adults  (Purcell  et  al., 


Jung  and  Houde:  Recruitment  and  spawning-stock  biomass  distribution  of  Anchoa  mitchil/i  75 


Early-stage  and  larvae 

Density-compensatory 

Prey  is  smaller 

Small  scale  (1  m-10  km) 


Densiy  of  early-stage  larvae 
(1  m-10  m  scale) 


Late-stage  larvae  and  juveniles 
Density-depensatory 
Predator  is  bigger 
Mesoscale(IO-lOOkm) 


Recruits 

Ontogenetic  migration 


Densiy  of  late-stage  larvae 
(10-100  km  scale) 


SSB 


Figure  7 

Hypotheses  and  conceptual  model  of  the  bay  anchovy  {Anchoa  mitchilli)  recruitment  process  in 
Chesapeake  Bay.  The  density-compensatory  process  acts  at  a  small  spatial  scale  during  the  early- 
larval  stages,  whereas  the  density-depensatory  process  acts  at  a  broader  spatial  scale  during  late-stage 
larval  and  juvenile  stages.  The  ontogenetic  migration  is  controlled  by  dissolved  oxygen  levels  and  other 
hydrological  factors. 


1994),  but  their  potential  role  with  respect  to  bay  anchovy 
recruitment  could  not  be  defined  in  our  study.  For  the 
present,  it  is  clear  that  most  spawning  occurs  in  the  lower 
and  mid  Chesapeake  Bay,  from  which  larval  and  juvenile 
anchovies  disperse  upbay.  We  hypothesize  that  food  avail- 
ability is  the  major  factor  controlling  production  of  bay 
anchovy  early-larval  stages  whereas  predation  becomes 
more  important  during  late-larval  and  juvenile  stages. 
Our  results  and  hypotheses  implicate  density-related  pro- 
cesses, operating  at  different  spatial  scales,  as  regulators 
of  recruitment  of  bay  anchovy  in  Chesapeake  Bay. 


Acknowledgments 

We  thank  S.  Leach,  E.  North,  J.  Hagy,  C.  Rilling,  J.  Cleve- 
land, A.  Madden,  D.  O'Brien,  B.  Pearson,  D.  Craige,  T.  Auth, 
and  the  able  crew  of  RV  Cape  Henlopen  for  assistance  in 
field  surveys.  T.  Miller  and  E.  Russek-Cohen  provided  com- 
ments and  assistance  on  statistical  analyses.  This  research 
was  supported  by  a  U.S.  National  Science  Foundation,  Land 
Margin  Ecosystem  Research  (LMER)  program  grant, 
"Trophic  interactions  in  estuarine  systems  (TIES)"  (grant 
DEB94-12113).  Additional  support  was  provided  by  NSF 
Grant  OCE-9521512  and  by  National  Oceanic  and  Atmo- 
spheric Administration  grant  NOAA,  NA170P2656. 


Literature  cited 

Able,  K.  W„  and  M.  P.  Fahay. 

1998.    The  first  year  in  the  life  of  estuarine  fishes  in  the 


Middle  Atlantic  Bight.  342  p.     Rutgers  Univ.  Press,  New 
Brunswick,  NJ. 
Alheit,  J. 

1987.     Egg  cannibalism  versus  egg  predation:  their  signifi- 
cance in  anchovies.     S.  Afr.  J.  Mar.  Sci.  5:467-470. 
Bailey.  K.  M.,  and  E.  D.  Houde. 

1989.     Predation  on  eggs  and  larvae  of  marine  fishes  and  the 
recruitment  problem.    Adv.  Mar.  Bio.  25:1-83. 
Baird,  D..  and  R.  E.  Ulanowicz. 

1989.    The  seasonal  dynamics  of  the  Chesapeake  Bay 
ecosystem.     Ecol.  Monogr.  59:329-364. 
Belsley,  D.  A.,  E.  Kuh,  and  R.  E.  Welsch. 

1980.     Regression  diagnostics:  identifying  influential  data 
and  sources  of  collinearity,  292  p.     John  Wiley  and  Sons, 
Inc.,  New  York,  NY. 
Bhattacharya,  C.  G. 

1967.    A  simple  method  of  resolution  of  a  distribution  into 
Gaussian  components.     Biometrics  23:115-135. 
Breitburg,  D.  L. 

1992.     Episodic  hypoxia  in  Chesapeake  Bay:  Interacting 
effects  of  recruitment,  behavior,  and  physical  disturbance. 
Ecol.  Monogr.  62:525-546. 
Cowan,  J.  H.,  K.  A.  Rose,  E.  D.  Houde,  S.  B.  Wang,  and  J.  Young. 
1999.     Modeling  effects  of  increased  larval  mortality  on  bay 
anchovy  population  dynamics  in  the  mesohaline  Chesa- 
peake Bay:  evidence  for  compensatory  reserve.     Mar.  Ecol. 
Prog.  Ser.  185:133-146. 
Cronin.  W.  B. 

1971.    Volumetric,  areal,  and  tidal  statistics  of  the  Chesa- 
peake bay  estuary  and  its  tributaries,  15  p.     Chesapeake 
Bay  Institute,  Johns  Hopkins  Univ.,  Baltimore,  MD. 
Dorsey,  S.  E.,  E.  D.  Houde.  and  J.  C.  Gamble. 

1996.  Cohort  abundances  and  daily  variability  in  mortality 
of  eggs  and  yolk-sac  larvae  of  bay  anchovy,  Anchoa  mitchilli, 
in  Chesapeake  Bay.     Fish.  Bull.  94:257-267. 


76 


Fishery  Bulletin  102(1) 


Dovel,  W.  L. 

1971.     Fish  eggs  and  larvae  of  the  upper  Chesapeake 
Bay.  Chesapeake  Biological  Laboratory,  special  report  4, 
71  p.     Natural  Resource  Institute,  Univ.  Maryland.  College 
Park,  MD. 
Harding,  L.  W.,  M.  E.  Mallonee.  and  E.  S.  Perry. 

2002.    Toward  a  predictive  understanding  of  primary  produc- 
tivity in  a  temperate,  partially  stratified  estuary.     Estuar. 
Coast.  Shelf  Sci.  55:437-463. 
Hartman.  K.  J.,  and  S.  B.  Brandt. 

1995.     Predatory  demand  and  impact  of  striped  bass,  blue- 
fish,  and  weakfish  in  the  Chesapeake  Bay:  applications 
of  bioenergetics  models.     Can.  J.  Fish.  Aquat.  Sci.  52: 
1667-1687. 
Hilborn,  R..  and  C.  J.  Walters. 

1992.     Stock  and  recruitment.     In  Quantitative  fisheries 
stock  assessments:  choice,  dynamics,  and  uncertainty  (R. 
Hilborn  and  C.  J.  Walters,  eds. ).  p.  241-296.     Chapman  & 
Hall,  New  York.  NY. 
Houde,  E.  D. 

1987.     Fish  early  life  dynamics  and  recruitment  variability. 
Am.  Fish.  Soc.  Symp.  2:17-29. 
Houde,  E.  D.,  and  C.  E.  Zastrow. 

1991.  Bay  anchovy.  In  Habitat  requirements  for  Chesa- 
peake Bay  living  resources,  2nd  ed.  (S.  L.  Funderburk,  J. 
A.  Mihursky,  S.  J.  Jordan,  and  D.  Riley,  eds.),  p.  8.1  to  8.14. 
Living  Resources  Subcommittee,  Chesapeake  Bay  Program, 
Annapolis,  MD. 

Hunter,  J.  R.,  and  C.  A.  Kimbrell. 

1980.     Egg  cannibalism  in  the  northern  anchovy,  Engraulis 
mordax.     Fish.  Bull.  78:811-816. 
Jung,  S. 

2002.  Fish  community  structure  and  the  spatial  and  tem- 
poral variability  in  recruitment  and  biomass  production 
in  Chesapeake  Bay.  Ph.D.  diss.,  349  p.  Univ.  Maryland, 
College  Park,  MD. 

Jung,  S.,  and  E.  D.  Houde. 

2003.  Spatial  and  temporal  variability  of  pelagic  fish  com- 
munity structure  and  distribution  in  Chesapeake  Bay, 
U.S.A.     Estuar.  Coast.  Shelf  Sci.  58:335-351. 

Keister,  J.  E.,  E.  D.  Houde,  and  D.  L.  Breitburg. 

2000.     Effects  of  bottom-layer  hypoxia  on  abundances  and 
depth  distribution  of  organisms  in  Patuxent  River,  Chesa- 
peake Bay.    Mar.  Ecol.  Prog.  Ser.  205:43-59. 
Kemp,  W.  M.,  and  W.  R.  Boynton. 

1980.     Influence  of  biological  and  physical  processes  on  dis- 
solved oxygen  dynamics  in  an  estuarine  system:  implica- 
tions for  measurement  of  community  metabolism.     Estuar. 
Coast.  Mar.  Sci.  11:407-431. 
Kemp,  W.  M.,  P.  A.  Sampou.  J.  Garber.  J.  Tuttle.  ami 
W.  R.  Boynton. 

1992.  Seasonal  depletion  of  oxygen  from  bottom  waters  of 
Chesapeake  Bay:  roles  of  benthic  and  planktonic  respira- 
tion and  physical  exchange  processes.  Mar.  Ecol.  Prog. 
Ser.  85:137-152. 

Kimura,  R.,  D.  H.  Secor,  E.  D.  Houde,  and  P.  M.  Piccoli. 

200D      Cp-estuar\  dispersal  ofyoung-of-the-year bay  anchovy 
Anchoa  mitchilli  in  the  Chesapeake  Bay:  inferences  from 
microprobe  analysis  of  strontium  in  otoliths.     Mar.  Ecol. 
Prog  Ser  208:217-227. 
King,  M. 

1995.     Fisheries  biology,  assessment  and  management,  341  p. 
Fishing  News  Books,  Oxford. 
Klebasko.  M.J. 

1991.     Feeding  ecology  and  daily  ration  of  bay  anchovy. 


Anchoa  mitchilli  in  the  mid-Chesapeake  Bay.     M.S.  thesis, 
103  p.    Univ.  Maryland.  College  Park,  MD. 

Loos,  J.  J.,  and  E.  S.  Perry. 

1991.  Larval  migration  and  mortality  rates  of  bay  anchovy 
in  the  Patuxent  River.  In  Larval  fish  recruitment  and 
research  in  the  Americas:  proceedings  of  the  13th  annual 
fish  conference,  21-26  May  1989,  Merida.  Mexico  (R.  D. 
Hoyt,  ed.),  p.  65-76.    NOAA Technical  Report  NMFS  95. 

Luo,  J.,  and  S.  B.  Brandt. 

1993.  Bay  anchovy  Anchoa  mitchilli  production  and  con- 
sumption in  mid-Chesapeake  Bay  based  on  a  bioenergetics 
model  and  acoustic  measurement  offish  abundance.  Mar. 
Ecol.  Prog.  Ser.  98:223-236. 

Luo,  J.,  and  J.  Musick. 

1991.     Reproductive  biology  of  the  bay  anchovy  in  Chesa- 
peake Bay.     Trans.  Am.  Fish.  Soc.  120:701-710. 
MacGregor,  J.  M.,  and  E.  D.  Houde. 

1996.  Onshore-offshore  pattern  and  variability  in  distribu- 
tion and  abundance  of  bay  anchovy  Anchoa  mitchilli  eggs 
and  larvae  in  Chesapeake  Bay.  Mar.  Ecol.  Prog.  Ser.  138: 
15-25. 

Murdy,  E.  O,  R.  S.  Birdsong,  and  J.  A.  Musick. 

1997.  Fishes  of  Chesapeake  Bay,  324  p.  Smithsonian  Insti- 
tution Press,  Washington,  D.C.  and  London. 

Newberger,  T.  A.,  and  E.  D.  Houde. 

1995.  Population  biology  of  bay  anchovy  Anchoa  mitchilli  in 
the  mid  Chesapeake  bay.     Mar.  Ecol.  Prog.  Ser.  116:25-37. 

North,  E.  W„  and  E.  D.  Houde. 

In  press.  Transport  and  dispersal  of  bay  anchovy  {Anchoa 
mitchilli)  eggs  and  larvae  in  Chesapeake  Bay.  Estuar. 
Coast.  Shelf  Sci.. 

Parker,  K. 

1985.  Biomass  model  for  the  egg  production  method.  In  An 
egg  production  method  for  estimating  spawning  biomass  of 
pelagic  fish:  application  to  the  northern  anchovy,  Engraulis 
mordax  (R.  Lasker,  ed. ).  p.  5-7.    NOAA  Tech.  Rep.  NMFS  36. 

Peebles,  E.  B..  J.  R.  Hall,  and  S.  G  Tolley. 

1996.  Egg  production  by  the  bay  anchovy  Anchoa  mitchilli 
in  relation  to  adult  and  larval  prey  fields.  Mar.  Ecol.  Prog. 
Ser.  131:61-73. 

Purcell,  J.  E.,  D.  A.  Nemazie,  S.  E.  Dorsey,  E.  D.  Houde,  and 
J.  C.  Gamble. 

1994.  Predation  mortality  of  bay  anchovy  Anchoa  mitchilli 
eggs  and  larvae  due  to  scyphomedusae  and  ctenophores  in 
Chesapeake  Bay.    Mar.  Ecol.  Prog.  Ser.  1 14  :47-58. 

Ricker,  W.  E. 

1975.     Computation  and  interpretation  of  biological  statis- 
tics of  fish  population.     Bull.  Fish.  Res.  Board  Can.  191: 
1-382. 
Rilling,  G.  C,  and  E.  D.  Houde. 

1999a.     Regional  and  temporal  variability  in  distribution  and 
abundance  of  bay  anchovy  (Anchoa  mitchilli  i  eggs,  larvae, 
and  adult  biomass  in  the  Chesapeake  Bay.     Estuaries  22: 
1096-1109. 
1999b.     Regional  and  temporal  variability  in  growth  and 
mortality  of  bay  anchovy.  Anchoa  mitchilli.  larvae  in  Chesa- 
peake Bay.     Fish.  Bull.  97:555-569. 
Rose,  K.  A..  J.  H.  Cowan,  M.  E.  Clark.  E.  D.  Houde.  and 
S.  B.  Wang. 

1999.     An  individual-based  model  of  bay  anchovy  population 
dynamics  in  the  mesohaline  region  of  Chesapeake  Bay. 
Mar.  Ecol.  Prog.  Ser.  185:113-132. 
SAS  Institute  Inc. 

1989.  SAS/STAT  user's  guide,  version  6,  4th  ed.,  1686  p. 
SAS  Institute  Inc..  Gary,  NC. 


Jung  and  Houde:  Recruitment  and  spawning-stock  biomass  distribution  of  Anchoa  mitchilli 


77 


Samuels,  M.  L. 

1989.    Statistics  for  the  life  sciences,  p.  409-42.    Prentice- 
Hall.  Inc.,  Upper  Saddle  River,  NJ. 
Schultz.  E.  T.,  R.  K.  Cowen,  K.  M.  M.  Lwiza.  and 
A.  M.  Gospodarek. 

2000.  Explaining  advection:  Do  larval  bay  anchovy  lAnchoa 
mitchilli)  show  selective  tidal-stream  transport?  ICES  J. 
Mar.  Sci.  57:360-371. 

Sissenwine,  M.  P. 

1984.     Why  do  fish  populations  vary?     In  Exploitation  of 
marine  communities  (R.  M.  May,  ed.l,  p.  59-94.     Springer- 
Verlad,  Berlin. 
Smith,  E.  M.,  and  W.  M.  Kemp. 

2001.  Size  structure  and  the  production/respiration  balance 
in  a  coastal  plankton  community.  Limnol.  Oceanogr.  46: 
473-485. 

Steel,  R.  G.  D.,  and  J.  H.  Torrie. 

1980.  Principles  and  procedures  of  statistics.  A  biometrical 
approach.  2nd  ed.,  633  p.     McGraw-Hill  Inc.  New  York,  NY. 


Vazquez-Rojas,  A.  V. 

1989.     Energetics,  trophic  relationships  and  chemical  compo- 
sition of  bay  anchovy,  Anchoa  mitchilli  in  the  Chesapeake 
Bay.     M.S.  thesis,  166  p.     Univ.  Maryland,  College  Park, 
MD. 
Wang,  S.  B.,  J.  H.  Cowan,  K.  A.  Rose,  and  E.  D.  Houde. 

1997.     Individual-based  modelling  of  recruitment  variability 
and  biomass  production  of  bay  anchovy  in  mid-Chesapeake 
Bay.     J.  Fish  Biol.  51  (suppl.  A):121-134. 
Wang,  S.  B.,  and  E.  D.  Houde. 

1995.     Distribution,  relative  abundance,  biomass  and  produc- 
tion of  bay  anchovy  Anchoa  mitchilli  in  the  Chesapeake 
Bay.     Mar.  Ecol.  Prog.  Ser.  121:27-38. 
Zastrow,  C.  E.,  E.  D.  Houde,  and  L.  G.  Morin. 

1991.  Spawning,  fecundity,  hatch-date  frequency  and  young- 
of-the-year  growth  of  bay  anchovy  Anchoa  mitchilli  in  mid- 
Chesapeake  Bay.    Mar.  Ecol.  Prog.  Ser.  73:161-171. 


78 


Abstract— Increasing  interest  in  the 
use  of  stock  enhancement  as  a  man- 
agement tool  necessitates  a  better 
understanding  of  the  relative  costs  and 
benefits  of  alternative  release  strate- 
gies. We  present  a  relatively  simple 
model  coupling  ecology  and  economic 
costs  to  make  inferences  about  optimal 
release  scenarios  for  summer  flounder 
(Paralichthys  dentatus),  a  subject  of 
stock  enhancement  interest  in  North 
Carolina.  The  model,  parameterized 
from  mark-recapture  experiments, 
predicts  optimal  release  scenarios  from 
both  survival  and  economic  standpoints 
for  varyious  dates-of-release,  sizes-at- 
release,  and  numbers  of  fish  released. 
Although  most  stock  enhancement 
efforts  involve  the  release  of  relatively 
small  fish,  the  model  suggests  that 
optimal  results  (maximum  survival 
and  minimum  costs)  will  be  obtained 
when  relatively  large  fish  (75-80  mm 
total  length!  are  released  early  in  the 
nursery  season  (April).  We  investigated 
the  sensitivity  of  model  predictions  to 
violations  of  the  assumption  of  den- 
sity-independent mortality  by  includ- 
ing density-mortality  relationships 
based  on  weak  and  strong  type-2  and 
type-3  predator  functional  responses 
(resulting  in  depensatory  mortality 
at  elevated  densities).  Depending  on 
postrelease  density,  density-mortality 
relationships  included  in  the  model  con- 
siderably affect  predicted  postrelease 
survival  and  economic  costs  associated 
with  enhancement  efforts,  but  do  not 
alter  the  release  scenario  (i.e.  combina- 
tion of  release  variables )  that  produces 
optimal  results.  Predicted  (from  model 
output)  declines  in  flounder  over  time 
most  closely  match  declines  observed 
in  replicate  field  sites  when  mortality 
in  the  model  is  density-independent 
or  governed  by  a  weak  type-3  func- 
tional response.  The  model  provides  an 
example  of  a  relatively  easy-to-develop 
predictive  tool  with  which  to  make 
inferences  about  the  ecological  and 
economic  potential  of  stock  enhance- 
ment of  summer  flounder  and  provides 
a  template  for  model  creation  for  addi- 
tional species  that  are  subjects  of  stock 
enhancement  interest,  but  for  which 
limited  empirical  data  exist. 


Manuscript  approved  for  publication 
17  July  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:78-93  (2004). 


Coupling  ecology  and  economy: 
modeling  optimal  release  scenarios  for 
summer  flounder  (Paralichthys  dentatus) 
stock  enhancement 

G.  Todd  Kellison 

David  B.  Eggleston 

Department  ol  Marine,  Earth,  and  Atmospheric  Sciences, 

North  Carolina  State  University 

Raleigh,  North  Carolina  27695-8208 

Present  address  (for  G  T.  Kellison,  contact  author):  National  Park  Service/ Biscayne  National  Park 

9700  SW  328th  St,  Homestead,  Florida  33033 
E-mail  address  (for  G  T  Kellison)  todd_kellison  5  nps  gov 


Commercially  important  marine  fish 
and  invertebrate  populations  are 
declining  worldwide  in  response  to 
overexploitation  and  habitat  degrada- 
tion (Rosenberg  et  al„  1993;  FAO  1998). 
This  reduction  in  harvestable  fishery 
resources  has  stimulated  increasing 
interest  in  the  use  of  hatchery-reared 
(HR)  animals  to  enhance  wild  stocks 
(Munro  and  Bell,  1997;  Travis  et  al., 
1998;  Cowx,  1999;  Kent  and  Draw- 
bridge, 1999).  Unfortunately,  many  stock 
enhancement  programs  proceed  before 
ecological  concerns  are  adequately 
addressed  (Blankenship  and  Leber, 
1996),  and  without  the  identification 
of  goals  or  the  evaluation  of  the  success 
of  enhancement  efforts  (Cowx,  1999). 
If  fishery  managers  can  satisfactorily 
determine  that  enhancement  efforts 
will  have  no  ecologically  significant 
negative  ramifications,  then  managers 
should  establish  specific,  quantifiable 
goals  and  objectives  of  enhancement 
efforts  as  part  of  a  responsible  approach 
to  stock  enhancement  (Blankenship 
and  Leber,  1996;  Heppell  and  Crowder, 
1998).  Once  such  goals  have  been 
established,  managers  should  identify 
stocking  approaches  that  will  lead  to 
the  most  cost-efficient  realization  of 
enhancement  goals — a  process  that 
can  be  accomplished  with  the  aid  of 
coupled  ecological  and  economic  models. 
Although  numerous  (advanced)  models 
(conceptual  and  species-specific)  exist 
to  predict  the  biological  and  ecological 
impact  of  alternative  enhancement 
scenarios  (e.g.  Botsford  and  Hobbs, 
1984;  Salvanes  et  al„  1992;  Barbeau 
and  Caswell,  1999;  Sutton  et  al.,  2000), 


there  are  few  models  ( of  which  we  are 
aware)  that  have  attempted  to  link  the 
biological  and  ecological  results  of  stock- 
ing efforts  (e.g.  addition  of  biomass  to  a 
stocked  population)  with  the  economic 
costs  associated  with  various  release 
scenarios  (e.g.  Botsford  and  Hobbs,  1984; 
Hobbs  et  al.,  1990;  Hernandez-Llamas, 
1997;  Kent  and  Drawbridge,  1999).  Such 
a  link  is  critical  to  the  responsible  use 
of  funding  to  rebuild  or  manage  fisher- 
ies, and  for  the  comparison  of  predicted 
costs  of  enhancement  versus  alternative 
management  techniques. 

In  North  Carolina,  there  has  been 
recent  interest  in  stock  enhancement 
with  summer  flounder  (Paralichthys 
dentatus)  (Waters,  1996;  Rickards, 
1998;  Waters  and  Mosher,  1999;  Burke 
et  al.,  2000;  Copeland  et  al. ' )  because  of 
a  combination  of  heavy  commercial  and 
recreational  exploitation,  established 
techniques  for  mass  hatchery-rearing 
(Burke  et  al.,  1999),  and  considerable 
knowledge  of  summer  flounder  life  his- 
tory (Powell  and  Schwartz,  1977;  Burke 
et  al.,  1991;  Burke,  1995).  Nevertheless, 
there  have  been  no  large-scale  release 
experiments  ( and  subsequent  collection 
of  data)  by  which  to  make  empirical 
inferences  about  stock  enhancement 
potential  for  this  species.  We  present 
a  compartmental  model,  parameterized 
from  mark-recapture  field  experiments, 


Copeland,  B.  J.,  J.  M.  Miller,  and  E.  B. 
Waters.  1998.  The  potential  for  flounder 
and  red  drum  stock  enhancement  in  North 
Carolina.  Summary  of  workshop,  30-31 
March.  1998,  22  p.  '  (Available  from  North 
Carolina  State  Univ,  Raleigh.  NC  27695.] 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Paralichthys  dentatus 


79 


Table  1 

Range  of  numbers  of  summer  flounder  (Paralichthys  dentatus)  released  (and  resulting  postrelease  densities),  sizes-at-release,  and 
dates  of  release  simulated  in  the  model. 


Number  released 


Postrelease  density 


Size-at-release 


Dates  of  release 


100-400,000 


0.001-4.0 


30-80  mm 


1  April-15  July 


that  incorporates  size  of  fish  released,  date-of-release,  and 
number  offish  released  to  calculate  1)  predicted  numbers  of 
survivors  and  2 )  economic  costs  associated  with  varying  re- 
lease scenarios  under  density-independent  mortality.  We  in- 
vestigated the  sensitivity  of  model  predictions  to  violations 
of  the  assumption  of  density-independent  mortality  because 
there  is  abundant  evidence  that  mortality  rates,  or  processes 
underlying  mortality  rates  (e.g.  growth),  are  affected  by  den- 
sity-dependent relationships  in  the  wild  ( see,  for  recent  ex- 
amples. Bucket  et  al.,  1999;  Bystroem  and  Garcia-Berthou. 
1999;  Jenkins  et  al,  1999;  Kimmerer  et  al.,  2000).  We  did 
so  by  repeating  model  simulations  under  varying  density- 
mortality  relationships  (depensatory  in  nature  at  elevated 
densities ),  using  experimental  evidence  from  our  own  field 
studies  and  published  observations  for  similar  species  to 
parameterize  density-mortality  relationships.  Additionally, 
we  used  a  scenario  in  which  the  density-mortality  relation- 
ship changed  over  time  to  make  inferences  about  the  effect  of 
more  complex  density-mortality  relationships  on  postrelease 
mortality  of  juvenile  summer  flounder.  Finally,  we  generated 
predicted  temporal  patterns  of  field  densities  under  vary- 
ing density-mortality  relationships  and  compared  them  with 
observed  (in  the  field)  patterns  to  determine  whether  model 
output  under  the  considered  density-mortality  relationships 
matched  actual  patterns  in  the  field.  The  model  provides 
an  example  of  a  relatively  easy-to-develop  predictive  tool 
with  which  to  make  inferences  about  the  ecological  and 
economic  potential  of  stock  enhancement  with  summer 
flounder  and  provides  a  template  for  model  creation  for 
additional  species  that  are  subjects  of  stock  enhancement 
interest,  but  for  which  limited  empirical  data  exist. 


Materials  and  methods 

Background 

In  North  Carolina,  wild  summer  flounder  recruit  to  shal- 
low-water estuarine  nursery  habitats  from  February  to 
May,  after  which  small  juvenile  (20-35  mm  total  length 
[TL] )  densities  range  from  -0.1  to  1.0  fish/m2  (Burke  et  al., 
1991;  Kellison  and  Taylor2).  Juveniles  subsequently  make 
an  ontogenetic  habitat  shift  to  deeper  waters  ( Powell  and 
Schwartz,  1977),  apparently  after  reaching  a  total  length 


2  Kellison,  G.  T.,  and  J.  C.  Taylor.  2000.  Unpubl.  data.  De- 
partment of  Marine,  Earth,  and  Atmospheric  Sciences,  North 
Carolina  State  University,  Raleigh,  NC  27695-8208. 


of  -80  mm  (Kellison  and  Taylor2).  By  mid-July,  densities 
of  juvenile  summer  flounder  in  the  shallow  water  nursery 
habitats  are  near  zero  (Kellison  and  Taylor2). 

Model  pathway 

Our  compartmental  model  simulated  the  daily  mortality 
and  growth  of  different-size  hatchery-reared  (HR)  fish 
released  in  the  field  over  a  105-day  period  ( 1  April  to  15 
July,  based  on  observed  field  abundances)  in  a  hypotheti- 
cal release  habitat  of  10  hectares.  The  model  predicted  the 
percentage  of  released  fish  surviving  and  economic  cost- 
per-survivor  under  2730  release  scenarios  for  a  specified 
number  offish  released  (see  below).  To  begin  the  model,  a 
value  of  number  offish  released  (NFR)  ranging  from  100  to 
400,000  (Table  1)  was  chosen  (Fig.  1),  resulting  in  postre- 
lease densities  (assuming  even  postrelease  distribution)  of 
0.001-4.0  fish/m2.  These  values  included  a  range  of  densi- 
ties of  juvenile  summer  flounder  observed  in  wild  nursery 
habitats  ( -0-1  fish/m2;  mean  -0.05  fish/m2;  Kellison  and 
Taylor2),  but  also  included  unusually  high  densities  (>1 
fish/m2)  in  order  to  examine  how  such  release  strategies 
would  affect  model  output  (we  did  not  examine  densities 
>4  fish/m2  because  of  a  lack  of  data  on  fish  response  to 
resource  limitation  likely  to  occur  as  densities  increased 
past  values  for  which  we  had  empirical  growth  data).  Each 
group  of  NFR  was  initially  assigned  a  "size-(TL)  at-release" 
of  30  mm  (the  smallest  size-at-release  simulated  in  the 
model),  after  which  a  size-dependent  economic  cost  associ- 
ated with  the  release  of  the  30-mm-TL  fish  was  calculated 
(see  below).  The  release  group  was  then  assigned  a  mini- 
mum Julian  "day  of  release"  of  92  (corresponding  to  1  April, 
the  earliest  release  date  simulated  in  the  model).  A  range 
of  Julian  days  of  release  was  included  in  the  model  because 
field-estimated  growth  rates  were  dependent  on  Julian  day 
(Kellison,  2000),  and  growth  rates  are  potentially  impor- 
tant to  the  determination  of  mortality  rates  (Rice  et  al. 
1993).  With  this  model,  we  then  calculated  daily  mortality 
and  growth  (described  below)  in  the  hypothetical  release 
habitat  over  the  number  of  days  at  large  (DAL),  where 

DAL  =  197  (the  Julian  day  corresponding  to  15  July)  -  92 
(Julian  release  day), 

and  output  a  number  of  survivors  and  a  calculated  cost- 
per-survivor  (CPS),  where 

CPS  =  cost  associated  with  release  -f 
predicted  number  of  survivors, 


80 


Fishery  Bulletin  102(1 I 


Input 

number  released 

(NR) 


'  assign  size-at-release  (SAR) 
*  calculate  cost  of  release 

(COR) 


<— 


Size-at-release 


N 


Density- 
independent 


Julian  day 


'  assign  date  of  release 
(DOR) 


< 


I 

'  determine 

number  of  survivors    (NOS) 

DAL 

at  the  beginning  of  the  day  (= 

«\ 

initial  #  of  fish  or  #  surviving 

from  previous  day) 

/ 
/ 
1 

da 

ly  mortality       ^ 

da 

ly  growth 

*M 

*  calculate 

number  of  survivors  and 

total  length  (TL) 

at  the  end  of  the  day 

1 

I  I 


*  output 

-  number  of  survivors 

-  cost  per  survivor  (CPS) 


\ 


/ 


Figure  1 

Model  flowchart.  Dashed  arrows  represent  model  "backloops"  to  the  indicated  compartment  where 
simulations  continue  with  the  next  value  of  the  arrow-labeled  variable.  Side  graphs  indicate  the  three 
relationships  between  density  and  mortality  (number  offish  consumed)  that  were  considered,  and  the 
general  relationship  between  growth  and  Julian  day. 


for  the  initial  release  scenario  of  fish  size  =  30  mm  TL. 
Julian  day  =  92,  and  an  NFR  input  determined  by  the  mod- 
eler). The  model  then  looped  back  to  the  "date-of-release" 
step  and  simulated  the  release  of  the  30-mm-TL  fish  for 
Julian  release  days  93-197,  outputting  a  predicted  number 
of  survivors  and  cost-per-survivor  for  each  release  date.  The 
model  then  repeated  all  previous  steps  under  sequentially 
larger  size-at-release  scenarios,  looping  back  to  the  "size- 
at-release"  step  and  simulating  the  release  of  fish  ranging 
in  size  from  32-80  mm  TL  fish  in  steps  of  2  mm  TL.  The 
model  output  was  a  predicted  number  of  survivors  and 
economic  cost-per-survivor  for  each  release  day  (92-197) 
for  each  size-at-release  (Fig.  1).  Thus,  for  each  input  NFR, 
there  were  26  size-at-release  possibilities  x  105  Julian  days 
of  release  possibilities,  which  resulted  in  2730  simulations, 
each  of  which  resulted  in  a  predicted  number  of  survivors 
and  cost-per-survivor  for  that  particular  release  scenario. 
For  each  input  NFR,  the  results  from  the  2730  simulations 
were  plotted  on  two  response  surfaces,  with  an  .v-axis  of 


size-at-release,  a  y-axis  of  date-of-release,  and  a  2-axis  of 
either  1)  predicted  number  of  survivors  (NOS),  or  2)  cost- 
per-survivor  ( CPS ),  to  identify  release  scenarios  resulting  in 
the  maximum  predicted  number  of  survivors  and  minimum 
cost-per-survivor,  respectively.  The  scenarios  resulting  in 
the  maximum  predicted  number  of  survivors  and  minimum 
cost-per-survivor  were  not  necessarily  identical. 

Calculation  of  mortality,  growth,  survival,  and  economic 
costs  associated  with  release 

During  each  day  at  large  (DAL),  released  fish  were  sub- 
jected to  a  density-independent  daily  mortality  rate  of 
0.02153,  derived  from  postrelease  mark-recapture  data 
of  HR  summer  flounder  (Kellison  et  al.,  2003b).  In  deriv- 
ing this  value,  mean  postrelease  densities  were  used  to 
estimate  a  total  number  of  survivors  from  experimental 
releases.  Daily  survival  was  then  calculated  with  the 
equation 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Paraltchthys  dentatus 


81 


NFR  x  SDDAL  =  NOS, 

where  NFR  =  number  released; 
SD  =  daily  survival; 
DAL  -  days  at  large  (from  release  date 

until  Julian  day  197);  and 
NOS  =  estimated  number  of  survivors. 

Daily  mortality  (MD)  was  then  calculated  from 
the  equation 


Mr 


1-Sr 


At  the  end  of  each  simulated  day,  all  fish  that 
were  alive  increased  in  growth  according  to  the 
equation 

GD  =  -0.0061  x  Julian  day  +  1.2487, 

which  was  derived  from  mark-recapture  data 
(Kellison,  2000),  and  in  which  GD  is  daily  growth 
in  millimeters.  Fish  reaching  80  mm  TL  during  the  model 
(i.e.  by  15  July)  were  considered  to  make  an  ontogenetic  hab- 
itat shift  to  deeper  waters.  These  fish  were  then  subjected 
to  one  half  year  of  natural  mortality  to  simulate  mortality- 
related  losses  from  deeper-water  habitats  (M=0.28;  Froese 
and  Pauley,  2001).  Remaining  fish,  now  having  survived 
-one  year  of  natural  mortality,  were  considered  to  be  sur- 
vivors (available  to  the  commercial  fishery),  which  is  a  con- 
servative assumption  because  1-yr-old  summer  flounder  are 
only  partially  recruited  to  the  commercial  fishery.  All  fish  not 
reaching  a  total  length  of  80  mm  were  assumed  to  perish. 

To  determine  size-dependent  economic  costs  offish  pro- 
duction, we  used  the  following  regression  equation  derived 
for  Japanese  flounder  (Paralichthys  olivaceus)  by  Sproul 
and  Tominaga  ( 1992 )  because  equivalent  economic  data  for 
summer  flounder  were  unavailable: 

CPF  =  14.24  +  1.234  x  TL, 

where  CPF  =  the  cost  per  fish  in  Japanese  yen  (¥);  and 
TL   =  the  total  length  of  the  HR  fish. 

Costs  were  then  converted  into  US$  by  using  an  exchange 
rate  of  106. 7¥  per  1  US$  (universal  currency  converter). 
We  feel  use  of  this  cost-of-fish-production  equation  is  appro- 
priate because  the  Japanese  flounder  is  closely  related 
and  similar  in  life  history  traits  to  the  summer  flounder 
(Tanakaet  al.,  1989;  Burke  etal.,  1991 ),  resulting  in  similar 
optimal  rearing  practices  for  hatchery-reared  Japanese  and 
summer  flounder  (Burke  et  al.,  1999),  and  thus  likely  simi- 
lar rearing  costs.  Additionally,  the  scale  of  Japanese  floun- 
der hatchery  production  is  similar  to,  or  greater  than,  other 
government  subsidized  hatchery  production  programs  (e.g. 
red  drum  in  Texas,  cod  in  Norway  [Svasand,  1998] ). 

Density-mortality  relationships 

We  tested  the  sensitivity  of  the  model  results  (optimal 
predicted  number  of  survivors  and  cost-per-survivor  esti- 
mates under  varying  NFRs)  to  violations  of  the  assumption 


0.50  -i 

~   0.40  ■ 

*  Type  2  -  weak 

k                                                                                      *  Type  2  -  strong 

E      0.30  - 

*        ^—^-^                                                            d  Type  3  -  weak 

ra 

k    f        ^^^^                                                   ■  Type  3  -  strong 

o 

t      0.20  ■ 

o 

Q. 

O 

*#                                            ^^^^ 

o-     0.10  -j 

|^»W                           °°OOnnn„ 

0                           12                          3                          4 

Density  (number  of  fish/m2) 

Figure  2 

Proportional  mortality  curves  for  juvenile  summer  flounder  corre- 

sponding to  weak  and  strong  type-2  and  type-3  mortality  responses. 

of  density-independent  mortality  by  incorporating  varying 
types  and  strengths  of  density-dependent  mortality  (depen- 
satory  in  nature  at  elevated  densities;  see  below)  into  the 
model.  As  a  basis  for  these  sensitivity  analyses,  we  assumed 
that  predation  was  the  driving  mechanism  underlying  the 
postrelease  mortality  of  HR  summer  flounder  under  the 
densities  examined  (Kellison  et  al.,  2000;  Kellison  et  al., 
2003b).  Thus,  we  made  daily  mortality  rates  correspond 
to  either  a  type-2  or  type-3  predator  functional  response 
(Holling,  1959;  see  Lindholm  et  al.,  2001  for  example),  in 
which  proportional  mortality  due  to  predation  decreases 
with  increasing  density  (type-2  response)  or  increases  ini- 
tially with  increasing  density,  reaches  a  zenith,  and  then 
decreases  with  increasing  density  (type-3  response)  (Fig. 
2).  Both  type-2  and  type-3  responses  result  in  decreasing 
(depensatory)  mortality  at  elevated  prey  densities  due  to 
predator  satiation.  We  did  not  include  scenarios  in  which 
mortality  increased  at  elevated  densities  (as  would  be 
expected  when  densities  reached  those  likely  to  result  in 
resource  limitation )  because  we  did  not  include  in  the  model 
elevated  release  densities  likely  to  result  in  resource  limita- 
tion. We  parameterized  the  daily  mortality  curves  so  that 
each  response  (type  2  or  3)  incorporated  the  daily  mortality 
rate  of  0.02153.  These  mortality  curves  contain  mortality 
values  that  are  within  ranges  reported  in  the  literature  for 
other  species  of  juvenile  marine  fishes  (Bax,  1983;  Houde, 
1987;  Nash,  1998;  Rose  et  al,  1999).  To  make  further  infer- 
ences about  the  importance  of  density-dependent  mortal- 
ity to  model  results,  we  included  a  1)  weak  and  2)  strong 
form  of  each  functional  response  (types  2  and  3)  (Fig.  2),  as 
well  as  scenarios  in  which  the  response  shifted  temporally 
from  3)  type  2  to  3,  and  4)  type  3  to  2  at  the  midpoint  of 
the  nursery  season  (Julian  day  145).  We  included  both  the 
weak  and  strong  forms  of  the  type-2  and  type-3  functional 
responses  to  determine  the  extent  to  which  variation  in  the 
strength  of  the  functional  response  would  affect  model  pre- 
dictions. The  strength  of  the  functional  response  could  vary 
because  of  annual  variation  in  the  presence  or  abundance 
of  prey  or  because  predators  could  affect  the  density-mor- 
tality relationship  (see,  for  example,  Hansen  et  al.,  1998). 


82 


Fishery  Bulletin  102(1) 


For  example,  a  strong  positive  (compensatory)  density- 
mortality  relationship  driven  by  predators  might  become 
weaker  in  years  when  predator  abundance  was  lower  than 
average.  We  included  the  temporally  shifting  functional 
response  scenarios  to  determine  the  extent  to  which  tem- 
poral variation  in  the  form  of  the  functional  response  would 
affect  model  predictions.  Temporal  variation  in  the  form  of 
the  functional  response  might  occur  because  of  temporal 
changes  in  the  predator  community,  or  because  of  changing 
predator-prey  size  dynamics  (e.g.  Stoner,  1980;  Black  and 
Hairston,  1988).  For  example,  as  the  nursery  season  for 
summer  flounder  progresses,  proportionately  greater  num- 
bers of  juveniles  grow  to  sizes  at  which  they  are  capable 
of  preying  on  smaller  juveniles  (Kellison,  personal  obs. ).  If 
cannibalistic  summer  flounder  exhibit  a  different  predatory 
functional  response  from  that  of  the  predator  guild  commu- 
nity predominating  earlier  in  the  season,  then  the  density- 
mortality  relationship  may  change  seasonally. 

We  replicated  all  model  simulations  over  each  of  the  six 
density-mortality  relationships  (weak  and  strong  types  2 
and  3,  and  shifting  patterns  [type  2  to  3  and  type  3  to  2] ) 
to  determine  optimal  release  scenarios  (maximum  num- 
ber of  survivors,  minimum  cost-per-survivor)  under  each 
relationship.  We  then  compared  results  to  those  obtained 
under  density-independent  mortality  to  make  inferences 
about  the  importance  of  density-mortality  relationships  to 
model  results. 

Correspondence  between  predicted  and 
observed  temporal  abundance  patterns 

Different  density-mortality  relationships  may  result  in 
distinct  temporal  patterns  of  abundance  (e.g.  rapid  versus 
more  gradual  declines  in  abundance)  depending  on  initial 
densities.  We  generated  predicted  patterns  of  temporal 
field  abundance  of  juvenile  summer  flounder  under  den- 
sity-independent mortality  and  four  additional  density- 
mortality  relationships  (governed  by  weak  and  strong  type 
2  and  3  functional  responses)  and  under  varying  initial 
densities  (0.1,  0.3,  and  0.5  fish/m2)  to  examine  whether  the 
different  density-mortality  relationships  would  result  in 
distinct  temporal  patterns  of  abundance.  We  used  1998-99 
field  data  and  logarithmic  or  polynomial  regression  models 
to  generate  curves  that  best  fitted  (based  on  r2  values) 
observed  (from  natural  nursery  sites)  temporal  declines  in 
abundance  under  varying  initial  densities.  We  compared 
the  best-fit  curves  to  those  predicted  by  the  model  under 
density-independent  and  four  additional  density-mortal- 
ity relationships.  These  comparisons  allowed  us  to  make 
qualitative  inferences  about  which  density-mortality 
relationship* s)  resulted  in  the  best  match  between  pre- 
dicted and  observed  temporal  patterns  of  abundance. 

Model  assumptions 

The  assumptions  of  the  model  are  the  following: 

1  Daily  mortality  is  independent  of  size.  Although  there 
is  strong  evidence  that  mortality  of  fishes  in  the  wild  is 
size-dependent  (Lorenzen,  2000 ),  particularly  in  regard 


to  the  importance  of  size  to  susceptibility  to  predation 
(see,  for  example,  Elis  and  Gibson,  1995;  Furuta,  1999; 
Manderson  et  al.,  1999),  we  found  no  evidence  (from 
recaptures  of  released  hatchery-reared  fish )  of  size- 
selective  daily  mortality  for  juvenile  summer  flounder 
ranging  in  size  from  -30-80  mm  TL  in  shallow-water 
nursery  areas  (Kellison  et  al.,  2003a).  Implications  for 
violations  of  this  assumption  are  addressed  in  the  "Dis- 
cussion" section. 

2  Daily  growth  is  independent  of  fish  density.  We  based 
this  assumption  on  field  experiments  that  indicated 
no  growth  limitation  at  densities  roughly  equal  to  the 
maximum  densities  explored  in  the  model  (Kellison 
et  al.,  2003b).  Similar  findings  (i.e.  no  food-limitation 
or  density-dependent  growth)  have  been  reported  for 
similar-size  plaice  in  shallow-water  nursery  habitats 
(van  der  Veer  and  Witte,  1993). 

3  Economic  cost  per  fish  (CPF)  is  independent  of  the 
number  of  fish  acquired  for  release  (i.e.  within  the 
range  of  numbers  offish  released  in  model  simulations, 
there  is  no  decrease  in  cost  per  fish  as  the  number  of 
fish  acquired  from  the  production  hatchery  for  release 
increases).  This  assumption  is  likely  to  be  valid  over 
changes  in  numbers  of  fish  released  common  to  stock 
enhancement  programs  (Sproul  and  Tominaga,  1992) 
but  may  not  be  valid  as  numbers  released  change 
over  orders  of  magnitude  because  of  economy  of  scale 
(Adams  and  Pomeroy  1991;  Garcia  et  al.,  1999). 

4  There  is  no  emigration  from  the  release  habitat  until 
fish  exhibit  an  ontogenetic  shift  in  habitat  at  80  mm  TL. 
Although  pre-ontogenetic  habitat  shift  emigration  may 
not  truly  be  zero,  we  feel  that  it  is  also  unlikely  that  pre- 
ontogenetic  habitat-shift  emigration  accounts  for  more 
than  a  minimal  amount  of  loss  of  released  fish  from 
the  habitat  of  release,  as  supported  by  several  points. 
First,  rates  of  pre-ontogenetic  shift  emigration  in  wild 
juveniles  are  apparently  low  (Kellison  and  Taylor2), 
suggesting  that  large-scale  spatial  migrations  may  not 
be  part  of  the  behavioral  repertoire  of  early  juvenile 
summer  flounder.  Second,  irregular  temporally  repli- 
cated sampling  outside  of  experimental  release  sites 
resulted  in  zero  captures  of  emigrating  hatchery-reared 
fish  (Kellison  et  al.,  2003b).  Third,  emigration  rates  of 
closely  related  HR  Japanese  flounder  {Paralichthys 
olivaceus)  are  reported  to  be  very  low  (Tominaga  and 
Watanabe,  1998).  In  combination,  these  points  suggest 
that  our  zero  emigration  assumption  is  appropriate. 

5  Fish  that  do  not  grow  to  80  mm  TL  during  the  model 
period  (i.e.  by  15  July)  do  not  survive.  Although  this 
assumption  cannot  be  examined  with  our  field  data, 
data  do  show  that  juvenile  summer  flounder  are 
absent  from  shallow-water  nursery  habitats  by  mid 
to  late  July  (Kellison  et  al.3).  Thus,  all  fish  have  either 
perished  or  made  ontogenetic  habitat  shifts  to  deeper 
habitats  by  this  time.  Our  field  observations  suggest 
that  the  deeper  habitats  to  which  larger  flounder 


:t  Kellison,  G.  T.,  J.  C.  Taylor,  and  J.  S.  Burke.  2000.  Unpubl. 
data.  Department  of  Marine,  Earth,  and  Atmospheric  Sciences, 
North  Carolina  State  Univ.,  Raleigh,  NC  27695-8208. 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Paralichthys  dentatus 


83 


make  ontogenetic  habitat  shifts  are 
inhabited  by  relatively  high  densities  of 
potential  predators  (e.g.  blue  crabs,  age  1+ 
flounders,  red  drum  [Sciaenops  ocellatus], 
searobin  [Prionotus  sp.],  and  lizardfish 
[Synodus  sp.] ),  which  may  be  considerably 
less  abundant  in  shallow-water  habitats. 
These  relatively  large  and  abundant 
predators  would  presumably  expose  small 
migrating  fish  to  high  rates  of  predation 
(see,  for  example,  Elis  and  Gibson,  1995; 
Furuta,  1999;  Manderson  et  al„  1999).  This 
assumption  is  supported  by  research  with 
the  congener  Japanese  flounder  (Paralich- 
thys olivaceus).  Although  a  range  of  sizes 
of  hatchery-reared  Japanese  flounder  may 
survive  within  relatively  shallow  nursery 
habitats,  fishes  less  than  90  mm  TL  moving 
into  relatively  deep  waters  are  poorly  rep- 
resented in  subsequent  age  classes,  most 
likely  due  to  predation-induced  mortality 
(Yamashita  et  al.,  1994;  Furuta,  1999). 
There  is  no  relationship  between  length 
of  rearing  period  (time  spent  in  the 
hatchery  environment)  and  probability  of 
postrelease  mortality  related  to  behavioral 
deficits  (Olla  et  al.,  1998).  Hatchery-specific 
selection  pressures  may  result  in  HR  fish 
that  are  behaviorally  selected  to  survive  in 
the  hatchery  and  not  in  the  wild  (see  Olla 
et  al.,  1998;  Kellison  et  al.,  2000;  for  discus- 
sion). We  assume  that  behavioral  deficits 
are  not  exacerbated  with  time  spent  in  the 
hatchery  (i.e.  behavioral  deficits  are  equal 
for  all  sizes-at-release). 


Results 

The  most  important  factor  affecting  the 
number  of  survivors  (and  therefore  percent 
survival)  was  size-at-release  because  the 
greatest  numbers  and  percentages  of  survi- 
vors were  always  produced  by  releasing  the 
largest  fish  possible  (80  mm  TL  in  the  model). 
Number  of  survivors  decreased  with  decreas- 
ing size-at-release  and  with  increasing  Julian 
day  of  release  (Fig.  3A).  The  cost-per-survivor 
( CPS )  was  also  most  affected  by  size-at-release, 
such  that  CPS  decreased  with  increasing  size- 
at-release  (Fig.  3B).  CPS  generally  increased 
with  increasing  Julian  day  of  release  (Fig.  3B),  although 
this  effect  was  less  important  than  the  effect  of  size-at- 
release.  Because  mortality  was  originally  assumed  to  be 
density-independent,  the  optimal  cost-per-survivor  did 
not  vary  with  the  number  offish  released  (Fig.  4),  and  the 
relationship  between  number  offish  released  and  number 
of  survivors  was  linear  (Fig.  4),  such  that  the  maximum 
number  of  survivors  were  generated  from  the  greatest 
number  offish  released  (NFR=400,000). 


220 


20      80 


90     220 


Figure  3 

Response  surfaces  of  iAi  number  offish  survivors  (summer  flounder  I 
and  (Bi  cost-per-survivor  (CPS)  as  a  function  of  date  of  release  and  size 
at  release  at  number  released  (NR)  =  5000  (postrelease  density=0.05) 
under  density-independent  mortality.  CPS  values  greater  than  $10  were 
set  equal  to  $10  for  ease  of  presentation. 


Sensitivity  of  model  predictions  to  violations  of 
density-independent  mortality  assumption 

Model  results  varied  considerably  under  the  various  den- 
sity-mortality relationships  (Fig.  5,  A  and  B),  indicating 
the  importance  of  knowledge  of  the  relationship  between 
numbers  of  fish  released  (density)  and  mortality  in  the 
wild  to  predicting  optimal  release  scenarios.  Variation  in 
model  output  was  dependent  on  the  type  and  strength  of 


Fishery  Bulletin  102(1) 


the  density-mortality  relationship.  For  example,  at  postre- 
lease densities  of  0.5  fish/m2  (NFR=50,000),  survival  of 
released  flounder  under  density-independent  mortality 
was  ~28%  higher  than  that  predicted  under  strong  type-3 
mortality,  but  only  -2%  higher  than  that  predicted  under 
weak  type-2  mortality  (Fig.  5A).  At  postrelease  densities 
of  0.001  fish/m2  (NFR=100),  survival  of  released  flounder 
under  density-independent  mortality  was  ~41%  higher 


450000  -I 
m      400000  • 
§      350000  • 
£      300000  • 
«      250000  ; 
°      200000  ■ 
E      150000  ■ 
|      100000  ■ 
z        50000  ■ 

— ■ —  optimal  number  of  survivors                      : 
— o—  optimal  CPS                             ^^^" 

r  1  60 

:  1  50      O 

■  1.40      g 

■  1  30    -g 
-1,20      5 
[110      <§ 
-  1  00      g 
■0.90      < 
■0  80      - 

■  0.70      C 

■  0-60      W 

0        50000    10000    15000    20000    25000    30000    35000 

40000 

Number  released 

Figure  4 

Optimal  number  of  fish  survivors  and  cost-per-survivor  as  a  function  of 
varying  numbers  of  summer  flounder  released  under  density-indepen- 
dent mortality. 

0  12  3  4  5 

Density  (number  of  fish/m2) 

Figure  5 

l A i  Optimal  percent  survival  and  iBi  optimal  cost-per-survival  (US$)  as  a  func- 
tion of  postrelease  density  undci  density-independent  and  varying  density- 
dependent,  mortality  relationships  for  summer  flounder. 


than  that  predicted  under  strong  type-2  mortality,  but  -2% 
less  than  that  predicted  under  strong  type-3  mortality  ( Fig. 
5A).  In  contrast,  when  postrelease  densities  were  relatively 
high,  there  was  less  of  an  impact  of  density-mortality  rela- 
tionship on  postrelease  survival  and  costs  associated  with 
stock  enhancement.  For  example,  at  postrelease  densities 
of  three  fish/m2  (NFR=300,000),  survival  of  released  floun- 
der differed  by  less  than  4%  between  density-independent, 
weak  or  strong  type-2,  and  weak  type-3  mor- 
tality, although  survival  under  strong  type-3 
mortality  was  ~99c  less  than  that  predicted 
under  density-independent  mortality  and 
-11%  less  than  that  predicted  under  strong 
type-2  mortality  (Fig.  5A).  Thus,  the  model 
results  were  most  sensitive  to  violations  of  the 
assumption  of  density-independent  mortality 
at  low  densities  offish  released  in  the  field. 


Type-2  mortality  As  with  density-indepen- 
dent mortality,  the  most  important  factor 
affecting  number  of  survivors  and  cost  per 
survivor  under  type-2  mortality  was  size-at- 
release  (Fig.  6,  A  and  B).  In  all  simulations, 
the  greatest  number  of  survivors  was  pro- 
duced by  releasing  the  largest  fish  possible. 
Number  of  survivors  decreased  with  increas- 
ing Julian  day  of  release  (Fig.  6A).  There  was 
a  considerable  interaction  between  size- 
at-release  and  number  of  fish  released, 
such  that  low  postrelease  densities  were 
subjected  to  relatively  high  proportional 
mortality.  Thus,  when  fish  were  released 
in  low  numbers  and  at  small  sizes,  the 
fish  were  subjected  to  relatively  high 
proportional  mortality  rates  for  long 
periods  of  time  (while  they  grew  towards 
the  80-mm-TL  ontogenetic  shift  size)  and 
consequently  produced  few  or  no  survi- 
vors (Fig.  6A).  Optimal  release  scenarios 
under  strong  type-2  mortality  produced 
substantially  lower  (>40%  in  some 
cases)  percent  survival  (and  therefore 
substantially  higher  cost-per-survivor) 
estimates  at  low  to  moderate  numbers 
released  (NFR=  100-50,000;  postrelease 
density=0.001-0.5  fish/m2)  than  under 
density-independent  mortality  (Fig.  5,  A 
and  B).  Differences  in  percent  survival 
estimates  (and  thus  cost-per-survivor 
estimates)  between  density-indepen- 
dent survival  and  weak  or  strong  type-2 
mortality  declined  to  less  than  5ri  when 
the  numbers  released  increased  to 
25,000  (postrelease  density=0.25  fish/m2) 
under  weak  type-2  mortality  and  75.000 
(postrelease  density=0.75  fish/m2)  under 
strong  type-2  mortality  (Fig.  5A).  Thus, 
model  predictions  under  density-inde- 
pendent mortality  differed  most  from 
predictions  under  mortality  governed  by 


-  density-independent 
-type  2  -  weak 

-  type  2  -  strong 
-type  3  -  weak 
■type  3  -  strong 


density-independent 
type  2  -  weak 
type  2  -  strong 
type  3  -  weak 
type  3  -  strong 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Paralichthys  dentatus 


85 


B 


a  type-2  predator  functional  response  when 
postrelease  densities  were  relatively  low. 

Type-3  mortality  As  in  all  other  simulations, 
the  most  important  factor  affecting  number 
of  survivors  under  type-3  mortality  was  size- 
at-release,  such  that  the  greatest  numbers  of 
survivors  were  always  produced  by  releasing 
the  largest  fish  possible  (Fig.  7A).  Number  of 
survivors  decreased  with  increasing  Julian 
day  of  release  (Fig.  7A).  Percent  survival 
was  considerably  lower  (>25%  in  some  cases) 
under  type-3  mortality  than  under  density- 
independent  mortality  at  moderate  to  high 
numbers  released  (NFR=10, 000-400, 000) 
(Fig.  5 A). 

In  nearly  all  simulations,  the  lowest  CPS 
values  were  produced  by  releasing  the  larg- 
est fish  possible  (Fig.  7B).  The  exceptions  to 
the  "large  size  =  optimal  CPS"  rule  occurred 
when  postrelease  densities  were  small  (cor- 
responding to  numbers  released  of  100,  500, 
and  1000)  and  the  mortality  curve  was  type  3 
(weak  or  strong).  In  these  instances,  mortality 
was  sufficiently  low  at  low  release  densities 
( Fig.  7B )  so  that  the  difference  in  overall  sur- 
vival between  small-  and  large-released  fish 
was  small  enough  to  be  overridden  by  the  in- 
creased cost  of  the  larger  fish,  and  the  mini- 
mum CPS  was  obtained  when  small  (42-44 
mm  TL)  fish  were  released  (e.g.  Fig.  7B). 

At  low  numbers  released  (NFR=100-1000), 
optimal  cost-per-survivor  was  considerably 
lower  (>45%  in  some  cases)  under  type-3 
mortality  than  under  density-independent 
mortality  (Fig.  5A).  As  NFR  increased,  CPS 
under  type-3  mortality  became  greater  ( -40^ 
in  some  cases)  than  that  achieved  under  den- 
sity-independent mortality  (Fig.  5B). 


Temporal  shift  in  functional  response  from 
type  2  to  type  3,  and  from  type  3  to  type  2 
The  optimal  numbers  of  survivors  under 
varying  numbers  released  were  identical,  and 
optimal  CPS  values  nearly  identical,  when 
the  form  of  the  functional  response  changed 
from  a  type  2  to  a  type  3,  and  from  a  type  3  to 
a  type  2,  midway  through  the  juvenile  nurs- 
ery season  (Fig.  8,  A  and  B).  The  differences 
at  low  postrelease  densities  between  optimal 
CPS  values  under  shifting  type  2  to  type  3  and  type  3  to 
type  2  scenarios  (Fig.  8A)  occurred  because  initial  mortality 
under  the  type-3  functional  response  was  sufficiently  low 
that  the  difference  in  overall  survival  between  small-  and 
large-released  fish  was  small  enough  to  be  overridden  by 
the  increased  cost  of  the  larger  fish  (Fig.  8A).  The  minimum 
CPS  was  obtained  when  small  (42-44  mm  TL)  fish  were 
released  (in  all  other  cases,  optimal  results  were  obtained 
when  size-at-release  was  maximized)  (Fig.  8A).  The  major 
difference  between  the  two  shifting  scenarios  is  that  the 


re/ease 


Figure  6 

Response  surfaces  of  (A)  number  offish  (summer  flounder  I  survivors  and 
(B)  cost-per-survivor  (CPS)  as  a  function  of  date  of  release  and  size  at 
release  at  number  released  (NR)  =  5000  (postrelease  density=0.05l  under 
a  strong  type-2  functional  response.  CPS  values  greater  than  $10  were  set 
equal  to  $10  for  ease  of  presentation. 


release  dates  producing  optimal  results  for  a  given  number 
of  fish  released  varied  depending  on  the  direction  of  the 
shifting  functional  response.  For  example,  when  the  func- 
tional response  shifted  from  a  type  2  to  a  type  3,  a  release 
of  100,000  HR  organisms  achieved  optimal  results  when 
release  occurred  early  in  the  season  (Julian  day  <145) 
(Fig.  9A).  When  the  functional  response  shifted  from  a 
type  3  to  a  type  2,  a  release  of  100,000  HR  summer  floun- 
der achieved  optimal  results  only  when  releases  occurred 
later  in  the  season  (Julian  day  >145)  (Fig.  9B).  When  the 


86 


Fishery  Bulletin  102(1) 


functional  response  shifted  from  a  type  3  to  a  type  2,  releas- 
ing 100,000  HR  organisms  prior  to  Julian  day  146  resulted 
in  markedly  decreased  survival  (and  therefore  increased 
CPS )  compared  to  results  obtained  from  releases  after  day 
146  (e.g.  releasing  on  Julian  day  92  resulted  in  a  decrease 
in  number  of  survivors  and  an  increase  in  CPS  of  22.8% 
and  29.7%,  respectively)  (Fig.  9B).  Thus,  date-of-release 
had  a  significant  effect  on  the  results  (and  therefore  in 
determining  optimal  release  strategies)  when  the  relation- 
ship between  density  and  mortality  changed  temporally, 
suggesting  that  the  presence  of  a  temporal  shift  in  the  func- 


500 


£     400 


300 


200 


E 
z 


100 


220 


OaV' 


Size  at  re/ease 


Figure  7 

Response  surfaces  of  (A)  number  offish  (summer  flounder)  survivors  and 
(B)  cost-per-survivor  (CPS)  as  a  function  of  date  of  release  and  size  at 
release  at  number  released  (NR)  =  500  (postrelease  density=0.005)  under 
a  strong  type-3  functional  response.  CPS  values  greater  than  $10  were  set 
equal  to  $10  for  ease  of  presentation. 


tional  response  of  the  predator  guild  would  have  consider- 
able effects  on  the  number  of  survivors  and  CPS  for  stock 
enhancement  efforts  with  juvenile  summer  flounder. 

Correspondence  between  predicted  and 
observed  temporal  abundance  patterns 

Under  the  assumption  of  a  type-2  functional  response, 
predicted  declines  in  juvenile  summer  flounder  density 
over  time  were  rapid  when  initial  density  was  relatively 
low  (i.e.  0.1  fish/m2)  (Fig.  10,  A  and  B).  These  predictions 
contrast  with  those  observed  in  the  field, 
in  which  declines  at  relatively  low  initial 
densities  were  gradual  (compare  Fig.  10A 
and  10B  to  Fig.  10F).  Under  the  assumption 
of  a  type-3  functional  response,  predicted 
declines  were  rapid  when  initial  density  was 
relatively  high  (i.e.  0.5  fish/m2)  I  Fig.  10,  C 
and  D).  These  results  generally  contrast  with 
those  observed  in  the  field,  in  which  declines 
at  relatively  high  densities  were  much  less 
rapid  than  those  predicted  under  a  strong 
type-3  functional  response,  and  somewhat 
less  rapid  than  those  predicted  under  a  weak 
type-3  functional  response  (Figs.  10F  and  11). 
Under  density-independent  mortality,  there 
was  little  difference  in  predicted  declines  in 
juvenile  summer  flounder  density  over  time 
between  the  three  initial  density  levels  (0.1, 
0.3,  and  0.5  fish/m2);  in  each  case  there  was 
a  gradual  decrease  in  density  over  time  (Fig. 
10E).  These  results  were  similar  to  those 
observed  in  the  field,  although  declines  at  rel- 
atively high  densities  in  the  field  were  some- 
what more  rapid  than  those  predicted  under 
density-independent  mortality  ( compare  Figs. 
10E  and  10F).  Thus,  a  density-mortality  rela- 
tionship lying  between  that  generated  under 
density-independence  and  that  generated 
under  the  weak  type-3  functional  response 
in  the  model  would  most  closely  predict  the 
temporal  declines  observed  in  the  field. 


Discussion 

Implications  for  stock  enhancement  of 
summer  flounder 

Regardless  of  the  relationship  between  den- 
sity and  mortality,  size-at-release  was  the 
most  important  variable  in  the  model  affect- 
ing survival  and  costs  associated  with  stock 
enhancement  of  summer  flounder.  The  model 
predicts  that  under  all  release  scenarios,  1) 
survival  will  be  maximized  and  2)  costs  asso- 
ciated with  stock  enhancement  (i.e.  cost  per 
survivor)  will  be  minimized  when  HR  fish  are 
released  at  the  largest  size  possible.  From  a 
survival  standpoint,  these  results  are  not 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Paralichthys  dentatus 


87 


surprising.  Larger  fish  spend  fewer  days  than  smaller  fish 
in  the  wild  nursery  habitats  before  making  an  ontogenetic 
habitat  shift  to  deeper  waters  and  thus  are  susceptible  to 
daily  natural  mortality  for  fewer  numbers  of  days  than  are 
smaller  fish.  Thus,  total  mortality  of  smaller  fish  is  greater 
than  that  of  larger  fish.  Additionally,  although  we  chose  to 
make  mortality  independent  of  size  in  the  model,  abundant 
literature  suggests  that  natural  mortality  (especially  due 
to  predation )  may  decrease  with  increasing  size  by  mecha- 
nisms such  as  enhanced  resistance  to  starvation,  decreased 
vulnerability  to  predators,  and  better  tolerance  of  environ- 
mental extremes  (Sogard,  1997;  Hurst  and  Conover,  1998; 
Lorenzen,  2000).  Thus,  the  difference  in  predicted  survival 
between  1 )  relatively  large  and  relatively  small  fish  and  2 ) 
fish  released  early  versus  late  in  the  season  in  our  model 
would  be  even  greater  if  larger  summer  flounder  suffered 
lower  natural  mortality  than  smaller  fish.  Furthermore, 
the  daily  mortality  estimate  used  in  the  density-inde- 
pendent simulations  and  to  parameterize  the  different 
types  of  density-mortality  relationships  may  have  been 
an  underestimate  of  daily  mortality  (Kellison,  2000).  If  a 
greater  estimate  of  daily  mortality  had  been  used,  the  dif- 
ference in  predicted  survival  between  relatively  large  and 
relatively  small  fish  in  our  model  would  have  been  further 
exacerbated  because  smaller  fish  spend  longer  amounts  of 
time  in  the  model  growing  to  the  80-mm-TL  ontogenetic 
shift  size.  These  conclusions  are  supported  by  empirical 
research  demonstrating  that  relatively  large  released  HR 
fish  suffer  lower  mortality  than  relatively  small  HR  fish 
released  in  the  field  (e.g.  Yamashita  et  al.,  1994;  Leber, 
1995;  Willis  et  al.,  1995;  Tominaga  and  Watanabe,  1998; 
Svasandetal.,2000). 

Although  the  survival  predictions  of  the  model  (total 
mortality  decreases  with  increasing  size-at-release)  are 
not  surprising,  the  economic  (cost-per-survivor)  predic- 
tions were  unexpected.  The  paradigm  for  stock  enhance- 
ment strategy  is  that  the  rearing  of  relatively  large  fish 
for  release  is  cost  prohibitive,  so  that  mass  releases  of 
relatively  small,  inexpensive-to-rear  fish  are  a  better 
strategy  than  the  release  of  larger,  expensive-to-rear  fish 
(Kellison,  personal  obs.).  Thus,  relatively  small  juveniles 
are  released  in  virtually  all  current  stock  enhancement 
programs  (e.g.  Russell  and  Rimmer,  1997;  Masuda  and 
Tsukamoto,  1998;  McEachron  et  al.,  1998;  Svasand,  1998; 
Serafy  et  al.,  1999).  Nevertheless,  large-scale  hatcheries 
and  grow-out  facilities  are  using  ever-increasing  technol- 
ogy to  minimize  the  costs  associated  with  the  production 
of  relatively  large  fishes  (Sproul  and  Tominaga,  1992). 
Thus,  for  species  for  which  1)  hatcheries  are  capable  of 
producing  relatively  large  fish  at  relatively  low  costs  (as 
is  likely  for  summer  flounder),  and  2)  postrelease  survival 
rates  increase  with  release  size,  release  scenarios  utilizing 
the  largest  fish  possible  may  maximize  the  potential  (i.e. 
produce  maximum  survival  at  minimum  costs )  of  stock  en- 
hancement efforts.  In  these  cases,  the  "small  fish  maximize 
stock  enhancement  potential"  paradigm  might  be  replaced 
with  a  "large  fish  maximize  potential"  paradigm.  As  a  ca- 
veat, this  "large  fish"  strategy  may  be  limited  by  spatial 
limitations  of  hatcheries  in  producing  large  numbers  of 
relatively  large  fish.  Because  reared  fish  generally  must 


1  40  i 

*/» 

^_ 

1.20- 

o 

2 

1  00- 

w 

0  80- 

(1> 
Q. 

0  60- 

If) 
O 

0  40- 

O 

Type  2  to  3 
Type  3  lo  2 


0  20-1— -! , 1 r 


Postrelease  density 

Figure  8 

Optimal  lA)  economic  cost-per-survivor  and  (B)  per- 
cent survival  of  released  hatchery-reared  summer 
flounder  under  temporally  shifting  functional  re- 
sponses of  type  2  to  type  3  and  type  3  to  type  2. 


be  kept  below  critical  densities  in  hatchery  environments 
because  of  water  quality  and  fish  interaction  issues  (e.g. 
cannibalism),  larger  fish  necessarily  require  more  space 
than  smaller  fish  for  rearing.  If  the  demand  for  space  to 
rear  large  quantities  of  large  fish  can  be  realized,  then  the 
model  simulations  indicate  that  stock  enhancement  strat- 
egies in  which  size-at-release  is  maximized  will  produce 
the  maximum  number  of  survivors. 

Although  not  as  important  as  size-at-release,  Julian  day 
of  release  had  a  significant  effect  on  survival  and  cost-per- 
survivor  in  the  model,  such  that  enhancement  efforts  were 
always  more  successful  (more  survivors,  lower  costs)  when 
fish  were  released  at  the  earliest  Julian  day  possible.  These 
results  occurred  because  growth  in  the  model  decreased 
with  increasing  Julian  Day.  Although  the  mechanisms  un- 
derlying this  decrease  in  growth  with  increasing  Julian  day 
are  unknown,  they  may  be  related  to  decreased  prey  avail- 
ability or  metabolic  efficiency  as  temperatures  increase 
with  increasing  Julian  day  (Malloy  and  Targett,  1994a, 
1994b;  Fujii  and  Noguchi,  1996;  Howson,  2000).  Thus,  for 
a  given  size-at-release,  fish  released  earlier  in  the  season 
experienced  greater  growth  rates  than  fish  of  the  same 
size-at-release  released  later  in  the  season  and  therefore 
reached  the  80-mm-TL  ontogenetic  shift  size  faster  (over  a 
period  of  fewer  days)  than  fish  released  later  in  the  season. 
Thus,  fish  released  earlier  in  the  season  were  susceptible 
to  natural  mortality  for  fewer  days  than  fish  released  later 
in  the  season  and  therefore  suffered  lower  total  mortality. 
These  results  emphasize  the  importance  of  knowledge  of 
possible  time-dependent  growth  in  the  field  prior  to  stock 
enhancement  efforts. 


Fishery  Bulletin  102(1) 


Is  density  important?  Effects  of  varying  density-mortality 
relationships 

Our  results  suggest  that  the  relationship  between  density 
and  mortality  has  the  potential  to  significantly  affect  opti- 
mal release  scenarios  associated  with  stock  enhancement 
efforts.  Because  the  original  simulations  were  performed 
under  density-independent  mortality,  the  number  of 
survivors  originally  increased  linearly  with  the  number 


B 


1e+5 


(/> 

8e+4 

o 

> 

> 

6e+4 

tfl 

n 

m 

4e+4 

a 

h 

3 

2e+4 

z 

0 

80 

released,  resulting  in  a  density-independent  cost-per- 
survivor.  Thus,  when  mortality  is  independent  of  density 
(over  a  given  range  of  densities)  for  a  target  species  for 
stock  enhancement,  managers  will  maximize  the  number 
of  survivors  produced  by  releasing  the  greatest  number  of 
fish  possible  within  that  range  for  a  given  size  class.  When 
mortality  varied  with  density  of  released  fish,  the  number 
of  survivors  and  cost-per-survivor  depended  on  the  den- 
sity-mortality relationship.  In  some  cases,  optimal  results 
(maximum  survival  and  minimum  cost)  differed 
depending  on  whether  the  response  variable  was 
number  of  survivors  or  cost-per-survivor.  Under 
the  assumption  of  a  strong  type-3  functional 
response  and  under  relatively  low  postrelease 
densities,  survival  was  optimized  (maximized) 
by  releasing  the  largest  fish  ( 80  mm  TL)  possible; 
however,  cost-per-survivor  was  optimized  (mini- 
mized) by  releasing  smaller  fish  (42-44  mm  TL). 
This  result  occurred  because  mortality  at  low 
postrelease  densities  was  sufficiently  low  that 
the  difference  in  total  mortality  attributed  to  the 
longer  "susceptibility"  period  of  the  smaller  fish 
was  insufficient  to  override  the  economic  advan- 
tage of  releasing  smaller  fish.  Simulations  under 
shifting  functional  responses  (type  2  to  type  3 
and  type  3  to  type  2)  produced  optimal  results 
similar  to  those  obtained  when  nonshifting  type- 
2  or  type-3  functional  responses  were  employed 
because  densities  were  generally  reduced  to  such 
low  numbers  by  the  time  the  shift  occurred  that 
the  changing  density-mortality  relationship  was 
inconsequential.  Importantly,  when  functional 
responses  shifted  temporally,  the  predicted 
number  of  survivors  and  economic  cost  per 
survivor  was  at  times  very  dependent  on  date  of 
release,  suggesting  that  identifying  or  ruling  out 
shifting  functional  responses  in  the  wild  may  be 
critical  to  accurate  prediction  of  response  vari- 
ables (survivors  and  economic  costs)  associated 
with  stock  enhancement.  Although  we  are  not 
aware  of  reports  in  the  literature  of  shifting 
functional  responses  in  the  wild,  we  are  also 
not  aware  of  studies  that  have  tested  for  such 
a  phenomenon,  possibly  because  of  the  logisti- 
cal difficulties  inherent  in  identifying  a  shifting 
functional  response. 

Correspondence  between  predicted  and 
observed  temporal  abundance  patterns 


Figure  9 

(A)  Response  surface  of  optimal  number  of  summer  flounder  survivors 
as  a  function  of  date  of  release  and  size  at  release  at  number  released 
(NR)  =  100,000  (postrelease  density=1.0  fish/m2 1  under  the  assumption 
of  a  temporally  shifting  functional  responses  from  type  2  to  type  3. 
<B>  Response  surfaces  of  optimal  number  of  survivors  as  a  function  of 
date  of  release  and  size  at  release  at  number  released  (NR)  =  100. 000 
(postrelease  density=1.0  fish/m2  >  under  the  assumption  of  a  temporally 
shifting  functional  responses  from  type  3  to  type  2. 


Predictions  of  field  abundance  patterns  of  juve- 
nile flounder  density  over  time  were  noticeably 
different  under  density-independent  mortality 
and  density-dependent  mortality  governed  by 
type-2  and  type-3  functional  responses.  For 
example,  our  simulations  predict  that  fish  den- 
sity should  decrease  rapidly  under  relatively 
low  initial  densities  if  the  functional  response  is 
type  2,  decrease  rapidly  at  relatively  high  initial 
densities  if  the  functional  response  is  type  3,  and 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Parahchthys  dentatus 


89 


OS- 

04- 

0.3 

0.2 

01 

00 


E 


Strong  type  2 


Strong  type  3 


150 


Dl 


B 


Weak  type  2 


05 
0.4 

0  3 
02 


110     130      150     170     190     210 


D 


Weak  type  3 


Julian  day 

Figure  10 

Predicted  temporal  trends  in  summer  flounder  abundance  under  initial  densities  of  0.5,  0.3,  and  0.1  fish/m2 
under  the  assumption  of  a  functional  response  that  is  a  (A)  strong  type  2.  IB)  weak  type  2,  (C)  strong  type 
3.  i D i  weak  type  3.  and  under  the  assumption  of  (E)  density-independent  iDIl  mortality.  The  curves  in  iFi 
are  best  fitted  (highest  r2  value)  to  data  collected  in  Duke  Beach  1999  (curve  a,  r2=0.82).  Haystacks  Marsh 
1999  (curve  b,  r2=0.73),  Prytherch  Marsh  1999  (curve  c.  ;-2=0.82),  Towne  Beach  1999  (curve  d,  r2=0.91). 
Radio  Beach  1999  (curve  e,  r2  =  0.27),  Duke  Beach  1998  (curve  f,  r2=0.31),  and  Prytherch  1998  (curve  g, 
r2=0.16)  (see  Fig.  11  for  data). 


gradually  decrease  regardless  of  initial  density  if  mortal- 
ity is  density  independent.  From  examinations  of  tempo- 
ral abundance  patterns  from  several  nursery  sites  (see 
Kellison  et  al.,  2003b,  for  site  descriptions),  it  is  evident 
that  observed  declines  at  relatively  low  initial  densities 
are  similar  to  predicted  declines  under  both  density-inde- 
pendent mortality  and  a  weak  type-3  functional  response; 
whereas  observed  declines  at  relatively  high  initial  densi- 
ties are  somewhat  less  gradual  than  predicted  under  den- 
sity-independent mortality,  but  somewhat  more  gradual 
than  predicted  under  the  weak  type-3  functional  response. 
These  results  suggest  that  model  predictions  made  under 
the  assumption  of  a  weak  type-3  response  may  give  rea- 
sonably accurate  but  conservative  predictions  of  juvenile 
summer  flounder  mortality  and  economic  costs  associated 
with  stock  enhancement  for  comparison  with  alternative 
management  methods.  As  a  caveat,  although  we  found  no 
evidence  of  size-dependent  daily  mortality  over  the  range 
of  fish  sizes  examined  in  this  study,  it  is  very  likely  that 


such  a  relationship  exists  to  some  extent  (Sogard,  1997; 
Lorenzen,  2000).  Incorporating  size-dependent  mortality 
into  the  model  would  decrease  the  slopes  of  the  predicted 
temporal  abundance  curves  but  should  not  change  the 
conclusion  that  the  observed  data  lie  somewhere  between 
values  predicted  under  density-independent  mortality 
and  those  governed  by  a  weak  type-3  functional  response, 
respectively.  Additionally,  because  the  portions  of  the 
curves  used  to  delineate  between  temporal  abundances 
expected  under  density-independent  versus  varying  den- 
sity-mortality relationships  are  from  early  in  the  growth 
season  (later  parts  of  the  curve  converge  on  very  low  den- 
sities) and  because  nearly  all  fish  in  these  portions  of  the 
curves  are  at  sizes  well  below  that  at  which  ontogenetic 
emigration  occurs,  the  exclusion  of  emigration  from  these 
simulations  should  not  affect  the  general  conclusions 
reached.  These  issues  could  be  clarified  with  further  field 
trials  to  investigate  the  dependence  of  daily  mortality 
rates  on  fish  size. 


90 


Fishery  Bulletin  102(1) 


E 


E 


Prytherch  1999 


Radio  1999 


003 

♦ 

Prytherch  1998 

*                    r*  =  0.1575 

0  02 

♦ 

001 

* 

_. ♦                                ♦♦ 

95      105     115      125     135     145      155      165 


B 


03 

♦ 

Duke  1999 

0  2 

• 

♦ 

r  =  0  8162 

01 

♦ 

9* 

♦ 

*     ♦*4U**TMT     *'     •*♦ 

D 


Haystacks  1999 


Towne  1999 

r"  i  0  9063 

^s^    • 

♦  "" 

"Y —                                  ♦              ♦♦ 
»                   '«W.  W. LT» 

95         115        135        155        175        195 


Duke  1998 

0  1 

♦ 

♦ 

* 

r2  =  03113 

0.05 

• 

* 

**> 

♦. 

• 

♦ 

♦  ~~7 

• 

♦    ♦      ♦ 

95        115       135       155       175       195 


Julian  day 


Julian  day 

Figure  11 

Temporal  density  patterns  from  (A)  Duke  Beach,  1999;  (B)  Haystacks  Marsh,  1999;  (C)  Prytherch  Marsh, 
1999;  (D)  Towne  Beach,  1999;  (E)  Radio  Beach,  1999;  (F)  Duke  Beach,  1998;  and  (G)  Prytherch  Marsh  1998. 
Densities  are  corrected  for  gear  bias  (see  Kellison,  2000). 


Model  utility  and  implications 

Although  model  results  varied  considerably  under  the 
various  density-mortality  relationships,  the  overall  pre- 
dictions that  survival  would  be  maximized  and  economic 
costs  minimized  when  relatively  large  fish  were  released 
early  in  the  season  were  unaffected  by  the  density- 
mortality  relationship.  These  results  suggest  that  manag- 
ers may  use  this  model  to  make  inferences  about  optimal 
release  scenarios  even  if  density-mortality  relationships 
are  unknown.  Additionally,  these  results  have  important 
implications  for  the  cost  efficiency  of  stock  enhancement 
programs.  Managers  can  use  the  model  to  determine 


the  release  scenarios  under  which  they  can  1)  maxi- 
mize the  number  of  survivors,  given  a  financial  limit 
(e.g.  given  a  budget  of  x  dollars,  what  release  scenario 
or  scenarios  will  produce  the  greatest  number  of  survi- 
vors?), and  2)  minimize  costs,  given  a  goal  of  number-of- 
survivors-produced  (e.g.  given  a  goal  of  producing 
.v  survivors,  what  release  scenario  or  scenarios  will  be  most 
cost  efficient?). 

In  conclusion,  the  compartmental  model  used  in  this 
study  provides  an  example  of  a  relatively  easy-to-develop 
predictive  tool  with  which  to  make  inferences  about  the 
ecological  and  economic  potential  of  stock  enhancement,  in 
relation  to  alternative  management  approaches,  to  rebuild 
depleted  fisheries. 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Paraltchthys  dentatus 


91 


Acknowledgments 

We  thank  Brian  Burke  (NCSU)  for  tutelage  in  the  use  of 
Visual  Basic.  Mike  Denson  (South  Carolina  Department 
of  Natural  Resources)  and  Pete  Schuhmann  (UNC-Wilm- 
ington )  greatly  contributed  to  the  editing  of  an  earlier  ver- 
sion of  this  manuscript.  Mark  Wuenschel,  Michael  Martin, 
Brian  Degan,  Lisa  Etherington.  and  Mikael  Currimjoe  pro- 
vided valuable  laboratory  and  field  assistance  necessary 
for  parameter  estimation.  This  project  was  partially  funded 
by  the  University  of  North  Carolina  at  Wilmington/North 
Carolina  State  University  Cooperative  Ph.D.  Program,  and 
a  grant  from  the  National  Science  Foundation  (OCE  97- 
34472)  to  D.  Eggleston. 


Literature  cited 

Adams,  C.  M..  and  R.  S.  Pomeroy. 

1991.     Scale  economies  in  hard  clam  aquaculture.     J.  Shell- 
fish Res.  10:307-308. 
Barbeau.  M.  A.,  and  H.  Caswell. 

1999.    A  matrix  model  for  short-term  dynamics  of  seeded 
populations  of  sea  scallops.     Ecol.  Appl.  9:266-287. 
Bax,  N.  J. 

1983.  Early  marine  mortality  of  marked  juvenile  chum 
salmon  iOncorhynchus  keta)  released  into  Hood  Canal, 
Puget  Sound,  Washington,  in  1980.  Can.  J.  Fish.  Aquat. 
Sci.  40:426-435. 

Black,  R.  W.  II,  and  N.  G.  Hairston  Jr. 

1988.  Predator  driven  changes  in  community  structure. 
Oecologia  77:468-479. 

Blankenship,  H.  L.,  and  K.  M.  Leber. 

1996.  A  responsible  approach  to  marine  stock 
enhancement.  In  Developing  and  sustaining  world  fish- 
eries resources:  the  state  of  science  and  management  (D.  A. 
Hancock,  D.  C.  Smith,  A.  Grant,  and  J.  P.  Beumer,  eds.  I,  p. 
485^91.     CSIRO,  Collingwood,  Australia. 

Botsford,  L.  W„  and  R.  C.  Hobbs. 

1984.  Optimal  fishery  policy  with  artificial  enhancement 
through  stocking:  California's  white  sturgeon  as  an 
example.    Ecol.  Model.  23:293-312. 

Buckel,  J.  A.,  D.  O.  Conover,  N.  D.  Steinberg,  and  K.  A.  McKown. 

1999.  Impact  of  age-0  bluefish  iPomatomus  saltatrix)  preda- 
tion  on  age-0  fishes  in  the  Hudson  River  estuary:  evidence 
for  density-dependent  loss  of  juvenile  striped  bass  iMorone 
saxatilis).    Can.  J.  Fish.  Aquat.  Sci.  56:275-287. 

Burke,  J.  S. 

1995.     Role  of  feeding  and  prey  distribution  on  summer  and 
southern  flounder  in  selection  of  estuarine  nursery  habitats. 
J.  Fish  Biol.  47:355-366 
Burke.  J.  S.,  J.  M.  Miller,  and  D.  E.  Hoss. 

1991.     Immigration  and  settlement  pattern  of  Paralichthys 
dentatus  and  P.  lethostigma  in  an  estuarine  nursery  ground. 
North  Carolina,  U.S.A.     Neth.  J.  Sea  Res.  27:393^405. 
Burke,  J.  S.,  J.  P.  Monaghan,  and  S.  Yokoyama. 

2000.  Efforts  to  understand  stock  structure  of  summer 
flounder  {Paralichthys  dentatus)  in  North  Carolina,  USA. 
J.  Sea  Res.  44:111-122. 

Burke,  J.  S.,  T.  Seikai,  Y.  Tanaka,  and  M.  Tanaka. 

1999.     Experimental  intensive  culture  of  summer  flounder, 
Paralichthys  dentatus.    Aquaculture  176:  135-144. 
Bystroem,  P.,  and  E.  Garcia-Berthou. 

1999.  Density  dependent  growth  and  size  specific  competi- 
tive interactions  in  young  fish.     Oikos  86:  217-232. 


Cowx,  I.  G. 

1999.    An  appraisal  of  stocking  strategies  in  the  light  of  devel- 
oping country  constraints.     Fish.  Manag.  Ecol.  6:  21-34. 
Elis,  T.,  and  R.  N.  Gibson. 

1995.  Size-selective  predation  of  0-group  flatfishes  in  a 
Scottish  coastal  nursery  ground.  Mar.  Ecol.  Prog.  Ser. 
127:27-37. 

FAO  (Food  and  Agriculture  Organization  of  the  United  Nations). 

1998.  Summary:  The  state  of  world  fisheries  and  aquacul- 
ture. http://www.fao.org/WAICENT/FAOINFO/FISHERY/ 
FISHERY.HTM.     [Accessed:  June  1998.1 

Froese,  R.,  and  D.  Pauly,  eds. 

2001.     FishBase.     World  wide  web  electronic  publication. 
http://www.fishbase.org.     [Accessed:  January  2001.] 
Fujii,  T,  and  M.  Noguchi. 

1996.  Feeding  and  growth  of  Japanese  flounder  {Paralich- 
thys olivaceus)  in  the  nursery  ground.  In  Proceedings  of 
an  international  workshop:  survival  strategies  in  early  life 
stages  of  marine  resources,  p.  141-154.  A.  A.  Balkema 
Publishers,  Brookfield,  VT. 

Furuta,  S. 

1999.  Seasonal  changes  in  abundance,  length  distribution, 
feeding  condition  and  predation  vulnerability  of  juvenile 
Japanese  flounder,  Paralichthys  olivaceus,  and  prey  mysid 
density  in  the  Tottori  coastal  area.  Nippon  Suisan  Gak- 
kaishi  65:167-174. 

Garcia,  L.  MaB..  R.  F.  Agbayani,  M.  N.  Duray, 
G.  V.  Hilomen-Garcia,  A.  C.  Emata,  and  C.  L.  Marie. 

1999.  Economic  assessment  of  commercial  hatchery  produc- 
tion of  milkfish  (Chanos  chanos  Forsskal)  fry.  J.  Applied 
Ichthyol.  15:70-74. 

Hansen,  M.  J.,  M.  A.  Bozek,  J.  R.  Newby,  S.  P.  Newman,  and 
M.  D.  Staggs. 

1998.     Factors  affecting  recruitment  of  walleyes  in  Escanaba 
Lake,  Wisconsin.  1958-1996.     North  Am.  J.  Fish.  Manag. 
18:764-774. 
Heppell,  S.  S.,  and  L.  B.  Crowder. 

1998.     Prognostic  evaluation  of  enhancement  programs  using 
population  models  and  life  history  analysis.     Bull.  Mar.  Sci. 
62:495-507. 
Hernandez-Llamas,  A. 

1997.  Management  strategies  of  stocking  density  and  length 
of  culture  period  for  the  Catarina  scallop  Argopecten  circu- 
lars (Sowerby):  A  bioeconomic  approach.  Aquacult.  Res. 
28:223-229. 

Hobbs,  R.  C,  L.  W  Botsford,  and  R.  G.  Kope. 

1990.     Bioeconomic  evaluation  of  the  culture/stocking  concept 
for  California  halibut.  The  California  halibut.  Paralichthys 
californicus,  resource  and  fisheries,  1990.  p.  417^150.     Cal. 
Dep.  Fish  Game  Fish.  Bull.,  vol.  174. 
Holling,  C.  S. 

1959.     Some  characteristics  of  simple  types  of  predation  and 
parasitism.     Canadian  Entomologist  91:  385-398. 
Houde,  E.  D. 

1987.     Fish  early  life  history  dynamics  and  recruitment 
variability.    Am.  Fish.  Soc.  Symp.  2:17-29. 
Howson,  U.  A. 

2000.  Nursery  habitat  quality  for  juvenile  paralichthyid 
flounders:  Experimental  analysis  of  the  effects  of  physico- 
chemical  parameters.  Ph.D.  diss.,  129  p.  LTniv.  Delaware, 
Newark,  DE.  129  p. 

Hurst,  T  P..  and  D.  O.  Conover. 

1998.  Winter  mortality  of  young-of-the-year  Hudson 
River  bass  (Morone  saxatilis):  Size-dependent  patterns 
and  effects  on  recruitment.  Can.  J.  Fish  Aquat.  Sci.  55: 
1122-1130. 


92 


Fishery  Bulletin  102(1) 


Jenkins,  T.  M.  Jr.,  S.  Diehl.  K.  W.  Kratz.  and  S.  D.  Cooper, 

1999.  Effects  of  population  density  on  individual  growth  of 
brown  trout  in  streams.     Ecol.  80:941-956. 

Kellison,  G.  T. 

2000.  Evaluation  of  stock  enhancement  potential  for 
summer  flounder  (Paraliehthys  dentatus):  an  integrated 
laboratory,  field  and  modeling  study.  Ph.D.  diss.,  196  p. 
NC  State  Univ.,  Raleigh,  NC. 

Kellison,  G.  T.,  D.  B.  Eggleston,  and  J.  S.  Burke. 

2000.     Comparative  behaviour  and   survival  of  hatch- 
ery-reared versus  wild  summer  flounder  {Paraliehthys 
dentatus).     Can.  J.  Fish.  Aquat.  Sci.  57:1870-1877. 
Kellison,  G.  T.,  D.  B.  Eggleston,  J.  C.  Taylor,  and  J.  S.  Burke. 

2003a.    An  assessment  of  biases  associated  with  caging, 
tethering,  and  habitat-specific  trawl  sampling  of  summer 
flounder  {Paraliehthys  dentatus).     Estuaries  26:64-71. 
Kellison  G.  T„  D.  B.  Eggleston.  J.  C.  Taylor,  J.  S.  Burke,  and 
J.  A.  Osborne. 

2003b.     Pilot  evaluation  of  summer  flounder  stock  enhance- 
ment potential  using  experimental  ecology.     Mar.  Ecol. 
Prog.  Ser.  250:263-278. 
Kent,  D.  B.,  and  M.  A.  Drawbridge. 

1999.  Developing  a  marine  ranching  programme:  A  multi- 
disciplinary  approach.  Marine  ranching:  Global  perspec- 
tives with  emphasis  on  the  Japanese  experience.  FAO 
fisheries  circular  943:66-78.     FAO,  Rome. 

Kimmerer,  W.  J.,  J.  H.  Cowman  Jr.,  L.  W.  Miller,  and  K.  A.  Rose. 

2000.  Analysis  of  an  estuarine  striped  bass  iMorone  saxa- 
tilis)  population:  influence  of  density-dependent  mortality 
between  metamorphosis  and  recruitment.  Can.  J.  Fish. 
Aquat.  Sci.  57:478-486. 

Leber,  K.  M. 

1995.     Significance  of  fish  size-at-release  on  enhancement 
of  striped  mullet  fisheries  in  Hawaii.     J.  World.  Aquacult. 
Soc.  26:143-153. 
Lindholm,  J.  B.,  P.  J.  Auster,  M.  Ruth,  and  L  Kaufman. 

2001.  Modeling  the  effects  of  fishing  and  implications  for  the 
design  of  marine  protected  areas:  juvenile  fish  responses  to 
variation  in  seafloor  habitat.     Conserv.  Biol.  15:424-437. 

Lorenzen,  K. 

2000.    Allometry  of  natural  mortality  as  a  basis  for  assessing 
optimal  release  size  in  fish  stocking  programmes.     Can.  J. 
Fish.  Aquat.  Sci.  57:2374-2381. 
Malloy,  K.  D.,  and  T.  E.  Targett. 

1994a.    The  use  of  RNA:DNA  ratios  to  predict  growth  limita- 
tion of  juvenile  summer  flounder  {Paraliehthys  dentatus) 
from  Delaware  and  North  Carolina  estuaries.     Mar.  Biol. 
118:367-375. 
1994b.     Effects  of  ration  limitation  and  low  temperature 
on  growth,  biochemical  condition,  and  survival  of  juve- 
nile summer  flounder  from  two  Atlantic  Coast  nurseries. 
Trans.  Am.  Fish.  Soc.  123:182-193. 
Manderson,  J.  P.,  B.  A.  Phelan,  A.  J.  Bejda,  L.  L.  Stehlik,  and 
A.  W.  Stoner. 

1999.  Predation  by  striped  searobin  (Prionotus  evolans, 
Triglidae)  on  young-of-the-ycar  winter  flounder  (Pscinln 
pleuronectes  americanus,  Walbaum):  examining  prey  size 
selection  and  prey  choice  usinj;  field  observations  and  labo- 
ratory experiments.  ,1.  Exp.  Mar.  Biol.  Ecol.  242:211-231. 
Masuda.  R.  and  K.  Tsukamoto. 

1998.     Stock  enhancement  in  Japan:  review  and  perspective. 
Bull.  Mar.  Sci.  62:337-358. 
McEachron,  L.  W..  R.  L.  Colura,  B.W.  Bumguardner,  and  R.  Ward. 
1998.     Survival  of  stocked  red  drum  in  Texas.     Bull.  Mar. 
Sci.  62:359-368. 


Munro,  J.  L.,  and  J.  D.  Bell. 

1997.  Enhancement  of  marine  fisheries  resources.  Rev. 
Fish.  Sci.  5:185-222. 

Nash,  R.  D.  M. 

1998.  Exploring  the  population  dynamics  of  Irish  Sea 
plaice,  Pleuronectes  platessa  L..  through  the  use  of  Paulik 
diagrams.     J.  Sea  Res.  40:1-18. 

Olla,  B.  L.,  M.  W.  Davis,  and  C.  H.  Ryer. 

1998.     Understanding   how    the    hatchery    environment 
represses  or  promotes  the  development  of  behavioural  sur- 
vival skills.    Bull.  Mar.  Sci.  62:531-550. 
Powell,  A.  B.,  and  F  J.  Schwartz. 

1977.     Distribution  of  paralichthid  flounders  (Bothidae: 
Paraliehthys)  in  North  Carolina  estuaries.     Chesapeake 
Sci.  18:  334-339. 
Rice,  J.  A.,  T.  J.  Miller,  K.  A.  Rose,  L.  B.  Crowder.  E.  A.  Marschall, 
A.  S.  Trebitz,  and  D.  L.  DeAngelis. 

1993.     Growth  rate  variation  and  larval  survival:  Inferences 
from  an  individual-based  size-dependent  predation  model. 
Can.  J.  Fish.  Aquat.  Sci.  50:133-142. 
Rickards,  W  L. 

1998.  Sustainable  flounder  culture  and  fisheries:  a  regional 
approach  involving  Rhode  Island,  New  Hampshire,  Vir- 
ginia, North  Carolina,  and  South  Carolina.  In  Nutrition 
and  technical  development  of  Aquaculture,  Nov.  1998. 
p.  17-20.     Sea  Grant,  Durham,  NH. 

Rose,  K.  A.,  J.  H.  Cowan  Jr,  M.  E.  Clark,  E.  D.  Houde,  and 
S-B  Wang. 

1999.  An  individual-based  model  of  bay  anchovy  popula- 
tion dynamics  in  the  mesohaline  region  of  Chesapeake 
Bay.     Mar.  Ecol.  Prog.  Ser.  185:113-132. 

Rosenberg,  A.  A.,  M.  J.  Fogarty,  M.  P.  Sissenwine, 
J.  R.  Beddington,  and  J.  G.  Shepard. 

1993.    Achieving  sustainable  use  of  renewable  resources. 
Science  262:828-829. 
Russell,  D.  J.,  and  M.  A.  Rimmer. 

1997.  Assessment  of  stock  enhancement  of  barramundi  Lates 
ealearifer  (Bloch)  in  a  coastal  river  system  in  far  northern 
Queensland.  Australia.  Developing  and  sustaining  world 
fisheries  resources.  In  Developing  and  sustaining  world 
fisheries  resources:  the  state  of  science  and  management, 
p.  498-503.  CSIRO,  Collingwood  (Australia  I. 
Salvanes,  A.  G.  V.,  D.  L.  Aksnes,  and  J.  Giske. 

1992.     Ecosystem  model  for  evaluating  potential  cod  produc- 
tion in  a  west  Norwegian  fjord.     Mar.  Ecol.  Prog.  Ser.  90: 
9-22. 
Serafy,  J.  E„  J.  S.  Ault,  T.  R.  Capo,  and  D.  R.  Schultz. 

1999.  Red  drum,  Sciaenops  ocellatus  L.,  stock  enhance- 
ment in  Biscayne  Bay.  FL,  USA:  assessment  of  releasing 
unmarked  early  juveniles.    Aquacult.  Res.  30:737-750. 

Sogard,  S.  M. 

1997.     Size-selective  mortality  in  the  juvenile  stage  of  tele- 
ost  fishes:  a  review.     Bull.  Mar.  Sci.  60:  1 129-1 1 57. 
Sproul.  J.  T.  and  O.  Tominaga. 

1992.    An  economic  review  of  the  Japanese  flounder  stock 
enhancement  project  in  Ishikari  Bay.  Hokkaido.     Bull.  Mar. 
Sci.  50:75-88. 
Stoner,  A.W. 

1980.    Feeding  ecology  of  Lagodon  rhomboides  (Pisces:  Spari- 
dae):  variation  and  functional  responses.     Fish.  Bull.  78: 
337-352. 
Sutton,  T.  M„  K.  A.  Rose,  and  J.  J.  Ney. 

2000.  A  model  analysis  of  strategies  for  enhancing  stocking 
success  of  landlocked  striped  bass  populations.  N.  Am.  J. 
Fish.  Manag.  20:841-859. 


Kellison  and  Eggleston:  Modeling  release  scenarios  for  Paralichthys  dentatus 


93 


Svasand.  T. 

1998.     Cod  enhancement  studies  in  Norway — background 
and  results  with  emphasis  on  releases  in  the  period  1983- 
1990.    Bull.  Mar.  Sci.  62:313-324. 
Svasand,  T.  S.,  T.  Kristiansen,  T.  Pedersen.  A.  G.  V  Salvanes, 
R.  Engelsen,  G.  Naevdal,  and  M.  Nodtvedt. 

2000.     The  enhancement  of  cod  stocks.     Fish  Fisheries  1: 
173-205. 
Tanaka,  M.,  T.  Goto,  M.  Tomiyama,  and  H.  Sudo. 

1989.     Immigration,  settlement  and  mortality  of  flounder 
Paralichthys  olivaceus  larvae  and  juveniles  in  a  nursery 
ground,  Shijiki  Bay,  Japan.     Neth.  J.  Sea  Res.  24:57-67. 
Tominaga,  O.,  and  Y.  Watanabe. 

1998.     Geographical  dispersal  and  optimum  release  size  of 
hatchery-reared  Japanese  flounder  Paralichthys  olivaceus 
released  in  Ishikari  Bay.  Hokkaido,  Japan.     J.  Sea  Res.  40: 
73-81. 
Travis.  J.,  F.  C.  Coleman,  C.  B.  Grimes,  D.  Conover,  T.  M.  Bert, 
and  M.  Tringali. 

1998.     Critically  assessing  stock  enhancement:  an  introduc- 
tion to  the  Mote  Symposium.     Bull.  Mar.  Sci.  62:  305-311. 
van  der  Veer,  H.  W.,  and  J.  I.  Witte. 

1993.     The  "maximum  growth/optimal  food  condition" 


hypothesis:  A  test  for  0-group  plaice  Pleuronectes  platessa 
in  the  Dutch  Wadden  Sea.     Mar.  Ecol.  Prog.  Ser.  101: 
81-90. 
Waters,  E.  B. 

1996.     Sustainable  flounder  culture  and  fisheries.     NC 
Sea  Grant  Publication  UNC-SG-96-14,  12  p.     Sea  Grant, 
Raleigh,  NC. 
Waters.  E.  B..  and  K.  Mosher.  eds. 

1999.     Flounder  aquaculture  and  stock  enhancement  in  North 
Carolina:  issues,  opportunities  and  recommendations.     NC 
Sea  Grant  Publication,  UNC-SG-99-02,  24  p.    Sea  Grant, 
Raleigh,  NC. 
Willis,  S.  A..  W.  W  Falls,  C.  W.  Dennis,  D.  E.  Roberts,  and 
P.  G.  Whitechurch. 

1995.    Assessment  of  season  of  release  and  size-at-release 
on  recapture  rates  of  hatchery-reared  red  drum.  Uses  and 
effects  of  cultured  fishes  in  aquatic  ecosystems.    Am.  Fish. 
Soc.  Symp.  15:354-365. 
Yamashita,  Y,  S.  Nagahora,  H.  Yamada,  and  D.  Kitigawa. 

1994.  Effects  of  release  size  on  survival  and  growth  of  Japa- 
nese flounder  Paralichthys  olivaceus  in  coastal  waters  off 
Iwate  Prefecture,  northeastern  Japan.  Mar.  Ecol.  Prog. 
Ser.  105:269-276. 


94 


Abstract— Sex-specific  demography 
and  reproductive  biology-  of  stripey  bass 
[Lutjanus  carponotatus l  I  also  known  as 
Spanish  flag  snapper.  FAO )  were  exam- 
ined at  the  Palm  and  Lizard  island 
groups,  Great  Barrier  Reef  ( GBR).  Total 
mortality  rates  were  similar  between 
the  sexes.  Males  had  larger  L .  at  both 
island  groups  and  Lizard  Island  group 
fish  had  larger  overall  L_,,  Female:male 
sex  ratios  were  1.3  and  1.1  at  the  Palm 
and  Lizard  island  groups,  respectively. 
The  former  is  statistically  different 
from  1,  but  is  unlikely  significantly 
different  in  a  biological  sense.  Females 
matured  on  average  at  2  years  of  age 
and  190  mm  fork  length  at  both  loca- 
tions. Female  gonadal  lipid  body  indices 
peaked  from  August  through  October, 
preceding  peak  gonadosomatic  indices 
in  October,  November,  and  December 
that  were  twice  as  great  as  in  any 
other  month.  However,  ovarian  stag- 
ing revealed  50^  or  more  ovaries  were 
ripe  from  September  through  February, 
suggesting  a  more  protracted  spawning 
season  and  highlighting  the  different 
interpretations  that  can  arise  between 
gonad  weight  and  gonad  staging  meth- 
ods. Gonadosomatic  index  increases 
slightly  with  body  size  and  larger  fish 
have  a  longer  average  spawning  season, 
which  suggests  that  larger  fish  produce 
greater  relative  reproductive  output. 
Lizard  Island  group  females  had 
ovaries  nearly  twice  as  large  as  Palm 
Island  group  females  at  a  given  body 
size.  However,  it  is  unclear  whether 
this  reflects  spatial  differences  akin 
to  those  observed  in  growth  or  effects 
of  sampling  Lizard  Island  group  fish 
closer  to  their  date  of  spawning.  These 
results  support  an  existing  250  mm 
minimum  size  limit  for  L.  carponotatus 
on  the  GBR,  as  well  as  the  timing  of  a 
proposed  October  through  December 
spawning  closure  for  the  fishery.  The 
results  also  caution  against  assessing 
reef-fish  stocks  without  reference  to 
sex-,  size-,  and  location-specific  biologi- 
cal traits. 


Sex-specific  growth  and  mortality,  spawning 
season,  and  female  maturation  of  the  stripey  bass 
{Lutjanus  carponotatus)  on  the  Great  Barrier  Reef 


Jacob  P.  Kritzer 

School  of  Marine  Biology  &  Aquaculture 

and  CRC  Reef  Research  Centre-Effects  of  Line  Fishing  Project 

James  Cook  University 

Townsville.  Queensland  4811,  Australia 

Present  address:  Department  of  Biological  Sciences 

University  of  Windsor 

401  Sunset  Avenue 

Windsor,  Ontario  N9B  3P4,  Canada 
E-mail  address  kntzenSuwindsorca 


Manuscript  approved  for  publication 
22  July  2003  by  Scientific  Editor. 

Manuscript  received  22  July  2003  at 
NMFS  Scientific  Publications  Office. 

Fish  Bull.  102:94-107  (2004). 


Lutjanid  snappers  are  among  the 
most  prominent  species  comprising 
the  catch  of  hook-and-line  fisheries 
on  tropical  reefs  worldwide  (Dalzell, 
1996).  A  notable  exception  is  the  line 
fishery  on  Australia's  Great  Barrier 
Reef  (GBR).  There,  the  finfish  catch, 
and  therefore  the  majority  of  fisheries 
research,  is  dominated  by  coral  trouts 
of  the  genus  Plectropomus  (Mapstone  et 
al.1).  However,  the  GBR  finfish  harvest 
is  diverse  and  the  catch  of  many  sec- 
ondary species  has  risen  steadily  since 
the  early  1990s  (Mapstone  et  al.1). 
Furthermore,  over  the  past  decade, 
the  GBR  fishery  has  changed  with  the 
advent  of  the  lucrative  Asian  live  reef- 
fish  market.  At  present,  only  a  handful 
of  the  many  species  harvested  on  the 
GBR  are  exported  to  the  live  reef-fish 
market.  However,  continued  expansion 
of  the  trade  coupled  with  the  depletion 
of  fish  stocks  in  other  source  nations 
(Bentley2)  has  the  potential  to  intro- 
duce demand  for  a  wider  range  of  spe- 
cies. Even  in  the  absence  of  changes  in 
the  species  composition  of  live  reef-fish 
exports,  increased  demand  for  second- 
ary species  due  to  changes  in  either 
domestic  preferences  or  availability  of 
primary  species  has  the  potential  to 
elevate  harvest  of  currently  nontarget 
species  (Kritzer,  2003). 

Effective  multispecies  management 
of  the  GBR  fishery  will  ultimately  re- 
quire understanding  the  biology  of  more 
than  simply  the  primary  target  species. 
For  example,  spawning  closures  of  the 
fishery  have  been  proposed  for  nine-day 


periods  around  the  new  moon  in  Octo- 
ber, November,  and  December  on  the 
rationale  that  this  will  protect  spawn- 
ing activity  of  a  wide  range  of  harvested 
species  (Queensland  Fisheries  Manage- 
ment Authority3).  Yet,  spawning  season 
information  for  species  beyond  the  com- 
mon coral  trout  {P.  leopardus )  ( Ferreira, 
1995;  Samoilys.  1997 )  is  nearly  nonexis- 
tent. The  GBR  fishery  is  in  a  fortunate 
position  with  respect  to  management 
of  many  species  for  which  exploitation 
is  still  at  relatively  low  levels  because 
baseline  biological  characteristics  can 
be  estimated  before  stock  structure  is 
drastically  altered  by  fishing.  These  da- 
ta can  then  be  used  in  both  formulating 
management  strategies  and  monitoring 
effects  of  fishing. 


1  Mapstone.  B.  D..  J.  P.  MacKinlay,  and  C.  R. 
Davies.  1996.  A  description  of  the  com- 
mercial reef  line  fishery  log  book  data  held 
by  the  Queensland  Fisheries  Management 
Authority.  Report  to  the  Queensland 
Fisheries  Management  Authority.  480  p. 
Primary  Industries  Building,  GPO  Box  4(i. 
Brisbane.  Queensland  4001.  Australia. 

2  Bentley.  N.  1999.  Fishing  for  solutions: 
can  the  live  trade  in  wild  groupers  and 
wrasses  from  Southeast  Asia  be  managed? 
TRAFFIC  Southeast  Asia  report.  143  p. 
Unit  9-3A,  3rd  Floor.  Jalan  SS23/11, 
Taman  SEA.  47400  Petaling  Java,  Selan- 
gor,  Malaysia. 

3  Queensland  Fisheries  Management  Auth- 
ority. 1999.  Queensland  coral  reef  fin 
fish  fishery.  Draft  management  plan  and 
regulatory  impact  statement,  80  p.  Pri- 
mary Industries  Building.  GPO  Box  46, 
Brisbane,  Queensland  4001,  Australia. 


Kritzer:  Sex-specific  growth  and  mortality,  spawning  season,  and  female  maturation  of  Lut/anus  carponotatus 


95 


One  of  the  most  prominent  secondary  species  in  the 
GBR  fishery  is  the  stripey  bass  (Lutjanus  carponotatus) 
(Spanish  flag  snapper.  FAO).  In  relation  to  other  large 
predators  on  the  GBR,  L.  carponotatus  is  highly  abundant 
on  inshore  reefs,  common  on  mid-continental  shelf  reefs, 
and  absent  from  outer-shelf  reefs  (Newman  and  Williams, 
1996;  Newman  et  al.,  1997;  Mapstone  et  al.4).  Although 
this  affinity  for  inshore  reefs  has  the  potential  to  make  the 
species  more  susceptible  to  recreational  fishing,  the  limited 
available  data  do  not  suggest  that  it  is  heavily  exploited 
by  the  recreational  fleet  (Higgs,  1993)  in  relation  to  the 
commercial  fleet  (Mapstone  et  al.1).  Lutjanus  carponota- 
tus has  a  broad-based  diet,  consuming  a  wide  variety  of 
smaller  reef  fishes  and  invertebrates  (Connell,  1998).  Its 
role  as  a  predator  coupled  with  its  abundance,  particularly 
on  inshore  reefs,  suggests  that  the  species  might  have  an 
important  ecological  function  on  the  GBR  in  addition  to  its 
role  as  a  fishery  resource. 

Davies  (1995)  and  Newman  et  al.  (2000)  have  collected 
basic  demographic  data  for  L.  carponotatus  on  the  north- 
ern and  central  GBR,  respectively.  They  both  reported  a 
pronounced  asymptote  in  the  growth  trajectory  and  that 
most  growth  occurred  over  the  first  three  to  five  years  and 
little  subsequent  growth  over  a  lifespan  that  can  reach  15 
to  20  years.  Newman  et  al.  (2000)  also  reported  a  heavily 
male-biased  sample  and  larger  body  sizes  among  males. 
Unlike  age  and  growth  data,  no  information  on  reproduc- 
tion of  L.  carponotatus  has  been  available  despite  that  fact 
that  existing  (minimum  size  limits)  and  proposed  (spawn- 
ing closures)  fisheries  regulations  are  based  largely  on 
reproductive  traits  (Queensland  Fisheries  Management 
Authority3). 

Specific  aims  of  this  study  were  1)  to  estimate  sex  ra- 
tios and  sex-specific  schedules  of  growth  and  mortality; 
2)  to  estimate  age-  and  size-specific  schedules  of  female 
maturation;  3)  to  identify  the  spawning  season;  and  4)  to 
determine  whether  reproductive  output  is  proportional 
to  body  size  by  examining  the  ovary  weight-body  weight 
relationship  and  the  average  spawning  duration  of  large 
and  small  fish.  All  traits  were  estimated  at  the  Palm  Island 
group  on  the  central  GBR.  Additionally,  sex-specific  growth 
and  female  maturity  schedules  were  also  examined  at  the 
Lizard  Island  group  on  the  northern  GBR  to  develop  spa- 
tial comparisons. 


Materials  and  methods 

Field  methods 

Size,  age,  and  reproductive  data  were  obtained  for  465 
L.  carponotatus  collected  by  spear  fishing  on  fringing  reef 
slopes  during  monthly  fishery  independent  sampling  at 


4  Mapstone,  B.  D.,  A.M.  Ayling,  and  J.  H.Choat.  1998.  Habitat, 
cross  shelf  and  regional  patterns  in  the  distributions  and  abun- 
dances of  some  coral  reef  organisms  on  the  northern  Great  Bar- 
rier Reef.  Great  Barrier  Reef  Marine  Park  Authority  research 
publication  48,  71  p.  GPO  Box  1379,  Townsville,  Queensland 
4810,  Australia. 


Pelorus,  Orpheus,  and  Fantome  Islands  in  the  Palm  Island 
group  on  the  central  GBR  ( Fig.  1 )  from  April  1997  through 
March  1998.  No  sampling  took  place  in  January  1998 
because  of  severe  flooding  in  the  area.  To  develop  spatial 
comparisons,  samples  of  118  and  18  fish  were  obtained  in 
October  1997  and  April  1999,  respectively,  by  spear  fishing 
at  the  Lizard  Island  group  approximately  400  km  north  of 
the  Palm  Island  group  (Fig.  1).  Fish  were  collected  from 
depths  of  2  to  15  m  by  teams  of  two  to  four  scuba  divers. 
Lutjanus  carponotatus  most  commonly  inhabits  depths  less 
than  15  m  (Newman  and  Williams,  1996);  therefore  sam- 
pling efforts  encountered  the  majority  of  the  population. 
Fish  were  targeted  as  encountered,  without  preference 
based  on  size,  in  order  to  collect  as  representative  a  sample 
as  possible.  Fish  <150  mm  fork  length  (FL)  were  rare  in 
the  samples  because  they  were  infrequently  observed  on 
reef  slopes  (Kritzer,  2002).  Therefore,  supplemental  spear 
fishing  on  reef  flats  targeting  smaller  fish  was  conducted 
at  the  Palm  Island  group  (n=24)  in  April  and  December 
1999  and  at  the  Lizard  Island  group  (n=25)  in  May  1999 
to  obtain  growth  data  for  size  classes  against  which  the 
primary  sampling  was  biased. 

Total  weight  (TW,  g)  and  FL  (mm)  of  each  specimen 
were  recorded.  Ovaries  and  testes  of  small  lutjanids  on 
the  GBR  are  characterized  by  a  lipid  body  running  along 
the  length  of  each  lobe,  akin  to  that  found  in  tropical 
acanthurids  (Fishelson  et  al.,  1985).  Gonads  and  these 
associated  lipid  bodies  were  removed  and  preserved  in 
FAAC  (formaldehyde  4%,  acetic  acid  5%,  calcium  chloride 
1.3%).  Sagittal  otoliths  were  removed,  cleaned,  and  stored 
for  later  analyses. 

Gonad  processing  and  ovarian  staging 

The  lipid  body  was  removed  from  each  ovary  or  testis  after 
fixation  and  the  weight  of  the  gonad  (GW)  and  lipid  body 
(LW)  were  measured  to  the  nearest  0.01  g.  A  gonadoso- 
matic  index  (GSI)  and  lipidsomatic  index  (LSI;  after  Lobel, 
1989)  were  calculated  for  each  sample  as  the  percentage  of 
TW  represented  by  GW  and  LW,  respectively.  Features  of 
whole  fixed  ovaries  including  color,  speckling,  and  surface 
texture  were  noted  as  potential  criteria  for  macroscopic 
staging  after  comparison  with  samples  processed  histologi- 
cally. Sex  of  the  April  1999  Lizard  Island  group  samples 
was  determined  macroscopically  only,  and  was  therefore 
used  in  sex-specific  growth  analyses  but  not  in  analysis  of 
maturity.  Fish  <150  mm  FL  had  undeveloped  gonads  and 
sex  of  these  specimens  was  not  determined  or  assigned  a 
reproductive  stage. 

A  subsample  of  131  ovaries  spanning  the  range  of  gonad 
sizes  and  external  appearances  were  prepared  for  histo- 
logical examination.  Samoilys  and  Roelofs  (2000)  found 
that  medial  gonad  sections  were  adequate  for  determina- 
tion of  reproductive  status.  Therefore,  a  medial  section  was 
removed  from  one  gonad  lobe,  dehydrated,  and  embedded 
in  paraffin.  Embedded  ovarian  tissues  were  sectioned  at 
5  nm  and  stained  with  hematoxylin  and  eosin.  Ovaries  were 
staged  on  the  basis  of  the  most  advanced  oocyte  stage  pres- 
ent (West,  1990).  Additional  features  used  in  histological 
staging  included  the  presence  of  brown  bodies  and  atretic 


96 


Fishery  Bulletin  102(1 


120° 


130° 


N 

4 


Australia 


Great 
LG      Barrier     15° 
Reef 


PG 

Queensland 


35° 


Lizard  Island  group  (LG) 


J\ 


Lizard 
Island  <\ 

V-— V^"        ,25km, 
Palfrey  Q  .  '  ' 

Island  _       Seabird  Islet 

South  Island 


Palm  Island  group  (PG) 


Pelorus  Island 


Brisk  Island 
<V,Havannah  Island 


Figure  1 

Location  of  the  Palm  Island  group  (PG)  and  Lizard  Island  (LG)  group  within  the  Great  Barrier  Reef 
off  the  coast  of  Queensland,  Australia. 


oocytes  and,  in  the  case  of  inactive  ovaries,  the  relative 
thickness  of  the  gonad  wall  and  the  compactness  of  the 
ovarian  lamellae  (Samoilys  and  Roelofs,  2000).  For  those 
samples  processed  histologically,  macroscopic  features 
were  compared  within  and  between  reproductive  stages  to 
determine  whether  any  macroscopic  characteristics  could 
be  used  to  accurately  stage  ovaries 

Age  determination 

Ages  offish  were  determined  in  order  to  estimate  age-based 
schedules  of  growth,  mortality,  and  maturation.  Otoliths 
lacking  broad,  opaque  macro-increments  were  processed 
to  enumerate  finer  presumed  daily  micro-increments. 
These  otoliths  were  ground  by  hand  first  from  the  anterior 
end  and  then  the  posterior  end  through  a  progressively 
finer  series  of  P1200  sandpaper,  12-um  lapping  film,  and 
9-um  lapping  film  until  a  thin  section  through  the  nucleus 
remained.  Micro-increments  were  enumerated  on  two  inde- 
pendent occasions.  If  the  two  counts  for  one  specimen  did 
not  deviate  by  more  than  5%  of  their  mean,  the  mean  was 
used  as  the  age  estimate.  Otherwise,  a  third  reading  was 
performed  and  the  mean  of  this  and  the  more  similar  of 
the  first  two  readings  was  used  as  the  age  estimate,  again 
provided  that  the  counts  differed  by  no  more  than  5'  <  of 
their  mean. 

Macro-increments  in  the  otoliths  of  L.  carponotatus  have 
been  validated  as  annuli  by  tetracycline  labeling  (Cappo  et 
al.,  2000).  A  pilot  analysis  indicated  that  age  estimates  did 


not  differ  between  readings  of  whole  left  and  right  otoliths 
(paired  t-test:  df=59;  r=0.60;  P=0.55);  therefore  one  otolith 
was  randomly  selected  from  each  sample  for  age  determi- 
nation. All  otoliths  were  initially  read  whole.  A  second  pilot 
analysis  compared  whole  and  sectioned  age  estimates  for 
a  subsample  of  L.  carponotatus  otoliths.  This  comparison 
suggested  that  whole  readings  began  to  drastically  under- 
estimate age  beyond  approximately  sectioned  age  12  (see 
Kritzer,  2002 ).  To  capitalize  on  both  the  greater  efficiency  of 
whole  readings  and  the  greater  accuracy  of  sectioned  read- 
ings, whole  readings  were  used  for  all  fish  except  for  those 
for  which  any  whole  reading  exceeded  10  or  for  which  there 
was  not  agreement  in  at  least  two  out  of  three  independent 
whole  readings.  If  at  least  two  out  of  three  independent 
readings  of  either  whole  or  sectioned  otoliths  (as  appropri- 
ate) agreed,  then  that  value  was  used  as  the  age  estimate. 
Ferreira  and  Russ  (1994)  have  described  the  whole-  and 
sectioned-otolith  preparation  and  reading  methods  used 
in  the  present  study. 

Sex-specific  demography 

Early  growth  of  L.  carponotatus  was  estimated  by  linear 
regression  of  FL  on  age  for  those  samples  processed  to 
read  subannual  micro-increments.  Separate  regressions 
were  performed  for  the  Palm  and  Lizard  island  groups  and 
these  were  compared  by  analysis  of  covariance  ( ANCOVA). 
Because  of  the  undeveloped  nature  of  the  gonads  of  the 
smallest  fish,  early  growth  was  estimated  without  refer- 


Kritzer:  Sex-specific  growth  and  mortality,  spawning  season,  and  female  maturation  of  Lutjanus  carponotatus 


97 


ence  to  sex.  Sex  ratios  at  each  island  group  were  compared 
with  an  expected  ratio  of  1:1  by  x2  goodness-of-fit  tests  by 
using  all  specimens  (i.e.  immature  and  mature)  whose  sex 
could  be  determined. 

Lifetime  growth  parameters  were  estimated  for  males 
and  females  from  each  island  group  by  fitting  the  von  Ber- 
talanffy  growth  function  (VBGF), 


2»=A.(1" 


exp( 


-Kit 


■'„>)). 


where  Lt  =  FL  at  age  t\ 

L^=  the  mean  asymptotic  FL; 

A'  =  the  Brody  growth  coefficient;  and 

t0  =  the  age  at  which  fish  have  theoretical  FL  of  0. 

Growth  functions  were  fitted  by  nonlinear  least-squares 
regression  of  FL  on  age  by  using  samples  for  which  sex  was 
determined.  Because  VBGF  parameter  estimates  can  be 
sensitive  to  the  range  of  ages  and  sizes  used  (see  Ferreira 
and  Russ.  1994,  for  an  empirical  example),  a  common  t0 
equivalent  to  the  .v-intercept  of  the  early  growth  estimates 
was  used  in  all  models  (see  "Results"  section).  Although 
the  sex-specific  sample  sizes  at  the  Lizard  Island  group 
were  smaller  (n=65  for  females;  n=62  for  males),  VBGF 
parameter  estimates  achieved  high  precision  at  sample 
sizes  between  50  and  100  (Kritzer  et  al.,  2001);  therefore 
the  Lizard  Island  group  data  were  included  in  the  analy- 
sis. Growth  parameters  were  compared  by  plotting  959c 
confidence  regions  of  the  parameters  K  and  Lx  (Kimura, 
1980)  for  each  sex  from  each  location  and  assessing  the 
degree  of  overlap. 

Sex-specific  total  mortality  rates,  Z,  were  estimated  by 
using  the  age-based  catch  curve  of  Ricker  (1975)  as  the 
slope  of  a  linear  regression  of  natural  log-transformed  fre- 
quency on  age  class.  Everhart  and  Youngs  ( 1981 )  proposed 
that  catch  curve  analysis  should  exclude  age  classes  with 
n<5  and  Murphy  ( 1997)  proposed  that  age  structures  used 
in  catch  curves  should  be  truncated  at  the  first  age  class 
with  n<5.  Alternatively,  Kritzer  et  al.  (2001)  proposed  that 
a  sample  should  contain  an  average  of  at  least  ten  fish  per 
age  class  irrespective  of  age  class-specific  sample  sizes. 
Therefore  catch  curves  were  fitted  by  two  different  methods 
for  each  sex  at  the  Palm  Island  group.  The  first  catch  curve 
began  at  the  modal  age  class  and  stopped  before  the  first 
age  class  with  n  <  5.  The  second  catch  curve  likewise  began 
at  the  modal  age  class  but  included  all  age  classes  that 
were  thereafter  represented  in  the  data  set.  Sex-specific 
sample  sizes  for  the  Lizard  Island  group  were  too  small  by 
any  of  these  criteria  and  this  location  was  excluded.  Mor- 
tality estimates  for  Palm  Island  group  fish  were  compared 
between  the  fitting  methods  within  each  sex  as  well  as 
between  sexes  by  ANCOVA. 

Reproductive  biology 

Maturation  schedules  of  female  fish  were  estimated  for 
each  island  group  by  fitting  a  logistic  model, 

P,  =  l/(l  +  exp(a-W)), 


where  P-  =   the  proportion  of  mature  fish  in  age  or  20-mm 
size  class  i; 
a        adjusts  the  position  of  the  curve  along  the 

abscissa;  and 
r       determines  its  steepness. 

Age-  and  size-specific  maturity  functions  were  used  to 
estimate  the  mean  age,  r50,  and  size,  L50,  at  which  50%  of 
females  are  mature  at  each  island  group. 

Monthly  mean  LSI  and  GSI  values  of  mature  Palm 
Island  group  fish  were  plotted  separately  for  males  and 
females  to  determine  seasonal  patterns  of  energy  storage 
and  the  peak  spawning  period  of  L.  carponotatus.  The  pro- 
portion of  specimens  at  each  mature  female  reproductive 
stage  in  each  month  was  also  plotted  to  examine  ovarian 
development  patterns  throughout  the  year  and  the  degree 
of  spawning  activity  occurring  outside  of  peak  months. 

To  examine  whether  relative  reproductive  output  in- 
creases with  body  size,  GW  and  GSI  for  stage-IV  ovaries 
collected  during  peak  spawning  months  were  regressed 
against  TW.  Residual  plots  were  used  to  assess  deviation 
from  a  linear  relationship  and  to  identify  three  outliers, 
which  were  removed  from  the  regression  analysis.  Regres- 
sion slopes  were  compared  between  the  two  island  groups 
by  ANCOVA.  Also,  mean  GSI  values  and  the  proportion  of 
Palm  Island  group  females  with  stage-IV  ovaries  during 
spawning  months  were  compared  between  females  <230 
mm  FL  and  those  >230  mm  FL  to  examine  whether  the 
duration  of  spawning  varies  between  size  classes  (nota 
bene:  230  mm  FL  is  approximately  the  mean  size  of  mature 
Palm  Island  group  females  and  splits  each  month's  sample 
approximately  in  half). 


Results 

Ovarian  staging 

Five  female  reproductive  stages  were  identified  through 
histological  analysis  (Table  1)  and  were  based  largely 
on  the  scheme  of  Samoilys  and  Roelofs  (2000).  Ovarian 
stages  I  (immature)  and  II  (resting  mature)  have  similar 
oocyte  stages.  These  can  be  distinguished  by  the  presence 
of  brown  bodies  or  atretic  oocytes,  which  are  typically  prod- 
ucts of  prior  spawning  (e.g.  Ha  and  Kinzie,  1996;  Adams 
et  al.,  2000)  and  are  usually  absent  from  stage-I  ovaries. 
However,  these  structures  will  not  necessarily  persist  in 
ovaries  that  have  spawned,  and  in  fact  were  rare  among 
the  samples;  therefore  identification  of  immature  females 
was  based  primarily  on  structural  organization  of  the 
ovary.  Stage-I  ovaries  typically  have  a  thin  ovarian  wall 
and  more  compacted  oocytes,  whereas  ovaries  that  have 
previously  spawned  tend  to  have  a  thicker  ovarian  wall 
and  a  more  disorganized  arrangement  of  oocytes  (Table  1). 
Also,  there  were  distinct  size  differences  between  stage-I 
ovaries  and  other  stages.  The  mean  GW  of  stage-I  ovaries 
was  approximately  one-third  that  of  stage-II  ovaries,  and 
mean  GSI  was  approximately  one-half  of  that  at  stage  II 
(Table  1),  and  the  distribution  of  body  sizes  offish  at  stage 
I  had  much  lower  minimum,  maximum,  and  modal  size 


98 


Fishery  Bulletin  102(1) 


Table  1 

Description  of  histological  and  macroscopic  features  (after  fixation  in  a  formaldehyde,  acetic  acid,  calcium  chloride  solution  I  of 
ovarian  developmental  stages  of  Lutjanus  carponotatus.  Stage  definitions  and  descriptions  are  largely  a  modification  of  the  scheme 
proposed  by  Samoilys  and  Roelofs  (2000).  Mean  ovary  weight  (GW)  and  gonosomatic  index  (GSI)  for  the  larger  Palm  Island  group 
sample  are  provided. 


Stage 


Histological  features 


Macroscopic  features 


Inactive      I  Immature 


II  Resting 


Active         III  Ripening 


IVa  Ripe 


IVb  Running  ripe 


Relatively  thin  ovarian  wall;  lamellae  well 
packed;  only  darkly  purple  staining  previ- 
tellogenic oocyte  stages  (oogonia  and  peri- 
nucleolar stages)  present. 

Relatively  thick  ovarian  wall;  spaces  be- 
tween lamellae  common;  only  previtellogenic 
oocyte  stages  and  possibly  brown  bodies  and 
few  atretic  vitellogenic  oocytes  present. 


Most  advanced  oocytes  are  at  yolk  globule  or 
migratory  nucleus  stage;  atretic  oocytes  or 
brown  bodies  possibly  present. 


Most  advanced  oocytes  at  yolk  vesicle  stage; 
atretic  oocytes  or  brown  bodies  possibly 
present. 


Similar  to  stage  IVa  but  large,  irregularly 
shaped,  clear  to  lightly  coloured  hydrated 
oocytes  are  present. 


Always  even  white  color  over  entire  surface; 
smooth  surface  texture;  lobes  quite  small  (typi- 
cally <2  cm  long)  and  thin  (mean  GW=0.33  g; 
meanGSI=0.24^). 

Even  white  to  cream  or  tan  color  over  gonad  sur- 
face; surface  may  be  smooth  or  somewhat  convo- 
luted; small  white  stage  II  ovaries  are  difficult  to 
distinguish  from  stage  I  without  histology  I  mean 
GW=1.01  g;  mean  GSI=0.43%). 

Color  sometimes  white  but  more  often  cream  to 
tan;  surface  is  commonly  convoluted;  difficult  to 
distinguish  from  stage  II  without  histology  (mean 
GW=1.18  g;  mean  GSI=0.53%  I. 

Color  tan  to  brown  or  mustard  with  opaque  speck- 
les that  become  larger  and  more  dense  as  late 
stage  oocytes  become  more  numerous;  convoluted 
surface  sometimes  with  prominent  vasculariza- 
tion (mean  GW=4.04  g;  mean  GSI=1.399S  \. 

External  appearance  identical  to  stage  IVa  and 
can  only  be  differentiated  histologically  (no  sam- 
ples found  at  Palm  Island  group). 


classes  compared  with  the  distribution  of  body  sizes  offish 
at  stage  II  (Fig.  2). 

Stage-Ill  (ripening)  ovaries  contain  oocytes  at  the  yolk 
vesicle  vesicle  stage,  which  some  authors  classify  as  vitel- 
logenic (e.g.  Samoilys  and  Roelofs,  2000)  and  others  classify 
as  previtellogenic  (e.g.  West  1990).  Like  stage-II  ovaries, 
stage-Ill  ovaries  can,  but  do  not  necessarily,  contain  brown 
bodies  or  atretic  oocytes  as  evidence  of  probable  prior 
spawning.  Although  the  fish  might  not  have  spawned  pre- 
viously, stage  III  is  considered  to  be  a  mature  stage  in  the 
present  study  because  the  appearance  of  yolk  vesicles  is 
associated  with  the  initial  development  of  the  yolk  globule 
and  represents  advanced  development  of  the  oocyte  beyond 
perinucleolar  stages  (West,  1990).  Therefore,  the  fish  is  pre- 
paring for  spawning  and  will  soon  be  part  of  the  mature 
population  if  it  is  not  already.  Mean  age  and  size  of  stage-II 
(4.4.  years  and  219  mm  FL),  stage-Ill  (5.0  years  and  222 
mm  FL),  and  stage-IV  (6.5  years  and  261  mm  FL)  females 
were  much  more  similar  to  one  another  than  they  were 
to  stage-I  females  (1.9  years  and  119  mm  FL).  Moreover, 
size-frequency  distributions  of  fish  at  stages  II,  III,  and 
IV  showed  considerable  overlap  and  similarity  with  one 
another  and  were  all  quite  distinct  from  the  size-frequency 
distribution  for  stage-I  females  (Fig.  2).  This  suggests  a 
division  between  immature  fish  and  those  that  are  spawn- 
ing or  are  nearly  ready  to  do  so.  The  pronounced  difference 


in  GW  and  GSI  between  stage-I  and  stage-Ill  ovaries  and 
similarity  in  these  metrics  between  stage-II  and  stage-Ill 
fish  (Table  1)  further  support  this  division. 

Most  immature  ovaries  and  all  ripe  ovaries  could  be 
identified  macroscopically.  Because  certain  macroscopic  fea- 
tures were  common  to  multiple  ovarian  stages,  additional 
histological  features  was  required  to  separate  the  largest 
immature  from  the  smallest  resting  ovaries  and  all  ripening 
from  resting  ovaries  among  the  samples  remaining  after 
the  initial  comparison  betw-een  histological  and  macroscopic 
features.  Only  one  ovary  with  fully  hydrated  oocytes,  col- 
lected at  the  Lizard  Island  group,  was  found  among  the 
samples  prepared  for  histological  analysis;  therefore  stages 
IVa  and  IVb  were  treated  as  a  single  stage.  Stage  IV  suf- 
ficiently represents  final  development  toward  spawning  on 
the  broad  seasonal  time  scale  adopted  in  this  study  but 
encompasses  a  wide  range  of  ovarian  characteristics  and 
would  need  to  be  divided  into  more  detailed  stages  for  finer 
temporal  scale  studies  of  lunar  or  diel  spawning  patterns. 
No  samples  exhibited  features  of  truly  "spent"  ovaries. 

Sex-specific  demography 

Differences  were  not  apparent  in  early  growth  of  L.  car- 
ponotatus  between  the  island  groups  (ANCOVA:  df=l, 
46;  F=1.07;  P=0.301);  therefore  the  data  were  pooled  to 


Kntzer:  Sex-specific  growth  and  mortality,  spawning  season,  and  female  maturation  of  Lut/anus  carponotatus 


99 


80 
70 
60 
50  H 
40 
30  H 
20 
10 
0 


rzL 


estimate  an  early  growth  rate  of 
0.76  mm/d,  assuming  daily  period- 
icity of  micro-increments  (Fig.  3). 
This  rate  of  growth  represents  quite 
rapid  growth,  given  that  fish  are 
adding  100  mm  of  length  in  around 
4  months,  increasing  from  approxi- 
mately 20  to  120  mm  FL  (Fig.  3). The 
x-intercept  of  the  early  growth  curve 
(=-17.98  d)  was  divided  by  365  d/yr 
to  estimate  a  common  t0  (=-0.049  yr) 
for  all  VBGF  models. 

Although  size  at  age  for  both  sexes 
at  both  island  groups  was  character- 
ized by  substantial  individual  vari- 
ability, different  growth  trajectories 
were  evident  for  males  and  females 
(Fig.  4,  A  and  B).  Estimates  (Table 
2)  and  95%  joint  confidence  regions 
(Fig.  4C)  for  the  VBGF  parameters 
indicated  that  the  primary  differ- 
ences in  these  trajectories  at  each 
island  group  lay  in  LM  (which  indi- 
cated that  males  grow  larger  than  females).  In 
contrast,  the  common  range  of  K  values  spanned 
by  the  sexes  within  each  island  group  indicated 
similar  curvature  (Table  2,  Fig.  4C).  However,  use 
of  a  common  t0  restricts  the  range  of  possible  fitted 
lvalues  (Kritzer  et  al.,  2001).  In  addition  to  the 
differences  between  the  sexes,  the  data  revealed  a 
general  pattern  of  larger  body  sizes  at  the  Lizard 
Island  group  (Table  2,  Fig.  4). 

Mortality  estimates  at  the  Palm  Island  group 
were  slightly  higher  when  all  age  classes  beyond  1 
year  were  included  compared  with  exclusion  of  age 
classes  with  n  <  5  (Fig.  5).  These  higher  mortality 
estimates  contrast  with  Murphy's  (1997)  finding 
that  truncation  of  the  age  structure  results  in 
higher  least-squares  estimates  of  Z.  The  differ- 
ences between  mortality  rates  estimated  with 
and  without  age  classes  with  n  <  5  were  minor 
for  both  males  (ANCOVA:  df=l,  20;  F=0.009; 
P=0.92)  and  females  (ANCOVA:  df=l,  23;  F=1.35; 
P=0.26).  Therefore,  for  comparisons  between  the 
sexes,  the  estimates  that  included  all  age  classes 
greater  than  1  yr  were  used.  In  contrast  to  the  sex- 
specific  growth  differences,  Z  estimates  of  0.26/yr 
and  0.29/yr  (Fig.  5)  corresponding  to  annual  survivorship 
of  77%  and  75%  for  females  and  males,  respectively,  at  the 
Palm  Island  group  were  similar  between  the  sexes  (ANCO- 
VA: df=l,  27;  F=0.505;  P=0.483).  Murphy's  (1997)  results 
also  suggested  that  least-squares  mortality  estimates  are 
likely  to  be  around  30%  less  than  the  true  mortality  rate 
when  n  =  200  and  the  true  Z  =  0.2/yr.  Correcting  these  mor- 
tality estimates  based  upon  this  potential  bias  results  in  Z 
estimates  up  to  0.37/yr  and  0.41/yr  for  females  and  males, 
respectively,  with  corresponding  annual  survivorship  of 
69%  and  66%-.  However,  the  catch  curve  estimates  (Fig. 
5)  corresponded  well  with  estimates  based  upon  Hoenig's 
(1983)  empirically  derived  relationship  between  Z  and 


n 


In 


I 


I 


□  Stage  I 
■  Stage  II 

□  Stage  III 
El  Stage  IV 


I 


■  ■ 


Bfl 


51 


1         I      i '  '  i  ii,     mi  i  m 


co,,3"ir>cor--ooa>o-<-<Mco-si-ir><or^coa>OT- 

t-t-t-t-^t-t-CM<M<MCMCN(>J(N(MCM<MCOCO 

Size  class  midpoint  (mm  FL) 

Figure  2 

Size-frequency  distributions  of  female  Palm  Island  group  Lutjanus  carponotatus  at 
each  of  the  four  stages  of  ovarian  development.  Stage  descriptions  are  provided  in 
Table  1. 


20  40  60  80         100        120        140        160 

Number  of  increments  (age  in  days) 

Figure  3 

Fork  length  at  microincrement  count  for  Lutjanus  carponotatus  lack- 
ing the  first  annulus  at  the  Palm  (□;  n =24 )  and  Lizard  (■;  n=25)  island 
groups.  Periodicity  of  increments  is  presumed  to  be  daily.  Data  from 
each  island  group  are  distinguished,  but  a  pooled  linear  growth  curve 
is  presented  as  separate  growth  curves  did  not  differ  (ANCOVA:  df=l, 
46;P=1.07;P=0.301). 


maximum  age,  tmax  (females:  tmax=18  yr,  Z=0.23/yr;  males: 
U,=16yr;Z=0.26/yr). 

The  observed  female-to-male  sex  ratios  of  1.3  and  1.1 
were  close  to  unity  at  the  Palm  and  Lizard  Island  groups, 
respectively  ( Table  2 ).  However,  x2  tests  suggest  this  ratio 
is  statistically  different  from  1  at  the  Palm  Island  group 
( df=  1;  ^2=7.74;  P=0.005 )  but  not  at  the  Lizard  Island  group 
(df=l;/2=0.031;P=0.86). 

Age  and  size  at  maturity 

Although  there  was  some  indication  that  Palm  Island  group 
females  mature  at  slightly  younger  ages  and  smaller  sizes 


100 


Fishery  Bulletin  102(1) 


than  Lizard  Island  group  females,  maturation  schedules 
were  generally  similar  (Fig.  6).  At  both  island  groups,  age 
2  was  the  age  at  both  earliest  maturity  and  50%  maturity, 
and  93-100%  of  females  had  matured  by  age  4  (Fig.  6A, 
Table  3).  Thus,  maturation  was  rapid,  beginning  early  in 
life  and  ending  within  a  2-year  period  with  nearly  all  mem- 
bers of  a  cohort  mature.  Length-specific  maturation  sched- 
ules also  exhibited  similarity  between  the  island  groups 
with  mature  fish  first  appearing  in  the  160-179  mm  FL 
size  class,  estimated  50%  maturation  in  the  180-199  mm 
FL  size  class,  and  93-100%  maturity  at  the  220-239  mm 
FL  size  class  (Fig.  6B,  Table  3). 


310 
300  -| 
290 
280 
270  -| 
260 
250  - 
240  - 
230 


0.4 


0.5 


0.6 


0.7 


0.8 


0.9 


K 


Figure  4 

Fork  length  at  age  and  estimated  von  Bertalanffy  growth 
turves  for  male  ■  solid  lines)  and  female  (□.  broken  lines) 
Lutjuiius  larpinuiliiliis  at  the  Palm  iA>  and  Lizard  iBl  island 
groups  and  estimated  959!  joint  confidence  regions  of  the 
parameters  A"  and  l.  (C),  Parameter  estimates  are  presented 
in  Table  2. 


Spawning  season 

Mature  female  LSI  values  were  highest  in  August  through 
October  with  a  maximum  in  September  ( Fig.  7A).  The  peak 
in  GSI  lagged  that  of  LSI  by  two  months  with  the  high- 
est values  occurring  from  October  through  December  and 
with  a  maximum  in  November  (Fig.  7A>.  The  absence  of  a 
January  sample  unfortunately  leaves  some  ambiguity  as  to 
whether  GSI,  and  therefore  presumably  spawning  activity, 
would  still  be  high  at  this  time  or  if  it  would  have  begun 
to  decline.  Male  GSI  values  also  exhibited  a  November 
maximum  (Fig.  7B).  Male  LSI  values,  however,  did  not 
show  any  clear  trend  of  increase  and  decline  throughout 
the  year  and  peaks  in  April,  May,  and  August  that  did 
not  correlate  with  future  GSI  values  as  clearly  as  seen  in 
the  female  data  (Fig.  7).  Unlike  LSI  values  for  females, 
monthly  mean  male  LSI  values  were  always  greater  than 
the  corresponding  GSI  values. 

The  seasonal  pattern  of  L.  carponotatus  spawning 
activity  suggested  by  monthly  trends  in  the  proportions 
of  mature  ovarian  stages  can  be  interpreted  as  differ- 
ent from  that  suggested  by  GSI  values.  The  lowest  GSI 
values  in  the  October-December  peak  period  were  close 
to  twice  as  great  as  the  next  highest  values  in  Septem- 
ber and  February  < Fig.  7A).  However,  the  percentage  of 
stage-IV  ovaries  in  the  September  sample  was  greater 
than  50%.  which  is  well  over  half  the  percentage  of  the 
October  sample;  whereas  the  February  sample  comprised 
approximately  the  same  percentage  of  stage-rV  ovaries  as 
October  (Fig.  8).  Also,  more  than  50%  of  the  March  sample 
was  stage-rV  ovaries  (Fig.  8).  whereas  its  GSI  value  was 
close  to  that  of  the  months  with  relatively  few  ripe  ovaries 
(Fig.  7A).  Furthermore,  September  and  March  had  the 
highest  proportions  of  ripening  (stage-Ill I  females  and 
thus  far  fewer  resting  mature  (stage-II)  females  than  the 
April  to  August  period  of  limited  spawning  activity  I  Fig. 
8).  Therefore,  regardless  of  whether  September,  February, 
and  March  are  defined  as  nonspawning  months  or  months 
of  limited  spawning  activity  based  upon  GSI,  analysis  of 
ovarian  stage  frequencies  suggests  these  to  be  periods 
of  greater  spawning  activity  than  might  be  predicted 
with  GSI.  Clearly,  the  presence  of  advanced  oocytes  is  a 
much  better  indication  of  imminent  spawning  than  any 
measure  of  gonad  size;  therefore  the  reproductive  stage- 
frequency  data  undoubtedly  provide  the  more  accurate 
picture  of  L.  carponotatus  spawning  patterns. 

Of  59  ovaries  staged  from  the  October  1997  Lizard 
Island  group  sample,  eight  were  at  stage  I,  two  were  at 
stage  II,  and  49  (96%  of  mature  females  in  the  sample) 
were  at  stage  PV.  This  finding  suggests  that  the  island 
groups  share  at  least  October  as  a  common  period  of  ac- 
tive spawning. 

Reproductive  differences  between  locations  and  among 
size  classes 

The  variation  in  GW  among  females  of  like  body  sizes 
during  peak  spawning  months  increased  to  some  degree 
with  increasing  TW,  but  there  was  a  generally  homoge- 
neous spread  of  data  around  the  predicted  regression 


Kntzer:  Sex-specific  growth  and  mortality,  spawning  season,  and  female  maturation  of  Lut/anus  carponotatus 


101 


male  ages  2+: 
y  =  -0.289x  +  4.319 
r2  =  0.896 
male  ages  with  n 


0.844 


female  ages  2+: 
y  =  -  0.261  x  +  4.557 
r2  =  0.872 
4:  female  ages  with  n  >  4: 


0.272X  +  4.282      y  =  -  0.203x  +  4.289 


0.879 


6         8        10       12 
Age  class  (years) 


18 


Figure  5 

Age-based  catch  curves  for  female  ■  higher  elevation  lines  i  and  male  (♦.  lower 
elevation  lines)  Lutjanus  carponotatus  at  the  Palm  Island  group  fitted  to  all  age 
classes  >1  (solid  lines  I  and  age  classes  >1  with  n  >4  (dashed  lines  I.  Open  symbols 
represent  age  class  1,  which  was  not  used  in  the  analysis. 


Table  2 

Sex-specific  von  Bertalanffy  growth  parameters 

for  Lu 

tja 

ms  carpom 

tatus  at  the  Palm  and  Lizard  Island  groups 

,  Great  Barrier 

Reef,  n  is  sample  size;  LF 

is  the  mean  fork  length 

( mm ) 

K 

is  the  Brody  growth 

:oefficient 

per  yr) 

L. 

is  the  mean  asymptotic  fork 

length  (mm);  a  common  t 

-,  of -0.049  yr  was  used 

n  all  growth  models. 

Standarc 

errors  are 

provided  below  parameter  estimates. 

n 

LF 

A" 

L. 

r2 

Palm  Island  group 

females 

263 

224.2 
12.11) 

0.77 
(0.032) 

246.3 

(2.25) 

0.515 

males 

202 

224.7 

(2.78) 

0.69 
(0.028) 

264.3 
(3.26) 

0.629 

sex  ratio 

1.3:1 

Lizard  Island  group 

females 

65 

239.9 

(4.76) 

0.56 
(0.043) 

263.5 

(4.24) 

0.618 

males 

62 

256.4 

(4.77) 

0.51 
(0.032) 

284.8 
(4.03) 

0.714 

sex  ratio 

1.1:1 

lines  across  body  sizes  (Fig.  9A).  This  suggests  that  on 
average  GW  at  stage  IV  during  peak  spawning  months  is 
a  linear  function  of  TW.  Lizard  Island  group  fish  generally 
had  larger  ovaries  at  a  given  size  than  did  Palm  Island 
group  fish  (Fig.  9A),  a  difference  supported  by  ANCOVA 
(df=l,  125;  F=34.7;  P<0.001).  In  fact,  regression  slopes  of 
0.25  and  0.52  suggest  relative  ovary  weights  at  the  Lizard 
Island  group  were  approximately  twice  as  large  as  those 
at  the  Palm  Island  group.  There  were  no  differences  in 
the  GW-TW  relationship  among  October,  November,  and 
December  at  the  Palm  Island  group,  and  therefore  the  dif- 
ferences in  this  relationship  between  the  island  groups  was 


consistent  whether  only  the  Palm  Island  group  October 
data  were  used  or  whether  the  October  through  December 
data  were  used. 

Although  GW  is  a  linear  function  of  TW,  the  nonzero 
regression  constants  (Fig.  9A)  mean  that  GW  is  not  a  con- 
stant proportion  of  TW.  Consequently,  GSI  increases  with 
increasing  TW  ( Fig.  9B ).  The  relationship  between  TW  and 
GSI  is  not  strong,  with  regression  slopes  close  to  zero  and 
low  r2  values  at  both  island  groups  (Fig.  9B).  Despite  this, 
the  relationship  is  statistically  strong  at  both  the  Palm 
( ANOVA:  df=l,82;  F=12.70;  P=0.006)  and  Lizard  (ANOVA: 
df=  1,42;  F=22.95;  P<0.0001)  Island  groups.  Also,  there  is 


102 


Fishery  Bulletin  102(1) 


some  suggestion  that,  like  the  GW-TW  relationship,  the 
GSI-TW  relationship  varies  between  the  island  groups, 
although  to  a  much  lesser  extent  (ANCOVA:  df=  1,125; 
F=7.44;P=0.007). 


o 

o 


A 

40 

34  22    19    12 16    18   12    4    5 

2    1       1 

4 

1 

1.0  - 

8 

33      6      29      5741 

2 

1 

56  /o 

E'"B 

0.8  - 

5/    .' 

0.6  - 

a    ■* 

0.4  - 

15 

/"-' 

0.2  - 

3/ 

00  - 

There  is  some  indication  that  larger  fish  spawn  over 
a  longer  period  at  the  Palm  Island  group.  During  the 
September-February  spawning  season,  mean  GSI  values 
were  always  higher  for  mature  Palm  Island  group  females 
>230  mm  FL  compared  with  mature  fe- 
males <230  mm  FL  at  the  same  location 
(Fig.  10).  This  pattern  is  likely  due  in  part 
to  the  higher  relative  gonad  weights  of 
larger  fish  (Fig.  9B)  but  also  seems  to  be 
driven  by  greater  proportions  of  stage-IV 
ovaries  among  larger  mature  females  in 
September,  October,  and  February  com- 
pared with  fish  <230  mm  FL  (Fig.  10). 
During  these  months,  13%,  13%  and  25% 
more  large  fish  were  at  stage  IV,  respec- 
tively, than  were  small  fish. 


0      1      2     3     4     5     6      7 


B 


9    10  11    12  13  14  15  16   17   M 
Age  class  (years) 

53    44    25    11    3      2 


1.0 
0.8 
0.6 
0.4 
0.2 
0.0 


10  50  90  130         170        210        250        290        330 

Size  class  midpoint  (mm  fork  length) 

Figure  6 

Proportion  of  mature  female  Lutjanus  carponotatus  and  estimated  age-spe- 
cific (A)  and  size-specific  (B)  logistic  maturation  schedules  at  the  Palm  ■ 
solid  lines)  and  Lizard  (□,  broken  lines)  island  groups.  Sample  sizes  for  the 
Palm  (top  value)  and  Lizard  (lower  value)  Island  groups  are  presented  above 
the  data  for  each  age  or  size  class.  Parameters  of  the  maturity  functions  are 
provided  in  Table  3. 


Discussion 

Demography  and  reproduction  of 
L.  carponotatus 

Growth  of  L.  carponotatus  is  rapid  for  the 
first  two  years  of  life,  slows  over  the  next 
two  years,  and  nearly  ceases  by  age  4.  The 
slowing  and  cessation  of  growth  coincide 
with  the  ages  at  50%  and  100%  maturity, 
respectively,  and  support  the  argument  of 
Day  and  Taylor  (1997)  that  maturation 
represents  a  pivotal  physiological  trans- 
formation and  consequently  a  fundamen- 
tal shift  in  the  growth  trajectory.  Further 
supporting  the  idea  that  reproductive 
development  occurs  at  the  expense  of 
somatic  growth  is  the  apparently  longer 
average  spawning  season  among  larger 
fish  that  have  ceased  most  somatic  growth. 
The  limited  growth  over  much  of  the  lifes- 


Table  3 

Parameters  of  age-  and 

size- 

specific  logistic  maturation  schedules  anc 

estimated  ages 

and  fork  1 

engths 

at 

50', 

maturity  of  female 

Lu  tja n  11  s  ca rpon otatus 

at  the  Palm  and  Lizar 

d  Island  gi 

oups,  Great  Barrier  Reef. 

a 

adjusts  th 

e  posit 

on 

of  the  logi 

stic  function 

along  the  abscissa;  r  determ 

ines  the  steepness 

of  the  logistic  function. 

f  i;n  is  the  age 

at  50% 

maturity;  Lr 

0  is  the 

fork  length  at  509S 

maturity.  Standard  errors  are  provided  below  parameter 

estimates. 

a 

r 

r2 

*S0  OI"  £50 

Age-specific 

Palm  Island  group 

6.40 
(1.42) 

3.42 
(0.12) 

0.985 

1.9  years 

Lizard  Island  group 

4.16 
(0.48) 

1.73 
(0.19) 

0.990 

2.4  years 

Size-specific 

Palm  Island  group 

14.72 
(1.49) 

0.081 
(0.008) 

0.994 

182  mm 

Lizard  Island  group 

11.61 
(3.84) 

0.061 
(0.020) 

0.908 

189  mm 

Kritzer:  Sex-specific  growth  and  mortality,  spawning  season,  and  female  maturation  of  Lut/anus  carponotatus 


103 


pan  of  L.  carponotatus  can  explain  the 
apparently  constant  mortality  rate 
over  many  age  classes  (evidenced  by 
high  catch  curve  r2  values)  given  that 
mortality  is  often  largely  a  function  of 
body  size  (Roff,  1992). 

The  development  and  regression  of 
visceral  fat  stores  preceding  increases 
in  ovary  weight  is  a  pattern  that  has 
been  observed  in  other  reef  fishes,  in- 
cluding tropical  surgeonfishes  (Acan- 
thuridae:  Fishelson  et  al.,  1985)  and 
groupers  (Serranidae:  Ferreira,  1995) 
and  temperate  rockfishes  (Scorpaeni- 
dae:  Guillemot  et  al.,  1985).  These  pat- 
terns suggest  that  the  stored  lipid  is 
fuelling  the  energetic  costs  of  spawning. 
The  lack  of  a  similar  pattern  for  males 
supports  the  idea  that  energetic  costs 
associated  with  production  of  sperm  are 
low  in  relation  to  eggs  (Wootton,  1985) 
thus  enabling  male  L.  carponotatus  to 
attain  larger  sizes,  as  also  reported  by 
Newman  et  al.  (2000).  Alternatively, 
males  might  spawn  more  frequently 
throughout  the  year  than  females  and 
the  lack  of  seasonal  patterns  in  lipid 
storage  among  males  might  reflect  a 
more  regular  energetic  demand  that 
precludes  energy  storage.  In  any  case, 
these  sex-specific  growth  patterns, 
coupled  with  similar  mortality  rates 
between  the  sexes  and  sex  ratios  that 
are  at  unity  or  that  are  at  most  only 
slightly  female-biased  (see  below),  sug- 
gest that  females  are  limiting  reproduction  of 
this  species.  Therefore  stock  dynamics  should  be 
modeled  in  terms  of  female  biology  (Hilborn  and 
Walters,  1992). 

The  apparently  female-biased  sex  ratio  at 
the  Palm  Island  group  starkly  contrasts  with 
the  heavily  male-biased  sex  ratio  reported  for 
mid-shelf  reefs  of  the  central  GBR  by  Newman 
et  al.  (2000).  However,  neither  a  male-  nor  fe- 
male-biased sex  ratio  would  be  expected  from  a 
nonhermaphrodite  that  is  not  known  to  possess  a 
complex  mating  system  such  as  defense  of  females 
or  territories.  It  is  possible  that  the  spawning  sex 
ratio  (i.e.  excluding  juveniles)  is  closer  to  unity  if 
males  mature  earlier  than  females,  but  this  ratio 
is  not  possible  to  assess  because  male  maturation 
has  not  yet  been  examined  for  this  species.  The 
difference  between  the  sex  ratio  reported  in  this 
study  and  that  by  Newman  et  al.  (2000)  might  be 
due  to  variation  in  mating  systems  across  a  cross- 
shelf  density  gradient  (Newman  and  Williams, 
1996).  Alternatively,  the  sampling  by  traps  and 
line  fishing  conducted  by  Newman  et  al.  (2000)  could  be 
more  heavily  biased  toward  males  than  the  sampling  by 
spear  fishing  used  in  the  present  study  because  of  larger 


%     25 

to 

8      2.0 
1.5 

1.0 
0.5 
0.0 


Figure  7 

Monthly  mean  gonadosomatic  index  IGSI  ±SE;  ■)  and  lipidsomatic  index  (LSI 
±SE;  □)  values  for  mature  female  (A)  and  all  male  (Bl  Lutjanus  carponotatus  at 
the  Palm  Island  group. 


B 

-1.0     P 

CO 

- 

-  0.8 

■ 

-  0.6 

v 

-H" 

-  0.4 

-  0.2 

April 

June 

Aug            Oct 

i          i 
Dec 

Feb 

Month  (1997-98) 

100% 
80% 
I"     60%  - 

CD 

f     40°= 

20% 

0% 


li 


i 


i 


I 


D  Stage  II 
■  Stage  III 
D  Stage  IV 


April      June      Aug       Oct       Dec      Feb 
Month  (1997-98) 

Figure  8 

Monthly  frequencies  of  ovarian  stages  of  mature  Lutjanus  carpono- 
tatus at  the  Palm  Island  group.  Stage  descriptions  are  provided  in 
Table  1. 


size,  wider  gape,  or  more  aggressive  behavior  toward  bait 
among  males  (Cappo  and  Brown,  1996).  Furthermore,  it 
is  likely  that  a  female-biased  sex  ratio  as  observed  at  the 


104 


Fishery  Bulletin  102(1) 


30  -i 

Lizard  group: 

25  - 

y  =  0.052x  -6.33 

Ol 

£       20  J 
1       15  - 

.-'•"          "                   r2  =  0.711 

5      10  - 
o 

5  - 

°     -■"  .     °^^<M^                            Palm  group: 
'$'**T}e^*°^             '                    y  =  0.025x  -  1.62 

■J!i^^^,  "                                   r"  =  °-691 

^&S*,> 

0                   200                 400                 600                 800                1000 

6.0  - 
(J      5.0  - 

B 

Lizard  group: 
_»--'               y  =  0.0071X  +  0.64 

«      4.0- 

■   °              aS°a!-''                                f2  =  0353 

c 

g      3.0  - 

to 

o      2.0  - 

c/i 

o 

13      1.0- 

c 

<§     0.0- 

□                                          D* 

u    m             a        £    . '     °           a                            ° 

'      °  •*!..■■''         -          ^—~—~' 

■          n    ^4  "          ■         D            □ ■ 

m         '  •  °~        "                  — m 

•  ■ '  <£J-zi~*~^°         '                             Palm  group: 
_*^*~""^°  ■    ■  '                     .               y  =  0.0029x  +  1.02 
."^1  "■  "                                                          r2  =  0.134 

0                   200                400                 600                800                1000 

Whole  body  weight  (g) 

Figure  9 

Fixed  ovary  weight  (A)  and  gonadosomatic  index  iBi  at  fresh  whole  body 

weight  for  mature  female  Lutjanus  carponotatus  at  ovarian  stage  IV  (see 

Table  1)  collected  during  peak  spawning  months  (Oct-Dec)  at  the  Palm 

■  solid  lines)  and  Lizard  (□,  dashed  lines)  island  groups. 

Sep        Oct         Nov        Dec        Jan 
Month  (1997-98) 


Feb 


Mar 


Figure  10 

Mean  gonadosomatic  index  (GSI  ±SE)  for  mature  female 
Lutjanus  carponotatus  at  the  Palm  Island  group  during  the 
September  through  March  spawning  season  for  small  (<230  mm 
fork  length;  ■>  and  large  (<230  mm  fork  length;  □)  size  classes. 
The  percentage  of  fish  at  stage  IV  (see  Table  1)  is  indicated  above 
each  data  point. 


Palm  Island  group  is  not  a  prevalent  feature 
of  L.  carponotatus  populations.  Rather,  the 
strong  statistical  suggestion  of  a  sex  ratio 
quite  different  from  unity  might  be  due  to 
the  fact  that  sex  ratios  often  show  temporal 
variability  (e.g.  Stergiou  et  al.,  1996)  coupled 
with  the  propensity  to  achieve  statistically 
significant  differences  when  using  large 
sample  sizes  (Johnson,  1999). 

Maturation  schedules  and  sex-specific 
growth  differences  were  consistent  between 
the  island  groups,  but  overall  growth  pat- 
terns differed,  with  Lizard  Island  group  fish 
reaching  larger  asymptotic  body  sizes.  Given 
the  vast  distance  between  the  island  groups, 
these  differences  might  be  due  to  inherent 
genetic  differences  between  the  populations. 
Or,  effects  of  temperature  (the  Palm  Island 
group  sits  at  a  higher  latitude),  turbidity, 
freshwater  run-off  (the  Palm  Island  group 
sits  closer  to  a  river  mouth  and  has  more 
developed  mangrove  systems),  or  other 
environmental  factors  could  be  driving  the 
differences.  Of  course,  these  possibilities  are 
not  mutually  exclusive. 

The  larger  ovaries  observed  among  Liz- 
ard Island  females  might  be  due  to  further 
spatial  differences  or  might  be  an  effect  of 
timing  of  sampling.  The  temporal  resolution 
of  sampling  aimed  to  identify  the  extent  of 
the  spawning  season  but  was  too  coarse  to 
account  for  intramonth  differences  in  ovar- 
ian development.  Large  changes  in  ovary 
size  might  occur  within  stage  IV,  and  the 
final  progression  to  immediate  prespawning 
stages  can  be  rapid  (e.g.  Davis  and  West, 
1993).  The  Lizard  Island  group  sample  was  collected 
from  17  to  23  October  1998,  whereas  the  corresponding 
Palm  Island  group  sample  was  collected  from  11  to  12 
October  1998.  The  October  1998  new  moon  was  on  the 
20th,  and  P.  leopardus,  the  only  GBR  species  for  which 
lunar  spawning  patterns  have  been  reported,  spawns 
primarily  around  the  new  moon  (Samoilys,  1997).  If 
L.  carponotatus  spawning  is  also  centered  around  the 
new  moon,  the  spatial  differences  in  ovary  weight  at 
body  weight  might  be  due  to  more  advanced  develop- 
ment toward  full  hydration  within  the  Lizard  Island 
group  sample.  In  fact,  the  higher  proportion  of  stage-FV 
ovaries  within  the  October  Lizard  Island  group  sample 
(96%)  compared  with  the  October  Palm  Island  group 
sample  (78'i ),  coupled  with  the  higher  relative  ovary 
weights  at  the  Lizard  Island  group  in  October,  can  be 
taken  as  preliminary  evidence  that  L.  carponotatus 
spawns  at  the  new  moon. 

Comparison  with  other  reef  fishes 

The  growth  differences  between  male  and  female  L. 
carponotatus  contrast  with  a  general  trend  of  larger 
body  sizes  among  female  lutjanids  observed  in  Atlan- 


Kritzer:  Sex-specific  growth  and  mortality,  spawning  season,  and  female  maturation  of  Lut/anus  carponotatus 


105 


tic,  Caribbean,  and  Hawaiian  species  (Grimes,  1987). 
However,  the  pattern  observed  in  the  present  study  seems 
common  in  the  Indo-Pacific  where  males  frequently  ( Davis 
and  West,  1992;  McPherson  and  Squire,  1992;  Newman  et 
al.,  1996,  2000).  but  not  universally  (Hilomen,  1997),  are 
the  larger  sex.  As  noted  above,  these  differences  are  consis- 
tent with  predictions  based  on  energetic  costs  of  producing 
sperm  and  eggs. 

Lutjanus  carponotatus  spawning  patterns  identified  by 
using  both  GSI  and  ovarian  stage  frequencies  show  pro- 
nounced seasonal  differences:  there  are  at  least  five  months 
of  very  limited  or  no  spawning  activity  from  April  through 
August.  This  finding  supports  Grimes's  ( 1987)  observation 
that  continental  lutjanid  populations  tend  to  have  more 
restricted  spawning  seasons  than  populations  associated 
with  oceanic  islands,  which  spawn  more  or  less  continu- 
ously throughout  the  year.  Although  seasonal  patterns  ex- 
ist, the  prominence  of  ripe  gonads  over  seven  months  from 
September  through  March  suggests  an  extended  spawning 
season  and  supports  the  general  observation  that  tropical 
reef  fishes  spawn  over  longer  periods  within  the  year  than 
do  cooler  water  species  (Lowe-McConnell,  1979).  However, 
a  study  with  finer  temporal  resolution  is  needed  to  verify 
that  spawning  actually  occurs  in  months  with  a  high  pro- 
portion of  stage-IV  ovaries. 

Female  L.  carponotatus  mature  on  average  at  approxi- 
mately 75%  of  their  mean  asymptotic  size,  54%  of  their 
maximum  observed  size,  and  119c  of  their  maximum 
longevity.  The  relative  size  at  maturity  contrasts  with 
Grimes's  ( 1987)  observations  that  shallow-water  continen- 
tal lutjanid  populations  like  those  of  L.  carponotatus  on  the 
GBR  typically  mature  at  smaller  relative  sizes  (=42%  maxi- 
mum size)  compared  to  deep-water  populations  associated 
with  oceanic  islands  (=50%  maximum  size).  Two  sympatric 
shallow -water  species,  L.  russelli  (Sheaves,  1995)  and  L. 
fulviflamma  (Hilomen,  1997),  likewise  contrast  with  the 
general  familial  trend  and  mature  at  approximately  50% 
and  75%  of  their  maximum  size,  respectively.  Hence,  a 
general  pattern  of  relative  size  at  maturity  might  exist 
among  shallow-water  lutjanids  in  the  GBR  region  that 
is  different  from  those  regions  covered  by  Grimes's  ( 1987 ) 
review.  Lutjanids  on  the  GBR  are  generally  lightly  fished 
(Mapstone  et  al.1);  therefore  the  geographic  difference  in 
sizes  at  maturity  might  be  due  to  fishing  pressure  selecting 
for  smaller  sizes  at  maturity  in  other  regions. 

The  relative  age  at  maturity  of  L.  carponotatus  cannot  be 
as  readily  placed  in  a  broader  familial  context  given  that 
ages  at  maturity  were  not  widely  estimated  for  lutjanids 
at  the  time  of  Grimes's  (1987)  review.  However,  an  array 
of  published  studies  suggests  that  many  tropical  and  sub- 
tropical demersal  fishes  share  the  absolute,  but  not  relative, 
ages  of  L.  carponotatus  at  50%  and  100%  maturity  at  2  and 
4  years,  respectively.  These  include  other  small  gonochores 
on  the  GBR  (Sheaves,  1995;  Hart  and  Russ,  1996;  Hilomen, 
1997),  as  well  as  a  range  of  gonochores  in  other  regions 
(Grimes  and  Huntsman,  1980;  Davis  and  West,  199.3;  Ross 
et  al.,  1995 )  and  hermaphrodites  on  the  GBR  and  elsewhere 
(Ferreira,  1993,  1995;  Bullock  and  Murphy,  1994).  The 
ubiquity  of  this  maturity  schedule,  despite  a  wide  array  of 
maximum  body  sizes  (160-1200  mm)  and  longevities  (6-56 


years)  among  these  species,  perhaps  suggests  a  common 
physiological  threshold  toward  which  many  species  gravi- 
tate in  order  to  maximize  lifetime  reproductive  success. 
More  comprehensive  analysis  of  life  history  trade-offs  (e.g. 
Roff,  1992)  is  needed  to  test  this  hypothesis. 

Fisheries  management 

Harvest  of  L.  carponotatus  is  currently  restricted  to  fish 
greater  than  250  mm  total  length  ( approximately  233  mm 
FL)  with  the  aim  of  allowing  50%  offish  to  spawn  at  least 
once,  and  this  regulation  is  proposed  to  remain  after  revi- 
sion by  the  GBR  fishery  management  plan  (Queensland 
Fisheries  Management  Authority3).  The  estimated  size 
at  50%  maturity  of  190  mm  FL  suggests  that  the  regula- 
tion is  meeting  its  objective.  However,  the  objective  itself 
might  not  adequately  protect  the  reproductive  potential  of 
L.  carponotatus  and  similar  species  if  individuals  require 
multiple  spawning  years  to  ensure  sufficient  replenish- 
ment of  the  stock.  The  extensive  longevities  of  many  reef 
fishes  have  been  hypothesized  to  be  a  mechanism  for  coping 
with  low  and  irregular  recruitment  rates  through  a  process 
dubbed  the  "storage  effect"  (Warner  and  Chesson,  1985). 
The  rationale  behind  the  storage  effect  hypothesis  is  that 
fish  must  reproduce  during  many  breeding  seasons  in  order 
to  endure  poor  recruitment  years  and  realize  high  repro- 
ductive success  during  the  unpredictable  and  intermittent 
good  recruitment  years.  If  this  process  is  important  for 
population  dynamics  of  L.  carponotatus  and  other  species, 
management  will  need  to  protect  an  intact  natural  popula- 
tion structure  in  some  areas  within  the  fishery.  Protecting 
older  age  classes  cannot  be  achieved  by  using  maximum 
size  limits  for  species  like  L.  carponotatus  that  have  a  pro- 
nounced asymptote  in  the  growth  trajectory  because  body 
sizes  are  similar  over  a  broad  range  of  age  classes  and  size 
is  therefore  poorly  correlated  with  age.  Protecting  natural 
age  structure  could  be  accomplished  through  a  system  of 
strategically  designed  marine  protected  areas  that  allow 
some  populations  to  experience  natural  survival  free  of 
fishing  mortality. 

Proposed  closures  of  the  GBR  line  fishery  during  nine- 
day  periods  around  the  new  moon  in  October,  November, 
and  December  are  aimed  at  protecting  spawning  activity 
and  particularly  spawning  aggregations  of  P.  leopardus 
and  other  harvested  species  (Queensland  Fisheries  Man- 
agement Authority3).  Lutjanus  carponotatus  shares  a  peak 
spawning  period  during  these  months  with  P.  leopardus 
(Ferreira,  1995;  Samoilys  1997)  and  several  other  sym- 
patric exploited  species  (McPherson  et  al.,  1992;  Sheaves, 
1995;  Hilomen,  1997;  Brown  et  al.5).  In  addition,  the 
larger  ovaries  of  the  Lizard  Island  group  fish,  which  were 
collected  closer  to  the  new  moon,  may  indicate  that,  like 
P.  leopardus  (Samoilys,  1997),  L.  carponotatus  spawns  at 


5  Brown,  I.  W.,  P.  J.  Doherty,  B.  Ferreira,  C.  Keenan,  G.  McPher- 
son, G.  Russ,  M.  Samoilys,  and  W.  Sumpton.  1994.  Growth, 
reproduction  and  recruitment  of  Great  Barrier  Reef  food  fish 
stocks.  Final  report  to  the  Fisheries  Research  and  Development 
Corporation,  FRDC  Project  90/18,  Queensland  Department  of 
Primary  Industries,  154  p.  Southern  Fisheries  Centre,  GPO 
Box  76.  Deception  Bay,  Queensland  4508,  Australia. 


106 


Fishery  Bulletin  102(1 


the  new  moon.  Therefore,  the  timing  of  the  proposed  spawn- 
ing closures  seems  appropriate.  However,  it  is  not  known 
whether  L.  carponotatus  aggregate  to  spawn;  therefore  the 
goal  of  protecting  spawning  aggregations  might  not  be  rel- 
evant for  this  species.  In  fact,  the  prevalence  and  ecological 
importance  of  spawning  aggregations  for  any  species  on 
the  GBR  is  largely  unknown;  therefore  the  efficacy  of  the 
proposed  closures  is  difficult  to  predict. 

Beyond  the  implications  for  management  regulations, 
these  data  have  implications  for  modeling  L.  carponotatus 
stock  dynamics.  In  particular,  the  results  suggest  that 
reproductive  output  by  a  unit  of  L.  carponotatus  biomass 
cannot  be  predicted  on  the  basis  of  that  biomass  alone. 
Relative  ovary  weight  increases  slightly  with  increasing 
body  size  and  there  is  evidence  that  larger  fish  spawn  more 
frequently.  The  greatest  difference  in  the  proportion  of  ripe 
ovaries  between  size  classes  occurred  in  February  1998  af- 
ter severe  flooding  in  January.  It  is  possible  that  the  lower 
proportion  of  ripe  ovaries  among  small  fish  in  February  was 
due  to  stresses  caused  by  changes  in  salinity  or  increased 
run-off  and  is  not  a  regular  trait.  However,  increased  resil- 
ience to  environmental  stresses  that  allows  more  frequent 
spawning  would  also  increase  the  relative  reproductive 
success  of  large  fish.  Therefore,  a  population  comprising 
fewer  larger  fish  is  likely  to  show  greater  annual  egg  pro- 
duction than  a  population  with  equivalent  biomass  that 
comprises  more  numerous  but  smaller  fish.  Additionally, 
the  sex-specific  patterns  reported  in  this  study  further 
suggest  gross  biomass  might  be  an  inadequate  index  of 
replenishment  potential  and  that  female  biomass  needs 
to  be  considered.  Therefore,  stock  structure,  in  terms  of 
sex  ratio  and  the  frequency  of  size  classes,  and  not  simply 
overall  biomass  needs  to  be  considered  when  predicting 
reproductive  potential. 


Acknowledgments 

I  thank  the  numerous  assistants  who  participated  in 
fieldwork,  as  well  as  Sam  Adams  and  Sue  Reilly  for  assis- 
tance with  histological  examinations.  The  manuscript  was 
greatly  improved  by  comments  from  Howard  Choat,  Carl 
Walters,  Tony  Fowler,  Campbell  Davies,  Sam  Adams,  Bruce 
Mapstone,  an  anonymous  thesis  examiner,  and  two  anony- 
mous reviewers.  This  work  was  conducted  while  the  author 
was  supported  by  an  international  postgraduate  research 
scholarship  from  the  Commonwealth  of  Australia  and  a 
postgraduate  stipend  from  the  CRC  Reef  Research  Centre. 
Final  preparation  of  the  manuscript  took  place  while  the 
author  was  supported  by  a  postdoctoral  fellowship  funded 
jointly  by  the  University  of  Windsor  and  the  Canadian 
National  Science  and  Engineering  Research  Council  (col- 
laborative research  opportunity  grant  no.  227965-00)  to 
Peter  Sale  and  others). 


Literature  cited 

Adams,  S.,  B.  D.  Mapstone,  G.  R.  Russ,  and  C.  R.  Davies. 

2000.     Geographic  variation  in  the  sex  ratio,  sex  specific  size, 


and  age  structure  of  Plectropomus  leopardus  (Serranidael 
between  reefs  open  and  closed  to  fishing  on  the  Great  Bar- 
rier Reef.     Can.  J.  Fish.  Aquatic  Sci.  57:  1448-1458. 
Bullock,  L.  H..  and  M.  D.  Murphy. 

1994.  Aspects  of  the  life  history  of  the  yellowmouth  grouper. 
Mycteroperca  interstitialis,  in  the  eastern  Gulf  of  Mexico. 
Bull.  Mar.  Sci.  55:30-45. 

Cappo.  M..  and  I.  W.  Brown. 

1996.     Evaluation  of  sampling  methods  for  reef  fish  of  com- 
mercial and  recreational  interest.     CRC  Reef  Research 
Centre  Technical  Report  6.  72  p. 
Cappo,  M.,  P.  Eden,  S.  J.  Newman,  and  S.  Robertson. 

2000.    A  new  approach  to  validation  of  periodicity  and  timing 
of  opaque  zone  formation  in  the  otoliths  of  eleven  species  of 
Lutjanus  from  the  central  Great  Barrier  Reef.     Fish.  Bull. 
98:474-488. 
Connell,  S.  D. 

1998.     Patterns  of  piscivory  by  resident  predatory  reef  fish 
at  One  Tree  Reef,  Great  Barrier  Reef.    Mar.  Freshw.  Res. 
49:25-30. 
Dalzell,  P. 

1996.  Catch  rates,  selectivity  and  yields  of  reef  fishing.  In 
Reef  fisheries  (N.  V.  C.  Polunin  and  C.  M.  Roberts,  eds.), 
p.  161-192.     Chapman  and  Hall,  London. 

Davies,  C.  R. 

1995.  Patterns  of  movement  of  three  species  of  coral  reef  fish 
on  the  Great  Barrier  Reef.  Ph.D.  diss,  212  p.  James  Cook 
University.  Queensland,  Australia. 

Davis,  T  L.  0„  and  G.  J.  West. 

1992.  Growth  and  mortality  of  Lutjanus  vittus  (Quoy  and 
Gaimard  I  from  the  North  West  Shelf  of  Australia.  Fish. 
Bull.  90:395-404. 

1993.  Maturation,  reproductive  seasonality,  fecundity,  and 
spawning  frequency  in  Lutjanus  vittus  (Quoy  and  Gaimard i 
from  the  North  West  Shelf  of  Australia.  Fish.  Bull.  91: 
224-236. 

Day, T.,  and  P.  D.Taylor. 

1997.  Von  Bertalanffy's  growth  equation  should  not  be 
used  to  model  age  and  size  at  maturity.  Am.  Nat.  149: 
381-393. 

Everhart,  W  H.,  and  W.  D.  Youngs. 

1981.     Principles  of  fishery  science,  349  p.     Cornell  Univ. 
Press,  Ithaca,  NY. 
Ferreira.  B.  P. 

1993.  Reproduction  of  the  inshore  coral  trout  Plectropo- 
mus maculatus  (Perciformes:  Serranidael  from  the  central 
Great  Barrier  Reef,  Australia.     J.  Fish  Biol.  42:831-844. 

1995.     Reproduction  of  the  common  coral  trout  Plectropomus 
leopardus  (Serranidae:  Epinephelinael  from  the  central  and 
northern  Great  Barrier  Reef.  Australia.     Bull.  Mar.  Sci.  56: 
653-669. 
Ferreira.  B.  P.,  and  G.  R.  Russ 

1994.  Age  validation  and  estimation  of  growth  rate  of  the 
coral  trout.  Plectropomus  leopardus,  (Lacepede  18021  from 
Lizard  Island,  northern  Great  Barrier  Reef.  Fish.  Bull. 
92:46-57. 

Fishelson,  L.,  W.  L.  Montgomery,  and  A.  A.  Myrberg. 

1985.    A  new  fat  body  associated  with  the  gonad  of  surgeon- 
fishes  (Acanthuridae:  Teleosteil.     Mar.  Biol.  86:109-112. 
Grimes,  C.  B. 

1987.     Reproductive  biology  of  the  Lutjanidae:  a  review. 
In  Tropical  snappers  and  groupers:  biology  and  fisheries 
management  (J.  Polovina  and  S.  Ralston.,  eds.),  p.  239-294. 
Westview  Press,  London. 
Grimes.  C.  B.,  and  G.  R.  Huntsman. 

1980.     Reproductive  biology  of  the  vermilion  snapper. 


Kritzer:  Sex-specific  growth  and  mortality,  spawning  season,  and  female  maturation  of  Lut/anus  carponotatus 


107 


Rhomboplites  aurorubens,  from  North  Carolina  and  South 

Carolina.     Fish.  Bull.  78:137-146. 
Guillemot,  P.  J.,  R.  J.  Larson,  and  W.  H.  Lenarz. 

1985.     Seasonal  cycles  of  fat  and  gonad  volume  in  five  species 

of  northern  California  rockfish  (Scorpaenidae:  Sebastes). 

Fish.  Bull.  83:299-311. 
Ha,  P.  Y,  and  R.  A.  Kinzie. 

1996.     Reproductive  biology  of  Awaous  guamensis,  an  amphi- 

dromous  Hawaiian  goby.     Environ.  Biol.  Fish.  45:383-396. 
Hart,  A.  M.  and  G.  R.  Russ. 

1996.  Response  of  herbivorous  fishes  to  crown-of-thorns 
Acanthaster  planci  outbreaks.  III.  Age,  growth,  mortality 
and  maturity  indices  of  Acanthurus  nigrofuscus.  Mar. 
Ecol.  Prog.  Ser.  136:25-35. 

Higgs,  J.  B. 

1993.     A  descriptive  analysis  of  records  of  the  recreational 
reef-line  fishery  on  the  Great  Barrier  Reef.     M.Sc.  thesis, 
128  p.     James  Cook  Univ.,  Queensland,  Australia. 
Hilborn.  R.,  and  C.  J.  Walters. 

1992.     Quantitative  fisheries  stock  assessment:  choice, 
dynamics  and  uncertainty,  570  p.     Kluwer  Academic, 
Boston,  MA. 
Hilomen,  V.  V. 

1997.  Inter-  and  intra-habitat  movement  patterns  and 
population  dynamics  of  small  reef  fishes  of  commercial  and 
recreational  significance.  Ph.D.  diss.,  277  p.  James  Cook 
Univ.,  Queensland,  Australia. 

Hoenig,  J.  M. 

1983.     Empirical  use  of  longevity  data  to  estimate  mortality 
rates.    Fish.  Bull.  82:898-902. 
Johnson,  D.  H. 

1999.     The  insignificance  of  statistical  significance  testing. 
J.  Wildlife  Manag.  63:763-772. 
Kimura,  D.  K. 

1980.     Likelihood  methods  for  the  von  Bertalanffy  growth 
curve.    Fish.  Bull.  77:765-776. 
Kritzer.  J.  P. 

2002.  Variation  in  the  population  biology  of  stripey  bass 
Lutjanus  carponotatus  within  and  between  two  island 
groups  on  the  Great  Barrier  Reef.  Mar.  Ecol.  Prog.  Ser. 
243:191-207. 

2003.  Biology  and  management  of  small  snappers  on  the 
Great  Barrier  Reef.  In  Bridging  the  gap:  a  workshop 
linking  student  research  with  fisheries  stakeholders  (A.  J. 
Williams,  D.  J.  Welch,  G.  Muldoon.  R.  Marriott,  J.  P.  Kritzer, 
and  S.  Adams  eds.  1.  p.  62-80.  CRC  Reef  Research  Centre, 
Townsville,  Queensland,  Australia. 

Kritzer,  J.  P.,  C.  R.  Davies,  and  B.  D.  Mapstone 

2001.     Characterizing  fish  populations:  effects  of  sample 
size  and  population  structure  on  the  precision  of  demo- 
graphic parameter  estimates.     Can.  J.  Fish.  Aquat.  Sci.  58: 
1557-1568. 
Lobel.  P.  S. 

1989.     Ocean  current  variability  and  the  spawning  season  of 
Hawaiian  reef  fishes.     Env.  Biol.  Fish.  24:161-171. 
Lowe-McConnell,  R.  H. 

1979.     Ecological  aspects  of  seasonality  in  fishes  of  tropical 
waters.    Symp.  Zool.  Soc.  Lond.  44:219-241. 
McPherson.  G.  R.,  and  L.  Squire 

1992.    Age  and  growth  of  three  dominant  Lutjanus  species 
of  the  Great  Barrier  Reef  inter-reef  fishery.    Asian  Fish. 
Sci.  5:25-36. 
McPherson,  G.R.,  L.  Squire,  and  J.  O'Brien 

1992.     Reproduction  of  three  dominant  Lutjanus  species  of 


the  Great  Barrier  Reef  inter-reef  fishery.    Asian  Fish.  Sci. 
5:15-24. 

Murphy,  M.  D. 

1997.  Bias  in  Chapman-Robson  and  least-squares  estima- 
tors of  mortality  rates  for  steady  state  populations.  Fish. 
Bull.  95:863-868. 

Newman,  S.  J.,  and  D.  M.  Williams 

1996.  Variation  in  reef  associated  assemblages  of  the  Lut- 
janidae  and  Lethrinidae  at  different  distances  offshore  in 
the  central  Great  Barrier  Reef.  Environ.  Biol.  Fish.  46: 
123-138. 

Newman,  S.  J.,  D.  M.  Williams,  and  G.  R.  Russ 

1996.  Age  validation,  growth  and  mortality  rates  of  the  tropi- 
cal snappers  (Pisces:  Lutjanidae)  Lutjanus  adetii  (Castel- 
nau,  18731  and  L.  quinquelineatus  iBloch,  1790)  from  the 
central  Great  Barrier  Reef,  Australia.  Mar.  Freshw.  Res. 
47:575-584. 

1997.  Patterns  of  zonation  of  assemblages  of  the  Lutjanidae, 
Lethrinidae  and  Serranidae  (Epinephilinaei  within  and 
among  mid-shelf  and  outer-shelf  reefs  in  the  central  Great 
Barrier  Reef.     Mar.  Freshw.  Res.  48:119-128. 

Newman.  S.  J.,  M.  Cappo,  and  D.  M.  Williams 

2000.    Age,  growth  and  mortality  of  the  stripey,  Lutjanus 
carponotatus  (Richardson)  and  the  brown-stripe  snapper, 
L.  vitta  I  Quoy  and  Gaimard  i  from  the  central  Great  Barrier 
Reef,  Australia.     Fish.  Res.  48:263-275. 
Ricker,  W.  E. 

1975.     Computation  and  interpretation  of  biological  statis- 
tics offish  populations,  382  p.     Bull.  Fish.  Res.  Board  Can. 
191. 
Roff,  DA. 

1992.    The  evolution  of  life  histories,  535  p.     Chapman  and 
Hall.  New  York,  NY. 
Ross,  J.  L.,  T  M.  Stevens,  and  D.  S.  Vaughan 

1995.    Age.  growth,  mortality,  and  reproductive  biology  of 
red  drums  in  North  Carolina  waters.     Trans.  Am.  Fish. 
Soc.  124:37-54. 
Samoilys.  M.  A. 

1997.     Periodicity  of  spawning  aggregations  of  coral  trout 
Plectropomus   leopardus   (Pisces:  Serranidae)  on  the 
northern  Great  Barrier  Reef.     Mar.  Ecol.  Prog.  Ser.  160: 
149-159. 
Samoilys,  M.  A.,  and  A.  Roelofs. 

2000.     Defining  the  reproductive  biology  of  a  large  serranid, 
Plectropomus  leopardus.     CRC  Reef  Research  Centre  Tech- 
nical Report  31.  36  p. 
Sheaves,  M. 

1995.  Large  lutjanid  and  serranid  fishes  in  tropical  estuar- 
ies: are  they  adults  or  juveniles?  Mar.  Ecol.  Prog.  Ser.  129: 
31-40. 

Stergiou,  K.  I.,  P.  Economidis,  and  A.  Sinis. 

1996.  Sex  ratio,  spawning  season  and  size  at  maturity  of 
red  bandfish  in  the  western  Aegean  Sea.  J.  Fish.  Biol.  49: 
561-572. 

Warner,  R.  R.,  and  P.  L.  Chesson. 

1985.     Coexistence  mediated  by  recruitment  fluctuations:  a 
field  guide  to  the  storage  effect.    Am.  Nat.  125:769-787. 
West,  G. 

1990.     Methods  of  assessing  ovarian  development  in  fishes:  a 
review.    Aust.  J.  Mar.  Freshw.  Res.  41:199-222. 
Wootton,  R.  J. 

1985.  Energetics  of  reproduction.  In  Fish  energetics:  new 
perspectives  (P.  Tytler  and  P.  Calow,  eds.),  p.  231-254. 
Croom  Helm,  London. 


108 


Abstract— The  increase  in  harbor  seal 
(Phoca  vitulina  richardsi)  abundance, 
concurrent  with  the  decrease  in  sal- 
monid  [Oncorhynehus  spp.)  and  other 
fish  stocks,  raises  concerns  about  the 
potential  negative  impact  of  seals  on 
fish  populations.  Although  harbor  seals 
are  found  in  rivers  and  estuaries,  their 
presence  is  not  necessarily  indicative 
of  exclusive  or  predominant  feeding  in 
these  systems.  We  examined  the  diet 
of  harbor  seals  in  the  Umpqua  River, 
Oregon,  during  1997  and  1998  to  indi- 
rectly assess  whether  or  not  they  were 
feeding  in  the  river.  Fish  otoliths  and 
other  skeletal  structures  were  recov- 
ered from  651  scats  and  used  to  identify 
seal  prey.  The  use  of  all  diagnostic  prey 
structures,  rather  than  just  otoliths, 
increased  our  estimates  of  the  number 
of  taxa,  the  minimum  number  of  indi- 
viduals and  percent  frequency  of  occur- 
rence C^FO)  of  prey  consumed.  The 
*7fFO  indicated  that  the  most  common 
prey  were  pleuronectids,  Pacific  hake 
(Merluccius  produetus),  Pacific  stag- 
horn  sculpin  [Leptocottus  armatus), 
osmerids.  and  shiner  surfperch  (Cyma- 
togaster  aggregata ).  The  majority  ( 76%) 
of  prey  were  fish  that  inhabit  marine 
waters  exclusively  and  fish  found  in 
marine  and  estuarine  areas  (e.g.  anad- 
romous  spp. )  which  would  indicate  that 
seals  forage  predominantly  at  sea  and 
use  the  estuary  for  resting  and  opportu- 
nistic feeding.  Salmonid  remains  were 
encountered  in  39  samples  (6%);  two 
samples  contained  identifiable  otoliths, 
which  were  determined  to  be  from  Chi- 
nook salmon  (O.  tshawytscha).  Because 
of  the  complex  salmonid  composition  in 
the  Umpqua  River,  we  used  molecular 
genetic  techniques  on  salmonid  bones 
retrieved  from  scat  to  discern  species 
that  were  rare  from  those  that  were 
abundant.  Of  the  37  scats  with  salmo- 
nid bones  but  no  otoliths,  bones  were 
identified  genetically  as  chinook  or  coho 
(O.  kisutch)  salmon,  or  steelhead  trout 
(O.  mykiss)  in  90'?  of  the  samples. 


Examination  of  the  foraging  habits  of 

Pacific  harbor  seal  (Phoca  vitulina  richardsi) 

to  describe  their  use  of  the  Umpqua  River,  Oregon, 

and  their  predation  on  salmonids 


Anthony  J.  Orr 

Adria  S.  Banks 

Steve  Mellman 

Harriet  R.  Huber 

Robert  L.  DeLong 

National  Marine  Mammal  Laboratory 

Alaska  Fisheries  Science  Center,  NMFS,  NOAA 

7600  Sand  Point  Way  NE 

Seattle,  Washington  98115 

E-mail  address  (for  A.  J.  Orr,  contact  author)  tony.orr  gnoaa.gov 

Robin  F.  Brown 

Oregon  Department  of  Fish  and  Wildlife 
2040  S  E.  Marine  Science  Drive 
Newport,  Oregon  97365 


Manuscript  approved  for  publication 
9  October  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:108-117  (2004). 


The  Pacific  harbor  seal  (Phoca  vitulina 
richardsi)  is  found  along  the  west  coast 
of  North  America  from  the  Aleutian 
Islands,  Alaska,  to  the  San  Roque 
Islands.  Baja  California  (King,  1983; 
Reeves  et  al.,  1992).  Before  the  pas- 
sage of  the  Marine  Mammal  Protection 
Act  (MMPA)  of  1972,  harbor  seals  in 
Oregon  were  kept  at  relatively  low 
numbers  (fewer  than  500  animals  in 
1968)  because  of  bounties  offered  by  the 
state  and  harassment  from  commercial 
and  sport  fishermen  (Pearson  and  Verts, 
1970).  Since  passage  of  protective  leg- 
islation, harbor  seals  in  Oregon  have 
increased  an  average  of  6^  to  7%  annu- 
ally between  1978  and  1998,  although, 
in  recent  years,  numbers  appear  to 
be  leveling  at  about  8000  individuals 
(Brown  and  Kohlmann.  1998). 

The  rapid  increase  in  harbor  seal 
numbers  has  revived  fishery-manag- 
ers' interest  in  seal  diet  because  of  the 
potential  for  increased  consumption  of 
commercial  fish  species.  In  addition, 
there  has  been  a  heightened  concern 
about  greater  harbor  seal  abundance 
in  rivers  and  estuaries  during  migra- 
tions of  depressed  salmonid  popula- 
tions because  of  the  potential  negative 
impact  on  the  recovery  of  these  fishes 


(NMFS,  1997).  Because  of  the  tenuous 
status  of  many  salmonid  (Oncorhyn- 
ehus spp.  I  species  along  the  west  coast, 
the  National  Marine  Fisheries  Service 
( NMFS )  recommended  that  the  United 
States  Congress  modify  the  MMPA  to 
allow  lethal  removal  of  seals  from  river 
mouths  where  they  may  prey  on  de- 
pressed salmonid  populations  (NMFS. 
1997 ).  Predation  of  salmonids  by  harbor 
seals  in  Oregon  has  been  documented 
(Brown,  1980;  Harvey.  1987;  Brown 
et  al.,  1995;  Riemer  and  Brown,  1997; 
Beach  et  al.1).  The  proportion  of  salmo- 
nids in  the  diet  of  harbor  seals  varied 
from  1%  to  30'r  depending  on  area, 
season,  and  sampling  method  (NMFS, 
1997). 

Pinniped  prey  consumption  can  be 
determined  from  direct  observations 
in  some  systems,  if  prey  is  consumed  at 


1  Beach,  R..  A.  Geiger.  S.  Jefferies.  S.  Treacy, 
and  B.  Troutman.  1985.  Marine  mam- 
mals and  their  interactions  with  fisheries 
of  the  Columbia  River  and  adjacent  waters, 
1980-1982.  NWAFC  (Northwest  Alaska 
Fisheries  Science  Center)  processed  rep. 
NWAFC  85-04,  316  p.  NWAFC,  National 
Marine  Fisheries  Service,  Seattle,  WA, 
98115. 


Orr  et  al.:  Foraging  habits  of  Phoca  vitulma  richardsi  in  the  Umpqua  River,  Oregon 


109 


Pacific  Ocean 


A 


N 


the  surface  (Bigg  et  al.,  1990);  however, 
consumption  is  typically  determined  by 
examining  scat  (fecal)  samples.  In  the 
past,  species-specific  sagittal  otoliths 
found  in  scats  were  used  exclusively 
to  determine  the  identification  of  prey 
taxa.  However,  because  otoliths  can  be 
partially  or  completely  digested,  or  are 
not  present  in  scats  (because  the  head  of 
the  prey  was  not  consumed ),  they  are  not 
always  an  adequate  representation  of  di- 
et. Recently,  investigators  have  begun  to 
use  additional  structures  (e.g.  cranial  el- 
ements, vertebrae)  recovered  from  scats 
to  identify  prey  (e.g.  Olesiuk  et  al.,  1990; 
Cottrell  et  al.,  1996;  Riemer  and  Brown, 
1997;  Browne  et  al.,  2002;  Lance  et  al.2). 
These  structures  usually  are  more  com- 
mon than  otoliths  and  frequently  can  be 
identified  to  species;  however,  bones  of 
some  species  can  be  identified  to  family 
only  (e.g.  salmonids).  Consequently,  the 
National  Marine  Mammal  Laboratory 
(NMML)  collaborated  with  the  Conser- 
vation Biology  Molecular  Genetics 
Laboratory  (CBMGL;  Northwest  Fish- 
eries Science  Center,  Seattle,  WA)  to 
develop  molecular  genetic  identification 
of  salmonid  species  (Purcell  et  al.,  2004). 
Because  of  the  complex  salmonid  species 
composition  in  the  Umpqua  River,  genetic  identification 
was  vital  to  distinguish  species  that  were  rare  from  those 
that  were  abundant. 

The  original  impetus  of  this  study  was  to  assess  the 
impact  of  harbor  seal  predation  on  the  recovery  of  the 
Umpqua  River  sea-run  cutthroat  trout  (O.  clarkii)  that 
were  listed  as  endangered  under  the  Endangered  Species 
Act  (ESA)  during  1996  (Johnson  et  al.,  1999).  Umpqua 
River  cutthroat  trout  were  removed  from  the  ESA  in  2000 
because  they  were  identified  to  be  part  of  the  larger  Oregon 
Coast  evolutionary  significant  unit  (U.S.  Fish  and  Wildlife 
Service,  2000).  The  present  study  was  continued  despite 
the  "delisting"  of  cutthroat  trout  because  the  Umpqua  is 
inhabited  year-round  by  harbor  seals  that  haul  out  sev- 
eral kilometers  upriver  and  is,  thus,  ideal  for  determining 
whether  the  presence  of  a  pinniped  species  within  a  sys- 
tem is  indicative  of  substantial  feeding  on  fish  species  of 
concern  within  that  environment.  In  addition,  the  Umpqua 
River  contains  several  other  salmonid  species  whose  status 
is  precarious  (NMFS,  1997).  Therefore,  the  development  of 
genetic  identification  techniques  was  considered  valuable 
for  this  system,  as  well  as  for  future  foraging  studies  in 
which  species-specific  identification  may  be  desirable  but 
impossible  by  way  of  conventional  identification  methods. 


Oregon 


L  mpquu  River 


hauiouts 


2  Lance,  M.,  A.  Orr,  S.  Riemer,  M.  Weise,  and  J.  Laake.  2001. 
Pinniped  food  habits  and  prev  identification  techniques  pro- 
tocol. AFSC  Proc.  Rep.  2001-04,  36  p.  AFSC,  NMFS,  NOAA. 
7600  Sand  Point  Way  NE,  Seattle.  WA  98115. 


Figure  1 

Map  of  the  lower  section  of  the  Umpqua  River,  Oregon,  where  scat  samples  were 
collected  at  two  haulout  sites  during  1997  and  1998. 


The  objectives  of  this  study  were  1 )  to  determine  by  an 
examination  of  diet  if  harbor  seals  that  haul  out  in  the 
Umpqua  River  feed  primarily  in  the  river  or  elsewhere, 
and  2)  to  apply  genetic  techniques  to  identify  salmonid 
prey  species. 


Materials  and  methods 

Study  area 

The  Umpqua  River,  located  in  southern  Oregon  ( Fig.  1 ).  is 
a  natal  river  for  sea-run  cutthroat  trout,  as  well  as  chinook 
(O.  tshawytscha),  coho  (O.  kisutch)  salmon,  and  steelhead 
trout  (O.  mykiss).  The  Umpqua  estuary  is  also  inhabited 
year-round  by  approximately  600-1000  harbor  seals  and 
has  been  designated  as  an  area  where  pinnipeds  and  sal- 
monids significantly  co-occur  (NMFS,  1997).  Scat  samples 
for  this  study  were  collected  from  two  hauiouts  located 
within  4.8  km  of  the  river's  mouth  and  within  1.6  km  of 
each  other  (Fig.  1). 

Scat  collection  and  analysis 

Samples  were  collected  during  two  seasons:  "spring" 
(March  through  June)  and  "fall"  (August  to  December). 
"Spring"  corresponded  to  the  migration  of  anadromous 
cutthroat  trout  adults  and  some  juveniles  to  the  ocean  and 
"fall"  coincided  approximately  with  the  freshwater  return 
of  spawning  anadromous  adults.  The  migratory  and  spawn- 


110 


Fishery  Bulletin  102(1) 


Table  1 

Collection  dates  of  harbor  seal  scats  and  numbers  of  scats  wi 

th  identifiable  prey  remains,  without 

identifiable  remains 

and  without 

remains  from  the  Umpqua 

River,  Oregon, 

during  1997  and 

1998 

Fall  and  spring 

periods 

correspond 

to  timing 

of  cutthi 

oat  trout 

runs  on  the  Umpqua  River. 

Collection  dates                  With  identifiable  remains 

Without 

dentifiable  remains 

Without  remains 

Total 

Fall,  1997 

16-23  Sep 

26 

1 

2 

29 

27  Sep-6  Oct 

5 

0 

3 

8 

12-24  Oct 

31 

0 

7 

38 

31  Oct-lONov 

21 

0 

6 

27 

12-25  Nov 

36 

0 

10 

46 

Total 

119 

1 

28 

148 

Spring  1998 

24-25  Mar 

27 

5 

2 

34 

13-15  Apr 

59 

5 

7 

71 

26-27  Apr 

45 

4 

4 

53 

13-14  May 

41 

0 

4 

45 

27-28  May 

12 

0 

1 

13 

11-12  Jun 

35 

2 

1 

38 

Total 

219 

16 

19 

254 

Fall  1998 

5-6  Aug 

142 

1 

1 

144 

19-20  Aug 

111 

1 

3 

115 

6-9  Sep 

28 

3 

3 

34 

19-21  Sep 

13 

0 

0 

13 

7-8  Oct 

19 

0 

1 

20 

Total 

313 

5 

8 

326 

ing  periods  of  chinook  and  coho  salmon,  and  steelhead  trout 
also  occur  during  these  times. 

During  fall  1997,  all  harbor  seal  scats  present  at  the 
haulouts  were  collected  every  other  day  during  the  day- 
time low  tide,  weather  permitting  (Table  1).  In  1998.  bi- 
weekly attempts  were  made  to  pick  a  minimum  of  50  scats 
during  low  tides  at  the  haulout  sites  (Table  1).  Scats  were 
collected,  placed  in  individual  plastic  bags,  and  frozen  for 
later  processing.  At  the  laboratory  samples  were  thawed 
and  rinsed  in  nested  sieves  (1.0  mm,  0.71  mm,  and  0.5  mm 
in  1997;  1.4  mm,  1.0  mm,  and  0.5  mm  in  1998).  Fish  struc- 
tures were  dried  and  stored  in  glass  vials  and  cephalopod 
remains  were  stored  in  vials  with  70"*  isopropyl  or  ethyl 
alcohol. 

Prey  were  identified  to  the  lowest  possible  taxon  by  using 
sagittal  otoliths,  skeletal,  and  cartilaginous  remains  from 
fish  and  beaks  and  statoliths  from  cephalopods.  Other  in- 
vertebrate remains  were  discarded  from  analysis  because 
of  the  uncertainty  of  identifying  them  as  primary  or  sec- 
ondary prey.  Unknown  prey  were  categorized  as  "unidenti- 
fied" and  "unidentifiable"  (Browne  et  al.,  2002).  Items  that 
were  categorized  as  "unidentifiable"  were  excluded  from 
analyses  because  they  could  not  be  distinguished  from 
prey  already  identified  in  the  sample.  Otoliths,  beaks,  and 
diagnostic  bones  were  identified  by  using  an  extensive  ref- 
erence collection  at  the  NMML  and  voucher  samples  veri- 
fied by  Pacific  Identifications  (Victoria,  British  Columbia). 


After  identification,  otoliths  were  separated  by  side  (left, 
right,  or  unknown )  and  enumerated  to  determine  minimum 
number  of  specific  prey.  Unique  diagnostic  structures  (e.g. 
quadrates,  angulars,  basioccipitals,  vomers)  were  used  for 
identification  and  enumeration  offish.  Non-unique  skeletal 
structures  such  as  gillrakers  and  teeth  were  used  to  iden- 
tify but  not  enumerate  taxa  (i.e.  their  presence  indicated 
only  a  single  individual)  unless  the  structures  were  from 
different  size  classes.  Vertebrae  were  treated  like  other 
non-unique  structures;  however,  for  salmon,  if  the  number 
of  vertebrae  reflected  more  than  one  individual,  then  they 
were  used  for  enumeration.  Cephalopod  beaks  were  sepa- 
rated by  side  (upper,  lower,  or  unknown)  and  enumerated 
to  determine  number  of  prey. 

To  discern  where  harbor  seals  were  feeding,  identified 
prey  were  categorized  as  those  exclusively  found  in  rivers 
or  estuaries  (e.g.  gobiids,  cyprinids),  those  found  exclu- 
sively in  marine  waters  (e.g.  gadids,  mvxinids),  and  those 
that  could  potentially  be  found  in  either  environment  (e.g. 
anadromous  species,  osmerids,  petromyzontids)  by  using 
Eschmeyer  et  al.  (1983).  A  seal  was  considered  to  feed  in 
the  river-estuary  system  if  all  the  prey  taxa  identified  in 
the  scat  were  definitely  or  could  potentially  be  found  in  the 
system.  For  example,  a  sample  containing  remains  of  pea- 
mouth  chub  iMylocheilus  caurinus),  threespine  stickleback 
( Gasterosteus  aculeatus ),  river  lamprey  iLampetra  ayresii ), 
and  chinook  salmon  would  be  classified  as  a  riverine- 


Orr  et  al.:  Foraging  habits  of  Phoca  vitulina  richardsi  in  the  Umpqua  River,  Oregon 


111 


estuarine  species  because  these  prey  items  could  feasibly 
be  consumed  in  the  river.  It  was  assumed  that  the  seal  was 
feeding  in  the  marine  environment  if  a  sample  contained 
exclusively  marine  prey,  such  as  Pacific  hagfish  (Eptatretus 
stoutti).  Pacific  hake  (Merluceius  productus),  and  rockfish 
(Sebastes  spp. ).  If  a  scat  comprised  prey  taxa  that  poten- 
tially could  be  found  in  a  riverine-estuarine  system  or 
marine  waters  (e.g.  salmonids,  osmerids),  as  well  as  those 
found  exclusively  in  marine  waters,  then  it  was  assumed 
that  the  feeding  environment  was  marine  or  mixed. 

Salmonid  skeletal  remains  were  sent  to  the  CBMGL  for 
species  identification.  Remains  to  be  analyzed  genetically 
were  selected  by  number  or  size  (or  both)  to  represent  dif- 
ferent species  or  individuals  present  in  each  scat.  For  ex- 
ample, if  a  scat  had  95  approximately  equal-size  vertebrae 
(a  salmonid  has  approximately  65  vertebrae;  Butler,  1990). 
then  at  least  two  vertebrae  (potentially  representing  at 
least  two  individuals)  were  sent  for  genetic  identification. 
Also,  if  a  sample  had  a  very  large  gillraker  and  three  small 
vertebrae,  then  the  gillraker  and  one  vertebra  were  sent 
for  genetic  identification.  The  size  of  diagnostic  structures 
was  also  used  to  categorize  salmon  remains  as  juvenile  or 
adult,  when  possible.  The  CBMGL  identified  salmonid  spe- 
cies by  direct  sequencing  of  mitochondrial  DNA  or  analysis 
of  restriction  fragment  length  polymorphism  (Purcell  et  al., 
2004). 

The  abundance  of  prey  taxa  in  harbor  seal  diet  for  each 
period  was  described  by  using  the  minimum  number  of 
individuals  (MNI)  and  percent  frequency  of  occurrence 
(%FO).  We  compared  the  effect  of  including  bone  on  the 
number  of  prey  consumed  by  estimating  MNI  using  the 
greater  number  of  right  or  left  otoliths  and  then  again 
using  all  diagnostic  skeletal  remains.  Cephalopod  MNI 
was  estimated  from  the  greater  number  of  upper  or  lower 
beaks.  The  %  FO  of  prey  taxon  i  was  defined  as 


I°" 


%FO, 


x  100, 


where  Oll;  =  absence  (0)  or  presence  (1)  of  taxon  i  in  scat 
k\  and 
s  =  the  total  number  of  scats  that  contained 
identifiable  prey  remains. 

The  presence  of  taxon  ;'  in  scat  k  was  determined  by  using 
otoliths  and  then  again  using  all  structures.  To  account  for 
variability  in  diet,  point  estimates  of  %FO  for  a  prey  taxon 
were  determined  during  each  sampling  period  and  then 
averaged  for  each  season. 


Results 

Scats 

Over  725  scats  were  collected  during  all  periods.  The 
number  of  scats  collected  with  identifiable  remains  was 
119  (99%;  n=148)  in  fall  1997,  219  (93%;  ?z=254)  in  spring 
1998,  and  313  (98%;  n=326)  in  fall  1998  (Table  1).  Of  the 


651  samples  with  identifiable  prey  remains,  605  (93%)  con- 
tained fish  bones,  347  (53%)  had  fish  otoliths,  231  (36%) 
contained  remains  from  cartilaginous  fish,  and  41  (6% )  had 
cephalopod  beaks.  A  majority  (65%  fall  1997,  65%  spring 
1998,  63%  fall  1998)  of  scats  with  identifiable  remains  had 
one  to  three  prey  taxa  present  and  less  than  4%  contained 
more  than  ten  taxa.  Approximately  40  prey  taxa,  repre- 
senting at  least  25  families,  were  identified  throughout  the 
study  (Tables  2  and  3). 

For  nearly  all  prey  taxa,  MNI  was  greater  when  all  skel- 
etal remains  were  identified  than  when  otoliths  were  used 
exclusively  (Table  2).  For  several  species,  such  as  Pacific 
hake.  Pacific  herring  (Clupea  pallasii),  and  Pacific  sardine 
{Sardinops  sagax),  MNI  at  least  tripled  when  all  structures 
were  used  for  enumeration  (Table  2).  For  most  salmonids, 
cartilaginous  fishes,  three-spine  stickleback,  Irish  lords 
(Hemilepidotus  spp.),  and  Pacific  mackerel  {Scomber  ja- 
ponicus),  no  otoliths  were  recovered;  therefore  other  skel- 
etal elements  had  to  be  used  for  identification  (Table  2). 
For  a  few  prey,  such  as  cyprinids,  gobiids,  and  butter  sole 
(Isopsetta  isolepis),  only  otoliths  were  recovered  (Table  2). 

Foraging  habits 

The  %FO  for  most  prey  taxa  was  greater  when  all  struc- 
tures were  used  than  when  j  ust  otoliths  were  used  ( Table  3 ). 
The  %FO  indicated  that  the  prey  most  frequently  con- 
sumed were  pleuronectids.  Pacific  hake.  Pacific  staghorn 
sculpin  {Leptocottus  armatus),  osmerids,  and  shiner  surf- 
perch  (Cymatogaster  aggregata).  Prey  frequently  found 
in  scats  included  those  that  were  exclusively  marine  (e.g. 
Pacific  hake,  rex  sole  (Glyptocephalus  zachirus),  English 
sole  (Parophiys  vetulus),  and  myxinids),  and  those  that 
occur  in  both  marine  and  estuarine  waters  (e.g.  Pacific 
staghorn  sculpin.  and  shiner  surfperch  (Table  3] ).  Only  24% 
of  scats  were  composed  entirely  of  prey  taxa  that  could  be 
found  in  riverine-estuarine  systems  (Fig.  2).  Consequently, 
a  majority  of  the  scats  contained  prey  species  that  were 
exclusively  marine  (.v=25.3%)  or  were  a  mixture  of  marine 
and  potentially  marine  species  (x=50.8%\  Fig.  2). 

Salmonids 

Salmonid  remains  were  found  in  only  6%  (39/651)  of  the 
samples.  Five  chinook  smolts  were  identified  from  otoliths 
in  two  samples  collected  during  fall  1997;  in  the  remaining 
37  samples,  salmonid  bones  were  unidentifiable  to  species 
with  conventional  techniques.  With  the  cooperation  of 
CBMGL,  we  examined  116  salmonid  bones  using  molecular 
genetic  techniques.  Species  identification  was  successful 
for  67%  (78/116)  of  the  bones  and  teeth  from  90%  (35/39) 
of  the  scat  samples  that  contained  salmonid  structures.  In 
the  four  samples  that  remained  unidentified,  three  con- 
tained only  a  single  salmonid  bone  that  failed  to  produce 
any  DNA.  Most  of  the  other  bones  where  DNA  could  not  be 
extracted  were  small  or  fragmented  and  highly  digested. 
Seventeen  of  the  samples  contained  chinook  salmon  bones 
(including  the  two  samples  with  chinook  salmon  otoliths); 
11  contained  coho  salmon  bones,  four  contained  steelhead 
trout  bones,  and  three  contained  bones  from  two  salmonid 


112 


Fishery  Bulletin  102(1) 


Table  2 

Minimum  number  of  individuals  ( MNI )  offish  prey  derived  from  sagittal  otoliths  and  all  structures  retrieved  from 

harbor  seal  scats 

collected  at  the  Umpqua  River  during  1997  and  1998.  s  represents  the  number  of  scats  with  identifiable 

remains,  na  indicates  taxon 

did  not  have  sagittal  otoliths  to  be  used  for  identification. 

Fall  1997(s=119i 

Spring 

1998(s=219i 

Fall  1998(s=313) 

MNI 

MNI 

MNI 

MNI 

MNI 

MNI 

Family 

Species 

otoliths 

all  structures 

otoliths 

all  structures 

otoliths 

all  structures 

Ammodytidae 

Pacific  sand  lance 

205 

208 

317 

321 

3 

7 

Bothidae 

Pacific  sanddab 

12 

13 

9 

9 

1 

2 

Clupeidae 

American  shad 

1 

2 

4 

11 

1 

15 

Pacific  herring 

6 

22 

3 

10 

121 

345 

Pacific  sardine 

0 

0 

50 

235 

39 

185 

Cottidae 

Pacific  staghorn  sculpin 

44 

65 

25 

48 

30 

85 

unidentified  cottid 

0 

0 

0 

0 

0 

8 

Cyprinidae 

peamouth  chub 

1 

1 

4 

4 

4 

4 

Embiotocidae 

shiner  surfperch 

104 

109 

209 

274 

23 

104 

Engraulididae 

northern  anchovy 

1 

3 

0 

0 

1 

2 

Gadidae 

Pacific  hake 

1 

35 

10 

44 

58 

199 

Pacific  tomcod 

9 

21 

19 

52 

8 

26 

Gasterosteidae 

threespine  stickleback 

0 

1 

0 

0 

0 

0 

Gobiidae 

unidentified  gobiid 

2 

2 

1 

1 

0 

0 

Hexagrammidae 

lingcod 

0 

1 

0 

0 

1 

1 

Myxinidae 

Pacific  hagfish 

0 

20 

0 

13 

0 

61 

Ophidiidae 

spotted  cusk-eel 

0 

0 

4 

4 

2 

2 

Osmeridae 

unidentified  osmerid 

42 

54 

14 

41 

105 

132 

Petromyzontidae 

Pacific  lamprey 

na 

5 

na 

89 

na 

41 

river  lamprey 

na 

2 

na 

1 

na 

0 

Pholididae 

saddleback  gunnel 

3 

7 

1 

3 

0 

1 

Pleuronectidae 

English  sole 

38 

41 

37 

39 

75 

84 

Dover  sole 

1 

4 

5 

6 

27 

51 

slender  sole 

1 

1 

18 

24 

28 

42 

butter  sole 

1 

1 

15 

15 

2 

2 

rex  sole 

19 

44 

44 

53 

96 

125 

petrale  sole 

0 

0 

0 

0 

1 

1 

starry  flounder 

10 

17 

8 

12 

6 

31 

Rajidae 

unidentified  rajid 

na 

1 

na 

7 

na 

4 

Scombridae 

Pacific  mackerel 

0 

2 

0 

3 

0 

2 

Scorpaenidae 

Sebastes  spp. 

0 

15 

6 

19 

2 

3 

Trichodontidae 

Pacific  sandfish 

0 

0 

0 

1 

2 

3 

Zoarcidae 

unidentified  zoarcid 

0 

0 

0 

0 

2 

2 

Salmonidae 

coho  salmon 

unknown 

0 

4 

0 

0 

0 

0 

juvenile 

0 

1 

0 

4 

0 

2 

adult 

0 

0 

0 

1 

0 

3 

Steelhead  or  rainbow  trou 

t 

unknown 

0 

0 

0 

2 

0 

2 

juvenile 

0 

0 

0 

0 

0 

1 

chinook  salmon 

unknown 

5 

6 

0 

0 

0 

3 

juvenile 

0 

5 

0 

2 

0 

5 

adult 

0 

1 

0 

0 

0 

0 

unidentified  salmonid 

unknown 

0 

2 

0 

1 

0 

2 

juvenile 

0 

1 

0 

0 

0 

1 

Orr  et  al.:  Foraging  habits  of  Phoca  vitulina  richardsi  in  the  Umpqua  River,  Oregon 


113 


Table  3 

Mean  percent  frequency  of  occurrence  (%FO)  of  common  prey  recovered  from  harbor  seal  scat  samples  collected  at  haulout  sites  in 

the  Umpqua  River, 

Oregon,  during  1997  and  1998. 

SD  indicates  standard  deviation. 

Family 

Species 

Fall  1997 

Spring  1997 

Fall  1998 

Mean(±SD) 

Mean(±SD) 

Mean(±SDl 

Ammodytidae 

Pacific  sand  lance 

12.5  ±8.3 

12.6  ±8.3 

9.1  ±8.9 

Bothidae 

Pacific  sanddab 

11.4  ±7.5 

4.1  ±2.5 

3.0  ±3.2 

Clupeidae 

American  shad 

4.3  ±0.6 

13.0  ±2.3 

5.3  ±3.1 

Pacific  herring 

16.9  ±13.7 

7.3  ±6.9 

35.9  ±21.8 

Pacific  sardine 

0 

16.1  ±12.2 

17.9  ±9.1 

Cottidae 

Pacific  staghorn  sculpin 

23.9  ±8.5 

21.0  ±19.0 

11.8  ±4.5 

unidentified  cottid 

16.5  ±20.4 

3.2  ±0.7 

0.8  ±0.1 

Cyprinidae 

peamouth  chub 

3.8 

2.3  ±0.6 

2.8 

Embiotocidae 

shiner  surfperch 

18.2  ±8.2 

23.6  ±19.4 

7.0  ±2.9 

Engraulididae 

northern  anchovy 

5.5  ±3.2 

0 

2.1  ±2.0 

Gadidae 

Pacific  hake 

27.9+9.7 

17.0  ±5.7 

41.6  ±25.5 

Pacific  tomcod 

15.4  ±7.8 

16.1  ±7.0 

12.3  ±8.3 

Gasterosteidae 

threespine  stickleback 

2.8 

0 

0 

Gobiidae 

unidentified  gobiid 

7.7 

1.7 

0 

Hexagrammidae 

lingcod 

3.8 

0 

0.7 

Loliginidae 

market  squid 

12.8  ±10.2 

3.5  ±1.3 

0 

Myxinidae 

Pacific  hagfish 

17.5  ±7.9 

6.7  ±3.5 

16.5  ±9.4 

Octopodidae 

Octopus  rubescens 

3.8  ±1.4 

8.3  ±2.6 

8.4  ±7.0 

Ophidiidae 

spotted  cusk-eel 

0 

0 

0.9 

Osmeridae 

unidentified  osmerid 

20.8  ±11.3 

14.6  ±8.2 

19.5  ±10.0 

Petromyzontidae 

Pacific  lamprey 

7.7  ±8.2 

20.5  ±10.1 

8.2  ±2.9 

river  lamprey 

5.6 

3.7 

0 

Pholididae 

saddleback  gunnel 

14.7  ±16.9 

2.6  ±0.3 

5.3 

Pleuronectidae 

English  sole 

21.9  ±1.7 

8.7  ±5.2 

17.5  ±12.0 

Dover  sole 

7.4  ±5.9 

4.6  ±0.7 

13.5  ±13.6 

slender  sole 

0 

11.0  ±7.2 

14.9  ±14.9 

butter  sole 

3.8 

7.2  ±3.7 

1.4 

rex  sole 

27.4  ±12.1 

14.2  ±9.6 

19.9  ±20.5 

petrale  sole 

0 

0 

0.7 

starry  flounder 

15.8  ±7.4 

3.7  ±1.0 

5.8  ±1.2 

Rajidae 

unidentified  rajid 

2.8 

5.0  ±1.6 

2.8 

Scombridae 

Pacific  mackerel 

3.8  ±1.4 

4.6  ±4.0 

0.8  ±0.1 

Scorpaenidae 

Sebastes  spp. 

15.7  ±8.3 

9.1  ±2.6 

2.1 

Trichodontidae 

Pacific  sandfish 

0 

1.7 

2.1 

unidentifed  bothid/ 

unidentified  flatfish 

38.5  ±15.9 

20.2  ±10.3 

14.8  ±2.5 

pleui-onectid 

Zoarcidae 

unidentified  zoarcid 

0 

0 

1.4 

Salmonidae 

coho  salmon 

unknown 

5.8  ±3.6 

0 

0 

juvenile 

4.8 

3.3  ±2.3 

0.7 

adult 

0 

2.4 

6.2  ±6.2 

steelhead/rainbow  trout 

unknown 

0 

2.7  ±1.4 

0.7 

juvenile 

0 

0 

0.9 

adult 

0 

0 

0.9 

chinook  salmon 

unknown 

7.6  ±3.5 

0 

0.8  ±0.1 

juvenile 

4.0  ±1.1 

3.4 

3.6  ±3.0 

adult 

4.8 

0 

0 

unidentified  salmonid(s) 

unknown 

4.3  ±0.6 

2.4 

0.8  ±0.1 

juvenile 

4.8 

0 

7.7 

114 


Fishery  Bulletin  102(1) 


species  (two  with  coho  and  chinook  salmon  and  one  with 
coho  salmon  and  steelhead  trout,  Table  2).  No  cutthroat 
trout  were  identified  with  conventional  or  molecular 
genetic  techniques. 

Using  otoliths  and  other  diagnostic  skeletal  struc- 
tures, we  enumerated  at  least  54  individual  salmonids 
in  39  scats  (Table  2).  All  individuals  identified  as  adults 
I  n  =5 )  were  coho  salmon,  except  one  chinook  salmon  from 
spring  1997.  Individual  juveniles  identified  as  steelhead 
trout  (n=l),  coho  salmon  (re=7),  chinook  salmon  («=12), 
or  unidentified  salmonids  (/2=2)  were  present  during 
all  periods.  Because  of  the  difficulty  of  determining 
age  from  size-variable  structures  such  as  gillrakers 
and  teeth,  most  individuals  («=27)  were  designated  as 
"unknown  age." 


Discussion 

Investigating  diet  is  essential  to  assessing  the  role  of 
harbor  seals  in  marine  and  freshwater  ecosystems  in 
order  to  quantify  their  interactions  with  fisheries  and 
determine  their  impact  on  the  recovery  of  endangered 
species.  All  methods  used  to  investigate  diet  of  seals 
and  other  pinnipeds  have  some  limitations  (Murie 
and  Lavigne,  1985,  1986;  Harvey,  1989).  With  scats,  it  is 
assumed  that  the  relative  frequency  of  prey  identified 
from  undigested  remains  reflects  the  frequency  of  prey 
eaten  (Tollit  et  al.,  1997).  However,  several  investigators 
have  determined  that  this  assumption  may  be  seriously 
biased  in  several  ways  (Hawes,  1983;  da  Silva  and  Neilson, 
1985;  Jobling,  1987;  Dellinger  and  Trillmich,  1988;  Harvey, 
1989;  Pierce  and  Boyle,  1991;  Cottrell  et  al.,  1996;  Tollit 
et  al.,  1997;  Bowen,  2000;  Orr  and  Harvey,  2001).  No  diet 
study  can  estimate  detrimental  or  lethal  impacts  to  prey 
resulting  from  harassment  by  pinnipeds.  In  addition,  once 
a  prey  is  captured,  a  seal  might  consume  only  the  soft 
tissue  (especially  of  larger  prey),  which  would  not  leave 
identifiable  evidence  in  scats.  Additionally,  because  skel- 
etal remains  from  different  prey  species  pass  through  the 
alimentary  canal  and  erode  at  different  rates  they  may  not 
reflect  the  true  number  or  proportions  of  prey  consumed 
(Hawes,  1983;  Harvey,  1989;  Pierce  and  Boyle,  1991; 
Cottrell  et  al.,  1996;  Tollit  et  al.,  1997).  Therefore,  preda- 
tion  estimates  determined  from  scat  samples  should  be 
regarded  as  a  measure  of  minimum  impact.  Although  there 
are  complications  inherent  in  the  use  of  scats  to  describe 
the  diet  of  seals,  scat  analysis  remains  useful  because 
many  scats  can  be  collected  quickly,  with  minimum  effort 
and  without  harm  to  the  animals  (Harvey,  1989). 

Scats 

Recently,  skeletal  remains  other  than  otoliths  and  beaks 
have  begun  to  be  used  to  identify  and  enumerate  prey  of 
pinnipeds  (e.g.  Olesiuk  et  al.,  1990;  Cottrell  et  al.,  1996; 
Riemer  and  Brown,  1997;  Browne  et  al.,  2002).  There  are 
constraints,  however,  for  using  all  skeletal  elements  to 
identify  prey  species,  including  the  need  for  a  reference  col- 
lect ion  and  the  extensive  training  of  personnel  to  identify 


Fall  I9M7 


Q 


nverine-estuanne 


marine  or  mixed 


Scat  categorization 


Figure  2 

Mean  percentage  plus  standard  deviation  (SD)  of  scats  that 
were  classified  as "riverine-estuarine" (i.e.  samples  composed  of 
prey  taxa  that  are  exclusively  or  potentially  (e.g.  anadromous 
species,  osmerids)  found  in  rivers  or  estuaries),  "marine"  (i.e. 
samples  composed  exclusively  of  prey  that  inhabit  marine 
waters  l,  and  "marine  or  mixed"  (i.e.  samples  composed  of  prey 
taxa  exclusively  found  in  marine  waters  or  those  that  might 
inhabit  marine  waters  at  some  stage  in  their  life). 


digested  prey  structures  (Cottrell  et  al.,  1996).  Moreover, 
there  is  usually  a  bias  in  the  recovery  and  recognition  of 
prey  structures  from  different  taxa  (Cottrell  et  al.,  1996; 
Laake  et  al.,  2002).  This  bias  may  be  a  significant  problem 
in  estimating  relative  abundance  of  prey  or  biomass  con- 
sumption by  harbor  seals  and  is  the  reason  these  indices 
were  not  considered  in  this  study. 

Despite  these  complications,  the  use  of  all  available 
structures  increased  our  estimates  of  prey  diversity,  MNI, 
and  %  FO  for  most  prey  taxa.  Examination  of  all  diagnostic 
structures  also  allowed  us  to  consider  a  greater  sample  size 
because  93%  of  scats  with  identifiable  remains  contained 
bones,  whereas  only  53%  of  scats  contained  otoliths.  Spe- 
cies not  represented  by  otoliths,  such  as  salmonids  (during 
1998)  and  cartilaginous  fishes,  were  detected  because  all 
structures  were  used.  In  addition,  the  MNI  of  important 
prey  such  as  Pacific  hake.  Pacific  herring,  and  Pacific  sar- 
dine would  have  been  greatly  underestimated  had  otoliths 
been  used  exclusively  because  the  MNI  derived  by  using 
all  structures  was  at  least  threefold  greater.  Although  there 
are  complexities  associated  with  estimating  MNI  from  all 
structures,  this  method  avoids  the  use  of  numerical  correc- 
tion factors  determined  from  recovery  rates  of  otoliths  fed 
to  captive  seals  during  laboratory  experiments  (Browne 
et  al.,  2002).  Results  from  captive  experiments  are  highly 
variable  between  repeated  trials  for  the  same  individual 
and  among  different  individuals  (Harvey,  1989;  Bowen  et 
al.,  2000;  Orr  and  Harvey,  2001 1, 

Foraging  habits 

Harbor  seals  in  the  lower  Umpqua  River  consumed  prey 
from  over  35  taxa;  however,  only  a  few  prey  taxa  were 
dominant  in  their  diet,  as  reflected  by  %FO.  Overall,  the 
five  most  abundant  families  of  prey  were  Clupeidae,  Cot- 


Orr  et  al.:  Foraging  habits  of  Phoca  vitulina  nchardsi  in  the  Umpqua  River,  Oregon 


115 


tidae,  Embiotocidae,  Gadidae,  and  Pleuronectidae.  These 
are  similar  to  those  reported  in  other  studies  of  harbor 
seal  diet  in  Oregon  (Riemer  and  Brown,  1997;  Browne  et 
al.,  2002;  Riemer  et  al.3-4). 

It  was  evident  by  the  presence  of  prey  like  Pacific  hake. 
Pacific  sardine,  hagfish,  and  various  flatfishes  that  seals 
fed  offshore  in  pelagic  and  demersal  areas.  Harbor  seals 
also  consumed  prey  (e.g.  Pacific  staghorn  sculpin)  com- 
monly found  inshore  or  in  estuarine  waters.  The  NMFS 
recommendations  to  remove  pinnipeds  from  systems  where 
endangered  prey  also  occur,  rely  on  the  assumption  that 
pinnipeds  are  primarily  feeding  (on  ESA-listed  species) 
in  that  system.  Our  study  indicated  that  this  was  not  the 
case.  Although  the  seals  at  the  Umpqua  hauled  out  several 
kilometers  up  river,  they  foraged  primarily  at  sea. 

Because  of  the  life  histories  of  many  of  the  prey  taxa,  our 
foraging  habitat  categories  must  be  considered  estimations 
of  where  the  prey  might  have  been  consumed.  For  example, 
we  estimated  that  24%  of  scats  contained  prey  attributable 
to  the  riverine-estuarine  environment.  However,  this  may 
actually  be  an  overestimation  because  some  of  these  spe- 
cies potentially  inhabit  the  marine  environment  at  some 
time  in  their  life  and  may  have  been  consumed  there.  Ad- 
ditionally, scats  categorized  as  marine  or  mixed  may  reflect 
that  the  seal  fed  solely  in  the  marine  environment  (because 
all  the  taxa  can  potentially  be  found  in  marine  waters)  or 
fed  at  sea  and  within  the  river.  Nevertheless,  these  catego- 
ries are  useful  for  a  broad  apportioning  of  foraging  habitat. 
Even  though  we  were  able  to  determine  that  approximately 
76%  of  the  scats  contained  marine  and  potentially  marine 
prey  taxa,  we  were  unable  to  assess  whether  this  reflected 
a  seal  population  with  homogeneous  or  heterogeneous  for- 
aging patterns.  In  other  words,  because  the  scats  could  not 
be  attributed  to  a  particular  individual,  we  had  no  way  of 
discerning:  1)  whether  the  entire  seal  population  foraged 
roughly  three-fourths  of  the  time  at  sea  and  one-fourth  of 
the  time  in  the  river,  or  2)  whether  76%  of  the  seals  fed  at 
sea  whereas  24%  foraged  closer  to  shore  and  in  the  river. 
This  distinction  may  be  important  if  only  a  subgroup  of 
seals  is  feeding  in  the  river  and  preying  on  fish  that  are 
seasonally  abundant  in  the  estuary,  such  as  salmonids. 
Studies  that  incorporate  radio-  or  satellite-telemetry  or 
genetic  identification  of  individual  prey  items  in  scats  may 
reveal  these  distinctions  in  the  future. 

Because  the  seals  haul  out  almost  5  km  upriver  and 
have  been  observed  as  far  as  32  km  upriver,  it  is  clear  that 


3  Riemer,  S.  D.,  R.  F.  Brown,  and  M.  I.  Dhruv.  1999.  Monitoring 
pinniped  predation  on  salmonids  in  the  Alsea  and  Rogue  River 
estuaries:  fall.  1997.  //;  Pinniped  predation  on  salmonids:  pre- 
liminary reports  on  field  investigations  in  Washington,  Oregon, 
and  California,  p.  104-152.  Compiled  by  National  Marine 
Fisheries  Service,  Northwest  Region.  [Available  from  ODFW, 
7118  NE  Vandenberg  Avenue,  Corvallis,  OR  97330.] 

4  Riemer,  S.  D.,  R.  F.  Brown,  and  M.  I.  Dhruv.  1999.  Monitoring 
pinniped  predation  on  salmonids  in  the  Alsea  and  Rogue  River 
estuaries:  fall,  1998.  In  Pinniped  predation  on  salmonids:  pre- 
liminary reports  on  field  investigations  in  Washington,  Oregon, 
and  California,  p.  153-188.  Compiled  by  National  Marine 
Fisheries  Service,  Northwest  Region.  [Available  from  ODFW. 
7118  NE  Vandenberg  Avenue,  Corvallis,  OR  97330.] 


seals  use  the  river  environment.  However,  the  prevalence 
of  marine  fish  remains  in  the  scat  samples  indicates  that 
the  seals  that  haul  out  at  the  Umpqua  River  do  not  feed 
exclusively  in  the  river.  The  predominance  of  marine  prey 
may  reflect  a  foraging  strategy  in  which  the  effort  required 
to  find  marine  sources  of  food  is  offset  by  the  energy  gained 
by  exploiting  large  aggregations  of  marine  schooling  fish 
(e.g.  Pacific  hake  and  Pacific  sardine).  In  this  scenario, 
the  seals  in  the  Umpqua  estuarine-riverine  system  may 
depend  on  marine  resources  while  taking  advantage  of 
protected  estuarine  waters  that  provide  a  sheltered  place 
to  rest  and  occasionally  feed. 

Salmonids 

We  used  two  methods  to  estimate  the  number  of  salmonids 
eaten  by  harbor  seals:  prey  remains  and  genetic  analyses 
of  scat  samples.  Analysis  of  skeletal  remains  was  of  lim- 
ited value  because  the  majority  of  salmonid  structures 
recovered  from  scat  samples  were  bones,  which  could  be 
identified  only  to  family.  This  study  represents  a  novel 
application  of  genetic  techniques  to  identify  salmonid  spe- 
cies from  bones  found  in  scats.  These  techniques  allowed  us 
to  determine  species  for  a  majority  of  the  salmonid  samples 
that  would  have  otherwise  remained  unidentified  because 
they  did  not  contain  otoliths. 

Salmonid  bones  or  otoliths  were  found  in  6%  of  the  har- 
bor seal  scats  collected  during  our  study — a  finding  that  is 
comparable  to  the  5%  found  by  Laake  et  al.  (2002)  at  the 
Columbia  River.  However,  it  is  about  one-half  of  what  was 
found  by  Riemer  and  Brown  ( 13% ;  1997 1  at  selected  sites 
in  Oregon.  Brown  et  al.  (1995)  found  salmonids  in  12%  of 
gastrointestinal  tracts  of  harbors  seals  taken  incidentally 
by  commercial  salmon  gillnet  fishing  operations,  and  Roffe 
and  Mate  (1984)  observed  that  salmonids  made  up  30%  of 
the  prey  for  harbor  seals  surface  feeding  in  the  Rogue  Riv- 
er. Regardless  of  sampling  method,  in  these  studies,  most  of 
the  salmonids  could  be  identified  only  to  family  because  few 
otoliths  were  recovered  and  genetic  techniques  to  identify 
bones  to  species  had  not  yet  been  developed. 

Salmonids  are  present  in  the  Umpqua  River  year-round 
although  species  and  age  composition  change  throughout 
the  year.  In  this  study,  most  salmonid  prey  of  known 
age  were  juveniles;  however,  we  could  determine  age  of 
only  one-half  of  the  individuals.  Juveniles  are  found  in  the 
Umpqua  River  system  year-round  and  may  be  easier  for 
seals  to  catch  than  adults.  Alternatively,  perhaps  seals  did 
not  consume  many  adult  skeletal  elements  because  adult 
salmonids  are  large  fish,  which  may  be  ripped  apart  rather 
than  swallowed  whole. 

Our  sampling  seasons  encompassed  at  least  some  por- 
tion of  the  migrations  of  all  salmonids,  all  of  which  (except 
cutthroat  trout )  were  prey  of  harbor  seals.  The  fact  that 
portions  of  all  migrations  were  included  in  the  sampling 
design  was  noteworthy  because  there  were  a  large  num- 
ber of  seals  in  the  river  throughout  the  year  and  yet  we 
found  no  evidence  through  genetic  or  otolith  identification 
that  seals  consumed  cutthroat  trout  in  the  Umpqua  River. 
The  genetic  identification  tools  developed  and  applied  in 
our  collaboration  with  CBMGL  were  useful  in  discerning 


116 


Fishery  Bulletin  102(1) 


scarce  from  abundant  salmonids.  These  techniques  may 
be  useful  in  identifying  other  pinniped  prey  that  lack  spe- 
cies-specific structures  and  would  allow  managers  to  better 
assess  the  impact  of  pinniped  predation  on  threatened  or 
endangered  species. 


Acknowledgments 

This  study  was  proposed  and  initiated  in  collaboration 
with  Joe  Scordino.  Scat  collection  and  harbor  seal  counts 
were  conducted  by  Lawrence  Lehman,  Kirt  Hughes,  Mer- 
rill Gosho,  Sharon  Melin,  and  Robert  DeLong.  The  U.S. 
Coast  Guard  Umpqua  River  Station  provided  boat  storage 
and  a  location  for  keeping  a  chest  freezer  during  the  1997 
field  season.  We  would  like  to  thank  the  Oregon  Institute 
of  Marine  Biology,  Charleston,  OR,  where  the  samples  col- 
lected during  1997  were  processed.  We  greatly  appreciate 
the  collaboration  with  Conservation  Biology  Molecular 
Genetics  Laboratory,  which  resulted  in  the  identification 
of  our  salmon  remains  based  on  genetic  methods.  We  would 
also  like  to  thank  Susan  Reimer  who  kindly  helped  us  with 
difficult  identifications,  as  well  as  Lawrence  Lehman  and 
Jason  Griffith  for  their  verification  of  bone  and  otolith 
identifications.  We  thank  Patience  Browne,  Patrick  Gearin, 
John  Jansen,  Mark  Dhruv,  and  three  anonymous  review- 
ers for  providing  helpful  comments  on  earlier  drafts  of  this 
manuscript. 


Literature  cited 

Bigg,  M.  A..  G.  Ellis,  P.  Cottrell,  and  L.  Milette. 

1990.     Predation  by  harbour  seals  and  sea  lions  on  adult 
salmon  in  Comox  Harbour  and  Cowichan  Bay,  British 
Columbia.     Can.  Tech.  Rep.  Fish.  Aquat.  Sci.  1769,  31  p. 
Bowen,  W.  D. 

2000.     Reconstruction  of  pinnipeds  diets:  accounting  for 
complete  digestion  of  otoliths  and  cephalopod  beaks.     Can. 
J.  Fish.  Aquat.  Sci.  57:898-905. 
Brown,  R.  F. 

1980.     Abundance,  movements  and  feeding  habits  of  the 
harbor  seal,  Phoea  vitulina,  at  Netarts  Bay,  Oregon.     M.S. 
thesis,  69  p.     Oregon  State  Univ.,  Corvallis,  OR. 
Brown,  R.  F.,  S.  D.  Riemer,  and  S.  Jefferies. 

1995.     Food  of  pinnipeds  collected  during  the  Columbia  River 
Area  Commercial  Salmon  Gillnet  Observation  Program, 
1991-1994.    ODFW  (Oregon  Dep.  Fish  Wildlife),  Wildlife 
Diverstiy  Program  Tech.  Rep.  95-6-01,  16  p. 
Brown,  R.  F,  and  S.  Kohlmann. 

1998.     Trends  in  abundance  and  current  status  of  the 
Pacific  harbor  seal  {Phoea  vitulina  richardsi)  in  Oregon: 
1977-1998.     ODFW,  Wildlife  Diverstiy  Program  Tech. 
Rep.  98-6-01,  16  p. 
Browne,  P.,  J.  L.  Laake,  and  R.  L.  DeLong. 

2002.     Improving  pinniped  diet  analyses  through  identifica- 
tion of  multiple  skeletal  structures  in  fecal  samples.     Fish. 
Bull.  100:423-433. 
Butler.  U.  L. 

1990.  Distinguishing  natural  from  cultural  salmonid  dep- 
osits in  Pacific  Northwest  North  America.  Ph.D.  diss., 
218  p.  Univ.  of  Washington,  Seattle,  WA. 


Cottrell,  P.  E.,  A.  W.  Trites,  and  E.  H.  Miller. 

1996.  Assessing  the  use  of  hard  parts  in  faeces  to  identify 
harbour  seal  prey:  results  of  captive-feeding  trials.  Can. 
J.  Zool.  74:875-880. 

da  Silva,  J.,  and  J.  Neilson. 

1985.     Limitations  of  using  otoliths  recovered  in  scats  to 

estimate  prey  consumption  in  seals.     Can.  J.  Fish.  Aquat. 

Sci.  42:1439-1442. 
Dellinger,  T,  and  F.  Trillmich. 

1988.  Estimating  diet  composition  from  scat  analysis  in 
otariid  seals  (Otariidae):  is  it  reliable?  Can.  J.  Zool.  66: 
1865-1870. 

Eschmeyer.  W.  N,  E.  S.  Herald,  and  H.  Hammann. 

1983.    A  field  guide  to  Pacific  Coast  fishes  of  North  America, 
336  p.     Houghton  Mifflin  Co.,  Boston,  MA. 
Harvey,  J.  T. 

1987.  Population  dynamics,  annual  food  consumption,  move- 
ments and  dive  behaviors  of  harbor  seals,  Phoca  vitulina 
richardsi,  in  Oregon.  Ph.D.  diss.,  177  p.  Oregon  State 
Univ.,  Corvallis,  OR. 

1989.  Assessment  of  errors  associated  with  harbor  seal 
(Phoca  vitulina)  faecal  sampling.  J.  Zool.,  Lond.  219: 
101-111. 

Hawes,  S. 

1983.     An  evaluation  of  California  sea  lion  scat  samples  as 
indicators  of  prey  importance.     M.S.  thesis,  50  p.     San 
Francisco  State  Univ.  San  Francisco,  CA. 
Jobling,  M. 

1987.     Marine  mammal  faeces  samples  as  indicators  of  prey 
importance — a  source  of  error  in  bioenergetics  studies. 
Sarsia  72:255-260. 
Johnson,  O,  M.  Ruckelshaus,  W  Grant.  F.  Waknitz,  A.  Garrett. 
G.  Bryant,  K.  Neely,  and  J.  Hard. 

1999.     Status  review  of  coastal  cutthroat  trout  from  Washing- 
ton, Oregon,  and  California.     NOAA  Tech.  Memo.  NMFS- 
NWFSC-37,  292  p. 
King,  J. 

1983.     Seals  of  the  world,  240  p.     Comstock  Publishing 
Assoc,  Cornell  Univ.  Press,  New  York,  NY. 
Laake,  J.  L.,  P.  Browne.  R.  L.  DeLong,  and  H.  R.  Huber. 

2002.     Pinniped  diet  composition:  a  comparison  of  estimation 
models.     Fish.  Bull.  100:434-447. 
Murie,  D.,  and  D.  Lavigne. 

1985.  A  technique  for  the  recovery  of  otoliths  from  stomach 
contents  of  piscivorous  pinnipeds.  J.  Wildl.  Manag.  49: 
910-912. 

1986.  Interpretation  of  otoliths  in  stomach  content  analy- 
ses of  phocid  seals:  Quantifying  fish  consumption.  Can.  J. 
Zool.  64:1152-1157. 

NMFS  (National  Marine  Fisheries  Service). 

1997.  Investigation  of  scientific  information  on  the  impacts 
of  California  sea  lions  and  Pacific  harbor  seals  on  salmo- 
nids and  on  the  coastal  ecosystem  of  Washington,  Oregon, 
and  California.  NOAA  Tech.  Memo.  NMFS-NWFSC-28, 
172  p. 

Olesiuk,  P.  E,  M.  A.  Bigg,  G.  M.  Ellis.  S.  J.  Crockford,  and 
R.  J.  Wigen. 

1990.  An  assessment  of  the  feeding  habits  of  harbour  seals 
(Phoca  vitulina  i  in  the  Strait  of  Georgia.  British  Columbia, 
based  on  scat  analysis.  Can.  Tech.  Rep.  Fish.  Aquat.  Sci. 
1730.  135  p. 

Orr,  A.  J.,  and  J.  T.  Harvey. 

2001.  Quantifying  errors  associated  with  using  fecal  sam- 
ples to  determine  the  diet  of  the  California  sea  lion  {Zalo- 
phu*  califbrnianus).    Can.  J.  Zool.  79:1080-1087. 


Orr  et  al.:  Foraging  habits  of  Phoca  vitultna  richardsi  in  the  Umpqua  River,  Oregon 


117 


Pearson,  J.,  and  B.  Verts. 

1970.     Abundance  and  distribution  of  harbor  seals  and 
northern  sea  lions  in  Oregon.     Murrelet  51:1-5. 
Pierce,  G.  J.,  and  P.  R.  Boyle. 

1991.  A  review  of  methods  for  diet  analysis  in  piscivorous 
marine  mammals.     Ocean.  Mar.  Biol.  Ann.  Rev.  29:409-486. 

Purcell,  M,  G.  Mackey,  E.  LaHood,  H.  Huber,  and  L.  Park. 

2004.     Molecular  methods  for  the  genetic  identification  of 

salmonid  prey  from  Pacific  harbor  seal  ( Phoca  vitulina 

richardsi)  scat.     Fish.  Bull.  102:213-220. 
Reeves,  R.,  B.  Stewart,  and  S.  Leatherwood. 

1992.  The  Sierra  Club  handbook  of  seals  and  sirenians, 
359  p.     Sierra  Club  Books,  San  Francisco.  CA. 

Riemer,  S.  D.,  and  R.  F.  Brown. 

1997.     Prey  of  pinnipeds  at  selected  sites  in  Oregon  identi- 


fied by  scat  (fecal)  analysis,  1983-1996.     ODFW  Wildlife 
Diversity  Program  Tech.  Rep.  97-6-02.  34  p. 
Roffe,  T,  and  B.  Mate. 

1984.    Abundances  and  feeding  habits  of  pinnipeds  in  the 
Rogue  River,  Oregon.     J.  Wildl.  Manag.  48:1262-1274. 
Tollit,  D.  J.,  M.J.  Steward,  P.  M.  Thompson,  G  J.  Pierce, 
M.  B.  Santos,  and  S.  Hughes. 

1997.     Species  and  size  differences  in  the  digestion  of  oto- 
liths and  beaks:  implications  for  estimates  of  pinniped  diet 
composition.     Can.  J.  Fish.  Aquat.  Sci.  54:105-119. 
U.  S.  Fish  and  Wildlife  Service. 

2000.  Endangered  and  threatened  wildlife  and  plants:  final 
rule  to  remove  the  Umpqua  River  cutthroat  trout  from  the 
list  of  endangered  wildlife.  Federal  Register:  26  April 
2000,  65181:24420-24422. 


118 


Abstract— Larval  development  of  the 
sidestriped  shrimp  ^Pandalopsis  dis- 
par) is  described  from  larvae  reared 
in  the  laboratory.  The  species  has  five 
zoeal  stages  and  one  postlarval  stage. 
Complete  larval  morphological  charac- 
teristics of  the  species  are  described  and 
compared  with  those  of  related  species 
of  the  genus.  The  number  of  setae  on 
the  margin  of  the  telson  in  the  first  and 
second  stages  is  variable:  11+12, 12+12, 
or  11+11.  Of  these,  11+12  pairs  are  most 
common.  The  present  study  confirms 
that  what  was  termed  the  fifth  stage 
in  the  original  study  done  by  Berkeley 
in  1930  was  the  sixth  stage  and  that 
the  fifth  stage  in  the  Berkeley's  study 
is  comparable  to  the  sixth  stage  that 
is  described  in  the  present  study.  The 
sixth  stage  has  a  segmented  inner  fla- 
gellum  of  the  antennule  and  fully  devel- 
oped pleopods  with  setae.  The  ability  to 
distinguish  larval  stages  of  P.  dispar 
from  larval  stages  of  other  plankton  can 
be  important  for  studies  of  the  effect  of 
climate  change  on  marine  communities 
in  the  Northeast  Pacific  and  for  marine 
resource  management  strategies. 


Larval  development  of  the  sidestriped  shrimp 
(.Pandalopsis  dispar  Rathbun) 
(Crustacea,  Decapoda,  Pandalidae) 
reared  in  the  laboratory 


Wongyu  Park 

School  of  Fisheries  and  Ocean  Sciences 
University  of  Alaska  Fairbanks 
Juneau,  Alaska,  99801-8677 
E-mail  address:  wparkig'uaf  edu 


R.  Ian  Perry 

Pacific  Biological  Station, 

Fisheries  and  Oceans 

Nanaimo,  British  Columbia,  V9R  5K6,  Canada 


Sung  Yun  Hong 

Department  of  Marine  Biology 
Pukyong  National  University 
Pusan,  608-737,  Korea 


Manuscipt  approved  for  publication 
23  June  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:118-126  (2004). 


Sixteen  species  of  the  genus  Pandalop- 
sis have  been  recognized  in  the  South- 
western Atlantic  and  North  Pacific 
Oceans  (Komai,  1994;  Jensen,  1998; 
Hanamura  et  al.,  2000).  Most  members 
of  the  genus  attain  a  large  body  size 
and  are  valuable  as  commercial  fishery 
resources  (Holthuis,  1980;  Baba  et  al., 
1986).  In  the  North  Pacific,  P.  dispar, 
P.  ampla,  P.  aleutica,  P.  longirostris,  P. 
lucidirimicola,  and  P.  spinosior  have 
been  reported.  Of  these,  Pandalopsis 
dispar  is  an  important  component  of  the 
commercial  shrimp  fisheries  along  with 
several  species  of  the  genus  Pandalus. 
Commercial  landings  of  shrimp  during 
1999  totaled  approximately  19  million 
tonstPSMFC,  1999). 

Knowledge  of  the  life  histories  of 
these  species,  including  the  duration 
and  growth  of  their  larvae,  is  important 
for  stock  assessment  and  management. 
However,  remarkably  little  is  known 
about  their  early  life  histories  because 
most  species  of  the  genus  live  at  con- 
siderable depths.  Of  the  16  Pandalopsis 
species,  the  larvae  of  only  three  species 
have  been  described  partly  or  com- 
pletely from  plankton  samples  or  from 


larvae  reared  in  the  laboratory.  The 
larvae  of  Pandalopsis  japonica  were 
described  completely  from  specimens 
reared  in  the  laboratory  by  Komai 
and  Mizushima  (1993).  Kurata  (1964) 
described  the  first  stage  of  P.  cocci nata 
from  plankton  samples  and  from  larvae 
hatched  in  the  laboratory.  Thatje  and 
Bacardit  (2000)  assumed  that  larvae  of 
P.  ampla  occurring  in  Argentine  waters 
were  similar  to  those  of  P.  dispar  and 
Pandalopsis  coccinata.  Berkeley  ( 1930) 
described  four  larval  stages  of  P.  dispar 
based  on  samples  collected  in  British 
Columbia  coastal  waters.  The  first  stage 
was  obtained  from  ovigerous  females, 
whereas  the  larvae  of  the  other  stages 
were  separated  from  plankton  samples. 
In  addition,  the  stage  described  as  the 
fifth  stage  was  not  clearly  defined. 

In  this  study,  we  describe  the  complete 
series  of  larval  stages  of  P.  dispar  using 
specimens  reared  in  the  laboratory. 


Materials  and  methods 

Ovigerous    females    of  Pandalopsis 
dispar  were  collected  on  25  March 


Park  et  al.:  Larval  development  of  Pandalopsis  dispor 


119 


1999  by  using  a  small  shrimp 
trawl  fished  at  depths  of  about 
40  m  near  Gabriola  Island  in 
the  vicinity  of  the  Pacific  Bio- 
logical Station,  Nanaimo,  Brit- 
ish Columbia  (latitude  49°13', 
longitude  123°55').  Water  tem- 
perature at  the  collection  site 
was  around  9°C,  and  salinity 
was  29.0<?c  The  females  were 
each  kept  in  a  20-L  jar  with 
seawater.  The  larvae  hatched 
on  1  April  1999.  Hundreds  of 
larvae  hatched  from  one  female. 
Of  these,  one  hundred  newly 
hatched  larvae  were  transferred 
into  individual  250-mL  jars.  To 
obtain  samples  for  drawing  and 
descriptions,  a  total  of  150  larvae 
from  the  female  were  reared  in  a 
20-L  jar.  Newly  hatched  Artemia 
nauplii  were  used  to  feed  the 
larvae.  We  used  filtered  natu- 
ral seawater  from  40  m  depth 
without  adjusting  the  water 
temperature  or  salinity.  Water 
temperature  during  the  rearing 
experiments  ranged  from  8.7°C 
to  12.2°C  (mean  10.5°C).  Salin- 
ity during  the  rearing  experi- 
ments ranged  from  26.0%c  to 
31.0%f  (mean  28.9%o).  The  water 
in  each  jar  was  changed  daily. 

All  drawings  were  made  with 
a  drawing  tube  attached  to  a  mi- 
croscope. Carapace  length  (CD 
was  measured  with  an  ocular 
micrometer  from  the  posterior 
edge  of  the  orbital  arch  to  the 
middorsal  posterior  edge  of  the 
carapace.  The  anatomical  terms 
used  in  this  paper  are  from 
Pike  and  Williamson  ( 1969)  and 
Haynes  (1985). 


Measurement  bars  represent  1  mm. 
Figure  1 


Results 

In  the  complete  larval  development  of  Pandalopsis  dispar 
there  are  five  zoeal  stages.  In  addition,  there  is  one  post- 
larval  stage.  The  duration  of  each  larval  stage  and  the 
survival  rate  of  P.  dispar  are  shown  in  Table  1. 

Larval  description 
First  stage 

Carapace  (Fig.  1A)  Carapace  length  (CL),  1.6  mm  (SD: 
0.06  mm,  n=89);  with  concave  lateral  margin;  rostrum  long. 
well  developed  and  directed  forward  and  upward;  weak 


Table  1 

Duration 

of  each  larval  stage  of  Pandalopsis  dispar  at 

8.7-12.2°C  (mean  10.5°C)  and  26.0-31.0% 

(mean28.9%c). 

Stage  6  is  a  postlarval 

stage. 

Mean  duration 

Range 

Number  of 

Stage 

(day) 

(day) 

] 

arvae  observed 

1 

10.7 

9-15 

89 

2 

8.9 

8-11 

81 

3 

9.5 

8-14 

67 

4 

10 

9-12 

59 

5 

10.8 

10-13 

51 

6 

10.5 

9-12 

48 

120 


Fishery  Bulletin  102(1) 


dorsal  denticles  and  bare  ventral  tubercles  on  rostrum; 
rostrum  about  0.7  times  as  long  as  carapace. 
Eyes  (Fig.  1  A)     Sessile. 
Abdomen  (Fig.  1A)     5  somites  plus  telson. 
Antennule  (Fig.  1,  A  and  B)     Peduncle  unsegmented  with 
a  strong  seta  at  distromesial  margin;  outer  flagellum  with 
2+2  short  aesthetascs. 

Antenna  (Fig.  1,  A  and  C)  Longer  than  the  whole  body 
length;  flagellum  segmented  throughout  its  length;  outer- 
distal  corner  terminated  with  an  acute  spine  and  a  minute 
seta;  inner-distal  margin  5-segmented  with  1,  1,  1,  1,  2 
setae;  inner  margin  fringed  with  28-29  setae. 


Measurement  bars  represent  1  mm. 
Figure  2 


Mandible  (Fig.  1  D)  Asymmetrical;  without  the  same 
arrangement  of  denticles,  however,  almost  the  same  size; 
incisor  process  not  separated  from  molar  process;  armed 
with  several  teeth  on  molar  part. 

Maxillule  (Fig.  1  E)  Coxal  and  basal  endite  with  serially 
developed  strong  spines  and  multiple  setae;  endopod  with 
2+3  terminal  setae 

Maxilla  (Fig.  1  F)  Palp  with  2, 1, 1,  2  setae;  coxal  endite  with 
6  distal  setae;  basal  endite  with  8  distal  setae;  broad  scaphog- 
nathite  with  narrow  posterior  lobes  having  long  naked  setae. 
First  maxilliped  (Fig.  1G)  Endites  separated  by  shallow 
notch  and  with  multiple  setae;  bilobed  epipod;  endopod  4-seg- 
mented  with  6,  3,  3,  4  setae;  exopod 
with  14  plumose  natatory  setae; 
terminal  segment  with  3  terminal 
spines  and  1  subterminal  spine. 
Second  maxilliped  (Fig.  1  H)  Coxa 
with  7  setae;  basis  with  3  setae;  no 
epipod;  endopod  5  segmented  with 
5,  5.  2.  4,  4  setae;  exopod  with  27 
plumose  natatory  setae. 
Third  maxilliped  (Fig.  II)  Coxa 
with  1  seta;  basis  with  2  setae;  endo- 
pod 5-segmented  with  various 
number  of  setae;  exopod  with  34-36 
plumose  natatory  setae. 
Pereiopods  (Fig.  1,  J— N)  1st  pereio- 
pod  (Fig.  1J)  not  chelate;  3  long  ter- 
minal spines  on  dactylus;  dactylus 
short;  propodus  longer  than  carpus; 
exopod  without  natatory  setae; 
endopod  of  2nd  pereiopod  chelated 
(Fig.  IK);  chela  with  numerous 
small  spines;  ischium  and  carpus  of 
pereiopods  3-5  (Fig.  1,  L-N)  longer 
than  1st  and  2nd  ones;  propodus 
armed  with  several  minute  spines. 
Pleopods  (Fig.  lO)  Bilobed  buds, 
not  functional. 

Telson  (Fig.  1  P)  Triangular  form; 
broadened  at  the  end,  posterolat- 
eral margin  with  11  +  12  (12+12. 
11+11)  marginal  spines;  each  spine 
with  fine  hairs. 

Second  stage 


Carapace  (Fig.  2A)     CL,2.2mm(SD: 

0.11,  n=81>:  rostrum  not  strongly 

curved  upwards;  5-6  prominent 

dorsal  denticles  and  3-4  weak 

ventral  spines;  rostrum  shorter 

than  carapace;  supraorbital  spine 

present. 

Eyes  (Fig.  2A)     Stalked;  separated 

from  carapace. 

Antenna  (Fig.  2A)     General  shape 

unchanged;  longer  than  that  of  1st 

Antennule  (Fig.  2B)     Peduncle  3- 


Park  et  al.:  Larval  development  of  Pandalopsis  dispar 


121 


segmented;  inner  flagellum  with 
2  distal  setae;  outer  flagellum 
with  2,  3,  4,  2  aesthetascs  on 
inner  margin. 

Mandible  (Fig.  2C)  General 
shape  unchanged;  bigger  than 
that  of  1st  stage. 
Maxillule  (Fig.  2D)  Coxal 
and  basal  endite  with  serially 
developed  strong  spines  and 
multiple  setae;  endopodite  with 
2+3  spines;  a  strong  subtermi- 
nal  seta. 

Maxilla  (Fig.  2E)  Palp  with  2, 
2,  2,  3  setae;  broad  scaphogna- 
thite  with  narrow  posterior  lobe 
having  a  long  naked  seta;  coxal 
endite  with  6  distal  setae;  basal 
endite  with  7  distal  setae. 
First  maxilliped  (Fig.  2F)  Epi- 
pod  bilobed;  endopod  with 
3+1,  2+1,  2+1,  3  setae;  exopod 
unsegmented  with  14  plumose 
natatory  setae. 

Second  maxilliped  (Fig.  2G) 
One  long  and  several  intermedi- 
ate sized  spines  in  basal  endite; 
endopod  5-segmented;  exopod 
with  24  plumose  natatory  setae. 
Third  maxilliped  (Fig.  2H)  En- 
dopod 5-segmented,  armed  with 
many  spines;  exopod  of  36  plu- 
mose natatory  setae. 
Pereiopods  (Fig.  2,  I— M)  Not 
chelate;  1st  pereiopod  of  4 
spines  in  basal  endite;  3  strong 
and  two  weak  spines  in  dactylus 
of  1st  pereiopod;  general  shape 
unchanged  from  2nd  pereiopod 
through  5th  pereiopod. 
Pleopods  (Fig.  2N)  Bilobed 
buds,  not  functional;  no  seta  and 
hair  on  buds;  no  further  devel- 
opment from  the  1st  stage. 
Telson  (Fig.  20)     Unchanged. 

Third  stage 


Measurement  bars  represent  1  mm. 


Figure  3 


Carapace  (Fig.  3A)     CL,  2.7  mm  (SD:  0.12,  rc=67);  longer 

rostrum  than  that  of  2nd  stage;  almost  0.9  times  as  long  as 

carapace;  rostrum  with  5-6  dorsal  spines  and  1-2  ventral 

spines. 

Antennule  (Fig.  3B)     Inner  flagellum  2-segmented  with 

0,  2  setae;  outer  flagellum  2-segmented  with  3+3+3,  3+3 

aesthetascs. 

Antenna  (Fig.  3A)     General  shape  unchanged;  longer  than 

2nd  stage. 

Mandible  (Fig.  3C)     Molar  and  incisor  processes  present; 

incisor  process  with  6-9  teeth;  molar  process  with  heavy 

teeth  on  biting  edge. 


Maxillule  (Fig.  3D)     Palp  with  2+3  setae;  a  small  subtermi- 

nal  spine;  basal  and  coxal  endite  with  numerous  spines. 

Maxilla  (Fig.  3E)     Protopodite  unsegmented;  palp  with  1,  2, 

2,  1+2  setae  and  with  4  lobes;  broad  scaphognathite  with 

narrow  posterior  lobe  bearing  numerous  setae. 

First  maxilliped  (Fig.  3F)     Epipod  bilobed;  endopod  4-seg- 

mented  with  4,  2,  2,  2  setae;  exopod  with  15-16  plumose 

natatory  setae; 

Second  maxilliped  (Fig.  3G)     Coxal  endite  with  an  epipod 

and  a  strong  spine;  endopod  5-segmented  with  3,  2,  2,  4,  6 

setae  exopod  with  setae. 

Third  maxilliped  (Fig.  3H)     Coxal  endite  with  one  long  and 


122 


Fishery  Bulletin  102(1) 


one  short  spine;  basal  endite  with  2  long  and  2  interme- 
diate sized  spines;  endopod  5-segmented  with  numerous 
setae;  exopod  with  25-26  plumose  natatory  setae. 
Pereiopods  (Fig.  3, 1— M)     General  shape  unchanged  except 
addition  of  setae. 

Pleopods  (Fig.  3N)     Buds  biramous;  much  longer  than  that 
of  2nd  stage. 

Uropods  (Fig.  30)  Biramous;  endopod  with  a  fused  spine 
at  distal  quarter  of  outer  margin  and  numerous  setae  on 
inner  distal  margin;  exopod  with  4  spines  on  outer  margin 
and  numerous  setae  on  inner  distal  margin. 
Telson  (Fig.  30).  With  12  pairs  of  posterolateral  spines 
plus  a  median  spine. 


,F,I-M,Q, 
,B,G,H,N 


Measurement  bars  represent  1  mm 


Figure  4 


Fourth  stage 

Carapace  (Fig.  4A)     CL.  3.1  mm  (SD:  0.13.  re=59);  Rostrum 
slightly  longer  than  carapace  and  directed  forward;  ros- 
trum with  15  dorsal  spines  and  6  ventral  spines. 
Antenna  (Fig.  4A)     General  shape  unchanged;  longer  than 
that  of  3rd  stage. 

Antennule  (Fig.  4B)  Much  longer  inner  flagellum  than 
that  of  3rd  stage;  inner  flagellum  about  0.9  times  as  long 
as  outer  flagellum,  2-segmented  with  0,  2  setae;  outer  fla- 
gellum 6-segmented  with  1,  2,  2,  3,  4  aesthetascs. 
Mandible  (Fig.  4C)  Similar  to  third  stage. 
Maxillule  (Fig  4D)  General  shape  unchanged  except  addi- 
tion of  setae  on  endites. 

Maxilla  (Fig.  4E)  Palp  with  3,  2, 
2,  3  setae;  endites  and  scaphog- 
nathite  added  numerous  setae. 
First  maxilliped  (Fig.  4F)  Expod 
with  15  plumose  natatory  setae. 
Second  maxilliped  (Fig.  4G) 
Basal  endite  with  an  epipod  and 
a  long  spine;  exopod  with  28-29 
plumose  natatory  setae. 
Third  maxilliped  (Fig.  4H)  Ex- 
popod  with  36-37  plumose  nata- 
tory setae. 

Pereiopods  (Fig.  4,  l-M)  Num- 
ber of  spines  increased. 
Pleopods  (Fig.  4N)  Lobes  much 
longer  than  those  of  third  stage. 
Uropods  (Fig.  40)  Endopod 
and  exopod  with  numerous  setae 
on  inner  distal  margin. 
Telson  (Fig.  40)  With  12  pairs 
of  spines  on  posterolateral 
margin;  a  pair  of  lateral  spines 
at  distal  third. 


Fifth  stage 

Carapace  (Fig.  5A)  CL.  3.6 
mm  (SD:  0.15,  re=51);  Rostrum 
directed  forward  and  upward, 
slightly  longer  than  cara- 
pace; rostrum  with  17-18 
dorsal  spines  and  7-8  ventral 
spines. 

Antennule  (Fig.  5B)  Inner 
flagellum  3-segmented  and 
about  0.9  times  as  long  as  outer 
flagellum:  outer  flagellum  with 
2+2+3+3+3+2+3+5  aesthetascs 
and  distal  third  6-segmented. 
Mandible  (Fig.  5C)  More  ad- 
vanced development  than  that 
of  6th;  not  much  change  in  biting 
surface. 

Maxillule  (Fig.  5D)  General 
shape  unchanged  except  addi- 
tion of  setae  on  endites. 


Park  et  al.:  Larval  development  of  Pandalopsis  dispar 


123 


Maxilla  (Fig.  5E)  General  shape 
unchanged. 

First  maxilliped  (Fig.  5F)  Exo- 
pod  with  16  plumose  natatory 
setae. 

Second  maxilliped  (Fig.  5G) 
Exopod  with  31-33  plumose 
natatory  setae. 

Third  maxilliped  (Fig.  5H)  Exo- 
pod with  46-48  plumose  nata- 
tory setae. 

Pereiopods  (Fig.  5,  l-M)  Is- 
chium slightly  expanded  in  first 
pereipod. 

Pleopods  (Fig.  5N)  Much  more 
developed  than  pleopods  of  4th 
stage;  exopod  with  13,  1  setae; 
endopod  with  6  setae  and  ves- 
tiges of  appendix  interna. 
Uropod  (Fig.  50)  Exopod  with 
numerous  minute  spines  on 
outer  margin 

Telson  (Fig.  50)  Both  lateral 
margins  parallel;  19  termi- 
nal spines;  2  pairs  of  lateral 
spines. 

Sixth  stage 


Carapace  (Fig.  6A)     CL,  4.0  mm 
(SD:  0.21,  n=48);  adult-like. 
Antennule     (Fig.     6B)     Inner 
flagellum  as  long  as  outer  fla- 
gellum;  inner  flagellum  with 
multisegments;  outer  flagellum 
with  numerous  segments. 
Mandible  (Fig.  6C)     Incisor  part 
separated  from  molar  process 
and  extended  anteriorly. 
Maxillule  (Fig.  6D)     9  terminal 
spines  on  basal  endite. 
Maxilla  (Fig.  6E)     Palp  with  2, 2, 
2,  1+2  spines;  broad  scaphogna- 
thite  with  narrow  posterior  lobe 
bearing  3  long  setae. 
First  maxilliped  (Fig.  6F)     Exop- 
odite  with  4+2,  2,  2  3  long  and 
1  short  spines. 

Second  maxilliped  (Fig.  6G) 
spines;  vestigial  dactylus. 

Third  maxilliped  (Fig.  6H)  Propodus  armed  with  many 
spines;  dactylus  with  2  spines. 

Pereiopods  (Fig.  6,  l-M)  1st  pereipods  with  subchelated 
terminal  segment;  1st  pereiopod  with  slightly  expanded 
ischium. 

Pleopods  (Fig.  6N)  Endopod  and  exopod  with  numer- 
ous plumose  natatory  setae;  endopod  with  epipod  almost 
adult-like. 

Uropods  (Fig.  60)  Biramous;  larger  than  those  of  fifth 
stage;  adult-like. 


Measurement  bars  represent  1  mm. 


Figure  5 


Basal  endite  with  2  long 


Telson  (Fig.  60)     Telson  with  20  terminal  spines  and  4 
pairs  of  lateral  spines. 


Discussion 

The  first  stage  larva  of  Pandalopsis  dispar  described  by 
Berkeley  (1930)  is  identical  to  the  larva  described  in  the 
present  study.  However,  we  found  that  she  overlooked 
some  important  characteristics.  She  described  the  first 
stage  larva  as  having  24  setae  on  the  margin  of  the  telson. 
We  found,  however,  that  the  number  of  setae  is  variable, 
and  that  the  larvae  have  11+12,  12+12,  or  11+11  marginal 


124 


Fishery  Bulletin  102(1) 


setae.  Of  these,  11+12  pairs  are  more  common  than  the 
others. 

Berkeley  ( 1930)  described  the  fifth  stage  based  on  plank- 
ton materials.  In  the  present  study,  what  was  described  by 
Berkeley  ( 1930)  as  the  fifth  stage  larva  turned  out  to  be  the 
sixth  stage  because  the  larvae  of  this  stage  have  fully  devel- 
oped pleopods.  Although  the  larvae  of  the  fifth  stage  have 
somewhat  natatory  setose  on  their  pleopods,  they  appear 
not  to  be  completely  functional.  Compared  to  the  larvae  of 
P.  japonica,  P.  dispar  has  one  more  stage  than  that  of  P. 
japonica.  The  pleopod  development  of  P.  japonica  from  the 
fourth  stage  to  the  fifth  stage  is  very  obvious,  whereas  that 
off!  dispar  has  another  stage  and  the  changes  in  its  fea- 
tures between  the  fourth  and  sixth  stages  are  easily  seen. 


Measurement  bars  represent  1  mm. 


Figure  6 


The  major  characteristics  of  the  six  larval  stages  of  P. 
dispar  are  summarized  in  Table  2.  This  table  can  be  used 
for  the  identification  of  the  larval  stages  of  this  species. 
Komai  ( 1994)  reviewed  the  morphological  characters  of  the 
first  larval  stage  of  three  Pandalopsis  spp.:  P.  dispar.  P.  coc- 
cinata,  and  P.japonica.The  larvae  of  P.  dispar  at  this  stage 
are  quite  different  from  those  of  the  other  two  species.  The 
larvae  of  P  dispar  have  a  triangular  telson,  whereas  those 
of  P  coccinata  and  P.  japonica  have  a  semicircular  telson. 
The  adults  of  the  genus  Pandalopsis  differ  from  those  of 
other  pandalid  shrimps  by  having  a  laminated  expansion 
on  the  first  pereiopod  (Schmit,  1921;  Butler,  1980).  This 
character  is  also  present  in  larvae  of  P  coccinata  and  P. 
japonica,  whereas  it  is  not  present  in  larvae  of  P.  dispar. 

From  the  third  stage  the  is- 
chium does  indicate  expansion, 
however,  it  is  not  distinctive.  It 
is  assumed  that  in  P  dispar.  the 
expansion  should  be  distinctive 
after  the  larval  stages. 

In  P.  coccinata  and  P.  japonica 
f  )     'vi_—  the  ischium  of  the  first  pereio- 

pod has  a  laminated  expansion; 
however,  in  P.  dispar  it  has  no 
lamination.  The  structure  of  the 
ischium  of  the  first  pereiopod 
can  be  a  diagnostic  feature  of  P. 
dispar  in  addition  to  the  shape 
of  the  telson. 

Interspecific  variation  in  the 
larval  stages  of  pandalid  shrimp 
is  large,  ranging  from  three  to 
thirteen  stages  (Rothlisberg, 
1980;  Komai  and  Mizushima, 
1993).  Haynes  (1980,  1985) 
assumed  that  P.  dispar  might 
have  seven  pelagic  stages,  or  at 
least  more  than  four.  The  pres- 
ent study  has  determined  that 
P.  dispar  has  five  zoeal  stages 
prior  to  the  juvenile  stage. 

Pandalopsis  dispar  is  one  of 
the  four  principal  target  species 
of  shrimp  trawl  fisheries  in  both 
offshore  and  inshore  areas  of 
the  NE  Pacific  Ocean  (PICES, 
2001)  but  has  undergone  very 
large  fluctuations  in  abundance, 
particularly  in  Alaska  where  it 
was  reduced  to  extremely  low 
levels  during  the  late  1980s  and 
through  the  1990s.  These  fluc- 
tuations appear  to  have  been 
associated  first  with  climate 
fluctuations  (Anderson,  2000), 
and  second  with  intense  har- 
vesting (Oresanz  et  al.,  1998). 
Anderson  (2000)  has  suggested 
that  pandalid  shrimp  population 
changes  are  one  of  the  early  in- 


i       i< 


Park  et  al.:  Larval  development  of  Pandalopsis  dispor 


125 


Table  2 

Major  characters  of  Pandalopsis 

dispar  larvae. 

Characters 

Larval  stage' 

1 

2 

3 

4 

5 

6  (postlarva) 

Antenna       Inner 
flagellum 

One  strong 
spine 

One  peduncle 
with  a  few 
small  spines 

2  segments 

2  segments 

3  segments 

Multisegmented 
over  14 

Outer 

Not  segmented 

Slightly 
developed 

2  segments 

6  segments 

7  segments 

12  segments 

Telson 

12+11,  12+12, 
or  11+11 

10  pairs  of 
terminal  spines, 
2  pairs  of 
uropods 

One  spine  on 
each  midlateral 
margin 

2  spines  on  each 
lateral  margin 

4  spines  on  each 
lateral  margin 

Pleopod 
development 

Wide  as  much 
as  long 

Longer  than 
wide 

Almost 
separated  lobes 

Longer  lobes 
than  those  of 
stage  3 

Lobes  separated 
completely  with 
natatory  setae 

Adult-like,  with 
many  natatory 
setae  on  both 
lobes 

'  Eyes  of  the  first  stags 

are  sessile  on  carapace,  whereas  those  of  the  second  and  later  stages  are  stalked. 

dicators  of  shifts  in  marine  communities  in  this  region. 
Orensanz  et  al.  (1998)  have  suggested  it  is  important  to 
recognize  that  crustacean  stocks  can  have  multiscale 
spatial  structures;  species  have  possibly  both  widely  dis- 
tributed populations  (such  as  in  the  oceanic  offshore)  and 
populations  with  discrete  and  localized  distributions  (as 
may  occur  in  the  nearshore  inlets). 

The  ability  to  distinguish  the  larval  stages  of  Pandalopsis 
dispar  from  routine  plankton  samples  is  therefore  of  use 
in  studying  both  these  problems  of  population  fluctuations 
and  population  distributions.  Early  identification  of  trends 
in  strong  versus  weak  year  classes  can  provide  rapid  indica- 
tions of  possible  changes  in  large-scale  climate  conditions. 
Unambiguous  identification  of  planktonic  stages  of  P.  dis- 
par is  also  essential  for  studies  of  the  spatial  structure  of  its 
populations,  for  studies  of  transport  pathways  and  potential 
mixing  rates  among  populations,  and  ultimately  for  under- 
standing the  metapopulation  structure  of  these  populations. 
This  latter  point  is  critical  for  the  development  of  improved 
management  approaches,  which  may  include  identification 
of  reproductive  refugia  (Orensanz  et  al.,  1998). 


Acknowledgments 

We  wish  to  thank  Jim  Boutillier  and  Steve  Head  for  their 
support  with  this  study.  This  study  was  supported  by 
the  Korea  Research  Foundation  Grant  (KRF-2002-013- 
H00005). 


Literature  cited 

Anderson,  P.  J. 

2000.     Pandalid  shrimp  as  indicators  of  ecosystem  regime 
shift.     J.  Northw.  Atl.  Fish.  Sci.  27:1-10. 


Baba,  K.,  K.  Hayashi,  and  M.  Toriyama. 

1986.     Decapod  crustaceans  from  continental  shelf  and  slope 
around  Japan,  336  p.     Japan  Fisheries  Resource  Conserva- 
tion Association,  Tosho  Printing  Co.  Ltd.,  Tokyo. 
Berkeley,  A. 

1930.     The  post-embryonic  development  of  the  common  pan- 
dalids  of  British  Columbia.     Cont.  Can.  Bio.  Fish.,  New 
Series  6:79-163. 
Butler,  T  H. 

1980.     Shrimps  of  the  Pacific  coast  of  Canada.     Can.  Bull. 
Fish.  Aquat.  Sci.  202,  280  p. 
Hanamura.  Y.,  H.  Khono.  and  H.  Sakai. 

2000.    A  new  species  of  the  deepwater  pandalid  shrimp  of  the 
genus  Pandalopsis  (Crustacea:  Decapoda:  Pandalidae)  from 
the  Kuril  Islands,  North  Pacific.     Crust.  Res.  29:27-34. 
Haynes,  E. 

1980.     Larval  morphology  of  Pandalus  tridens  and  summary 
of  the  principal  morphological  characteristics  of  North 
Pacific  pandalid  shrimp  larvae.     Fish.  Bull.  77:625-640. 
Haynes,  E. 

1985.     Morphological  development,  identification,  and  biol- 
ogy of  larvae  of  Pandalidae,  Hippolytidae  and  Crangoni- 
dae  (Crustacea:  Decapoda)  of  the  northern  North  Pacific 
Ocean.     Fish.  Bull.  83:253-288. 
Holthuis,  L.  B. 

1980.     Shrimps  and  prawns  of  the  world,  an  annotated 
catalogue  of  species  of  interest  to  fisheries.     FAO  species 
Catalogue.    FAO  Fish.  Syn.  125,  FIR/S  125  Vol.1,  271  p. 
Jensen,  G.  C. 

1998.    A  new  species  of  the  genus  Pandalopsis  (Decapoda: 
Caridea:  Pandalidae)  from  the  Eastern  Pacific,  with  notes 
on  its  natural  history.     Species  Diversity  3:81-88. 
Komai.  T. 

1994.     Deep-sea  shrimps  of  the  genus  Pandalopsis  (Decap- 
oda: Caridea:  Pandalidae)  from  the  Pacific  Coast  of  eastern 
Hokkaido,  Japan  with  the  description  of  two  new  species. 
J.  Crust.  Biol.  14:538-559. 
Komai,  T,  and  T.  Mizushima. 

1993.    Advanced  larval  development  of  Pandalopsis  japonica 


126 


Fishery  Bulletin  102(1) 


Balss,  1914  (Decapoda:  Caridea:  Pandalidae)  reared  in  the 
Laboratory.     Crustaceana  64:24-39. 
Kurata,  H. 

1964.     Larvae  of  deacapod   Crustacea  of  Hokkaido.  3. 
Pandalidae.     Bull.  Hokkaido  Reg.  Fish.  Res.  Lab.  28:23- 
34. 
Orensanz,  J.  M.,  J.  Armstrong,  D.  Armstrong,  and  R.  Hilborn. 

1998.     Crustacean  resources  are  vulnerable  to  serial  deple- 
tion— the  multifaceted  decline  of  crab  and  shrimp  fisheries 
in  the  greater  Gulf  of  Alaska.     Rev.  Fish  Biol.  Fisheries  8: 
117-176. 
Pike,  R.  B„  and  D.  I.  Williamson. 

1964.     The  larvae  of  some  species  of  Pandalidae  (Decapoda). 
Crustaceana  6:265-284. 
PICES. 

2001.  Commercially  important  crabs,  shrimps  and  lobsters 
of  the  North  Pacific  Ocean.  PICES  Scientific  Report  19. 
79  p.  North  Pacific  Marine  Science  Organization,  Sidney, 
B.C.,  Canada 


PSMFC  (Pacific  States  Marine  Fisheries  Commission). 

1999.  52nd  annual  report  of  the  Pacific  States  Marine  Fish- 
eries Commission  (A.  J.  Didier.  ed.l,  37  p.  Pacific  States 
Marine  Fisheries  Commission,  Gladston,  OR. 

Rothlisberg,  P.  C. 

1980.  A  complete  larval  description  of  Pandalus  jordani 
Rathbun  (Decapoda,  Pandalidae i  and  its  relation  to  other 
members  of  the  genus  Pandalus.     Crustaceana  38:19-48. 

Schmitt,  W.  L. 

1921.  The  marine  decapod  Crustacea  of  California  with 
special  reference  to  the  decapod  Crustacea  collected  by  the 
United  States  Bureau  of  Fisheries  Steamer  "Albatross"  in 
connection  with  the  biological  survey  of  San  Francisco  bay 
during  the  years  1912-1913.  LTniv.  California,  Publ.  Zool., 
vol.  23,  470  p. 

Thatje,  S.,  and  R.  Bacardit. 

2000.  Larval  development  of  Austropandalus  grayi  (Cun- 
ningham, 1871)  (Decapoda.  Caridea.  Pandalidae)  from  the 
southwestern  Atlantic  Ocean.     Crustaceana  73:609-628. 


127 


Abstract— This  study  was  undertaken 
to  resolve  problems  in  age  determina- 
tion of  sablefish  (Anoplopoma  fimbria). 
Aging  of  this  species  has  been  ham- 
pered by  poor  agreement  (averaging 
less  than  45%)  among  age  readers  and 
by  differences  in  assigned  ages  of  as 
much  as  15  years. 

Otoliths  from  fish  that  had  been 
injected  with  oxytetracycline  (OTC) 
and  that  had  been  at  liberty  for  known 
durations  were  used  to  determine  why 
age  determinations  were  so  difficult 
and  to  help  determine  the  correct  aging 
procedure.  All  fish  were  sampled  from 
Oregon  southwards,  which  represents 
the  southern  part  of  their  range.  The 
otoliths  were  examined  with  the  aid  of 
image  processing. 

Some  fish  showed  little  or  no  growth 
on  the  otolith  after  eight  months  at 
liberty,  whereas  otoliths  from  other  fish 
grew  substantially.  Some  fish  lay  down 
two  prominent  hyaline  zones  within  a 
single  year,  one  in  the  summer  and  one 
in  the  winter.  We  classified  the  otoliths 
by  morphological  type  and  found  that 
certain  types  are  more  likely  to  lay 
down  multiple  hyaline  zones  and  other 
types  are  likely  to  lay  down  little  or  no 
zones.  This  finding  suggests  that  some 
improvement  could  be  achieved  by 
detailed  knowledge  of  the  growth  char- 
acteristics of  the  different  types. 

This  study  suggests  that  it  may  not 
be  possible  to  obtain  reliable  ages  from 
sablefish  otoliths.  At  the  very  least, 
more  studies  will  be  required  to  under- 
stand the  growth  of  sablefish  otoliths. 


Sources  of  age  determination  errors  for  sablefish 
(Anoplopoma  fimbria)* 


Donald  E.  Pearson 

Santa  Cruz  Laboratory 

National  Marine  Fisheries  Service 

1 10  Shaffer  Road 

Santa  Cruz,  California  95060 

E-mail  address  Don  Pearsom&Noaa  Gov 


Franklin  R.  Shaw 

Alaska  Fisheries  Science  Center 
National  Marine  Fisheries  Service 
7600  Sand  Point  Way  NE 
Seattle,  Washington  98118 


Manuscript  approved  for  publication 
14  July  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish  Bull.  102:127-141  (2004). 


Sablefish  (Anoplopoma  fimbria)  are  a 
valuable  groundfish  resource  off  the 
west  coast  of  North  America.  The  fish- 
ery in  California,  Oregon,  and  Wash- 
ington is  tightly  regulated  according 
to  periodic  stock  assessments.  Between 
1990  and  1998  landings  averaged  more 
than  8000  metric  tons  per  year  and  an 
average  exvessel  (retail)  value  of  12.5 
million  dollars  per  year  (PFMC,  1999). 

Sablefish  are  distributed  in  the 
northeastern  Pacific  Ocean  from 
Baja  California  to  the  Bering  Sea  and 
southeast  to  northern  Japan  (Miller 
and  Lea,  1972).  Males  and  females  are 
sexually  mature  between  55  and  67  cm, 
although  there  is  considerable  variation 
(Fujiwara  and  Hankin  1988a;  Hunter  et 
al,  1989).  Off  Washington,  Oregon,  and 
California,  sablefish  spawn  from  Octo- 
ber through  April  and  spawning  peaks 
in  January  and  February.  Sablefish  are 
oviparous,  releasing  eggs  that  float 
near  the  surface  (Hunter  et  al.,  1989). 
After  hatching,  larvae  and  juveniles  in- 
habit surface  waters  offshore  for  several 
years  after  which  they  migrate  inshore 
and  settle  to  the  bottom. 

Sablefish  are  found  on  the  continen- 
tal slope  and  are  commercially  fished  at 
depths  from  200  to  1400  meters  (Leet  et 
al.,  1992).  Adult  sablefish  feed  on  fish, 
cephalopods,  and  crustaceans  (Laidig 
et  al.,  1998).  They  reach  a  maximum 
length  of  102  cm  (Miller  and  Lea,  1972) 
and  are  believed  to  be  a  very  long-lived 
species  (possibly  100  years  or  more). 

Many  physical  features  have  been 
used  to  age  this  species,  including 


scales,  finrays,  thin-sectioned  otoliths, 
and  broken  and  burned  otoliths,  but  all 
methods  have  resulted  in  less  than  45% 
agreement  among  readers  (Lai,  1985; 
Fujiwara  and  Hankin  1988b;  Kimura 
and  Lyons,  1991;  Heifetz  et  al.  1999). 
The  broken  and  burned  otolith  method 
(Chilton  and  Beamish.  1982)  is  the 
principal  method  used  in  aging  of  the 
species  in  both  the  United  States  and 
Canada.  Typically,  age  readers  agree 
on  ages  less  than  50%  of  the  time,  and 
for  fish  older  than  7  years,  agreement 
drops  to  less  than  15%  (Kimura  and 
Lyons.  1991). 

There  have  been  repeated  efforts  at 
validating  sablefish  ages  and  develop- 
ing aging  criteria.  Beamish  et  al.  ( 1983) 
successfully  used  oxytetracycline  (OTC ) 
marking  to  validate  ages  and  repeated 
his  experiment  in  1995  when  additional 
marked  fish  were  recovered  (MacFar- 
lane  and  Beamish,  1995).  Lai  (1985) 
validated  the  use  of  otoliths  for  aging 
sablefish.  Fujiwara  and  Hankin  ( 1988b) 
examined  otolith  growth  characteristics 
to  help  refine  aging  criteria.  Heifetz  et 
al.  (1999)  validated  the  currently  ac- 
cepted aging  practices  and  examined 
sources  of  error  in  the  aging  of  sablefish. 
Kastelle  et  al.  (1994)  used  radiometric 
methods  to  generally  validate  the  aging 
criteria  currently  used.  Even  with  all  of 
these  studies  that  have  validated  age 


*  Contribution  119  from  the  Santa  Cruz  La- 
boratory, National  Marine  Fisheries  Ser- 
vice, Santa  Cruz,  CA  95060. 


128 


Fishery  Bulletin  102(1) 


?8W  126"W  124°W  122  W  120W  118"W  116°W 

Figure  1 

Map  of  California  and  southern  Oregon  showing  the  locations  ( black  dots ) 
of  sablefish  sampling  and  tagging  in  September  and  October  of  1991. 


determinations,  independent  age  readings  seldom  are  in 
agreement.  This  suggests  that  the  methods  used  to  validate 
the  ages  were  insufficient  to  allow  development  of  precise 
aging  criteria.  The  lack  of  reliable  age  data  has  made  stock 
assessments  difficult  and  controversial  (Crone  et  al.,  1997 ). 
and  in  addition,  accurate  aging  is  needed  to  support  eco- 
logical and  habitat  studies. 

In  September  and  October  of  1991,  a  tagging  and  oxytet- 
racycline  (OTC)  injection  study  was  included  as  part  of  a 
fish  trap  survey  of  the  abundance  of  sablefish  in  southern 
Oregon  and  California.  The  purpose  of  this  study  was  to 
attempt,  once  more,  to  improve  our  ability  to  reliably  age 
sablefish,  thereby  improving  our  ability  to  manage  the 
species. 


Methods 


trarily  selected  fish  at  each  station,  and  the  rest  of  the  fish 
were  tagged  with  blue  spaghetti  tags.  Three  of  every  four 
tagged  fish  were  injected  intraperitoneally  with  30  mg  of 
OTC  per  kilogram  offish  (Beamish  et  al.,  1983)  and  the 
fourth  fish  was  used  as  a  control.  A  complete  description  of 
the  survey  can  be  found  in  Parks  and  Shaw  ( 1994 ). 

A  scientist  visited  the  major  commercial  fishing  ports 
in  California  and  southern  Oregon  to  make  port  samplers, 
commercial  dealers,  and  fishermen  aware  of  the  impor- 
tance of  the  study  and  to  explain  handling  procedures  in 
the  study.  A  $50.00  reward  was  offered  for  the  return  of 
whole  tagged  fish. 

When  a  tagged  fish  was  returned,  the  port  sampler 
measured  it  (fork  length  in  mm),  determined  the  sex,  and 
removed  the  otoliths.  The  otoliths  were  cleaned  and  stored 
in  painted  glass  vials  (because  the  OTC  mark  was  light 
labile)  with  a  5095  ethanol  solution. 


Capture,  tagging,  injection,  and  recovery 

In  September  1991,  the  fisheries  research  vessel  Alaska 
was  chartered  by  the  National  Marine  Fisheries  Service 
to  conduct  a  trap  survey  from  Coos  Bay,  Oregon,  to  Cortez 
Bank,  California  (Fig.  1).  A  total  of  nine  sites  were  visited. 
At  each  site  seven  strings  of  ten  traps  were  deployed  in 
various  depths  between  250  and  1900  meters.  The  traps 
were  retrieved  after  24  hours,  the  catch  was  removed,  and 
the  traps  reset  for  an  additional  24  hours.  All  the  sablefish 
were  counted,  otoliths  were  removed  from  the  first  20  arbi- 


Processing  of  the  otoliths 

Two  pairs  of  otoliths  were  initially  selected  to  develop  the 
procedures  to  be  used  in  the  study.  It  was  found  that  the 
OTC  mark  was  very  faint  and  upon  heating  (as  required 
by  conventional  age  determination  methods),  the  mark 
disappeared.  Accordingly,  we  developed  a  method  to 
obtain  images  of  the  otoliths  before  and  after  heating,  and 
to  superimpose  the  two  images  of  the  same  otolith;  the 
first  viewed  under  UV  light  and,  the  second,  after  heating, 
under  white  light. 


Pearson  and  Shaw:  Age  determination  errors  for  Anoplopomo  fimbria 


129 


OTC  Mark 


Alignment  point 


Alignment  point 


Pasted  UV  Image 


Figure  2 

Composite  image  of  a  sablefish  otolith.  The  otolith  was  first  viewed  under  UV  light 
and  an  image  was  captured.  It  was  then  baked  and  a  second  image  was  captured  by 
using  white  light.  Then  a  small  rectangle  from  the  UV  image  was  electronically  cut 
and  pasted  on  the  image  of  the  baked  otolith.  The  fiourescent  mark  produced  by  the 
OTC  appears  as  a  dark  line  on  the  UV  section.  Points  on  the  otolith  used  for  correct 
positioning  of  the  pasted  section  are  shown. 


The  otoliths  were  embedded  in  epoxy  casting  resin.  After 
the  resin  hardened,  the  blocks  containing  the  otoliths  were 
sliced  in  half  across  the  dorsoventral  axis  with  a  diamond 
saw. 

Images  were  captured  in  a  two-stage  process.  The  first 
stage  used  ultraviolet  light  to  reveal  the  OTC  mark,  and  the 
second  stage  used  white  light  to  reveal  the  growth  marks 
used  for  age  determination.  In  the  first  stage,  the  room  was 
completely  darkened  and  an  image  of  the  otolith,  including 
the  OTC  mark,  was  captured  by  using  a  video  camera  capa- 
ble of  capturing  images  under  low  light  conditions.  We  used 
an  ultraviolet  lamp  which  produced  a  strong  beam  of  light 
at  365  angstroms.  The  otolith  was  viewed  on  a  compound 
microscope  using  reflected  light.  The  camera  and  image  pro- 
cessing system  were  connected  to  a  PC  computer  equipped 
with  a  frame  grabber  card.  A  version  of  NIH  Image,  a  pub- 
lic domain  image  processing  software  (Scion  Corporation, 
Frederick,  MD),  was  used  to  process  the  images. 

The  embedded  otolith  was  placed  on  the  microscope  and 
a  drop  of  mineral  oil  was  placed  on  the  surface  of  the  oto- 
lith. The  limited  amount  of  UV  light  available  to  the  cam- 
era required  the  use  of  frame  averaging.  Usually  30  frames 
were  sufficient  to  produce  a  sharp  view  of  the  otolith  and 
the  fluorescing  mark.  In  some  cases,  the  mark  was  too  faint 
to  allow  an  image  to  be  captured.  When  there  was  sufficient 
fluorescence,  two  composite  images  were  captured,  one  at 
4x  and  one  at  40x. 

In  the  second  stage,  the  same  embedded  otolith  was 
placed  in  a  small  toaster  oven  at  270°C  and  heated  for  20 


to  25  minutes  until  it  had  turned  dark  brown.  This  baking 
process  enhanced  the  growth  rings  for  visual  analysis  and 
approximated  what  age  readers  see  using  the  break  and 
burn  method;  however,  the  latter  process  results  in  darker 
hyaline  zones  than  those  obtained  with  this  method.  After 
cooling,  the  otolith  was  viewed  under  white  light.  A  second 
set  of  images  was  then  captured.  A  section  of  each  UV  im- 
age was  then  electronically  cut  and  pasted  onto  the  image 
captured  under  visible  light.  With  some  experimentation  it 
was  found  that  the  pasted  sections  could  be  aligned  exactly 
over  the  visible  light  images,  creating  a  final  composite  im- 
age as  shown  in  Figure  2. 

Initial  examination  of  the  otoliths 

Initially,  all  OTC-marked  otoliths  were  examined  with 
knowledge  of  the  year  and  season  of  release,  but  without 
any  other  information  about  the  fish.  Composite  UV  and 
white  light  images  were  obtained  as  previously  described. 
The  age  reader  determined  the  following:  whether  or  not 
the  OTC  mark  was  visible;  whether  the  OTC  mark  was  in 
a  hyaline  or  opaque  zone;  the  number  of  annual  hyaline 
zones  visible  beyond  the  OTC  mark  (and  whether  or  not 
the  edge  was  included  in  the  count);  edge  type  (hyaline, 
narrow  opaque,  wide  opaque,  or  unidentifiable),  and  the 
shape  of  the  otolith.  In  some  cases  the  OTC  mark  could 
not  be  identified  or  the  mark  was  too  faint  to  be  captured 
as  a  composite  image;  these  specimens  were  excluded  from 
subsequent  analyses. 


130 


Fishery  Bulletin  102(1) 


:  v.;  -  ■*■-■ 


count  from  here 


Figure  3 

Example  of  an  image  of  a  baked  sablefish  otolith  which  has  been  annotated  with  a 
mark.  The  image  is  an  example  of  one  of  the  images  provided  to  three  researchers  in 
order  to  obtain  cross-reading  comparisons. 


Following  standard  age  determination  procedures  (Chil- 
ton and  Beamish.  1982),  if  a  hyaline  zone  was  not  visible 
on  the  edge  between  January  and  March,  then  the  edge 
was  counted.  If  a  mark  was  not  visible  on  the  edge  between 
April  and  May  and  there  was  a  wide  opaque  zone,  then  the 
edge  was  counted  as  a  mark.  If  a  mark  was  visible  on  the 
edge  and  the  month  was  after  May,  the  edge  was  not  count- 
ed. This  procedure  is  used  to  properly  assign  the  fish  to  an 
annual  cohort.  Because  the  reader  was  not  given  the  month 
of  recapture,  the  ages  were  adjusted  based  on  the  count  of 
hyaline  zones,  the  month  of  recapture,  and  whether  the 
edge  had  been  counted.  This  adjustment  provided  a  cor- 
rected reader  count  of  annual  marks.  The  corrected  count 
was  compared  to  the  number  of  annual  marks  that  would 
have  been  present  if  marks  were  laid  down  annually. 

Previous  experience  suggested  that  there  are  differ- 
ent patterns  of  sablefish  otolith  growth.  We  attempted  to 
classify  and  characterize  these  different  types  of  growth 
patterns  based  on  morphology  of  the  otoliths  as  seen  in 
cross  section.  After  the  otoliths  had  been  examined,  we 
developed  a  standard  classification  scheme  of  morphologi- 
cal classes  and  types  which  could  be  used  to  classify  the 
most  commonly  observed  morphological  types.  The  otoliths 
were  re-examined  and  reclassified  to  see  if  difficulties  and 
discrepancies  in  aging  were  associated  with  morphological 
type.  It  was  hoped  that  this  process  could  be  used  to  refine 
the  aging  criteria  and  improve  precision. 

Because  sample  size  was  small,  we  used  a  Fisher  exact 
test  (Agresti,  1990)  to  test  for  independence  of  morphological 
type  versus  tendency  to  over-estimate,  correctly  estimate,  or 
under-estimate  the  number  of  annual  marks.  The  columns 
in  the  test  indicated  whether  the  fish  had  been  over-aged. 


correctly  aged,  or  under-aged.  The  rows  in  the  test  were  the 
four  morphological  types  identified  in  this  study. 

Examination  of  the  otoliths  by  the  age  readers 

To  determine  how  age  readers  would  count  the  marks  on 
the  otoliths,  we  selected  a  subsample  of  25  otoliths  to  be 
aged  at  four  West  Coast  fisheries  laboratories.  The  otolith 
selection  was  based  on  having  good  quality  images  and 
otoliths.  The  images  of  the  baked  otoliths  (not  the  compos- 
ite images )  were  annotated  with  a  mark  ( Fig.  3 ).  The  mark 
was  placed  in  a  location  which  could  be  readily  located  on 
the  actual  otolith  by  the  readers — on  the  zone  just  inside  of 
the  OTC  mark.  Readers  were  given  the  following:  a  set  of 
printed  images,  an  electronic  file  of  the  images  for  viewing 
on  a  computer  screen,  the  embedded  otolith,  the  month  of 
capture,  the  size  and  sex  of  the  fish  from  which  the  otolith 
came,  and  a  set  of  instructions  for  examining  the  otoliths. 
Readers  were  not  told  where  the  mark  on  the  image  was 
placed  in  relation  to  where  the  OTC  mark  was  in  order  to 
reduce  bias  from  readers  who  may  have  known  when  the 
fish  were  injected  and  recaptured.  Readers  were  asked  to 
provide  the  following:  the  number  of  annual  marks  vis- 
ible outside  the  mark  on  the  image,  whether  the  edge  was 
counted,  how  confident  they  were  of  their  readings,  and  any 
comments  they  might  have. 

Three  readers  participated  in  this  analysis,  two  of  whom 
had  extensive,  long-term  experience  in  aging  sablefish. 
The  readings  and  age  determination  criteria  (including 
edge  count  criteria)  were  compared  to  each  other  and  to 
the  time  known  to  have  passed  between  OTC  marking  and 
recapture. 


Pearson  and  Shaw:  Age  determination  errors  for  Anop/opoma  fimbria 


131 


Figure  4 

Images  of  four  otolith  morphological  types.  (A)  Otolith  is  a  wide  type,  (B )  otolith  is  a  wide, 
wedge  subtype.  (C)  otolith  is  a  thick  type,  and  (D)  otolith  is  a  thick,  wedge  subtype. 


To  determine  if  age  determination  difficulties  were  relat- 
ed to  sex,  size,  area  of  capture,  depth  of  capture,  or  otolith 
morphological  type;  Fisher  exact  tests  were  performed.  In 
each  test,  the  variables  were  compared  to  whether  the  fish 
had  been  correctly  aged,  over  aged,  or  under  aged. 


Results 

Recoveries 

A  total  of  2575  fish  were  tagged  at  the  nine  sites,  and  368 
tagged  fish  were  recaptured.  Of  the  recaptured  fish,  284 
had  been  injected  with  OTC.  Of  the  284  injected  fish,  usable 


otoliths  were  recovered  from  191  fish;  for  the  remaining 
fish,  otoliths  either  were  not  recovered  or  were  too  badly 
damaged  during  removal  to  be  used. 

Otolith  morphological  types 

After  examination  of  all  the  otoliths,  "wide"  and  "thick" 
morphological  types  were  identified,  and  each  type  had  a 
"wedge"  subtype  ( Fig.  4 ).  Each  otolith  in  the  study  was  then 
classified  according  to  this  scheme. 

The  wide  type  (Fig.  4A)  is  characterized  by  new  growth 
that  steadily  increases  cross  sectional  width  along 
the  dorsal  and  ventral  surfaces.  In  the  wedge  subtype 
(Fig.  4B),  initial  growth  increases  the  width,  but  the  most 


132 


Fishery  Bulletin  102(1) 


recent  growth  is  concentrated  on  the  medial  or  lateral 
surface  at  the  sulcus,  decreasing  towards  the  dorsal  and 
ventral  surfaces,  resulting  in  a  wedgelike  appearance. 

The  thick  type  (Fig.  4C)  is  characterized  by  new  growth 
that  increases  the  thickness  of  the  otolith  without  increas- 
ing the  cross  sectional  width,  causing  the  annulii  to  appear 
closely  spaced  on  the  lateral  surfaces.  In  the  wedge  subtype 
(Fig.  4D),  the  most  recent  growth  is  concentrated  at  the  sul- 
cus and  narrows  towards  the  dorsal  and  ventral  surfaces, 
forming  a  wedge  shape. 

It  should  be  noted  that  these  types  and  subtypes  are  not 
always  clearly  defined.  It  should  also  be  noted  that  clas- 
sification to  the  subtype  is  based  on  the  most  recent  one 
or  more  hyaline  zones.  A  wedge  subtype  is  formed  when  a 
single  hyaline  zone  widens  near  the  sulcus  and  comes  to  a 
point  at  the  outer  edge. 


Of  the  191  otoliths  examined,  63  (33.0%)  were  classified 
as  "wide"  types,  76  (39.7% )  were  classified  as  "wide,  wedge 
subtypes,"  32  (16.8%)  were  classified  as  "thick"  types,  5 
(2.6%)  were  classified  as  "thick,  wedge  subtypes,'  and  15 
i  7.99?  ),  could  not  be  classified  by  this  scheme. 

Position  of  the  OTC  mark 

There  was  no  detectable  OTC  mark  in  22  of  191  otoliths. 
The  absence  of  marks  appeared  to  be  a  random  event, 
occurring  in  otoliths  from  several  different  recovery  years 
and  equally  likely  to  be  found  among  different  sexes,  otolith 
types,  different  depths,  and  locations. 

Of  the  169  otoliths  with  detectable  marks,  the  OTC  mark 
was  found  in  a  hyaline  zone  in  129  otoliths  (76.3%),  in  an 
opaque  zone  in  36  otoliths  (21.3%),  and  could  not  be  reli- 


Pearson  and  Shaw:  Age  determination  errors  for  Anoplopoma  fimbria 


133 


Table  1 

Frequency  of  otoliths  with  an  OTC  mark 

appearing  on  the 

edge 

versus  those  with  the  marks  inside  the  edge.  All  fish 

were 

injected  between 

September  and  October  of  1991. 

Mark 

Mark 

Year 

Month 

on  edge 

not  on  edge 

1991 

Oct 

2 

1 

Nov 

1 

3 

Dec 

4 

2 

1992 

Jan 

2 

4 

Feb 

1 

6 

Mar 

7 

4 

Apr 

3 

1 

May 

7 

26 

Jun 

2 

Jul 

1 

7 

Aug 

1 

4 

Sep 

1 

2 

Oct 

5 

Nov 

3 

Dec 

ably  determined  in  four  otoliths  (2Ac/c)  because  the  marks 
were  between  a  hyaline  and  opaque  zone.  Of  the  .36  otoliths 
with  the  mark  in  an  opaque  zone,  the  mark  occurred  just 
after  a  hyaline  zone  in  four  otoliths.  In  24  of  the  36  otoliths 
with  the  mark  in  an  opaque  zone,  the  mark  was  on  the 
edge  where  it  can  be  difficult  to  determine  whether  it  is 
opaque  or  hyaline.  In  no  case  did  the  reader  indicate  that 
the  mark  was  in  a  hyaline  zone  at  the  edge  and  thus  the 
edge  appeared  to  be  opaque  in  most  cases. 

The  OTC  mark  occurred  on  the  otolith  edge  in  30  of  the 
otoliths  recaptured  prior  to  1993  (up  to  16  months  after 
injection).  Examination  of  the  monthly  distribution  of  oto- 
liths with  marks  on  the  edge  ( Table  1 )  indicated  that  some 
fish  exhibited  little  or  no  otolith  growth  for  substantial 
lengths  of  time. 

Otoliths  from  fish  recaptured  in  1992  with  marks  on  the 
edge  (i.e.  showing  little  growth)  were  examined  and  classi- 
fied by  morphological  type  (Table  2).  This  examination  indi- 
cated that  the  thick  type  is  more  likely  to  have  little  growth 


Table  3 

Number  of  visible  hyaline  zones  occurring  after  an  OTC 
mark  on  otoliths  from  fish  recaptured  in  1992.  This  is 
shown  by  three-month  interval  to  show  the  progression  of 
development  of  the  hyaline  zones.  All  fish  were  injected  in 
September  and  October  of  1991. 

Interval 

No.  of  hyaline  zones 

0                            1                         2 

Jan-Mar 

12                            8                         1 

Apr-Jun 
Jul-Sep 
Oct-Dec 

5  14                         4 

6  2 
3                           1 

because  32"*  of  the  otoliths  with  marks  on  the  edge  were  the 
thick  type,  yet  they  made  up  only  Y1CA  of  the  otoliths  in  the 
study.  Conversely,  only  18%  of  the  otoliths  with  the  mark  on 
the  edge  were  of  the  wide  type;  however,  they  made  up  33^ 
of  the  otoliths  in  the  study.  This  trend  was  not  statistically 
significant,  however,  because  the  P-value  was  0.106. 

Number  of  visible  hyaline  zones 

The  number  of  prominent  hyaline  zones  after  the  OTC 
mark  for  fish  recaptured  in  1992  at  three-month  intervals 
is  shown  in  Table  3.  This  distribution  shows  the  otoliths 
that  had  no  detectable  growth  but  also  shows  that  a  hya- 
line zone  forms  in  many  fish  during  the  winter.  It  also 
shows  that  in  some  fish,  a  summer  hyaline  zone  is  formed; 
however,  the  sample  size  for  October-December  was  small 
and  this  is  a  period  when  a  summer  hyaline  zone  would  be 
expected  to  be  fully  visible. 

The  number  of  visible  and  prominent  hyaline  zones  after 
the  OTC  mark  for  fish  recaptured  after  1992  (Table  4),  com- 
pared with  the  number  of  zones  which  should  have  been 
counted,  showed  that  if  a  reader  had  counted  each  of  the 
prominent  hyaline  zones  as  an  annulus,  the  count  would 
have  overestimated  the  age  of  the  fish.  An  example  of  an 
otolith  with  a  larger  number  of  prominent  hyaline  zones 
than  expected  is  shown  in  Figure  5.  It  should  be  noted  that 
a  reader  would  not  necessarily  have  counted  each  of  the 


Table  2 

Number  of  otoliths  in  1992  with  OTC  marks  on  the  edge  by  otolith  morphological  type.  Also  shown  is 
morphological  types  in  the  present  study.  All  fish  were  injected  in  September  and  October  1991. 

the 

overall  percentage  of  the 

Otolith  type 

Wide                             Wide,  wedge 

Thick 

Thick,  wedge 

No. 

Percent            No.             Percent              No. 

Percent 

No.                  Percent 

1992  otoliths                                      4 
Otoliths  in  this  study                     63 

18                  10                  45                     7 
33                  76                  40                   32 

32 
17 

1                          5 
5                        3 

134 


Fishery  Bulletin  102(1) 


Figure  5 

Image  of  a  sablefish  otolith  having  more  prominent  hyaline  zones  than  should  have 
been  present.  The  fish  was  caught  after  eight  months  at  liberty.  A  single  hyaline  zone 
should  have  formed;  however,  there  is  a  zone  on  the  edge  and  one  midway  between 
the  dark  OTC  mark. 


Table  4 

Counts  of  the  number  of  prominent  hyaline  zones  versus  the  number  of  annual  hyaline  zones  that  should  have  been  present  after 
an  OTC  mark.  These  counts  are  for  fish  recaptured  more  than  15  months  after  initial  capture.  Agreement  between  counts  and 
number  of  expected  annual  hyaline  zones  is  shown  in  bold. 


Year 


Expected  number 


1993 
1994 
1995 
1996 

1997 


No.  of  prominent  hyaline  zones 


10 


3 

1 

3 

1 

5 

1 

2 

1 

1 

1 

2 

1 

1 

2 

Table  5 

Percent  and  number  (in  parentheses  I  of  sablefish  otoliths 
with  more  hyaline  zones  than  were  expected,  with  the 
expected  number  of  hyaline  zones  (correct  count),  and  with 
fewer  hyaline  zones  than  were  expected  for  each  otolith  type. 


Otolith  type 


More 
zones 


Expected 
number 
of  zones 


Fewer 
zones 


Thick  10.3%  (3)  41.4%(12)  48.3%  ( 14 1 

Thick,  wedge  0  (0)  40.09!     (2)  60.09!     (3) 

Wide  39.39!  (22)  48.29!  (27)  12.5%     (7) 

Wide, wedge  35.29!  (25)  45.19!  (32)  19.79!  (14) 


prominent  hyaline  zones  as  an  annulus  (they  might  have 
considered  them  to  be  checks).  In  many  of  these  otoliths, 
there  were  less  prominent  zones  that  were  not  counted  and 
which  were  interpreted  as  checks. 

Thick  type  otoliths  and  thick,  wedge  subtype  otoliths 
tend  to  have  fewer  visible  hyaline  zones  than  expected 
(Table  5).  In  contrast,  wide  type  and  the  wide,  wedge  sub- 
type otoliths  are  more  likely  to  have  more  hyaline  zones 
than  expected.  The  Fisher  exact  test  yielded  a  significant 
P-value  of  0.001. 

Blind  comparisons  of  reader  counts 

A  comparison  of  the  counts  of  annual  hyaline  zones  for  each 
reader  to  the  expected  number  of  annual  hyaline  zones 


Pearson  and  Shaw:  Age  determination  errors  for  Anoplopoma  fimbria 


135 


Table  6 

Comparison  of  number  of  annual  hyaline  zones 
by  reader  1  versus  the  expected  number  of  annua 
zones  that  should  have  been  counted.  Agreement 
expected  counts  are  shown  in  bold. 

counted 

hyaline 

with  the 

Expected  count 

Reade 

•  1  count 

1         2       3 

4 

5 

6        7 

1 

2         7        2 

1 

2 

2        4 

1 

1 

3 

1 

1 

4 

2 

5 

1 

Table  8 

Comparison  of  number  of  annual  hyaline  zones  counted 
by  reader  3  versus  the  expected  number  of  annual  hyaline 
zones  which  should  have  been  counted.  Agreement  with 
the  expected  counts  are  shown  in  bold. 

Reader  3  count 

Expected  count                1         2 

3        4         5         6        7 

1  10        2 

2  2         1 

3  1         1 

4  2 

5  1 

3         1         1 

Table  7 

Comparison  of  number  of  annual  hyaline  zones  counted 
by  reader  2  versus  the  expected  number  of  annual  hyaline 
zones  that  should  have  been  counted.  Agreement  with  the 
expected  counts  are  shown  in  bold. 

Reader  2  count 

Expected  count                12        3        4         5 

6       7 

1  5        2        3         1 

2  3        2         2 

3  1         1 
4 

5                                                                           1 

1 
1 

1     1 

after  the  OTC  mark  are  shown  in  Tables  6,  7,  and  8.  In 
these  tables,  it  is  assumed  that  the  readers  should  not  have 
counted  the  zone  in  which  the  OTC  mark  occurred  because 
that  mark  is  presumed  to  have  formed  in  the  summer  of 
1991.  Readers  1  and  2  tended  to  overestimate,  whereas 
reader  3  (the  least  experienced  age  reader)  had  generally 
good  agreement.  Reader  1  agreed  with  the  expected  count 
24%  of  the  time,  reader  2  agreed  with  the  expected  count 
4%  of  the  time,  and  reader  3  agreed  with  the  expected  count 
44%  of  the  time.  The  result  for  reader  3  is  deceptive,  how- 
ever, because  that  reader  did  not  follow  accepted  methods 
of  when  to  count  the  edge. 

Reader  1  and  reader  2  agreed  on  whether  to  count  the 
edge  of  the  otolith  in  24  of  25  otoliths  (Table  9).  Reader  3 
agreed  with  reader  1  on  whether  to  count  the  edge  in  16  of 
25  otoliths  and  17  of  25  otoliths  with  reader  2.  Had  reader 
3  followed  accepted  practice,  agreement  with  the  expected 
count  would  have  been  much  less. 

Efforts  to  determine  what  factors  (depth  of  capture,  loca- 
tion of  capture,  sex,  size  of  the  fish,  and  otolith  morphologi- 
cal type)  resulted  in  a  miscount  of  the  true  number  of  an- 
nual marks  were  inconclusive.  We  first  corrected  the  count 
for  the  fact  that  all  readers  counted  the  mark  in  which 
the  OTC  mark  had  occurred  by  subtracting  one  from  their 


counts,  and  we  then  eliminated  the  readings  from  reader 
3  because  of  his  lack  of  experience  and  anomalous  age  de- 
termination criteria.  Then  we  examined  the  relationship 
of  how  many  otoliths  had  been  over-aged,  correctly  aged, 
and  under-aged  to  the  above  factors.  Depth  of  capture  was 
divided  into  two  groups:  less  than  600  m  and  600  or  more 
m.  Location  was  divided  into  two  groups:  north  and  south 
of  latitude  39  north.  Sizes  were  divided  into  two  groups: 
<55  cm  FL  and  ;>55  cm  FL.  And  finally,  we  tested  each  of 
the  four  otolith  morphological  types. 

We  used  Fisher  exact  tests  to  determine  the  probability 
that  differences  were  due  to  chance  alone.  There  were  no 
detectable  differences  from  the  null  hypothesis  for  depth, 
sex,  or  location  of  capture  (Table  10);  however,  there  was 
some  evidence  that  fish  length  and  otolith  morphological 
type  might  be  related  to  miscounting.  Small  fish  showed  a 
slightly  greater  tendency  to  be  over  counted  (more  rings 
than  should  have  been  present)  than  larger  fish  (P=0.150). 
Otolith  morphological  type  showed  some  departure  from 
randomness:  thick  types  appeared  to  be  more  likely  to  be 
undercounted  (fewer  rings  than  should  have  been  pres- 
ent) and  wide  types  were  more  likely  to  be  over  counted 
(P=0.066). 


Discussion 

Position  of  mark 

There  was  no  visible  mark  on  22  of  the  191  otoliths  ( 11.5%). 
Beamish  et  al.  ( 1983 )  reported  that  14  of  129  OTC-injected 
fish  ( 10.9%)  had  no  detectable  mark.  They  attributed  this  to 
improper  handling  of  the  fish  after  recapture.  The  similar- 
ity in  the  number  of  otoliths  failing  to  show  the  OTC  mark 
between  their  study  and  our  study  suggests  that  some 
portion  of  the  population  may  not  absorb  sufficient  OTC  to 
produce  a  visible  mark. 

The  finding  that  most  of  the  OTC  marks  were  in  a 
hyaline  zone  is  important.  This  indicates  that  many  of 
the  sablefish  in  our  study  laid  down  a  prominent  hyaline 
zone  in  the  summer.  Age  readers  who  conventionally  as- 
sume that  an  annual  mark  is  laid  down  only  in  the  winter 


136 


Fishery  Bulletin  102(1) 


Table  9 

Blind  reading  results  of  25  s 

ablefish  otoliths  by  3  readers.  All  fish  had  been  captured  and  injected  with  OTC  in 

September 

and  Octo- 

ber  of  1991.  The  counts  thev  providec 

are  the  number  of  annual  marks 

outside  of  the  OTC  mark 

"Expected 

count" 

indicate 

s  how 

many  winter  hya 

ine  zones 

should  have  been  present. 

The  columns  labeled  "Edge' 

refer  to  whether  or  not  the  edge 

was 

included 

in  the  age  reader' 

s  counts. 

Fish  ID  no. 

Recapture  date 

Expected  count 

Reader  1 

Reader  2 

Reader 

3 

Count 

Edge 

Count 

Edge 

Count 

Edge 

10375 

4  May  92 

1 

2 

Y 

2 

Y 

N 

10030 

14  May 

92 

1 

1 

Y 

2 

Y 

N 

10267 

17  May 

92 

1 

2 

Y 

3 

Y 

N 

10408 

17  May 

92 

1 

2 

Y 

2 

Y 

Y 

10417 

18  May  92 

1 

2 

Y 

3 

Y 

N 

10630 

25  May 

92 

1 

4 

Y 

4 

Y 

N 

12148 

26  May  92 

1 

3 

Y 

4 

Y 

N 

12176 

26  May 

92 

1 

2 

Y 

5 

Y 

N 

12431 

26  May 

92 

1 

2 

Y 

2 

Y 

Y 

10568 

29  Jul 

92 

1 

1 

N 

2 

N 

N 

11121 

lOct 

92 

1 

2 

N 

6 

N 

N 

11117 

16  Oct 

92 

1 

3 

N 

4 

N 

N 

10400 

12  Jan 

93 

2 

3 

Y 

5 

Y 

3 

Y 

10370 

14  Jan 

93 

2 

4 

Y 

4 

Y 

3 

Y 

10870 

15  Feb 

93 

2 

2 

Y 

7 

Y 

5 

Y 

10246 

15  Apr 

93 

2 

3 

N 

3 

Y 

3 

Y 

11735 

16  May 

93 

2 

3 

Y 

3 

Y 

2 

Y 

11586 

18  May 

93 

2 

2 

Y 

4 

Y 

4 

Y 

11106 

3  Aug 

93 

2 

3 

N 

3 

N 

1 

N 

10617 

2  Dec 

93 

2 

5 

N 

5 

N 

1 

N 

10580 

23  Mav 

94 

3 

2 

Y 

4 

Y 

2 

Y 

10714 

9  Dec 

94 

3 

5 

N 

3 

N 

1 

N 

11516 

3  Aug 

95 

4 

4 

N 

7 

N 

1 

N 

11524 

16  Dec 

95 

4 

4 

N 

6 

N 

1 

N 

11761 

25  Apr 

96 

5 

3 

Y 

5 

Y 

3 

N 

Table  10 

Comparison  of  the 

lumber  offish  under  counted. 

correctly  counted,  and  over  counted  by  two 

experienced  age  readers  versus  depth 

of  capture,  location 

( north 

or  south  of  39  degrees 

latitude), 

sex,  fork  length 

and  otolith  moi 

•phological  type.  The  P-value 

from  the 

Fisher  exact  test  is 

shown 

indicating  the  level  of 

significance. 

LInder  counted 

Correctly  counted 

Over  counted 

P 

Depth 

<600  meters 

7 

14 

15 

0.987 

>600  meters 

3 

6 

5 

Location 

South 

4 

8 

10 

0.606 

North 

3 

12 

7 

Sex 

Male 

3 

2 

5 

0.381 

Female 

7 

18 

15 

Length 

<55  cm 

4 

1  1 

15 

0.150 

255  cm 

6 

6 

5 

Otolith  type 

Thick 

4 

2 

2 

0.066 

Thick,  wedge 

1 

1 

0 

Wide 

2 

5 

11 

Wide,  wedge 

3 

12 

7 

Pearson  and  Shaw:  Age  determination  errors  for  Anop/opoma  fimbria 


137 


Figure  6 

Image  of  a  baked  sablefish  otolith  with  an  electronically  pasted  section  taken  from 
an  image  captured  under  UV  light.  The  dark  OTC  mark  is  clearly  located  within  a 
hyaline  zone,  and  the  hyaline  zone  persists  through  the  entire  otolith.  The  fish  was 
injected  with  OTC  on  5  October  1991. 


would  probably  mis-age  these  fish.  Because  the  age  read- 
ers who  examined  the  otoliths  without  knowledge  of  the 
recapture  information  were  not  informed  that  the  point 
they  were  counting  from  was  just  inside  the  summer  mark, 
it  was  interesting  to  note  that  all  three  of  them  counted 
the  hyaline  zone  in  which  the  OTC  mark  had  occurred  as 
an  annual  hyaline  zone  in  all  cases.  In  other  words,  the 
summer  hyaline  zone  did  not  appear  to  be  a  check  to  the 
readers.  The  readers  indicated  that  the  manner  of  prepa- 
ration of  the  otoliths  (embedded  and  baked)  was  not  the 
manner  in  which  they  were  accustomed  to  view  otoliths 
and  may  have  influenced  their  results.  The  fact  that  the 
hyaline  zones  were  not  as  dark  with  the  baking  method 
as  opposed  to  the  burning  method  may  have  influenced 
the  readers  age  estimates;  however,  some  otolith  burns 
can  be  quite  light  and  experienced  readers  recognize  the 
various  levels  of  burning,  particularly  when  cross  reading 
otoliths  from  other  age  readers.  Readers  sometimes  use 
multiple  sections  and  are  free  to  manipulate  the  otolith 
to  improve  viewing,  which  was  not  possible  in  the  present 
study.  Beamish  et  al.  (1983)  indicated  that  when  readers 
knew  how  many  marks  to  look  for,  they  were  able  to  iden- 
tify false  annual  marks  (checks).  According  to  their  study,  a 
check  is  not  persistent  throughout  the  otolith.  In  Figure  6, 
the  hyaline  zone  in  which  the  OTC  mark  appeared  clearly 
persists  throughout  the  otolith.  If  the  hyaline  zone  which 
contained  the  OTC  mark  began  to  be  laid  down  in  the  win- 
ter, then  there  would  be  very  little  time  for  the  formation 
of  a  wide  opaque  zone  to  form  after  injection  in  the  fall. 
Because  the  age  readers  counted  the  hyaline  zone  in  which 
the  OTC  mark  occurred,  they  clearly  assumed  that  it  was 
not  a  check.  If  the  age  readers  had  known  that  the  hyaline 


zone  (in  which  the  OTC  mark  occurred)  had  formed  in  the 
summer,  then  they  presumably  would  not  have  counted 
it.  It  is  therefore  of  interest  to  see  the  effect  on  agreement 
between  reader  counts  minus  the  hyaline  zone  where  the 
OTC  mark  occurred  and  the  actual  number  of  hyaline 
zones  that  should  have  been  present.  When  we  adjusted  the 
reader  counts  by  subtracting  one  year  from  their  original 
counts  and  compared  their  adjusted  counts  to  the  expected 
number  of  annual  marks  (Table  11 ),  agreement  for  readers 
1  and  2  improved,  whereas  it  decreased  for  reader  3  (the 
least  experienced  reader). 

Also  of  importance  is  the  fact  that  on  some  otoliths,  even 
after  eight  months  at  liberty,  no  growth  had  occurred,  as 
evidenced  by  the  fact  that  the  OTC  mark  was  on  the  edge. 
For  example,  otoliths  from  two  fish,  recaptured  after  eight 
months  at  liberty  showed  marked  differences  in  otolith 
growth  ( Fig.  7 ).  On  otolith  A  there  was  no  detectable  growth 
with  the  OTC  mark  on  the  edge,  whereas  on  otolith  B  there 
was  substantial  growth.  The  OTC  marks  on  both  otoliths 
were  very  prominent.  These  otoliths  came  from  similar  fish; 
that  is,  otolith  A  came  from  a  597-mm  female  fish  caught 
in  680  meters  of  water  at  40°52'  latitude,  and  otolith  B 
came  from  a  610-mm  female  fish  caught  in  480  meters  of 
water  at  41°54'  latitude.  This  provides  strong  evidence 
that  otolith  growth,  and  presumably  fish  growth,  varies 
greatly  among  individual  sablefish.  Beamish  et  al.  (1983) 
reported  that  the  OTC  mark  was  on  or  near  the  edge  in 
28  otoliths  (18.1%)  of  154  fish  which  had  been  at  liberty 
for  two  to  three  years.  In  a  similar  time  interval,  we  found 
that  34  of  126  (27.0%)  had  the  OTC  marks  on  or  near  the 
edge.  Both  the  finding  of  a  summer  hyaline  zone  and  the 
differences  in  growth  of  the  otolith  among  individual  fish 


138 


Fishery  Bulletin  102(1) 


Figure  7 

Images  of  otoliths  from  two  sablefish  showing  differences  in  otolith  growth  rate.  Both 
fish  were  injected  with  OTC  in  early  October  of  1991  and  were  recaptured  in  May  of 
1992.  (Al  Otolith  was  from  a  597-mm  female  caught  in  680  meters  of  water  at  40°52' 
latitude.  iBl  Otolith  was  from  a  610-mm  female  fish  caught  in  480  meters  of  water  at 
41°52'  latitude.  The  OTC  mark  in  A  was  on  the  edge,  whereas  the  position  of  the  OTC 
mark  in  B  is  shown  on  the  insert. 


are  important  factors  in  developing  reliable  and  consistent 
age  determination  criteria. 

The  importance  of  using  the  same  age  determination 
criteria  among  readers  cannot  be  overestimated.  In  the 
blind  comparison,  the  readers  were  asked  whether  they 
had  included  the  edge  in  their  count  of  annual  zones.  With 
standard  age  determination  methods,  if  no  hyaline  mate- 
rial is  visible  on  the  edge  up  to  about  May,  then  the  edge  is 
counted.  This  procedure  is  based  on  the  assumption  that  a 
zone  is  in  the  process  of  forming  but  is  not  yet  clearly  vis- 


ible. On  the  other  hand,  if  hyaline  material  is  observed  on 
the  edge  after  May,  it  is  not  counted  because  it  is  assumed 
to  be  either  a  check  or  the  beginning  of  the  next  winter's 
hyaline  zone.  Reader  1  and  reader  2  (the  two  most  expe- 
rienced age  readers)  agreed  on  whether  to  count  the  edge 
96%  of  the  time,  indicating  that  they  were  using  the  same 
criteria.  Reader  3,  however,  agreed  with  reader  1  only  64% 
of  the  time  and  with  reader  2  only  68%  of  the  time  which 
suggests  that  reader  3  was  using  different  edge-interpreta- 
tion criteria. 


Pearson  and  Shaw:  Age  determination  errors  for  Anoplopoma  fimbria 


139 


Table  11 

Percent  agreement  between  number  of  hyaline  zones  counted  by  three  age  readers  and  the  number  which  should  have  been  pres- 
ent. Also  shown  is  the  effect  of  removing  the  count  of  a  hyaline  zone  which  formed  in  the  summer  and  which  should  not  have  been 
counted  as  an  annual  mark. 


Reader  1 


Reader  2 


Reader  3 


Original 


Corrected 


Original 


Corrected 


Original 


Corrected 


24% 


44% 


36% 


44% 


20% 


Effect  of  ages  on  stock  assessments 

Crone  et  al.  (1997)  noted  that  one  of  the  problems  with 
stock  assessments  of  sablefish  is  that  the  size  at  50%  sexual 
maturity  is  between  55  and  67  cm  ( age  5-7 )  and  that  there 
is  considerable  variability  in  the  these  estimates.  Further, 
they  noted  that  there  has  been  difficulty  in  determining 
age-specific  selectivity  because  of  problems  with  the  ages 
used  in  previous  assessments.  Crone  et  al.  (1997)  further 
noted  that  there  is  a  considerable  discrepancy  in  ages 
among  the  age  determination  laboratories  on  the  west 
coast.  Finally,  the  model  used  to  perform  stock  assessments 
has  estimated  that  in  order  to  obtain  a  good  fit  with  the 
data,  the  actual  level  of  aging  error  should  be  higher  than 
has  been  reported.  The  lack  of  reliable  age  data  has  been 
used  to  criticize  stock  assessments. 

Age  and  length  at  sexual  maturity  has  been  found  to 
vary  substantially  by  depth  (Fujiwara  and  Hankin,  1988a). 
Fujiwara  and  Hankin  found  that  both  males  and  females 
had  a  length  of  550  mm  for  the  length  at  50%  sexual  matu- 
rity in  shallow  water  (<600  meters ).  In  depths  greater  than 
600  m,  the  size  at  50%  sexual  maturity  was  450  mm  for 
males  and  500  mm  for  females.  To  determine  age,  they  used 
sectioned  otoliths  and  methods  that  may  not  have  been 
directly  comparable  to  the  methods  used  in  other  studies 
or  the  methods  used  in  the  present  study;  nonetheless,  they 
found  that  both  males  and  females  matured  at  a  younger 
age  in  deeper  water.  Saunders  et  al.  (1997)  also  reported 
differences  in  length  at  maturity  related  to  depth  and  loca- 
tion of  capture.  Methot1  found  that  ontogenetic  movement 
into  deeper  water  for  spawning  was  more  closely  related  to 
age  than  size.  If  sexual  maturity  is  more  closely  related  to 
age  than  length  as  suggested  by  Methot,  then  unreliable 
ages  may  explain  the  variable  maturity  schedule  for  sable- 
fish.  In  our  study,  fish  were  captured  over  a  900  nmi  range 
at  depths  from  200  to  more  than  1000  m.  If  depth  is  related 
to  growth  of  sablefish,  then  it  is  possible  that  the  different 
morphometric  types  of  otoliths  observed  in  our  study  may 
also  be  a  function  of  depth.  If  depth  is  responsible  for  the 
morphological  types,  it  also  suggests  that  reliability  of  ages 
may  be  a  function  of  the  depth  at  which  the  sablefish  are 
found.  Further,  if  depth  influences  growth,  a  fish  which 


1  Methot.  R.  D.  1995.  Geographic  patterns  in  growth  and 
maturity  of  female  sablefish  off  the  U.S.  west  coast.  Unpubl. 
manuscript,  39  p.  NOAA,  NMFS,  Northwest  Fisheries  Science 
Center,  Seattle,  WA. 


changes  its  depth  over  time,  may  exhibit  different  patterns 
of  growth  throughout  its  life  which  would  further  compli- 
cate the  problem  of  determining  reliable  ages. 

Potential  sources  of  error  in  this  study 

This  study  used  sablefish  caught  in  the  southern  part  of 
the  sablefish  range.  Many  species  show  latitudinal  varia- 
tion in  growth  (June  and  Reintjes,  1959;  White  and  Chit- 
tenden, 1977;  Leggett  and  Carscadden,  1978;  Shepherd  and 
Grimes,  1983;  Pearson  and  Hightower,  1991).  It  is  possible 
that  the  results  of  this  study  do  not  apply  to  the  northern 
portion  of  their  range. 

Another  potential  source  of  error  in  our  study  is  the  effect 
of  tagging  on  the  growth  of  the  sablefish.  MacFarlane  and 
Beamish  ( 1990 )  found  that  tagged  sablefish  grew  slower 
than  untagged  fish.  If  this  is  true,  then  the  results  of  this 
study  are  much  more  difficult  to  interpret.  MacFarlane  and 
Beamish  did  not  use  OTC  and  as  a  result  they  based  their 
ages  on  conventional  aging  methods.  If  they  had  injected 
the  fish,  it  would  have  been  interesting  to  note  whether 
the  ages  for  the  fish  in  their  study  would  have  been  inter- 
preted differently.  If  fish  do  grow  differently  after  tagging, 
many  age,  growth,  and  validation  studies  will  need  to  be 
re-evaluated. 


Conclusion 

Obtaining  accurate  ages,  with  reasonable  precision,  for 
sablefish  is  very  difficult.  Previous  aging  studies  of  sable- 
fish have  obtained  results  similar  to  ours,  even  when  the 
readers  knew  how  many  annual  marks  should  have  been 
present  (Beamish  et  al.  1983;  MacFarlane  and  Beamish, 
1995).  We  found  that  some  fish  lay  down  two  marks  a  year 
and  others  may  not  lay  down  any.  We  also  found  that  certain 
morphological  types  of  otoliths  may  be  indicative  of  slow 
growing  fish  and  others  may  be  indicative  of  rapidly  grow- 
ing fish  (assuming  otolith  growth  relates  to  fish  growth). 

The  fact  that  agreement  among  readers  or  with  the  cor- 
rect age  consistently  ranges  between  30%  and  45%  sug- 
gests that  this  imprecision  may  be  inherent  in  sablefish 
aging.  A  substantial  fraction  of  the  population  may  not  be 
able  to  be  reliably  aged:  some  otoliths  do  not  appear  to 
grow  and  others  grow  very  rapidly,  laying  down  prominent 
summer  hyaline  zones  that  even  experienced  age  readers 
cannot  differentiate  from  winter  hyaline  zones. 


140 


Fishery  Bulletin  102(1) 


We  believe  the  wide  type  and  wide,  wedge  subtypes  are 
often  over-aged,  and  the  thick  type  and  thick,  wedge  sub- 
types are  occasionally  under-aged  and  further  propose  that 
readers  be  made  aware  that  a  hyaline  zone  typically  forms 
in  the  winter,  but  that  it  is  not  uncommon  for  a  second 
mark  to  form  in  the  summer. 

Another,  less  desirable  approach,  would  be  for  age  read- 
ers to  record  the  morphological  type  of  otolith  as  a  routine 
part  of  aging.  Users  of  the  data  could  then  incorporate  this 
information  into  their  studies  by  using  a  correction  factor 
for  fish  likely  to  be  under-aged  and  for  fish  likely  to  be 
over-aged.  This  factor  could  be  in  the  form  of  an  aging  error 
matrix  as  suggested  by  Heifetz  et  al.  ( 1999 ).  This  approach 
may  not  be  practical  until  more  data  are  available  on  the 
true  effect  on  ages  for  the  morphological  types  described 
in  this  study,  including  how  many  years  would  need  to  be 
added  or  subtracted  for  each  type.  Finally,  a  more  complete 
description  of  the  morphological  types  would  be  needed  to 
assist  the  age  readers. 


Acknowledgments 

We  would  like  to  express  our  gratitude  to  Delsa  Anderl 
(Alaska  Fisheries  Science  Center),  Kristin  Munk  (Alaska 
Department  of  Fish  and  Game),  Shayne  MacLellan  (Pacific 
Biological  Station,  Canadian  Department  of  Fisheries  and 
Oceans),  and  Bruce  Pederson  (Oregon  Department  of  Fish 
and  Wildlife)  for  participating  in  the  otolith  blind  reading 
component  of  this  paper.  We  would  also  like  to  thank  Dan 
Kimura  and  Craig  Kastelle  of  the  Alaska  Fisheries  Science 
Center  for  their  assistance  in  developing  the  design  of  this 
study.  We  would  like  to  thank  Michael  Mohr  (Southwest 
Fisheries  Science  Center,  Santa  Cruz,  CA)  for  his  valuable 
contribution  to  the  statistical  analyses  used  in  this  study 
This  study  could  not  have  been  completed  without  the  sup- 
port of  Gary  Stauffer  (Alaska  Fisheries  Science  Center) 
who  provided  funding  for  the  recovery  of  the  sablefish. 
Additionally,  this  study  would  never  have  been  completed 
without  the  assistance  of  numerous  commercial  market 
samplers,  port  biologists,  commercial  fishermen,  and  deal- 
ers who  were  responsible  for  collecting  and  processing  the 
sablefish  when  they  were  caught.  And  finally,  we  would 
like  to  thank  William  Lenarz  (Southwest  Fisheries  Science 
Center,  retired)  for  his  support  of  this  study. 


Literature  cited 

Agresti,  A. 

1990.     Categorical  data  analysis,  558  p.     Wiley.  New  York. 
NY. 
Beamish.  R.  J..  G.  A.  MacFarlane,  and  D.  E.  Chilton. 

1983.     Use  of  oxytetracycline  and  other  methods  to  validate  a 
method  of  age  determination  for  sablefish.     In  Proceedings 
of  the  second  international  sablefish  symposium,  p.  95-116. 
Alaska  Sea  Grant  Rep.  83-8,  Univ.  Alaska,  Fairbanks,  AK. 
Chilton,  D.  E„  and  R.  ■).  Beamish. 

1982.  Age  determination  methods  for  fishes  studied  by 
groundfish  program  at  the  Pacific  Biological  Station.  Can 
Spec  Publ.  Fish.  Aquat.  Sci.  60:19-22. 


Crone,  P.  R.,  R.  D.  Methot,  R.  J.  Conser,  R.  R.  Lauth.  and 
M.  E.  Wilkins. 

1997.  Status  of  the  sablefish  resource  off  the  U.S.  Pacific 
Coast  in  1997.  Appendix  G  in  Status  of  the  Pacific  Coast 
groundfish  fishery  through  1997  and  recommended  accept- 
able biological  catches  for  1998,  135  p.  Pacific  Fisheries 
Management  Council  1998. 

Fujiwara,  S.,  and  D.  G.  Hankin. 

1988a.  Sex  ratio,  spawning  period,  and  size  and  age  at  matu- 
rity of  sablefish  Anoplopoma  fimbria  off  northern  California. 
Nippon  Suisan  Gakkaisha  54(8):1333-1338. 
1988b.  Aging  discrepancy  related  to  asymmetrical  otolith 
growth  for  sablefish  Anoplopoma  fimbria  in  northern 
California.  Nippon  Suisan  Gakkaishi  54(1):27-31. 
Heifetz,  J.,  D.  Anderl,  N.  E.  Maloney,  and  T.  L.  Rutecki. 

1999.     Age  validation  and  analysis  of  aging  error  from 
marked  and  recaptured  sablefish.  Anoplopoma  fimbria. 
Fish.  Bull.  97:256-263. 
Hunter,  J.  R.,  B.  J.  Macewicz,  and  C.  A.  Kimbrell. 

1989.  Fecundity  and  other  aspects  of  the  reproduction  of 
sablefish,  Anoplopoma  fimbria,  in  central  California  waters. 
CalCOFI  Rep.  30,  1989,  p.  61-72. 

June,  F.  C,  and  J.  W.  Reintjes. 

1959.  Age  and  size  composition  of  the  menhaden  catch  along 
the  Atlantic  coast  of  the  United  States,  1952-55:  with  a  brief 
review  of  the  commercial  fishery.  U.S.  Fish  and  Wildl. 
Serv.,  Spec.  Sci.  Rep.  Fish.  317,  65  p. 

Kastelle,  C.  R,  D.  K.  Kimura,  A.  E.  Nevissi,  and  D.  R.  Gunderson. 

1994.  Using  Pb-210/Ra-226  disequilibria  for  sablefish.  Ano- 
plopoma fimbria,  age  validation.     Fish.  Bull.  92:292-301. 

Kimura,  D.  K.,  and  J.  J.  Lyons. 

1991.  Between-reader  bias  and  variability  in  the  age  deter- 
mination process.     Fish.  Bull.  89:53-60. 

Lai,  H.  L. 

1985.  Evaluation  and  validation  of  age  determination  for 
sablefish.  pollock  and  yellowfin  sole;  optimum  sampling 
design  using  age-length  key;  and  implications  of  aging 
variability  in  pollock.  Ph.D.  diss.,  426  p.  Univ.  Washing- 
ton, Seattle,  WA 

Laidig,  T.  E.,  P.  B.  Adams,  and  W  M.  Samiere. 

1998.  Feeding  habits  of  sablefish,  Anoplopoma  fimbria,  off 
the  coast  of  Oregon  and  California.  NOAA  Tech.  Rep 
NMFS  130:65-79. 

Leet,  W.  S.,  C.  M.  Dewees,  and  C.  W.  Haugen. 

1992.  California's  living  marine  resources  and  their  utiliza- 
tion. Calif.  Seagrant  Publication  UCSGEP-92-12,  257  p. 
Univ.  California  Davis,  Davis,  CA. 

Leggett,  W.  C,  and  J.  E.  Carscadden. 

1978.  Latitudinal  variation  in  reproductive  characteristics 
of  American  shad  (Alosa  sapidissima):  evidence  for  popu- 
lation specific  life  history  strategies  in  fish.  J.  Fish.  Res. 
Board  Can.  35:1469-1478. 

Macfarlane,  G.  A.,  and  R.  J.  Beamish. 

1990.  Effect  of  an  external  tag  on  growth  of  sablefish  <A/io- 
plopoma  fimbria  I,  and  consequences  to  mortality  and  age  at 
maturity.     Can.  J.  Fish.  Aquat.  Sci.  47:1551-1557. 

1995.  Validation  of  the  otolith  cross-section  method  of  age 
determination  for  sablefish  [Anoplopoma  fimbria*  using 
oxytetracycline.  In  Recent  developments  in  fish  otolith 
research  ID.  H.  Secor,  J.  M.  Dean,  and  S.  E.  Campana,  eds.), 
p.  319-329.     Univ.  South  Carolina  Press,  Columbia,  SC. 

Miller,  D.  J.,  and  R.  N.  Lea. 

1972.     Guide  to  the  coastal  marine  fishes  of  California. 
CDFG  (Calif.  Dep.  Fish  and  Game)  Fish  Bull   157.  249  p. 
Parks,  N.B..  and  F.R.Shaw. 

1994.     Relative  abundance  and  size  composition  of  sablefish 


Pearson  and  Shaw:  Age  determination  errors  for  Anop/opoma  fimbria 


141 


iAnoplopoma  fimbria)  in  the  coastal  waters  of  California 
and  southern  Oregon,  1984-1991.  NOAA  Tech.  Memo 
NMFS-AFSC-35,  38  p. 

Pearson,  D.  E.,  and  J.  E.  Hightower. 

1991.  Spatial  and  temporal  variability  in  growth  of  widow 
rockfish  (Sebastes  entomelas).  NOAA  Tech.  Mem.  NOAA- 
TM-NMFS-SWSC-167,  43  p. 

PFMC  (Pacific  Fishery  Management  Council). 

1999.  Status  of  the  Pacific  Coast  groundfish  fishery  through 
1999  and  recommended  acceptable  biological  catches  for 
2000,  232  p.  Pacific  Fishery  Management  Council,  Port- 
land, Oregon. 


Saunders.  M.  W..  B.  M.  Leaman,  G.  A.  McFarlane. 

1997.     Influence  of  ontogeny  and  fishing  mortality  on  the 
interpretation  of  sablefish,  Anoplopoma  fimbria,  life 
history.     NOAA  Tech  Rep.  NMFS  130:81-92. 
Shepherd,  G.,  and  C.  B.  Grimes. 

1983.     Geographic  and  historic  variations  in  growth  of  weak- 
fish,  Cynoscion  regalis,  in  the  middle  Atlantic  bight.     Fish. 
Bull.  81:4:803-813. 
White,  M.  L.,  and  M.  E.  Chittenden. 

1977.  Age  determination,  reproduction,  and  population 
dynamics  of  the  Atlantic  croaker,  Micropogonius  undu- 
latus.     Fish.  Bull.  75:109-123. 


142 


Abstract— Life  history  aspects  of  larval 
and,  mainly,  juvenile  spotted  seatrout 
tCynoscion  nebulosus)  were  studied 
in  Florida  Bay.  Everglades  National 
Park,  Florida.  Collections  were  made 
in  1994-97,  although  the  majority  of 
juveniles  were  collected  in  1995.  The 
main  objective  was  to  obtain  life  history 
data  to  eventually  develop  a  spatially 
explicit  model  and  provide  baseline 
data  to  understand  how  Everglades  res- 
toration plans  (i.e.  increased  freshwater 
flows)  could  influence  spotted  seatrout 
vital  rates.  Growth  of  larvae  and  juve- 
niles (<80  mm  SL)  was  best  described 
by  the  equation  log,  standard  length 
=  -1.31  +  1.2162  (log,,  age).  Growth  in 
length  of  juveniles  (12-80  mm  SL)  was 
best  described  by  the  equation  standard 
length  =  -7.50  +  0.8417  (age).  Growth 
in  wet  weight  of  juveniles  (15-69  mm 
SL)  was  best  described  by  the  equation 
logc  wet-weight  =  -4.44  +  0.0748  (age). 
There  were  no  significant  differences 
in  juvenile  growth  in  length  of  spot- 
ted seatrout  in  1995  between  three 
geographical  subdivisions  of  Florida 
Bay:  central,  western,  and  waters  adja- 
cent to  the  Gulf  of  Mexico.  We  found  a 
significant  difference  in  wet-weight  for 
one  of  six  cohorts  categorized  by  month 
of  hatchdate  in  1995.  and  a  significant 
difference  in  length  for  another  cohort. 
Juveniles  (i.e.  survivors)  used  to  cal- 
culate weekly  hatchdate  distributions 
during  1995  had  estimated  spawning 
times  that  were  cyclical  and  protracted, 
and  there  was  no  correlation  between 
spawning  and  moon  phase.  Tem- 
perature influenced  otolith  increment 
widths  during  certain  growth  periods  in 
1995.  There  was  no  evidence  of  a  rela- 
tionship between  otolith  growth  rate 
and  temperature  for  the  first  21  incre- 
ments. For  increments  22-60,  otolith 
growth  rates  decreased  with  increas- 
ing age  and  the  extent  of  the  decrease 
depended  strongly  in  a  quadratic  fash- 
ion on  the  temperature  to  which  the 
fish  was  exposed.  For  temperatures  at 
the  lower  and  higher  range,  increment 
growth  rates  were  highest.  We  suggest 
that  this  quadratic  relationship  might 
be  influenced  by  an  environmental 
factor  other  than  temperature.  There 
was  insufficient  information  to  obtain 
reliable  inferences  on  the  relationship 
of  increment  growth  rate  to  salinity 


Manuscript  approved  for  publication 
23  June  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  ( >ffice. 

Fish.  Bull.  102:142-155  (2004). 


Growth,  mortality,  and  hatchdate  distributions 
of  larval  and  juvenile  spotted  seatrout 
{Cynoscion  nebulosus)  in  Florida  Bay, 
Everglades  National  Park 


Allyn  B.  Powell 
Robin  T.  Cheshire 
Elisabeth  H.  Laban 

National  Ocean  Service 

National  Oceanic  and  Atmospheric  Administration 

Center  for  Coastal  Fisheries  and  Habitat  Research 

101  Pivers  Island  Road 

Beaufort,  North  Carolina  28516 

E-mail  address  (for  A.  B  Powell):  allyn  powellgnoaa  gov 

James  Colvocoresses 

Patrick  O'Donnell 

Florida  Fish  and  Wildlife  Commission 
Florida  Marine  Research  Institute 
2796  Overseas  Highway,  Suite  119 
Marathon,  Florida  33050 

Marie  Davidian 

Room  209,  Patterson  Hall 
2501  Founder's  Drive 
North  Carolina  State  University 
Raleigh,  North  Carolina  27695 


The  spotted  seatrout  (Cynoscion  nebu- 
losus) is  an  important  recreational  fish 
in  Florida  Bay  and  spends  its  entire  life 
history  within  Florida  Bay  I  Rutherford 
et  al.,1989).  The  biology  of  adult  spotted 
seatrout  in  Florida  Bay  is  well  known 
(Rutherford  et  al.,  1982, 1989),  as  are  the 
distribution  and  abundance  of  juveniles 
in  the  bay,  including  a  description  of  the 
juvenile  habitats  and  their  diets  (Het- 
tler,  1989;  Chester  and  Thayer,  1990; 
Thayer  et  al.,  1999;  Florida  Department 
of  Environmental  Protection1).  The 
temporal  and  spatial  distribution  and 
abundance  of  larval  spotted  seatrout  in 
Florida  Bay  and  adjacent  waters,  and  the 
spatial  and  temporal  spawning  habits  of 
these  larvae  also  have  been  determined 
(Powell  et  al.,  1989;  Rutherford  et  al.. 
1989;  Powell,  2003). 

The  early  life  history  of  spotted 
seatrout  in  other  south  Florida  estu- 
aries also  has  been  well  documented. 
Peebles  and  Tolley  ( 1988)  described  the 
distribution,  growth,  and  mortality  of 
larval  spotted  seatrout  in  Naples  and 


Fakahatchee  Bays,  and  McMichael  and 
Peters  (1989)  described  the  size  distri- 
bution, growth,  spawning,  and  diet  of 
spotted  seatrout  in  Tampa  Bay. 

Information  on  growth  and  mortality 
of  larval  and  juvenile  spotted  seatrout 
in  Florida  Bay  is  lacking.  Research  on 
these  topics  would  enhance  our  under- 
standing of  the  entire  life  history  of  this 
valuable  species,  and  in  particular  aid 
in  eventually  developing  a  spatially  ex- 
plicit model  for  spotted  seatrout  that  is 
highly  desired  by  the  Program  Manage- 
ment Committee  for  the  South  Florida 
Ecosystem  Restoration  Prediction  and 
Modeling  Program.  In  addition,  these 
life  history  studies  could  help  clarify  ju- 
venile growth  and  survival  and  provide 
needed  information  for  the  restoration 


Florida  Department  of  Environmental 
Protection.  1996.  Fisheries-independent- 
monitoring  program.  1995  annual  report, 
58  p.  Florida  Department  of  Environmen- 
tal Protection,  Florida  Marine  Research 
Institute,  100  8th  Avenue  SE,  St.  Peters- 
burg, FL  33701. 


Powell  et  al.:  Growth,  mortality,  and  hatchdate  distributions  for  Cynoscion  nebulosus 


143 


25°20' 


25°  10' 


2.V00' 


—  24"50' 


81  "00' 


80"45' 


so  Mr 


Figure  1 

Location  of  sampling  sites  for  spotted  seatrout  (Cynoscion  nebulosus)  in  Florida  Bay, 
Everglades  National  Park,  Florida,  including  Florida  Bay  Subdivisions. 


of  the  Everglades,  including  a  return  of  historic  freshwater 
flows  into  Florida  Bay. 

Two  conceptual  frameworks  have  been  advanced  to  couple 
the  role  of  growth  and  mortality  in  influencing  cohort  dy- 
namics. Anderson  ( 1988),  in  a  review  of  hypotheses  relating 
survival  of  prerecruits  to  recruitment,  advocated  a  growth- 
mortality  hypothesis  as  a  rational  framework  for  early  life 
history  studies  that  address  recruitment  variability.  This 
concept  predicts  that  survival  of  a  cohort  is  directly  related 
to  growth  rates  during  the  early  life  stages.  The  growth- 
mortality  framework,  which  includes  several  important  in- 
tegrated components  and  is  based  on  bioenergetic  principles 
of  growth  and  ecological  theory  that  predict  growth  rate,  is 
directly  related  to  survival.  If  it  can  be  demonstrated  that 
survival  is  a  function  of  growth  during  the  early  life  stage, 
then  a  valuable  tool  becomes  available  for  examining  mecha- 
nisms influencing  recruitment  of  marine  fishes. 

Another  framework  suggests  that  the  mortality  rate  does 
not  operate  alone  in  determining  stage-specific  survival, 
but  it  is  the  mortality:growth  (M:G)  ratio  (mortality  per 
unit  of  growth)  that  determines  stage-specific  survival  (see 
citations  in  Houde,  1997 ).  Houde  ( 1997 )  advanced  the  idea 
of  using  the  M:G  ratio  as  an  estimator  of  production  and 
potential  survivorship  especially  in  early  life  stages  when 
both  mortality  and  growth  are  high  and  variable.  This  con- 
cept was  partly  based  on  the  strong  coupling  of  growth  and 
mortality  demonstrated  by  Ware  ( 1975 )  who  argued  that 
when  growth  rate  is  poorer  than  average,  larvae  would  be 
exposed  to  sources  of  mortality  over  a  longer  period  and 
hence  their  mortality  rate  would  increase.  Growth  and 


mortality  values  for  successive  cohorts  would  tend  to  form 
a  cluster  of  points  around  a  regression  of  mortality  on 
growth  based  on  average  values  for  a  particular  species. 

Our  intent  is  not  to  test  the  growth-mortality  hypothesis 
(sensu  Hare  and  Cowen,  1997)  as  outlined  by  Anderson 
(1988),  nor  fully  to  develop  the  M:G  ratio  concept  (Houde, 
1997),  but  rather  to  use  these  concepts  as  a  framework 
for  our  study.  The  major  goal  is  to  provide  information  on 
growth  and  survival  of  larval  and,  mainly,  juvenile  spotted 
seatrout  that  can  ultimately  be  used  to  develop  a  spatially 
explicit  model  that  can  be  linked  to  Everglades  restoration 
activities.  Therefore,  the  major  objectives  of  this  paper  are 
1 )  to  determine  overall  growth  rates  of  larval  and  juvenile 
spotted  seatrout  in  Florida  Bay;  2)  to  determine  and  com- 
pare juvenile  growth  rates  geographically;  3)  to  estimate 
natural  mortality  rates  of  juveniles;  4)  to  estimate  hatch- 
date  distributions;  5 )  to  compare  cohort  growth  and  mortal- 
ity rates  and  G:M  ratios  for  juveniles;  and  6)  to  evaluate 
the  effects  of  salinity  and  temperature  on  otolith  growth — a 
surrogate  for  somatic  growth. 


Methods  and  materials 

Field  collections 

Larval  fish  used  for  otolith  microstructure  analysis  were 
collected  from  September  1994  through  July  1997,  mainly 
in  the  Gulf  transition,  western,  and  central  subdivisions 
(Table  1,  Fig.  1).  These  subdivisions  designated  by  the 


144 


Fishery  Bulletin  102(1) 


Table  1 

Florida 

Bay  sampling  stations  where  otoliths  from  spotted  seatrout  were  collected 

Included  are  numbers  in  >  of  larvae  and  juveniles 

used  in 

the  otolith  microstructure  ar 

alysis.  and  subdivisions  as  defined  by  the  South  Florida  Ecosystem  Restoration  Prediction  and 

Modeling  Program,  Program  Management 

Committee. 

Station 

Latitude 

Longitude 

Florida  Bay 

Juveniles 

Larvae 

numbei 

(degrees  and  minutes) 

(degr 

ees  and  minutes) 

subdivisions 

Location 

(n) 

in) 

1 

25  06.81 

81  05.27 

Gulf  transition 

Cape  Sable 



4 

2 

25  06.37 

81  01.42 

Gulf  transition 

Middle  Ground 

1 

10 

3 

25  06.40 

80  58.58 

Gulf  transition 

Conchie  Channel 

— 

4 

4 

25  07.70 

80  56.90 

Gulf  transition 

Bradley  Key 

119 

— 

5 

25  07.12 

80  56.07 

western 

Murray  Key 

4 

8 

6 

25  08.11 

80  50.95 

central 

Snake  Bight 

3 

— 

7 

25  09.45 

80  53.42 

central 

Snake  Bight 

4 

— 

8 

25  07.50 

80  48.51 

central 

Rankin  Lake 

12 

— 

9 

25  05.06 

80  47.30 

central 

Roscoe  Key 

20 

— 

10 

25  02.30 

81  .1.12 

Gulf  transition 

Sandy  Key 

49 

— 

11 

25  02.90 

80  55.00 

western 

Johnson  Key  Basin 

125 

— 

12 

25  06.00 

80  52.50 

western 

Palm  Key  Basin 

110 

— 

13 

25  04.50 

80  45.15 

central 

Whipray  Basin 

2 

62 

14 

25  08.00 

80  43.20 

central 

Crocodile  Point 

9 

— 

15 

24  56.70 

80  57.20 

Gulf  transition 

Schooner  Bank 

2 

— 

16 

24  54.70 

80  56.31 

Gulf  Transition 

Sprigger  Bank 

— 

8 

17 

25  00.40 

80  47.68 

central 

Sid  Key  Bank 

6 

— 

18 

24  57.03 

80  47.52 

central 

Twin  Key  Basin 

6 

— 

19 

25  07.98 

80  40.48 

eastern 

Madeira  Point 

1 

— 

20 

25  11.85 

80  37.15 

northern 

Little  Madeira  Bay 

8 

— 

21 

25  13.00 

80  27.80 

eastern 

Shell  Key 

5 

— 

South  Florida  Ecosystem  Restoration  Prediction  and 
Modeling  Program,  Program  Management  Committee, 
were  based  on  modifications  of  the  benthic  mollusc  com- 
munity (Turney  and  Perkins,  1972).  In  1994  and  1995,  we 
used  60-cm  bongo  nets  fitted  with  0.333-mm  mesh  fished 
from  the  port  side  of  a  5.4-m  boat.  Beginning  in  1996,  we 
used  a  paired  60-cm  bow-mounted  push  net  with  0.333- 
mm  mesh  nets  similar  to  that  described  by  Hettler  and 
Chester  (1990). 

Juvenile  spotted  seatrout  were  obtained  from  monitor- 
ing programs  established  by  the  NOAA  Center  for  Coastal 
Fisheries  and  Habitat  Research  (NOAA)  and  Florida  Ma- 
rine Research  Institute  (FMRI).  NOAA  collections  were 
made  from  May  1995  through  September  1997.  Juveniles 
were  collected  with  an  otter  trawl  towed  between  two  5-m 
boats.  The  otter  trawl  measured  3.4  m  (headrope)  and  was 
fitted  with  a  3.2-mm  mesh  tailbag  with  6-mm  mesh.  FMRI 
collections  were  made  in  1995  with  a  seine  and  a  trawl.  The 
21.4-m  center-bag  drag  seine  was  fitted  with  a  1.8  m  x  1.8  m 
x  1.8  m  bag  of  3.2-mm  mesh.  The  6.1-m  (headrope)  otter 
trawl  was  fitted  with  a  body  of  38.1 -mm  stretch  mesh  and  a 
3.2-mm  mesh  tailbag.  The  majority  of  juveniles  (86ri  )  from 
NOAA  and  FMRI  collections  were  collected  in  1995. 

Otolith  microstructure  analysis 

Otolith  processing  Otolith  removal  and  preparation  gen- 
erally followed  the  methods  of  Secor  et  al.  (1991).  All  oto- 


liths, except  for  the  right  sagitta,  were  mounted  on  a  slide 
with  mounting  media  and  archived.  The  right  sagittal  oto- 
lith was  embedded  for  transverse  sectioning  or  polishing 
(or  both).  The  left  sagitta  was  embedded  for  transverse  sec- 
tioning if  the  right  was  damaged.  Sagittae  were  read  with 
a  light  microscope  at  lOOOx  magnification  under  oil  immer- 
sion. The  first  increment  was  determined  as  that  following 
the  core  increment;  which  was  defined  as  a  well-defined 
dark  increment  surrounding  the  core  (Powell  et  al.,  2000). 
Two  blind  counts  of  increments  were  made  by  one  reader 
and  if  the  counts  differed  by  more  than  5,  then  the  otolith 
was  read  again.  If  the  counts  were  within  the  acceptable 
range,  the  two  counts  were  averaged.  Based  on  a  previous 
validation  study  (Powell  et  al.,  2000),  2.5  days  were  added 
to  the  increment  counts  to  obtain  the  daily  age.  A  total 
of  582  sagittal  otoliths  were  aged.  This  total  included  96 
from  larval  collections  from  September  1994  through  July 
1997, 139  juveniles  from  NOAA  collections  from  June  1995 
through  September  1997,  and  347  from  FMRI  collections 
from  June  1995  through  December  1995. 

Increment  widths  were  measured  on  347  otoliths  from 
FMRI  collections  (1995)  by  using  image  analysis.  The 
measuring  path  consisted  of  two  segments:  a  ventral  path 
from  the  core  to  the  21st  increment  and  a  ventral-medial 
path  along  the  sulcus,  from  the  21st  increment  to  the  edge 
(Fig.  2).  The  21st  increment  was  selected  as  the  transition 
point  in  these  measuring  paths  by  test  reading  30  otolith 
sections  representing  the  entire  range  of  sample  fish 


Powell  et  al.:  Growth,  mortality,  and  hatchdate  distributions  for  Cynoscion  nebulosus 


145 


Distal  Edae 


Ventral 


Counting  Path  2 
(21  days  to  capture) 


Figure  2 

Transverse  polished  section  of  a  spotted  seatrout  ( Cynoscion  nebulosus)  ( 18  mm  SL;  age  48  days)  otolith 
showing  the  counting  paths. 


lengths.  In  all  samples,  the  21st  increment  could  easily  be 
traced  in  both  measuring  paths  and  in  all  samples  the  first 
21  increments  could  be  measured  within  the  same  image. 
Increment  widths  were  averaged  over  a  7-day  period.  Age 
estimates  were  also  obtained  and  we  eliminated  any  oto- 
lith used  to  measure  increment  widths  if  the  difference  in 
total  increment  count  between  the  two  methods  ( counts  ob- 
tained directly  from  the  microscope  versus  those  attained 
by  image  analysis)  was  greater  than  7  days  or  10%.  On 
this  basis,  117  otoliths  were  removed  from  the  increment 
width  analysis. 

We  believed  counts  obtained  directly  from  the  microscope 
were  more  accurate  than  those  obtained  by  summing  the 
number  of  increments  measured  on  the  computer  moni- 
tor with  the  image  analysis  system.  Counting  increments 
directly  through  the  microscope  lens  allows  the  reader  to 
optically  section  the  otolith  (by  varying  the  focus),  which 
helps  in  detecting  daily  increments.  "Frozen"  multiple  im- 
ages are  a  result  of  using  the  image  analysis;  hence  optical 
sectioning  is  not  possible. 

Data  analysis  Data  from  all  years  and  sources  were  used 
for  1)  overall  growth  (i.e.  larval  and  juvenile);  2)  juvenile 
growth;  and  3)  estimates  of  juvenile  mortality.  Data  from 
NOAA  larval  and  juvenile  collections  were  used  to  estimate 
a  body-length-otolith-radius  relationship.  Data  from  1995 
FMRI  and  NOAA  collections,  which  was  the  most  com- 
plete data  set,  were  used  for  growth  comparisons  between 
cohorts,  and  hatchdate  distributions.  Data  from  1995  FMRI 
collections  were  used  for  1)  growth  comparisons  between 
geographical  subdivisions;  2)  estimating  a  wet-weight-age 
relationship  to  compute  the  ratio  of  wet-weight  specific- 


growth  to  mortality  (G:M  ratios),  which  assesses  the  rela- 
tive recruitment  potential  of  individual  cohorts  (Houde, 
1996;  Rilling  and  Houde,  1999;  Rooker  et  al.,  1999);  and  3) 
determining  the  influence  of  temperature  on  otolith  incre- 
ment width.  We  used  the  FMRI  data  set  exclusively  for 
the  above  analyzes  because  collections  were  spatially  more 
localized  and  wet  weights  were  available. 

Natural  mortality  (M)  estimates  were  derived  by  regress- 
ing log.  unadjusted  numbers  on  age  classes  (5-day  bins); 
the  resulting  slope  provided  an  estimate  of  total  mortality 
(Ricker,  1975).  However,  on  the  basis  of  the  age-frequency 
distributions  (Fig.  3),  we  considered  juveniles  a40  days  old 
fully  recruited  to  our  gear  and  juveniles  >90  days  old  ap- 
peared to  avoid  our  gear.  Hence,  only  juveniles  between  40 
and  90  days  old  were  used  to  calculate  mortality. 

Hatchdate  distributions  were  computed  on  a  weekly  ba- 
sis and  adjustments  for  mortality  were  made  on  individual 
juveniles  by  the  equation 

N  =N,/e-z<, 


where  N0  =  estimated  number  at  hatching; 

Nt  =  number  at  time  t  (Nt=l  because  N0  was  calcu- 
lated for  each  individual  fish); 

Z  =  instantaneous  daily  mortality  coefficient; 
and 

t      =  age  in  days. 

Spotted  seatrout  cohorts  were  divided  into  weekly  units, 
but  comparisons  between  cohort  growth  was  done  on  a 
monthly  basis  because  of  inadequate  numbers  for  weekly 
comparisons.  A  test  of  heterogeneity  of  slopes  was  imple- 


146 


Fishery  Bulletin  102(1) 


CJ 

^f     •*    -a-     Tt 

n     't    m    to     n 

CO 

■5f 

o 
CM 

o     o    o    o     o 
co     -^-     in    co     r*~ 

Age  class  (days) 

o 

CO 

o 

CD 

Figure  3 

Frequency  distribution  of  spotted  seatrout 
{Cynoscion  nebulosus)  age  classes  used  in  deter- 
mining minimum  age  at  full  recruitment  to  the 
sampling  gear,  and  mortality. 


32 
30 

I      28 

CO 

ra      26 
a> 

Q. 

E 


CO 


merited  by  using  a  generalized  linear  model 
(SAS/STAT  software,  version  6.12,  SAS  Insti- 
tute, Cary,  NO  to  test  if  growth  differed  among 
cohorts.  A  general  linear  test  ( Neter  et  al.,  1983 ) 
was  used  to  compare  growth  between  three  geo- 
graphical subdivisions  (Gulf  transition,  western, 
and  central).  This  test  is  a  function  of  the  error 
sum  of  squares  of  the  reduced  model  minus  the 
error  sum  of  squares  of  the  full  model.  Adequate 
numbers  of  juveniles  were  not  available  to  com- 
pare growth  in  eastern  and  northern  subdivi- 
sions (Table  1).  Circular  statistics  (Batschelet, 
1981)  were  used  to  determine  if  spawning,  as 
determined  from  hatchdate  distributions,  was 
uniform  over  the  lunar  month.  The  phase  of  the 
moon  for  1995  was  identified  by  the  fraction  il- 
luminated (U.  S.  Naval  Observatory  Applications 
Department,  1997).  A  3-point  moving  average 
was  used  to  test  if  spawning  was  cyclical. 

Cohorts  (1995)  were  categorized  according 
to  the  following  hatchdates:  cohort  A,  29  March-2  May 
("April");  cohort  B,  3  May-6  June  ("May");  cohort  C,  7 
June-4  July  ("June");  cohort  D,  5  July-1  August  ("July"); 
cohort  E,  2  August-5  September  ("August");  cohort  F,  6 
September— 3  October  ("September"). 

Comparisons  of  the  relative  recruitment  potential  of 
individual  cohorts  (G:M  ratios)  between  all  cohorts  were 
unresolved.  Although  cohort  mortality  estimates  could 
be  generated,  they  were  appropriate  (by  analyzing  r2  and 
P-values  from  regression  analysis)  for  only  three  cohorts 
(cohorts  B,  D,  and  F). 

A  random  coefficient  model  was  used  to  investigate  the 
relationship  between  growth  rate  of  otoliths  with  age  and 


24 


22 


20  - 


(I 

O 

II 


4U  - 

35  - 

30  - 

< 

,              * 

\ 

i 

\ 

I 

25  - 

20  - 

15  - 

10  - 

< 

t 

5  - 

« 

» 

Station 


Figure  4 

Mean  and  ranges  of  temperature  and  salinity  data  by  station  used  in 
the  otolith  microstructure  longitudinal  analysis  (relationship  between 
increment  width  and  temperature  and  salinity).  For  station  locations 
relative  to  subdivisions,  see  Table  1  and  Figure  1. 


temperature  from  juveniles  collected  in  1995.  Most  fish 
were  exposed  to  salinities  in  a  narrow  range  between  28 
and  34  ppt;  only  9  fish  were  exposed  to  salinities  in  the  5-13 
ppt  range  (Fig.  4).  Consequently,  there  was  insufficient  in- 
formation to  obtain  reliable  inferences  on  the  relationship 
of  growth  rate  to  salinity  or  the  relationship  to  salinity 
and  temperature  for  growth  information  obtained  by  using 
either  otolith  measuring  path.  This  was  a  disappointment 
because  growth  responses  to  salinity  were  considered  an 
important  objective  in  relation  to  proposed  Everglades 
water  management  activities.  Thus,  investigation  was 
restricted  to  the  relationship  of  growth  with  temperature. 
A  separate  model  was  fitted  for  the  first  ( 1-21  increments) 


Powell  et  al.:  Growth,  mortality,  and  hatchdate  distributions  for  Cynoscion  nebulosus 


147 


and  second  (22-60  increments)  measuring  paths  because 
otolith  increment  width  changed  at  a  constant  (age-inde- 
pendent) rate  for  each  path.  We  did  not  include  fish  with 
>60  increments  because  the  relationship  past  this  number 
was  determined  for  only  10%  of  the  fish  and  included  obvi- 
ous outliers.  Letting  Y  be  the  otolith  width  measurement 
for  fish  ;  at  age  a  ,  where  y  indexes  time,  the  model  for  each 
path  was 

Y,j  =  «o,  +  «i 

where  a0l  and  «1;  are  the  fish-specific  intercept  and  slope 
describing  the  relationship  between  increment  width  and 
age  for  fish  i,  and  e  is  a  normally  distributed  error  term; 
thus,  ah  is  the  growth  rate  for  fish  i  over  the  measuring 
path.  Temperature  exhibited  only  negligible  change  for  any 
given  fish  over  the  measuring  path;  thus,  temperature  for 
fish  i  was  summarized  as  tr  the  average  temperature  over 
the  path  for  that  fish.  To  determine  an  appropriate  model 
for  the  relationship  between  intercept  and  growth  rate 
and  temperature,  a  preliminary  analysis  was  performed  in 
which  ordinary  least  squares  estimates  of  «0;  and  ah  were 
obtained  separately  for  each  fish  i  and  plotted  against  tem- 
perature. For  the  first  measuring  path  (1-21  increments), 
the  appropriate  model  was 

«o,  =  A)0  +  0oi'i  +  bor  «i,  =  Pw  +  Put,  +  P\4?  +  bii> 

where  b0l  and  bh  are  normally  distributed  random  effects, 
allowing  growth  rates  for  fish  at  the  same  temperature  to 
vary  across  fish.  For  the  second  measuring  path  (22-60 
increments),  the  appropriate  model  was, 

«o,  =  Poo  +  /V,  +  Po-i'r  +  V  «ii  =  Pw  +  0ii'i  +  Put?  +  bu- 

By  substitution,  these  considerations  yielded  models  1  and 
2  for  the  first  and  second  paths,  respectively; 


Y„  =  {Poo  +  fVP  +  Cfto  +  011*1  +  012*^  a„  +  bo 

blpii+eii 

(•Pw  +  Pnl,  +  012^  aa  +  bo,  +  bi,a„  +  e,j 


(1) 


(2) 


thus  representing  otolith  increment  width  in  each  case 
as  having  a  straight  line  relationship  with  age,  where  the 
slope  (age-independent  growth  rate)  depends  on  average 
temperature  according  to  a  quadratic  relationship.  The 
random  effects  allow  observations  on  the  same  fish  to 
be  correlated,  whereas  observations  across  fish  are  inde- 
pendent. Models  1  and  2  were  implemented  in  SAS  Proc 
Mixed  (SAS/STAT  software,  version  6.12,  SAS  Institute, 
Cary,NC). 

Daily  temperature  records  were  obtained  from  the  Unit- 
ed States  Department  of  Interiors  National  Park  Service, 
Florida  Bay  monitoring  stations  and  averaged  over  a  7-day 
period.  In  1995,  temperature  records  were  available  only 
for  Johnson  Key  Basin  ( JKB),  Whipray  Basin  (WB),  Little 
Blackwater  Sound  (LBS),  and  Little  Madeira  Bay  (LMB), 
but  spotted  seatrout  were  also  collected  at  other  sites 


(Table  1).  Daily  temperatures  were  estimated  for  Sandy 
Key  (SK)  and  Roscoe  Keys  (RK)  from  values  recorded  dur- 
ing sampling  trips  because  both  these  stations  are  not  in 
close  proximity  to  National  Park  Service  monitoring  sites. 
Sandy  Key  values  were  regressed  on  JKB  values  (same 
dates).  Sandy  Key  temperatures  were  collected  from  Janu- 
ary 1994  through  August  1996.  The  regression  model  for 
temperature  was  SK  =  0.76  +  0.9536  JKB  [r2=0.89;  w=25], 
Roscoe  Key  values  were  regressed  on  WB  values  (same 
dates).  Roscoe  Key  temperatures  were  collected  from  Janu- 
ary 1994  through  August  1996.  The  regression  model  for 
temperature  was  RK  =  5.60  +  0.7976  WB  [r2=0.87;  n=31). 
Temperature  values  were  available  at  Murray  Key  (MK)  in 
1997.  To  attain  values  for  our  1995  analysis  we  regressed 
MK  on  JKB  (same  dates).  The  temperature  regression 
model  was  MK  =  0.77  +  0.9680  JKB  [r2=0.99;  re=342]. 

We  reported  measurements  in  standard  length  (SL).  For 
preflexion  and  flexion  larvae,  standard  length  was  mea- 
sured from  the  tip  of  the  snout  to  the  tip  of  the  notochord. 
For  postflexion  larvae  and  juveniles,  standard  length  was 
measured  from  the  tip  of  the  snout  to  the  base  of  the  hy- 
pural  plate. 


Results 

Overall  growth  of  larvae  and  juveniles  (<80  mm  SL)  was 
best  described  by  the  equation  log,  standar-d  length  = 
-1.31  +  1.2162  (loge  age)  [«=582;  r2=0.97].  Growth  in  body 
length  of  juveniles  (12-80  mm  SL)  was  best  described  by 
the  linear  equation  standard  length  =  -7.50  +  0.8417  {age) 
[n=486;  /-2=0.84];  hence,  juveniles  between  approximately 
age  20-100  days  grew  on  average  0.84  mni/d.  There  were 
no  significant  differences  in  juvenile  growth  in  body  length 
among  three  geographical  subdivisions  [F*327=0.756; 
n=333]  (Table  2),  but  there  was  a  significant  growth  differ- 
ence in  length  for  one  of  six  1995  cohorts  (Table  3,  Fig.  5). 
Growth  in  wet  weight  of  juveniles  ( 15-69  mm  SL)  was  best 
described  by  the  equation  log(,  wet  weight  =  -AAA  +  0.0748 
(age)  [n=347,  r2=0.84].  There  was  a  significant  growth  dif- 
ference in  wet  weight  for  one  cohort  (Table  4,  Fig.  6). 

Weekly  1995  hatchdate  distributions,  determined  by  us- 
ing daily  instantaneous  mortality  ( 0.0585.  Fig.  7 ).  indicated 
juveniles  in  collections  (i.e.  survivors)  were  from  spawning 
that  was  cyclical  and  protracted  (Fig.  8).  The  most  intense 
successful  spawning  occurred  during  21-27  June  (9.2%  of 
total).  Using  a  3-point  moving  average,  we  observed  three 
similar  cycles  (Fig.  8).  From  data  on  survivors,  -25%  of  ju- 
veniles were  spawned  by  late  May,  50%  by  early  July,  and 
75%  by  late  August  and  from  data  on  cohorts,  three  cohorts 
(cohorts  C,  D,  and  E;  early  June-late  August)  comprised 
55%  of  the  total  estimated  spawn  of  spotted  seatrout.  There 
was  no  correlation  between  spawning  and  moon  phase  (pe- 
riodic regression  r2=0.019,  P=0.754)  (Fig.  8). 

The  relative  recruitment  potential  (G:M  ratio)  of  the  1995 
year  class  estimated  from  the  wet-weight  specific  growth 
coefficient  (0.0748)  and  the  instantaneous  daily  mortal- 
ity rate  (0.0585,  Fig.  7)  was  1.28.  The  G:M  ratio  for  three 
cohorts  (B,  May;  D,  July;  and  F,  September)  was  greater 
than  the  ratio  for  the  total  1995  year  class  because  mortal- 


148 


Fishery  Bulletin  102(11 


Table  2 

Summary  of  growth  data  used  to  compare 
best  described  by  the  linear  equation:  stan 

growth  ir 
dard  leng 

length  of  spotted 
th  =  a  +  b  ( age  in 

seatrout  among 
days). 

three  Florida 

Bay  subc 

ivisions.  Growth  was 

Subdivision 

Intercept 

Slope 

n 

r- 

Size  range  (mm  SD 

Gulf  transition 

-11.07 

0.8914 

139 

0.86 

16-69 

Central 

-12.23 

0.9298 

49 

0.80 

15-63 

Western 

-10.56 

0.8834 

145 

0.85 

17-69 

Table  3 

Summary  of  statistics  for  a  test  for  heterogeneity  of  slopes  for  cohort  somatic  growth  rates 
rized  according  to  month  of  hatchdate  (see  text).  The  base  parameter  is  cohort  F  and  all 
the  base  cohort.  For  growth  equations,  see  Figure  5. 

of  spotted  seatrout.  Cohorts  were  catego- 
parameter  estimates  are  deviations  from 

Parameter 

Estimate 

Standard  error 

/-value 

P-value 

Intercept 

-7.97270 

3.02829 

-2.63 

0.0088 

Cohort  A 

-0.86811 

4.31970 

-0.20 

0.8408 

Cohort  B 

-11.85849 

3.91387 

-3.03 

0.0026 

Cohort  C 

1.65094 

3.86470 

0.43 

0.6695 

Cohort  D 

-6.70820 

4.74936 

-1.41 

0.1586 

Cohort  E 

1.02077 

3.50931 

0.29 

0.7713 

Slope 

0.82088 

0.05613 

14.62 

<0.001 

Cohort  A 

0.04054 

0.08604 

0.47 

0.6378 

Cohort  B 

0.24578 

0.07106 

3.46 

0.0006 

Cohort  C 

-0.00113 

0.07091 

-0.02 

0.9873 

Cohort  D 

0.15741 

0.08730 

1.80 

0.0721 

Cohort  E 

0.04544 

0.06821 

0.67 

0.5058 

Table  4 

Summary  of  statistics 

for  a 

test  for  heterogeneity  of 

slopes  for  cohort  wet-weight 

growth  rate 

of  spotted 

seatrout. 

Cohorts  were 

categorized  according 

,o  month  of  hatch  date  (see  text ).  The 

base  parameter  is  cohort  F  and 

all  parameter 

estimates  are  deviations 

from  the  base  cohort.  For  growth  equations,  see 

Figure  6. 

Parameter 

Estimate 

Standard  error 

/-value 

P-value 

Intercept 

-4.27384 

0.23763 

-17.99 

<0.0001 

Cohort  A 

0.10201 

0.37014 

0.28 

0.7830 

Cohort  B 

-0.46348 

0.34092 

-1.36 

0.1749 

Cohort  C 

-0.27866 

0.32116 

-0.87 

0.3862 

Cohort  D 

-0.19540 

0.38352 

-0.51 

0.6108 

Cohort  E 

-0.38260 

0.29853 

-1.28 

0.2009 

Slope 

0.06974 

0.00439 

15.88 

<0.0001 

Cohort  A 

0.00195 

0.00729 

0.27 

0.7889 

Cohort  B 

0.00679 

0.00622 

1.09 

0.2754 

Cohort  C 

0.00502 

0.00575 

0.87 

0.3835 

Cohort  1) 

0.00759 

0.00702 

1.08 

0.2808 

Cohort  E 

0.01188 

0.00564 

2. 11 

0.0359 

Powell  et  al.:  Growth,  mortality,  and  hatchdate  distributions  for  Cynoscion  nebulosus 


149 


Cohort  A 

Length  =  -8.84+0  8614  (Age) 

60 

40 

20 

0 


20 


60 


Cohort  B 

Length  =  -19.83  +  1.0667  (Age) 

F 

B0 

^  =  0.80 

F 

n  =  57 

<9 

en 

60 

o 

'O 

a> 

40 

%    o<g> 

•n 

nftri'™ 

20 

--?-: 

c 

m 

rn 

40 


80 


100 


Cohort  C 
i  Length  =  -6.32  +  0.8198  (Age) 
?  =  0.90 

B0  |  n  =  55 

40 

20 

0 


80 
60 
40 
20 
0 


80 
60 
40 
20 
0 

80 
60 
40 
20 
0 


Cohort  D 

Length  =  -14  68  +  0.9783  (Age) 

r2  =  0.80 

n  =  69 


Cohort  E 

Length  =  -6.95  +  0.8663  (Age) 

^  =  0.89 

n  =99 


20 


40 


60 


Cohort  F 

Length  =  -7.97  +  0.8209  (Age) 

r  =  0.86 

n  =  50 


Age  (days) 

Figure  5 

Comparison  of  growth  in  standard  length  among  six  spotted  seatrout  (Cynoscion 
nebulosus)  cohorts  collected  in  1995.  See  text  for  cohort  hatchdates. 


Table  5 

Daily  gr 

owth  (wet  weight  in  grams) 

rates  and  daily  mort 

ality  rates  for  three  cohorts  in  Florida  Bay  in  1995.  Cohorts 

were 

cate- 

gorized  according  to  month  of  hatchdate  (see  text).  The  G:M  ratio  derived  from  the  growth  and  mortality 

rates  is  also 

presented. 

For  growth  equations  and  associated 

r-  values,  see  Figure 

6. 

Cohort 

Hatchdate  month 

Growth  rate 

Mortality  rate 

r2 

G:M  ratio 

Size  range  (mm  SL) 

B 

May 

0.0765 

0.0445 

0.54 

1.72 

28- 

-62 

D 

July 

0.0773 

0.0565 

0.82 

1.37 

37- 

-68 

F 

September 

0.0697 

0.0354 

0.67 

1.97 

37- 

-66 

ity  rates  appeared  relatively  low  compared  to  the  overall 
mortality  rate  (0.0585)  for  juveniles  (Table  5).  However, 
differences  in  mortality  rates  among  these  three  cohorts 
were  not  significant  (F4.;i=1.414).  There  were  no  significant 
differences  in  weight-specific  coefficients  among  the  three 
cohorts  (B,  D,  and  F)  (Table  4),  but  a  significant  difference 
in  length-specific  coefficients  among  the  three  cohorts  was 
found  (Table  3).  Cohort  B  (May)  had  a  significantly  higher 
growth  rate  than  the  other  two  cohorts. 


There  was  a  close  relationship  between  otolith  radius  and 
body  length  (Fig.  9).  A  linear  equation  with  the  sagittal  ven- 
tral radius,  had  a  similar  r2  as  a  curvilinear  equation  with 
the  sagittal  dorsal  radius.  However,  we  were  unable  to  mea- 
sure increment  widths  along  this  plane  and  instead  used  a 
combination  of  a  ventral  path  and  a  ventral  medial  path. 

As  an  initial  demonstration  that  otolith  increment  width 
increased  with  age  along  the  1-21  increment  measuring 
path  and  decreased  along  the  22-60  increment  path,  simpli- 


150 


Fishery  Bulletin  102(1) 


2  ! 

1 
o 
-1 

-2 
-3 


Cohort  A 
Loge  weight  = 


-4.17  +  0-0717  (age) 


20 


40 

Cohort  B 
Loge  weight 

z2  =  0.88 

n  =  47 


60 


80 


100 


:  -4.74  +  0.0765  (age) 


20 


40 


60 


80 


100 


Cohort  C 

Loge  weight  =  -4.55  +  0.0748  (age) 


^  =  0.94 

n  =  47 

JM 

o^°^ 

o 

20 


40 


60 


80 


100 


Cohort  D 

Log  B  weight  =  -4.47  +  0.0773  (age) 

^  =  0.81 
n  =  61 


20 


40 


60 


80 


100 


3 
2 
1 

0 
-1 
•2 
-3 
-4 
20 


Cohort  E 

Loge  weight  =  ^1.66  +  0.0816  (age) 

?  =  0.85 
n  =  66 


40  60  80  100 

Cohort  F 

Log  e  weight  =  -4.27  +  0.0697  (age) 

f2  =  0.86 
n  =  47 


20 


40 


60 


80 


100 


Age  (days) 


Figure  6 

Comparison  of  growth  in  wet-weight  (grams)  among  six  spotted  seatrout  \Cynoscion 
nebulosus)  cohorts  collected  in  1995.  See  text  for  cohort  hatchdates. 


fieri  versions  of  Equations  1  and  2  ( see  above )  were  fitted,  in 
which  all  coefficients  of  temperature  were  set  equal  to  zero, 
so  that  Equations  1  and  2  represent  simple  linear  relation- 
ships with  age.  For  the  first  path,  the  estimate  of  slope  was 
0.153  fjm/d  (P<0.0001);  that  for  the  second  path  was 
-0.065  fim/d  (P<0.0001>.  Addition  of  quadratic  terms  to 
each  model  was  not  supported  (P=0.81  and  0.12,  respec- 
tively). For  the  first  path,  whether  intercept  or  growth  rate 
were  associated  with  temperature  was  determined  by  test- 
ing whether  the  parameters  j301,  /3n,  and  j312  were  equal  to 
zero.  There  was  no  evidence  that  any  of  these  parameters 
were  different  from  zero  (P=0.45,  0.35,  and  0.42,  respec- 
tively); the  latter  two  may  indicate  that  the  data  do  not 
support  the  contention  that  growth  rate  depends  on  tem- 
perature in  this  range  (1-21  d).  For  the  second  path,  tests 
"I  /'„  ,=0  and  /i12=0  offered  strong  evidence  that  these  pa- 
rameters are  different  from  zero  (P<0.001  in  each  case).  In 
particular,  these  results  suggested  for  the  age  range  22-60 
d,  otolith  growth  rates  decrease.  The  extent  of  the  decrease 
is  strongly  associated  with  average  temperature  according 
to  a  quadratic  relationship  such  that  growth  rates  were 
more  steeply  decreasing  with  age  for  lower  temperatures 


and  then  became  shallower  at  higher  temperatures.  In 
summary,  for  temperatures  at  the  lower  and  higher  end  of 
the  observed  temperature  range,  otolith  growth  rates  for 
the  age  range  22-60  d  were  higher  than  they  were  in  the 
middle  of  the  observed  temperature  range. 


Discussion 

Growth  in  body  length  of  juvenile  spotted  seatrout  in  Flor- 
ida Bay  was  faster  than  growth  of  juveniles  from  Tampa 
Bay  (Table  6,  McMichael  and  Peters,  1989).  Florida  Bay  is 
generally  considered  an  oligotrophic  system  (Fourqurean 
and  Robblee,  1999).  Nevertheless,  seagrass  beds  in  west- 
ern Florida  Bay,  where  juvenile  spotted  seatrout  are  most 
common  (Chester  and  Thayer,  1990),  are  significantly 
more  dense  than  beds  in  northwestern  Florida  waters, 
slightly  north  of  Tampa  Bay  (Iverson  and  Bittaker,  1986). 
Increased  growth  of  juveniles  in  Florida  Bay  could  be 
attributed  to  the  dense  seagrass  beds  that  provide  habitat 
for  epifaunal  crustaceans  (Holmquist  et  al.,  1989;  Mathe- 
son  et  al.,  1999),  which  are  important  in  the  diet  of  juve- 


Powell  et  al.:  Growth,  mortality,  and  hatchdate  distributions  for  Cynosaon  nebulosus 


151 


5.0  - 

Log,  abundance  =  6.83  -0.0585  (age) 

^  =  094 

4.5  - 

n           n=  10 

4.0  - 

^^^\ 

8     3.5- 

c 

CO 
T3 

§     3.0  - 
.a 

CO 

Cu 

g1     2.5  - 

_j 

^^\.   0 

2.0  - 

\ 

1.5  - 

o 

^f                                 ^f                                 -^                                 Tf                                 r}- 

■3-                    in                    cd                    h~                    co 

o                   o                   o                  o                   o 
^f                    in                    cd                    r^.                    co 

Age  class  (days) 

Figure  7 

Catch  curve  of  juvenile  spotted  seatrout  (Cynoscion  nebulosus)  used  to 

estimate  daily  instantaneous  mortality  (Z).  Z  =  slope  =  -0.0585.  Spotted 

seatrout  were  fully  recruited  to  the  gear  at  age  40-44  days. 

Comparison  of  spotted  seatrout  growth  (size 
Florida  Bay,  Florida  (this  study). 

Table  6 

in  mm  SL  at  age)  betw 

;en  Tampa  Bay. 

Florida  (McMichaels  and  Peters, 

1989) and 

Age  (days) 

Area                                10                   20 

30                    40 

50 

60                    70 

80 

90 

Tampa  Bay                    5.1                 10.2 
Florida  Bay                   4.4                 10.3 

15.3                 20.3 
16.9                23.3 

25.4 
31.4 

30.5                35.6 
39.2               47.2 

40.7 
55.6 

45.8 
64.1 

nile  spotted  seatrout  (Hettler,  1989;  McMichael  and  Peters, 
1989).  Additionally,  warmer  water  temperatures  have  been 
observed  in  Florida  Bay  (Boyer  et  al.,  1999)  compared  to 
Tampa  Bay  (McMichael  and  Peters,  1989);  these  warmer 
temperatures  could  enhance  growth  if  adequate  food  is 
available  (Warren,  1971).  However,  our  study  and  that  of 
McMichael  and  Peters  ( 1989)  were  quite  a  few  years  apart; 
hence  differences  that  we  observed  could  also  be  accounted 
for  by  interannual  variability.  In  addition,  differences  in 
growth  could  also  be  attributed  to  differences  in  sampling 
gear  between  the  two  studies. 

Florida  Bay  is  a  heterogenous  ecosystem  and  consists 
of  ecologically  distinct  regions  (Phlips  and  Badylak,  1996; 
Fourqurean  and  Robblee,  1999);  however,  we  did  not  de- 
tect any  differences  in  growth  of  juvenile  spotted  seatrout 
among  our  three  subdivisions.  In  general,  juvenile  collec- 
tions from  the  central  subdivision  were  from  stations  that 
were  spatially  dispersed;  whereas,  juvenile  collections  in 


the  western  and  Gulf  transition  subdivisions  were  from 
relatively  few  stations  (Table  1 ).  Normally,  the  central  sub- 
division is  characterized  by  the  highest  salinities  in  the  bay 
and  the  western  and  the  Gulf  transition  are  characterized 
by  high  salinities  (Orlando  et  al.,  1997).  However,  in  our 
study,  salinities  in  the  three  subdivisions  were  moderate 
and  similar  (Fig  4).  and  growth  rates  estimated  for  the 
three  subdivisions  could  be  useful  as  baseline  rates,  par- 
ticularly in  the  central  subdivision  where  salinities  are 
commonly  hyperhaline  (Orlando  et  al.,  1997). 

The  spawning  habits  of  spotted  seatrout  throughout 
their  entire  range  are  generally  similar.  They  have  a 
protracted  spawning  season,  are  multiple  spawners,  and 
reach  sexual  maturity  at  an  early  age.  Initiation  of  spawn- 
ing might  be  temperature  dependent,  with  water  tempera- 
tures between  20°  and  23°C  necessary  to  initiate  repro- 
ductive development  (Brown-Peterson  and  Warren,  2001). 
Hatchdate  distributions  calculated  for  spotted  seatrout  in 


152 


Fishery  Bulletin  102(1) 


Birthweek 

Figure  8 

( Al  Spotted  seatrout  {Cynoscion  nebulosus)  («=417)  weekly 
hatchdate  distributions  adjusted  for  mortality,  including 
moon  phases  (#=new  moon;  0=full  moon),  and  3-point 
moving  average  (solid  line)  of  hatchdate  distributions.  (B) 
Cumulative  frequency  of  spotted  seatrout  (n=417)  hatch- 
date  distributions. 


Florida  Bay  in  this  study  along  with  early  stage  larval 
collections  (Powell,  2003)  indicate  that  spotted  seatrout 
spawn  between  March  and  October  (based  on  hatchdate 
distributions)  and  that  the  majority  of  spawning  occurs 
between  27°  and  35°C  ,  with  very  little  spawning  between 
20°  and  26°C  (based  on  early  stage  larval  collections). 
Spawning  peaks,  based  on  larval  collections  in  1994-96, 
occurred  in  June,  August,  and  September  (Powell,  2003), 
and  early  May,  late  June,  and  late  August  through  early 
September  based  on  1995  hatchdate  distributions  (this 
study).  However,  Stewart  (19611  reported  that  spotted 
seatrout  in  Florida  Bay  spawned  throughout  the  year  and 
that  spawning  peaked  in  spring  and  fall.  Another  larval 
fish  study  in  Florida  Bay  indicated  that  some  spawning 
occurs  as  early  as  February  and  continues  into  December 
(Rutherford  et  al.,  1989). 


Loge  Body  length  =  -1 .64  +  0.7821  (dorsal  radius) 

?  =  0.99 
n  =  232 


80  -i 


60 


40 


20  - 


E 

1 ' 1 ' 1 ' 1 ' ' 1 

B>     0    200   400   600   800   1000  1200 


Sagittal  dorsal  radius  (microns) 


£     80 

CO 

(7) 


au  -i 

Body  length  = 

=  0.75  +  0.0503  (ventral  radius) 

?  =  0.99 

60  - 

n  =  232 

rffO 

40  - 

CM§^> 

QqAO 

20  - 

n- 

kd 

I       '       I       '       I       '       I       '       I 

0         200       400       600       800      1000     1200 
Sagittal  ventral  radius  (microns) 

Figure  9 

The  relationship  between  sagittal  otolith  radius 
and  standard  length  (top),  and  sagittal  ventral 
radius  and  standard  length  (bottom)  for  spotted 
seatrout  (.Cynoscion  nebulosus). 


Peak  spawning  activity  of  spotted  seatrout  is  highly 
variable  (McMichael  and  Peters,  1989;  Brown-Peterson  and 
Warren,  2001).  McMichael  and  Peters  (1989)  observed  two 
spawning  peaks;  spring  and  summer.  Older  fish  participate 
in  two  peak  spawning  periods  ( Tucker  and  Faulkner,  1987 ), 
and  a  portion  of  the  larger  spring-spawned  fish  (age- 1+)  en- 
ter the  spawning  population  during  their  second  summer, 
augmenting  the  number  of  summer  spawning  fish. 

We  found  that  spawning  activity  and  moon  phase  were 
uncorrelated,  which  is  not  in  concordance  with  observations 
of  McMichael  and  Peters  (1989).  They  found  that  distinct 
peaks  in  spawning  (based  on  hatchdate  distributions  of  lar- 
val spotted  seatrout)  occurred  at  monthly  intervals,  and  this 
periodicity  might  coincide  with  moon  phase.  However,  this 
monthly  periodicity  was  not  observed  when  their  data  for 
juvenile  spotted  seatrout  were  examined.  Moreover,  statisti- 
cal tests  were  not  performed  on  the  data  in  their  study. 


Powell  et  al.:  Growth,  mortality,  and  hatchdate  distributions  for  Cynoscion  nebulosus 


153 


Our  inferences,  from  this  study,  in  relation  to  spotted 
seatrout  peak  spawning  are  based  on  hatchdate  distribu- 
tions and  should  be  viewed  with  caution  because  hatch- 
dates  are  based  on  survivors.  Differential  survival  for  early 
life  history  stages  can  bias  results.  Hatchdate  distributions 
are  valuable  when  compared  to  egg  or  recently  hatched  lar- 
val densities  and  might  suggest  processes  responsible  for 
differential  cohort  survivorship.  Because  spotted  seatrout 
undergo  a  protracted  spawning  period  and  because  there  is 
high  variation  associated  with  icthyoplankton  samples  ( Cyr 
et  al.,  1992),  intensive  and  extensive  sampling  of  recently 
hatched  larvae  would  be  required  over  a  long  duration  to 
answer  these  process-oriented  mortality  questions. 

The  daily  instantaneous  mortality  rate  of  juvenile  spot- 
ted seatrout  was  higher  in  Florida  Bay  than  those  reported 
from  northwestern  Florida  systems  (Nelson  and  Leffler, 
2001).  Mortality  rates  of  juvenile  spotted  seatrout  from 
Florida  Bay  were  5.7%/d;  whereas,  for  the  other  systems, 
rates  approximated  3%/d.  In  general,  mortality  rates  might 
increase  with  increasing  estuarine  temperatures  (Houde 
and  Zastrow,  1993).  Although  we  were  unable  to  estimate 
instantaneous  daily  mortality  rates  for  larval  spotted  seat- 
rout, these  data  have  been  estimated  for  larvae  (3.5-6.5 
mm)  in  two  southwestern  Florida  estuaries  (Peebles  and 
Tolly,  1988).  Highly  variable  rates  were  reported  between 
the  two  Florida  estuaries  (Naples  Bay:  0.70  or  50%/d;  and 
Fakahatchee  area:  0.37  or  31%/d).  Houde  ( 1996)  reported  a 
generalized  instantaneous  daily  mortality  rate  for  marine 
fish  larvae  of  0.239  (21%/d).  Estimating  mortality  rates  for 
larval  spotted  seatrout  in  Florida  Bay  will  be  critical  for 
calculating  G:M  ratios  in  order  to  evaluate  stage-specific 
survival  and  to  develop  credible  spatially  explicit  models. 

Mortality  rates  of  spotted  seatrout  cohorts  could  be  cal- 
culated for  only  three  of  six  cohorts  (B,  May;  D,  July;  and 
F,  September)  because  slopes  were  significantly  different 
from  zero  for  only  these  cohorts.  Furthermore,  mortality 
rates  of  two  of  the  three  cohorts  (B  and  F)  were  associated 
with  low  r2  values  (Table  5);  hence  the  G:M  ratios  along 
with  the  mortality  rates  for  these  three  cohorts  should 
be  considered  "rough"  estimates.  Attaining  more  accurate 
mortality  estimates  for  spotted  seatrout  would  be  valuable 
in  linking  cohort  variability  with  potential  recruitment  and 
stage-specific  survival.  For  example,  larval  cohorts  of  bay 
anchovy  [Anchoa  mitchilli)  from  Chesapeake  Bay,  a  tem- 
perate estuary,  exhibit  growth  rates  that  are  temporally 
variable  and  mortality  rates  that  are  spatially  and  tem- 
porally variable  (Rilling  and  Houde,  1999).  Temperature, 
zooplankton  prey  and  gelatinous  predators  are  believed  to 
influence  growth  and  mortality  rates  of  the  bay  anchovy. 
For  striped  bass  iMorone  saxatilis )  in  a  subestuary  of  Ches- 
apeake Bay,  cohorts  exhibited  highly  variable  seasonal  G:M 
ratios  that  were  strongly  influenced  by  temperature 
(Houde,  1997).  In  a  subtropical  estuary,  cohort-specific 
mortality  rates  for  juvenile  red  drum  varied  temporally; 
early  and  late  season  cohorts  exhibited  the  highest  mortal- 
ity rates,  which  coincided  with  highest  growth  rates  and 
G:M  ratios  for  midseason  cohorts  (Rooker  et  al.,  1999).  We 
agree  with  Houde  ( 1997)  that  future  research  should  focus 
on  the  variability  and  causes  of  variability  in  growth  and 
mortality,  both  of  which  interact  to  determine  stage-spe- 


cific survival.  The  developmental  stage  or  age  where  G:M 
variability  is  greatest,  along  with  the  relationship  of  this 
variability  to  recruitment,  need  to  be  determined  for  spot- 
ted seatrout  in  Florida  Bay.  No  doubt  a  relationship  exists 
between  G:M  ratios  and  recruitment.  Future  research 
should  also  determine  if  cohort  G:M  ratios  and  somatic 
growth  rates  are  seasonally  or  spatially  variable.  If  they 
are,  then  a  limited  spatial  and  temporal  sampling  program 
could  be  designed  to  annually  evaluate  G:M  ratios  at  highly 
variable  stages  or  ages  as  an  index  of  year-class  strength 
of  spotted  seatrout  in  Florida  Bay.  Such  an  index  could  be 
verified  by  examining  year-class  catch  rates  on  an  annual 
basis  or  by  virtual  population  analysis. 

In  our  study  there  was  little  temporal  difference  in 
growth  of  juvenile  spotted  seatrout  cohorts.  Larval  growth 
and  mortality,  which  was  not  treated  adequately  in  our 
study,  could  be  influenced  by  copepod  prey — an  important 
dietary  component  of  larval  spotted  seatrout  (McMichael 
and  Peters,  1989).  The  copepod  Acartia  tonsa  is  dominant 
in  Florida  Bay,  but  egg  production  rates  for  this  species  are 
low  in  the  bay  compared  to  those  in  other  systems  (Kleppel 
et  al.,  1998).  We  suspect  the  "bottleneck"  to  recruitment  of 
spotted  seatrout  could  occur  during  the  larval  stage.  Hence, 
future  research  should  examine  mortality  and  growth  of 
larval  and  recently  settled  spotted  seatrout;  in  particular 
the  patterns  of  larval  production  potential  (G:M  ratios). 
Research  in  these  areas  should  increase  our  understand- 
ing of  the  degree  of  variability  in  stage-specific  survival 
and  recruitment  of  spotted  seatrout  in  Florida  Bay  (Houde, 
1996). 

For  most  species,  especially  those  with  protracted  spawn- 
ing habits,  it  is  most  informative  to  analyze  cohort  growth 
and  mortality.  For  example,  striped  bass  and  bay  anchovy 
cohorts  in  Chesapeake  Bay  exhibit  highly  variable  growth 
rates,  mortality  rates,  and  stage  durations  (Rutherford 
and  Houde,  1995;  Rilling  and  Houde,  1999).  This  variabil- 
ity could  cause  differential  survival  for  cohorts  and  result 
in  frequency  distributions  of  survivor  hatchdates  that  do 
not  resemble  recently  hatched  larvae  or  egg-production 
frequency  distributions  (e.g.  Crecco  and  Savoy,  1985;  Rice 
etal.,  1987). 

We  are  unable  to  interpret  the  significance  of  the  abso- 
lute value  of  the  G:M  ratio  for  juvenile  spotted  seatrout, 
because  interannual  comparisons  were  not  made,  but  we 
presented  the  ratio  for  future  comparisons.  Generally,  the 
G:M  ratio  is  <1.0  during  the  early  larval  stage,  indicating 
a  decline  in  biomass.  However,  the  G:M  ratio  of  a  cohort 
will  eventually  exceed  1.0  as  a  result  of  a  relative  decline 
in  mortality  as  larvae  grow  (Houde  and  Zastrow,  1993). 
Clearly,  stage  specific  analysis  of  the  spotted  seatrout  from 
egg  through  juvenile  stage  would  have  been  more  informa- 
tive in  determining  when  the  maximum  G:M  ratio  occurs 
(when  cohort  biomass  increases  at  a  maximum  rate)  and 
in  providing  insight  into  stage-specific  dynamics  of  spotted 
seatrout  (Houde,  1997).  A  constraint  of  our  study  was  our 
inability  to  estimate  larval  mortality  rates;  hence  early  life 
history  stage  dynamics  could  not  be  examined. 

Size-selective  mortality  in  the  juvenile  life  history  stages 
can  have  important  consequences  for  recruitment.  Sogard 
( 1997 )  argued  that  "within-cohort  size-selective  mortality" 


154 


Fishery  Bulletin  102(1) 


is  more  evident  in  the  juvenile  stage  than  during  the  egg 
and  larval  stages  when  random  mortality  independent  of 
fish  size  is  more  likely  to  occur  (e.g.  dispersal  of  eggs  and 
larvae  away  from  suitable  nursery  areas).  In  addition,  vari- 
ation in  size,  which  provides  a  "template"  for  size-selective 
processes,  increases  during  the  juvenile  stage  as  larval  size 
is  constrained  by  egg  size.  Sogard  ( 1997)  cited  a  number  of 
recent  studies  that  suggest  the  early  juvenile  period  plays 
a  greater  role  in  determining  year-class  strength  than 
previously  thought. 

We  were  unable  to  determine  if  salinity  influenced  incre- 
ment width  (a  surrogate  for  somatic  growth)  at  early  life 
stages.  Understanding  the  relationship  between  salinity 
and  growth  is  critical  because  Everglades  restoration  will 
most  likely  result  in  increased  freshwater  flows  to  Florida 
Bay,  and  during  low  rainfall  periods,  salinities  in  the  north 
central  portion  of  the  bay  can  exceed  45  ppt  (Orlando  et 
al.,  1997;  Boyer  et  al.,  1999).  But,  salinities  were  moderate 
and  similar  at  most  stations  where  juvenile  trout  were  col- 
lected in  the  bay  during  1995  I  Fig.  4).  Very  few  fish  were 
collected  at  low  salinities;  in  fact,  juvenile  spotted  seatrout 
are  not  commonly  collected  at  low-salinity  stations  (Table 
1;  Florida  Department  of  Environmental  Protection1),  and 
hyperhaline  conditions  were  not  observed  in  1995.  There- 
fore, we  were  only  able  to  determine  if  temperature  could 
influence  increment  widths.  The  curvilinear  relationship 
between  otolith  growth  rate  and  temperature,  although 
a  statistically  strong  relationship,  is  difficult  to  explain 
biologically.  Temperature  could  mask  other  factors,  e.g. 
temporal  variability  in  prey  and  predator  availability,  and 
optimal  temperatures  for  growth  (Rooker  et  al.,  1999).  We 
were  able  to  demonstrate  that  one  cohort  grew  faster  than 
five  other  cohorts,  possibly  indicating  differential  prey 
availability  in  1995.  An  individual-based  bioenergetics 
model  for  spotted  seatrout  now  in  preparation  (Wuenschel 
et  al.2)  should  add  to  our  understanding  of  the  effects  of 
salinity  and  temperature  on  larval  and  juvenile  spotted 
seatrout 


Acknowledgments 

We  are  especially  grateful  to  Al  Crosby,  Mike  Greene,  Mike 
LaCroix,  and  other  Beaufort  staff  that  participated  in  the 
field  work.  We  thank  James  Waters  of  the  NMFS  Southeast 
Fisheries  Science  Center  for  computer  programing  assis- 
tance and  Jon  Hare  of  our  laboratory  for  performing  the 
circular  statistics.  We  are  grateful  to  Dean  Ahrenholz,  Jon 
Hare,  Patti  Marraro,  Joseph  Smith,  and  three  anonymous 
reviewers  for  their  valuable  reviews  of  the  manuscript.  We 
also  thank  Steve  Bobko  at  Old  Dominion  University  for 
the  image  analysis  macro  used  to  obtain  otolith  increment 
widths. 


2  Wuenschel,  M.  J.,  R.  G.  Werner,  D.  E.  Hoss,  and  A.  B.  Powell. 
2001.  Bioenergetics  of  larval  spotted  seatrout  (Cynoscion 
nebulosus)  in  Florida  Bav.  Florida  Bay  Science  Conference, 
April  23-26,  2001,  p.  215-216.  Westen  Beach  Resort,  Key 
Largo,  Florida.  Abstract.  Center  for  Coastal  Fisheries  and 
Habitat  Research,  Beaufort  Laboratory,  101  Pivers  Island  Road, 
Beaufort,  NC  28516. 


Literature  cited 

Anderson,  J.  T. 

1988.  A  review  of  size  dependent  survival  during  pre-recruit 
stages  of  fishes  in  relation  to  recruitment.  J.  Northw.  Atl. 
Fish.  Sci.  8:55-66. 

Batschelet,  E. 

1981.     Circular  statistics  in  biology,  371  p.     Academic  Press, 
New  York,  NY. 
Boyer,  J.  N,  J.  W.  Fourqurean,  and  R.  D.  Jones. 

1999.     Seasonal  and  long-term  trends  in  the  water  quality  of 
Florida  Bay  ( 1989-1997).    Estuaries  22:417-430. 
Brown-Peterson.  N.  J.,  and  J.  W  Warren. 

2001.    The  reproductive  biology  of  spotted  seatrout,  Cyno- 
scion nebulosus,  along  the  Mississippi  Gulf  Coast.     Gulf 
Mexico  Sci.  2001:61-73. 
Chester,  A.  J.,  and  G.  W  Thayer. 

1990.     Distribution  of  spotted  seatrout  iCynoseion  nebu- 
losus) and  gray  snapper  (Lutjanus  griseus)  juveniles  in 
seagrass  habitats  of  western  Florida  Bay.     Bull.  Mar.  Sci. 
46:345-357. 
Crecco,  V.,  and  T  Savoy. 

1985.  Effects  of  biotic  and  abiotic  factors  on  growth  and 
relative  survival  of  young  American  shad,  Alosa  sapidis- 
sima,  in  the  Connecticut  River.  Can.  J.  Fish.  Aquat.  Sci. 
42:  1640-1648. 

Cyr,  H.,  J.  A.  Downing,  S.  Lalonde,  S.  B.  Baines,  and  M.  L.  Price. 

1992.  Sampling  larval  fish:  choice  of  sample  number  and 
size.    Trans.  Am.  Fish.  Soc.  121:356-368. 

Fourqurean,  J.  W.,  and  M.  B.  Robblee. 

1999.     Florida  Bay:  a  history  of  recent  ecological  changes. 
Estuaries  22:345-357. 
Hare,  J.  A.,  and  R.  K.  Cowen. 

1997.     Size,  growth,  development,  and  survival  of  the  plank- 
tonic  larvae  of  Pomatomus  saltatrix  (Pisces:  Pomatomidael. 
Ecology  78:2415-2431. 
Hettler.W  F.,Jr. 

1989.  Food  habits  of  juveniles  of  spotted  seatrout  and 
gray  snapper  in  western  Florida  Bay.  Bull.  Mar.  Sci.  44: 
152-165. 

Hettler,  W.  E,  and  A.  J.  Chester. 

1990.  Temporal  distribution  of  ichthyoplankton  near 
Beaufort  Inlet,  North  Carolina.  Mar.  Ecol.  Prog.  Ser.  68: 
157-168. 

Holmquist,  J.  G„  G.  V.  N.  Powell,  and  S.  M.  Sogard. 

1989.  Decapod  and  stomatopod  communities  of  seagrass- 
covered  mud  banks  in  Florida  Bay:  inter-  and  intra- 
bank  heterogeneity  with  special  reference  to  isolated 
subenvironments.     Bull.  Mar.  Sci.  44:251-262. 

Houde,  E.  D. 

1996.  Evaluating  stage-specific  survival  during  the  early  life 
offish.  In  Survival  strategies  in  early  life  stages  of  marine 
resources  lY.  Watanabe.  Y  Yamashita.  and  Y.  Oozeki,  eds.), 
p.  51-66.     Balkema,  Brookfield.  VT. 

1997.  Patterns  and  consequences  of  selective  processes  in 
teleost  early  life  histories.  In  Early  life  history  and  recruit- 
ment offish  populations  (R.  C.  Chambers  and  E.  A.  Trippel, 
eds.),  p.  173-196.     Chapman  and  Hall,  London. 

Houde,  E.  D.,  and  C.  E.  Zastrow. 

1993.  Ecosystem-and  taxon-specific  dynamic  and  energetic 
properties  of  larval  fish  assemblages.  Bull.  Mar.  Sci.  53: 
290-335. 

Iverson,  R.  L.,  and  H.  F.  Bittaker. 

1986.  Seagrass  distributions  and  abundances  in  eastern 
Gulf  of  Mexico  coastal  waters.  Estuar.  Coast.  Shelf  Sci. 
1986:577-602. 


Powell  et  al.:  Growth,  mortality,  and  hatchdate  distributions  for  Cynosaon  nebulosus 


155 


Kleppel,  G.  S.,  C.  A.  Burkart,  L.  Houchin,  and  C.  Tomas. 

1998.  Egg  production  of  the  copepod  Acartia  tonsa  in  Florida 
Bay  during  summer.  1.  The  roles  of  food  environment  and 
diet.    Estuaries  21:328-339. 

Matheson,  R.  E..  Jr.,  S.  M.  Sogard,  and  K.  A.  Bjorgo. 

1999.  Changes  in  seagrass-associated  fish  and  crustacean 
communities  on  Florida  Bay  mud  banks:  the  effects  of 
recent  ecosystem  changes?     Estuaries  22:534-551. 

McMichael,  R.  H.,  Jr.,  and  K.  M.  Peters. 

1989.     Early  life  history  of  spotted  seatrout,  Cynoscion  nebu- 
losus (Pisces:  Sciaenidae),  in  Tampa  Bay,  Florida.     Estuar- 
ies 12:98-110. 
Nelson,  G.  A.,  and  D.  Leffler. 

2001.     Abundance,  spatial  distribution,  and  mortality  of 
young-of-the-year  spotted  seatrout  (Cynoscion  nebulosus) 
along  the  Gulf  Coast  of  Florida.    Gulf  Mex.  Sci.  19:30-42. 
Neter,  J.,  W.  Wasserman,  and  M.  H.  Kutner. 

1983.    Applied  linear  regression  models,  547  p.     Richard  D. 
Irwin,  Inc.,  Homewood,  IL. 
Orlando,  S.  P.,  Jr.,  M.  B.  Robblee,  and  C.  J.  Klein. 

1997.     Salinity  characteristics  of  Florida  Bay:  a  review  of 
the  archived  data  set  ( 1955-1995),  33  p.    Office  of  Ocean 
Resources  Conservation  and  Assessments.  NOAA.  Silver 
Spring,  MD. 
Peebles,  E.  B.,  and  S.  G.  Tolley. 

1988.  Distribution,  growth  and  mortality  of  larval  spotted 
seatrout,  Cynoscion  nebulosus:  a  comparison  between  two 
adjacent  estuarine  areas  of  southwest  Florida.  Bull.  Mar. 
Sci.  42:397-410. 

Phlips,  E.  J.,  and  S.  Badylak. 

1996.     Spatial  variability  in  phytoplankton  standing  crop  and 
composition  in  a  shallow  inner-shelf  lagoon.  Florida  Bay. 
Florida.    Bull.  Mar.  Sci.  58:203-216. 
Powell,  A.  B. 

2003.     Larval  abundance  and  distribution,  and  spawning 
habits  of  spotted  seatrout,  Cynoscion  nebulosus,  in  Florida 
Bay,  Everglades  national  Park,  Florida.     Fish.  Bull.  101: 
704-711. 
Powell,  A.  B„  D.  E.  Hoss,  W.  F.  Hettler,  D.  S.  Peters,  and 
S.  Wagner. 

1989.  Abundance  and  distribution  if  ichthyoplankton  in  Flor- 
ida Bay  and  adjacent  waters.     Bull.  Mar.  Sci.  44:35-48. 

Powell,  A.  B.,  E.  H.  Laban,  S.  A.  Holt,  and  G.  J.  Holt. 

2000.  Validation  of  age  estimates  from  otoliths  of  larval  and 
juvenile  spotted  seatrout,  Cynoscion  nebulosus.  Fish.  Bull. 
98:650-654. 

Rice,  J.  A..  L.  B.  Crowder,  and  M.  E.  Holey. 

1987.     Exploration  of  mechanisms  regulating  larval  survival 
in  Lake  Michigan  bloater:  a  recruitment  analysis  based  on 
characteristics  of  individual  larvae.    Trans.  Am.  Fish.  Soc. 
116:481-491. 
Ricker.W.  E. 

1975.     Computation  and  interpretation  of  biological  statistics 
of  fish  populations.     Bull.  Fish.  Res.  Board.  Can.  191,  382  p. 
Rilling,  G.  C,  and  E.  D.  Houde. 

1999.     Regional  and  temporal  variability  in  growth  and  mor- 


tality of  bay  anchovy,  Anchoa  mitchilli,  larvae  in  Chesa- 
peake Bay.    Fish.  Bull.  97:555-569. 
Rooker,  J.  R.,  S.  A.  Holt,  G  J.  Holt,  and  L.  A.  Fuiman. 

1999.     Spatial  and  temporal  variability  in  growth,  mortal- 
ity, and  recruitment  potential  of  postsettlement  red  drum, 
Sciaenops  ocellatus,  in  a  subtropical  estuary.     Fish.  Bull. 
97:581-590. 
Rutherford,  E.  S.,  and  E.  D.  Houde. 

1995.    The  influence  of  temperature  on  cohort-specific  growth, 
survival,  and  recruitment  of  striped  bass,  Morone  saxatilis, 
larvae  in  Chesapeake  Bay.     Fish.  Bull.  93:315-332. 
Rutherford,  E.  S.,  E.  B.  Thue,  and  D.  G.  Buker. 

1982.  Population  characteristics,  food  habits  and  spawning 
activity  of  spotted  seatrout.  Cynoscion  nebulosus,  in  Ever- 
glades National  Park,  48  p.  United  States  Park  Service, 
South  Florida  Research  Center  Report  T-668,  Homestead, 
Florida. 
Rutherford,  E.  S.,  T.  W.  Schmidt,  and  J.  T.  Tilmant. 

1989.     Early  life  history  of  spotted  seatrout  (Cynoscion 
nebulosus)  and  gray  snapper  (Lutjanus  griseus)  in  Florida 
Bay,  Everglades  National  Park,  Florida.     Bull.  Mar.  Sci. 
44:49-64. 
Secor,  D.  H.,  J.  M.  Dean,  and  E.  H.  Laban. 

1991.     Manual  for  otolith  removal  and  preparation  for 
microstructural  analysis,  85  p.     Belle  W.  Baruch  Institute 
for  Marine  Biology  and  Coastal  Research  Technical  Publica- 
tion 1990-01. 
Sogard,  S.  M. 

1997.     Size-selective  mortality  in  the  juvenile  stage  of  teleost 
fishes:  a  review.     Bull.  Mar.  Sci.  60:1129-1157. 
Stewart,  K.  W. 

1961.     Contributions  to  the  biology  of  the  spotted  seatrout 
(Cynoscion  nebulosus)  in  the  Everglades  National  Park, 
Florida.     M.S  thesis,  103  p.     Univ.  Miami,  Coral  Gables. 
FL. 
Thayer,  G  W„  A.  B.  Powell,  and  D.  E.  Hoss. 

1999.     Composition  of  larval,  juvenile  and  small  adult  fishes 
relative  to  change  in  environmental  conditions  in  Florida 
Bay.     Estuaries  22:518-533. 
Tucker,  J.  W.,  Jr.,  and  B.  E  Faulkner. 

1987.     Voluntary  spawning  patterns  of  captive  spotted 
seatrout.    Northeast  Gulf  Sci.  9:59-63. 
Turney,  W.  J.,  and  B.  F.  Perkins. 

1972.     Molluscan  distribution  in  Florida  Bay.  Sedimenta  III, 
37  p.     Rosenstiel  School  of  Atmospheric  Science,  Univ.  of 
Miami,  FL. 
U.  S.  Naval  Observatory  Applications  Department. 

1997.     Fraction  of  the  moon  illuminated.     [Available  at 
http//aa.usno. navy.mil/data/docs/Moon  Fraction.html.]  Last 
accessed  17  December  2001. 
Ware,  D.  M. 

1975.     Relation  between  egg  size,  growth,  and  natural  mortal- 
ity of  larval  fish.    J.  Fish.  Res.  Board  Can.  32:2503-2512. 
Warren,  C.  E. 

1971.     Biology  and  water  pollution  control,  434  p.     W.  B. 
Saunders  Company.  Philadelphia,  PA. 


156 


Abstract— Age  and  growth  of  the  night 
shark  (Carcharhinus  signatus)  from 
areas  off  northeastern  Brazil  were 
determined  from  317  unstained  ver- 
tebral sections  of  182  males  (113-215 
cm  total  length  [TLI>,  132  females 
(111.5-234.9  cm!  and  three  individuals 
of  unknown  sex  ( 169-242  cm ).  Although 
marginal  increment  (MI)  analysis  sug- 
gests that  band  formation  occurs  in  the 
third  and  fourth  trimesters  in  juve- 
niles, it  was  inconclusive  for  adults. 
Thus,  it  was  assumed  that  one  band 
is  formed  annually.  Births  that  occur 
over  a  protracted  period  may  be  the 
most  important  source  of  bias  in  MI 
analysis.  An  estimated  average  percent 
error  of  2.4'S  was  found  in  readings  for 
individuals  between  two  and  seventeen 
years.  The  von  Bertalanffy  growth 
function  (VBGF)  showed  no  significant 
differences  between  sexes,  and  the 
model  derived  from  back-calculated 
mean  length  at  age  best  represented 
growth  for  the  species  (1^=270  cm,  K= 
0.11/yr,  t0=-2.71  yr)  when  compared  to 
the  observed  mean  lengths  at  age  and 
the  Fabens'  method.  Length-frequency 
analysis  on  1055  specimens  (93-260 
cm)  was  used  to  verify  age  determina- 
tion. Back-calculated  size  at  birth  was 
66.8  cm  and  maturity  was  reached 
at  180-190  cm  (age  8)  for  males  and 
200-205  cm  (age  ten)  for  females.  Age 
composition,  estimated  from  an  age- 
length  key,  indicated  that  juveniles 
predominate  in  commercial  catches, 
representing  74.3%  of  the  catch.  A 
growth  rate  of  25.4  cm/yr  was  esti- 
mated from  birth  to  the  first  band  (i.e. 
juveniles  grow  38?<  of  their  birth  length 
during  the  first  year),  and  a  growth  rate 
of  8.55  cm/yr  was  estimated  for  eight-  to 
ten-year-old  adults. 


Age  determination  and  growth  of  the 
night  shark  (Carcharhinus  signatus) 
off  the  northeastern  Brazilian  coast 

Francisco  M.  Santana 

Rosangela  Lessa 

Universidade  Federal  Rural  de  Pernambuco  (UFRPE) 

Departamento  de  Pesca,  Laboratory  de  Dinamica  de  Populacoes  Mannhas  -  DIMAR 

Dois  Irmaos,  Recile-PE,  Brazil,  CEP  52171-900 

E-mail  address  (for  R.  Lessa.  contact  author)  rplessaigig.com  br 


Manuscipt  approved  for  publication 
26  June  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:156-167  (2004). 


The  night  shark  (Carcharhinus  sig- 
natus) is  a  deepwater  coastal  or  semi- 
oceanic  carcharhinid  that  is  found  in 
the  western  Atlantic  Ocean  along  the 
outer  continental  or  insular  tropical 
and  warm  temperate  shelves,  at  depths 
exceeding  100  meters  (Bigelow  and 
Schroeder,  1948).  The  species  has  been 
recorded  from  Delaware  to  Florida,  the 
Caribbean  sea  (Cuba),  and  northern 
South  America  (Guayana)  (Compagno, 
1984).  It  has  also  been  recorded  in 
southern  Brazil,  Uruguay,  and  Argen- 
tina (Krefft,  1968;  Compagno,  1984; 
Marin  et  al.,  1998),  and  on  the  sea- 
mounts  off  northeastern  Brazil  (02°16' 
to  04°05'S  and  033°43'  to  037°30'W. 
Menni  et  al.,  1995)  where  it  is  called 
"toninha." 

Since  1991,  tuna  longline  vessels  have 
targeted  the  night  shark  in  northeast- 
ern Brazil  (Hazin  et  al.,  1998)  because 
of  its  highly  prized  fins,  the  increasing 
value  of  shark  meat  in  the  local  market, 
and  their  relatively  large  abundance 
and  accessability  on  seamounts  (Menni 
et  al.,  1995).  This  species  is  most  im- 
portant in  the  area,  making  up  909; 
of  catches  over  shallow  banks  (CPUE, 
in  number,  is  2.94/100  hook),  and  only 
15%  of  catches  on  the  surrounding  deep 
area,  yielding  0.04/100  hook  (Amorim 
etal.,  1998). 

Information  on  this  species  is  re- 
stricted to  taxonomic  descriptions 
(Bigelow  and  Schroeder  1948;  Cadenat 
and  Blache,  1981;  Compagno,  1984, 
1988),  and  some  biological  aspects 
(Guitart  Manday,  1975;  Hazin  et  al., 
2000).  Night  sharks  reach  >270-280  cm 
maximum  total  length  (TL)  (Compagno, 
1984;  Branstetter,  1990).  Off  northeast- 
ern Brazil,  females  mature  at  200-205 


cm  TL,  males  at  185-190  cm.  Litter  sizes 
range  from  10  to  15  pups  and  the  gesta- 
tion period  may  last  one  year  ( Hazin  et 
al.,  2000).  The  assumed  size-at-birth  off 
the  United  States  is  60-65  cm  TL  (Com- 
pagno, 1984;  Branstetter,  1990).  Age  and 
growth  have  not  been  estimated. 

The  aim  of  this  study  is  to  present 
the  first  growth  curve  for  Carcharhinus 
signatus  from  vertebral  and  length-fre- 
quency analyses.  This  information  will 
permit  the  use  of  age-based  stock  as- 
sessment methods  for  the  management 
of  the  species  in  the  Exclusive  Economic 
Zone  (EEZ)  off  Brazil. 


Materials  and  methods 

Sampling  data  and  vertebrae  were  col- 
lected from  November  1995  to  Novem- 
ber 1999  from  commercial  landings 
(Natal,  Brazil)  caught  in  deep  (Aracati, 
Dois  Irmaos,  Fundo,  Sirius)  and  shallow 
(Pequeno,  Leste,  and  Sueste)  seamounts 
with  depths  between  38  to  370  m  at  the 
summits  (Fig.  1 ). 

Commercial  vessels  were  equipped 
with  -30  km  Japanese-style  multifila- 
ment longline  gear  (Suzuki  et  al.,  1977). 
On  average,  each  vessel  used  970-980 
hook  per  day;  mainline  sets  began  at 
-02:00  h  and  ended  at  -06:00  h.  The 
retrieval  of  gear  began  at  noon  and  fin- 
ished by  dusk.  The  Brazilian  sardinella 
(Sardinella  brasiliensis),  margined  fly- 
ingfish  (Cypselurus  cyanopterus),  and 
squid  [Loligo  sp. )  were  used  as  bait 
(Hazin  etal,  1998). 

A  total  of  1055  individuals,  landed 
whole,  eviscerated,  or  as  carcasses 
(headless  and  finless).  were  sampled. 
The  interdorsal  space  (posterior  dorsal 


Santana  and  Lessa:  Age  and  growth  of  Carcharhinus  signatus  off  the  northeastern  Brazilian  coast  157 


0° 


02°S  - 


04°S  - 


06°S 


ATLANTIC  OCEAN 


Archipelago  of 

Fernando  de 

Noronha 


40:W 


38=W 


36°W 


34 =W 


32W 


Figure  1 

Location  of  the  sampling  area  for  the  night  shark  iC.  signatus)  collected  off 
northeastern  Brazil. 


fin  base  to  origin  of  the  second  dorsal  fin  [IDS,  cm] ),  total 
length  (snout  to  a  perpendicular  line  from  the  tip  of  the  up- 
per caudal  fin  [TL,  cm] )  and  fork  length  (snout  to  fork  of  tail 
[FL,  cm])  were  measured.  In  carcasses,  only  IDS  was  mea- 
sured, and  IDS,  FL,  and  TL  were  recorded  for  eviscerated  or 
whole  individuals.  A  set  of  five  or  six  vertebrae  were  removed 
from  below  the  first  dorsal  fin  in  317  specimens.  Total  length 
was  measured  as  the  "natural  length"  (without  depressing 
the  tail)  according  to  Garrick  ( 1982). 

To  estimate  TL  for  carcasses,  relationships  from  sub- 
samples  of  IDS  versus  TL  and  FL  versus  TL  were  estab- 
lished for  males  and  females  separately.  Linear  regressions 
derived  for  each  sex  were  tested  for  homogeneity  and  ana- 
lyzed for  covariances  (ANCOVA),  resulting  in  TL=1.2049 
FL  +  1.7972  (r2=0.944.n=668,P=0.41)  and  TL  =  3.3467  IDS 
+  30.879  (r2=0.824;  rc=764,  P=0.161).  Whenever  length  is 
mentioned  hereafter,  we  always  refer  to  TL. 

Vertebrae  were  processed  by  removing  excess  tissue, 
fixed  in  49c  formaldehyde  for  24  hours,  and  preserved  in 
70%  alcohol.  Each  vertebra  was  embedded  in  polyester  resin 
and  the  resulting  block  was  cut  to  about  a  1-mm  thick  sec- 
tion containing  the  nucleus  by  using  a  Buehler®  low  speed 
saw.  Initially,  alizarin-red-s  stained  sections  (Gruber  and 
Stout,  1983)  were  compared  to  unstained  sections  from  the 
same  individuals  to  define  the  best  contrast  for  narrow  and 
broad  zones.  In  the  first  procedure,  sections  were  immersed 
overnight  in  an  aqueous  solution  of  alizarin  red  s  and  0.1% 
NaOH  at  a  ratio  of  1:9  and  then  rinsed  in  running  tap  water. 
In  stained  sections,  narrow  zones  were  visible  as  dark  red 
and  broad  zones  as  light  red,  whereas  in  unstained  sections 
translucent  (narrow)  and  opaque  (broad)  zones  were  visible 
under  transmitted  light.  Unstained  sections  produced  com- 


parable results  to  alizarin  stained  sections  and  were  used 
for  band  observation  in  the  study. 

Bands  counted  in  each  section  and  distances  from  the  focus 
to  the  margin  of  each  narrow  zone  were  recorded.  Vertebral 
radius  (VR)  was  measured  by  using  a  binocular  dissecting 
microscope  equipped  with  an  ocular  micrometer.  Measure- 
ments were  made  at  lOx  magnification  ( 1  micrometer  unit=l 
mm)  with  both  reflected  and  transmitted  light.  The  same 
reader  read  sections  from  the  same  specimen  twice  at  dif- 
ferent times  without  knowledge  of  the  individual  size  or 
previous  count.  Whenever  the  counts  differed  between  the 
two  readings,  a  third  reading  was  used  for  back-calculation 
of  size-at-age. 

The  index  of  average  percentage  error  (IAPE)  (Beamish 
and  Fournier,  1981)  to  compare  reproducibility  of  age  de- 
termination between  readings  was  calculated. 

IAPE  =  1  /  Ar]T  ( 1  / R^ (  |  Xtj -  Xj  \Xj)x  100, 

where  N  =  the  number  of  fish  aged; 
R  =  the  number  of  readings; 

Xt   -  the  mean  age  off1'  fish  at  the  i'h  reading;  and 
Xj  =  the  mean  age  calculated  for  the/,!  fish. 

Marginal  increment  ( MI )  analysis  to  determine  the  time 
of  band  formation  was  used.  The  analysis  was  restricted  to 
1995-97,  when  samples  were  collected  every  month.  The  dis- 
tance from  the  final  band  to  the  vertebral's  edge  (MI)  was 
expressed  as  a  percentage  of  the  distance  between  the  last 
two  bands  formed  on  vertebrae  (Crabtree  and  Bullock,  1998). 
The  distance  between  the  last  and  the  penultimate  band 
was  divided  by  the  distance  between  the  nucleus  and  the 


158 


Fishery  Bulletin  102(1) 


last  band  for  each  vertebra  that  was  measured,  and  we  then 
calculated  the  mean  of  this  number  for  the  entire  sample: 


IK*. 


,)-i?„)//2=0.13(SE  =  0.0009). 


The  expected  distance  between  the  last  (Rn )  and  the  pen- 
ultimate (i?n_! )  bands  was  estimated  as  a  function  of  the 
distance  between  the  vertebral  nucleus  and  the  last  band 
(MI).  The  percent  marginal  increment  (PMI)  was  calcu- 
lated as 

PMI  =  [MI  I  (0. 13  x  Rn )]  x  100. 

Analysis  of  variance  to  test  for  differences  in  PMI  by 
month  was  used.  Post-hoc  tests  (Tukey  honest  significant 
differences  ( [HSD] )  were  performed  to  indicate  which 
months  were  different. 

Characterization  of  the  vertebral  edge  was  used  to  de- 
termine the  time  period  of  band  formation  (Carlson  et  al., 
1999).  Under  reflected  light,  a  narrow  dark  zone  (MI  0),  a 
narrow  light  zone  ( MI  0. 1  to  0.5 ),  and  a  broad  light  zone  ( MI 
0.6  to  1 )  were  observed.  Absolute  marginal  increments  ( MI ) 
were  also  analyzed  by  trimester  for  juveniles  aged  four  and 
five  years,  and  for  adults  ( more  than  eight  years )  to  confirm 
the  time  of  translucent  zone  formation. 

The  relationship  between  VR  and  TL  was  calculated 
by  sex,  tested  for  normality,  and  compared  by  ANCOVA 
(Zar,  1996).  The  final  regression  in  both  sexes  did  not  pass 
through  the  origin,  thus  suggesting  that  the  Fraser-Lee 
method  was  the  most  appropriate  for  back-calculation 
(Ricker,  1969). 


[TL]„  =  (RJVR)({TL\-a)  +  a, 


where  [TL] 
R 


=  the  back-calculated  length  at  age  n; 

-  vertebral  radius  at  the  time  of  the  ring  n\ 
VR  =  the  vertebral  radius  at  capture; 
TL  =  the  length  at  capture;  and 
a  =  the  intercept  on  the  length  axis. 

A  von  Bertalanffy  growth  function  (VBGF)  (von  Berta- 
lanffy,  1938)  was  fitted  to  back-calculated  and  observed 
length-at-age  data  with  the  following  equation. 


L .  1- 


kit  („)i 


where  Lt  =  predicted  length  at  age  t; 

Lr  =  mean  asymptotic  total  length; 

K  =  growth  rate  constant;  and 

t0  =  the  age  when  length  is  theoretically  zero. 

To  obtain  parameters  of  VBGF,  data  were  analyzed  by 
using  FISHPARM  (Prager  et  al.,  1987)  for  nonlinear  least- 
squares  parameter  estimation.  The  Kappenman's  method 
(1981),  based  on  the  sum  of  squares  of  the  differences 
between  observed  and  predicted  lengths  from  a  growth 
model,  was  used  for  comparing  male  and  female  growth 
curves.  In  addition,  likelihood-ratio  tests  were  used  to  com- 
pare parameter  estimates  of  the  von  Bertalanffy  equation 
between  sexes  (Cerrato,  1990). 


Von  Bertalanffy  parameters  (Lx,  K)  were  also  estimated 
by  the  method  of  Fabens  ( 1965 )  usually  employed  for  recap- 
ture data  and  which  takes  into  account  the  size  at  birth  (L(l) 
instead  of  t0.  This  method  reconfigures  VBGF  and  forces  the 
regression  through  a  known  size  at  birth: 


L,  =Ljl-be- 


where  b  =  (L., 


-L0)/Lx 


We  used  Fabens  routine  for  growth  increment  data 
analysis  of  the  FAO-ICLARM  stock  assessment  tools  (FI- 
SAT)  program  (Gayanilo  et  al.,  1996),  assuming  that  the 
time  intervals  (=At)  for  each  size-at-age  class  were  equal 
and  had  a  periodicity  identical  to  that  obtained  from  the 
vertebral  analysis. 

The  lengths  of  1055  individuals  were  divided  into  5-cm 
intervals  and  analyzed  by  the  Shepherd  method  ( 1987 )  with 
the  length-frequency  data  analysis  program  ( LFDA ).  Initial 
values  of  Lv  were  based  on  results  from  maximal  lengths 
in  the  sample  and  from  literature  (Compagno.  1984).  K 
values  ranging  from  0.05  to  1.8  were  used  as  input  into  the 
program,  which  was  run  repeatedly  until  the  highest  score 
function  was  obtained.  The  Lx  and  /f  values  were  then  used 
to  calculate  t0  (Sparre  et  al.,  1989): 

tQ  =  t  +  {l/K)(\nlL.  -lt])/LJ. 

Using  an  age-length  key,  based  on  317  individuals  for 
which  vertebrae  were  read,  we  evaluated  the  age  composi- 
tion of  the  sample  (Bartoo  and  Parker,  1983).  Maximal  ages 
in  the  sample  were  calculated  by  employing  the  inverted 
VBGF  (Sparre  et  al..  1989).  Further,  the  formula  by  Fa- 
bens (1965)  [5(ln2)/AT  for  longevity  estimation  was  used. 
All  statistical  inferences  were  made  at  a  significance  level 
of  0.05. 


Results 

The  total  sample  size  consisted  of  1055  individuals:  (551 
males  [93-248  cm],  499  females  [110-252  cm],  and  5 
individuals  of  undetermined  sex  [169-260  cm])  (Fig.  2).  Of 
these,  vertebrae  were  removed  from  317  specimens  (182 
males  [113-215  cm],  132  females  [111.5-234.9  cm],  and 
3  individuals  of  undetermined  sex  [169-242  cm]). 

Differences  in  the  relationship  between  VR  and  TL 
between  sexes  were  not  found  to  be  significant  (P=0.81D. 
The  regression  for  the  overall  sample  showed  a  linear 
relationship:  TL  =  13.523V/?  +  41.824  <rM).89:  n=317>, 
indicating  that  vertebrae  are  suitable  structures  for  age 
determination,  and  methods  based  on  direct  proportion  are 
appropriate  for  back-calculation. 

The  average  percentage  error,  calculated  between  two 
readings,  ranged  from  098  to  4.5^  in  vertebrae  with  2  to 
17  bands  and  the  average  IAPE  for  the  overall  sample  was 
2.4'-.  Coefficient  of  variation  (CV)  between  readings  for 
total  sample  was  6.88'  < . 

Monthly  PMI  analysis,  for  the  entire  sample,  indicated 
that  bands  were  formed  from  June  to  October,  when  high- 


Santana  and  Lessa:  Age  and  growth  of  Carcharhinus  signatus  off  the  northeastern  Brazilian  coast 


159 


70  i 
60 
50 
>,     40 

CJ 

c 

CD 

=>      30  i 

cr 

CD 

it      20- 

10  ■ 

ii  n  i 

if 

1 

1 

1 

n=1055 
\\\]  Jl  1    r,   t  I  .          . 

Lengtl 
easter 
bars  = 

muimmmLfiifimiflminmmifimifimcn 

OUCVJCvjCvjC\ic\JC\iC\i<NC\icNiC\ic\ic\JCcicNC\ic\J 

cno*-c\jct3Ti/}(or^<oa>OT-cjpO'<fmc0 

Total  length  (cm) 

Figure  2 

l-frequency  distribution  for  the  night  shark  (C.  signatus)  caught  off  north- 
i  Brazil  between  1995-99  (black  bars=females;  white  bars  =  males;  grey 
undetermined  sex). 

250 


200 


-     150 


E     100- 


M  J  J 

Month 


A         S        O         N         D 


Figure  3 

Percent  marginal  increments  means  (-)  with  the  minimum  and  maximum  values 
for  the  night  shark  (C.  signatus)  caught  from  1995  to  1997  off  northeastern  Brazil 
(n=171).  The  number  of  individuals  sampled  per  month  is  shown  above  the  ver- 
tical bars. 


est  mean  values  are  reached  (Fig.  3).  These  values  are 
followed  by  the  lowest  mean  PMI  in  October,  indicating 
that  the  new  translucent  zone  forms  from  that  point  on. 
Monthly  PMIs  showed  significant  differences  throughout 
the  year  (P=0.0463)  and  post-hoc  comparisons  detected 
differences  in  February,  April,  September,  and  October. 
Furthermore,  monthly  categorization  of  vertebral  edges 
indicated  that  the  highest  frequency  of  broad  light  edges 
(MI  0.6-1)  appears  from  July  through  December  and  nar- 
row dark  edges  (MI  0)  from  March  through  December, 
with  the  exception  for  months  of  May  and  August  (Fig.  4). 
Trimonthly  frequency  distribution  of  absolute  marginal  in- 
crements (Mis)  was  carried  out  for  juveniles,  revealing  four 
and  five  bands,  and  for  adults,  revealing  more  than  eight 
bands.  For  the  former  group,  a  higher  number  of  broader 


increments  and  fully  formed  bands  in  the  third  and  fourth 
trimesters  were  observed  (Fig.  5).  For  adults,  an  unclear 
pattern  was  observerd  perhaps  because  a  smaller  sample 
size  was  obtained. 

Because  there  was  no  complete  agreement  on  the  time 
of  band  formation  among  different  MI  analysis  for  juve- 
niles and  adults,  age  was  assigned  by  assuming  an  annual 
pattern  of  band  deposition.  The  birth  mark  present  in  all 
analyzed  vertebrae  was  not  taken  into  account  for  age  as- 
signation. Under  this  assumption,  band  counts  indicate 
relative  age  (years). 

Mean  observed  lengths-at-age  were  higher  than  mean 
back-calculated  lengths  for  males  and  females  and  were 
likely  due  to  the  strong  variation  in  size  for  each  age  class 
(Table  1).  The  tendency  of  back-calculated  lengths  of  older 


160 


Fishery  Bulletin  102(1) 


Table  T 

Mean  back-calculated  (BC)  and  observed  length-at-age 
eastern  Brazil  (SD=standard  deviation  I. 

( OL  l  data  for  male  and  female 

night  sharks  (C.  signatus)  collected  off  north- 

Age  ( yr  1 

Females 

Males 

BClcm)±SD 

OL(cm)  ±SD 

BCicmi±SD 

OLicml±SD 

0 

66.8  ±1.78 

— 

67.3  ±1.41 

— 

1 

91.9  ±1.31 

— 

92.3  ±1.37 

— 

2 

113.4  ±2.13 

122.5  ±16.93 

113.3  ±1.48 

120.1  ±4.21 

3 

128.8  ±2.21 

132.9+9.77 

128.6  ±1.54 

135  ±8.91 

4 

142.7  ±2.41 

149.8  ±7.75 

142.4  ±1.94 

151.5  ±9.72 

5 

154.7  ±2.92 

160.7  ±7.21 

154.5  ±2.7 

157.5  ±7.86 

6 

165.9  ±3.46 

166.8  ±10.32 

166.3  ±3.25 

167.5  ±8.1 

7 

176.8  ±3.4 

179.8  ±9.56 

177.4  ±2.64 

177.6  ±9.34 

8 

185.9  ±3.71 

184.9  ±9.12 

187.4  ±2.22 

189.8  ±6.53 

9 

194.8  ±3.82 

197.1  ±6.49 

195.8  ±2.25 

199.9  ±5.26 

10 

202  ±4.75 

208.2  ±3.89 

202.4  ±2.78 

204.3  ±3.13 

11 

206.9  ±5.56 

202 

209.8 

212.5  ±3.54 

12 

215.7  ±2.4 

218 

— 

— 

13 

222.2 

— 

— 

— 

14 

226.9 

— 

— 

— 

15 

231.7 

234.4  ±0.63 

— 

— 

fish  in  the  early  years  to  be  systematically  lower  than 
younger  ones  at  the  same  age  (Lee's  phenomenon)  was 
not  evident  (Tables  1  and  2). 

Using  back-calculated  lengths-at-age  (Table  3),  we 
plotted  male  and  female  growth  curves  separately  and 
then  tested  the  data;  no  indication  of  significant  differ- 
ences in  growth  was  observed  between  sexes  with  both 
the  Kapenman's  (P>0.05)  and  likelihood  ratio  tests 
(Table  4).  Data  were  then  treated  together,  incorporat- 
ing individuals  of  undetermined  sex.  VBGFs  derived 
from  observed  length  at  age  were  not  tested  because 
of  missing  values  in  different  age  classes.  The  method 
of  Fabens  for  combined  sexes,  fitted  to  back-calculated 
data,  provided  L,  and  K,  by  using  b  =  0.781,  L0=  62.5 
cm  (Compagno,  1984)  and.  At  =  1  year  (Table  2). 

Parameters  from  back-calculation  were  close  to 
those  derived  from  length-frequency  analysis  for  1055 
specimens,  whereas  observed  lengths  and  the  Fabens 
method,  provided  the  most  varying  parameters  with 
lowest  correlation  and  highest  coefficients  of  variation 
(Table  2). 

The  smallest  specimen  in  the  vertebral  sample  show- 
ing two  complete  bands  in  sections  was  111.5  cm,  close 
to  the  estimated  mean  back-calculated  length  at  age 
two  of  113.7  cm  (Table  3).  Size  at  maturity,  185-190  cm  for 
males  and  200-205  cm  for  females,  corresponded  to  8-  and 
10-year-old  individuals,  respectively  (Fig.  6).  The  largest 
and  oldest  specimen  whose  vertebrae  were  used,  was  242 
cm,  which  corresponded  to  17-year-old  individual. 

A  growth  rate  of  25.4  cm/yr  was  estimated  from  birth  to 
the  first  band — a  rate  that  corresponded  to  389  of  the  birth 


0  • 

■  0 

□  0.1-0.5 

5! 

r 

1 

□  0.6-1 

33 

45 

8  ■ 

51 

6  ■ 

15 

2< 

\ 

43 

4  • 
2  ■ 

1 

7 

9 

1 
6 

5 

' 

% 

-i 

0  • 

-| 

i 

IJI 

1 

J 

■ 

1 

-X 

--I-1- 

M     A 


J        J 
Month 


Figure  4 

Categorization  of  edges  by  month  for  the  night  shark  iC. 
signal  its)  off  northeastern  Brazil. 


length  (the  length  at  birth  being  66.8  cm).  Also,  a  mean  rate 
of  8.55  cm/yr  was  calculated  for  8-  to  10-year-old  individu- 
als, when  maturity  is  achieved  (Table  3). 

Considering  mature  individuals  >185  cm.  the  age  com- 
position for  the  vertebral  samples  («=317)  indicated  that 
17.3%  of  specimens  were  adults  (Table  5).  Instead,  for  the 
total  sample  (ra=1055),  where  the  age  ranged  between  2  to 
al7  years,  adults  corresponded  to  25.3%  of  the  total  sample 


Santana  and  Lessa:  Age  and  growth  of  Carcharhinus  signatus  off  the  northeastern  Brazilian  coast 


161 


0  1       0.2      0  3      0  4      0  5      0.6      0.7      08      0  9        1 


12 
10  - 

8 

6 

4 

2 

0 


ill 


0  1       0  2      0  3      0  4      0  5      0.6      07      08      09        1 


12 
10 

8 

6 

4  - 

0 


3,d  Trimester 
n=36 


I..   I 


0  1       0  2      0  3      0.4      0  5      0.6      0  7      0  8      0  9        1 


12  - 

10 


4lh  Trimester 


Nihil.  . 

0  1       0  2      0  3      0  4      0.5      0.6      0  7      08      0  9        1 


B 


r'  Trimester 
n=  14 


II I  ■  - 


2nd  Trimester 

12  - 

n  =  24 

10  - 

8 

6 

4 

2 

■ 

o  - 

01       0.2      0.3     0.4      0.5      0.6      0.7      08      09        1 


12 
10  - 


3rd  Trimester 
n=  14 


ll 


01       0.2      03      0.4      05      0.6      0.7      0.8      09        1 


12 
10 
8 
6 

4 
2 
0 


4m  Trimester 
n=10 


Jl 


0  1       0.2      0  3      0  4      0.5      0  6      0.7      0  8      0.9        1 


Ml 


Figure  5 

Marginal  increments  (MI)  by  trimester  for  ages  4  and  5  (n  =  139)  (A)  and  a8  (Bl  (n=54)  for  the  night  shark  (C. 
signatus)  from  northeastern  Brazil. 


(Fig.  7).  According  to  the  inverted  back-calculated  VBGF 
the  oldest  specimen  in  the  sample  was  31.7  years  old  (260 
cm),  whereas  longevity  was  31.5  years. 


Discussion 

Validating  the  time  of  band  formation  is  considered  critical 
when  using  hard  parts  for  age  estimates  (Brothers,  1983), 
and  validation  is  successful  when  growth  zones  are  shown 
to  form  annually  in  all  age  groups  of  the  population  (Beam- 


ish and  McFarlane  1983).  Marginal  increment  analysis, 
carried  out  on  younger  and  faster  growing  individuals, 
cannot  always  be  used  for  validating  older  age  groups,  and 
therefore  all  ages  must  be  ascertained  (Brothers,  1983). 
In  the  present  study,  we  obtained  significant  differences 
in  marginal  increments  for  the  total  sample.  However,  the 
significance  level  of  the  test  (P=0.046)  was  close  enough 
to  0.05  to  cause  us  to  suspect  that  the  distributions  could 
have  been  similar.  The  time  of  band  formation  varied  when 
different  age  groups  were  analyzed  separately,  despite 
suggestions  that  bands  are  completed  in  the  third  and 


162 


Fishery  Bulletin  102(1) 


E       o 


150 
100 


B 


\ 


0  1   2  3   4  5   6  7  8   9  10  11  12  13  14  15  16  17  18         0   1   2  3   4  5  6  7  8  9  10  1 1  12  13  14  15  16  17  1 


250 
200 
150 
100 
50 
0 


i  !  V  '■ 


■  Back-calculated  Observed Fabens 


0      1       2     3      4      5      6     7     8      9     10    11    12    13    14    15    16    17    18 

Age  (years) 

Figure  6 

Growth  curves  generated  from  (A)  females.  (B)  males,  and  (Cl  sexes  combined  for  the  night  shark  (C.  signatus)  off  the  northeastern 
Brazil. 


Table  2 

Von  Bertalanffy  parameters  derived  from  back-calculated  lengths  (BC),  observed  lengths  (OL),  lengths 
the  length-frequency  data  analysis  (LFDA)  package  for  the  pooled  database  (SE  is  standard  error;  CV 

from  the  Fabens  method,  and 
is  coefficient  of  variation  i. 

Methods 

Sex 

L_,  (cm) 

SE 

CV 

/f(/year) 

SE 

CV 

f0(year) 

SE 

CV 

r- 

BC 

Males 

256.5 

5.56 

0.022 

0.124 

0.007 

0.055 

-2.538 

0.119 

0.047 

0.999 

Females 

265.4 

4.15 

0.016 

0.114 

0.005 

0.045 

-2.695 

0.127 

0.047 

0.999 

Both 

270 

2.78 

0.01 

0.112 

0.003 

0.031 

-2.705 

0.099 

0.037 

0.999 

OL 

Males 

306.1 

37.71 

0.117 

0.076 

0.02 

0.267 

-4.663 

0.882 

0.189 

0.995 

Females 

297.1 

26.71 

0.09 

0.077 

0.018 

0.235 

-4.853 

0.977 

0.201 

0.99 

Both 

289.9 

7.6 

0.026 

0.085 

0.006 

0.077 

-4.395 

0.348 

0.079 

0.998 

Fabens 

Both 

285.3 

15.69 

0.055 

0.08 

0.016 

0.2 

— 

— 

— 

— 

LFDA 

Both 

270.9 

— 

— 

0.106 

— 

— 

— 

— 

— 

— 

fourth  trimesters  (new  bands  begin  to  form  in  this  period) 
in  juveniles.  Results  were  inconclusive  for  adults.  For  C. 
obscurus  (Natanson  et  al.,  1995),  C.  plumbeus  (Sminkey 
and  Musick  1995),  C.  porosus  (Batista  and  Silva,  1995: 


Lessa  and  Santana,  1998),  C.  acronotus  (Carlson  et  al., 
1999).  and  /.  oxyrhynchus  (Lessa  et  al.,  2000),  inconclusive 
results  for  MI  analysis  were  obtained.  The  inability  to  dem- 
onstrate the  periodicity  of  band  deposition  in  adult  sharks 


Santana  and  Lessa:  Age  and  growth  of  Carcharhinus  signatus  off  the  northeastern  Brazilian  coast 


163 


in  the  present  study  is  similar  to  the 
outcome  for  C.  limbatus  older  than 
four  years  (Wintrier  and  Cliff,  1996). 
For  the  last  mentioned  species,  the 
problem  was  circumvented  by 
restricting  MI  analysis  to  juveniles 
(Killam  and  Parsons,  1989). 

Age  was  assigned  by  assuming 
an  annual  pattern  of  deposition,  as 
commonly  occurs  for  most  carcha- 
rhinids  like  C.  brevipinna  and  C. 
limbatus,  Rhizoprionodon  terraeno- 
vae  (Branstetter  et  al.,  1987;  Brans- 
tetter  and  Stiles,  1987),  Negaprion 
brevirostris  (Gruber  and  Stout, 
1983),  and  C.  longimanus  (Seki  et 
al,  1998;  Lessa  et  al.,  1999c).  Three 
sources  of  bias  generally  occur  with 
MI  analysis:  1)  sample  sizes  are 
small  for  any  particular  month  or 
for  any  age  class  (Cailliet,  1990);  2) 
data  are  collected  over  a  too  long 
a  period  causing  variability  on  ac- 
count of  annual  marks  that  are  not 
formed  at  the  same  time  ( Brothers, 
1983 )  and  3 )  births  occur  over  a  long 
period  (Brothers,  1983).  All  these 
may  have  biased  MI  analysis  in  the 
present  study. 

Research  carried  out  in  the  study 
area  by  Hazin  et  al.  (2000)  indi- 
cated that  copulation  takes  places 
throughout  the  austral  summer. 
Embryos  measuring  10  to  40  cm 
were  collected  in  February,  whereas 
31.8  to  37.2  cm  embryos  were  found 
in  June.  This  remarkable  variability 
in  embryo  size  during  the  gestation 
period  suggests  that  birth  period 
lasts  several  months.  Furthermore, 
with  an  estimated  back-calculated 
birth  length  of  66.8  cm,  individuals 
measuring  -40  cm  in  February  will 
be  born  long  before  individuals  that 
measured  37.2  cm  in  June.  Such  a 
protracted  parturition  period  could 
lead  to  differences  in  MI  of  the  same 
cohort.  Thus,  after  an  assumed  -12 
months  gestation  period,  individu- 
als are  born  with  birth  dates  vary- 
ing by  several  months.  Moreover, 
no  significant  differences  in  MI 
analysis  was  found  for  C.  porosus 
and  /.  oxyrhynchus,  which  also  have 
a  protracted  birth  seasons  (Lessa  et 
al.,  1999a,  1999b). 

A  comparison  of  growth  model 
parameters  by  using  known  size 
information,  such  as  size-at-birth 
and  maximum  observed  size,  can  be 


01 

t- 

CO 

CO 

1 

CM     O 

Hr-> 
03 

- 

— 

- 

1 

CM 

CN 

CO 

ed 

co 

<D 

CO 

2 

00 

CO 

1 

|         | 

03 

rH 

— 

[-' 

t> 

XI 

00 

CO 

CN 

CM 

a 

o> 

-— 

CS 

lO 

[^ 

r- 

l> 

CO 

■*     135 

a 

> — 1 

CO 

^ 

CO 

CN 

c- 

CO 

CO     «? 

CO 

CO 

CO 

CO 

(N 

CN 

CM 

'-' 

IN     O 

03 

0 

=3 

^t 

CO 

05 

d 

CN 

in 
ai 

<N 

CN 
00 
<N 

CO 

- 
t> 

1         1 

CN 

IN 

CN 

■^ 

_^ 

CO 

CO 

CN 
CN 

CO 

iri 

CO 
CO 

CO 
M 

1         1 

oa 

CN 

CM 

CN 

CN 

CN 

CM 

CN 

C 

'-, 

03 

CM 

5< 

■* 

■* 

00 

r-i 

CN 

CO    o 

CO 

a 

03 

CM 

t-' 

d 

t> 

CO 
CO 

CM 

CN 

CN 

CM 

CN 

'— 

o 

E> 

[^ 

CO 

CN 

CM 

05 

05     C- 

EC 
o 

'~l 

c— 

CO 

z 

CO 

- 

— 

: 

CM 
CM 

o    in 
in    in 

CN 

CN 

CN 

CN 

M 

CO 

d 

rC 

b£ 

3 

o 

rH 

£ 

CD 

•* 

CN 

CN 

iri 

CN 

m 

CO 

CO 

c 

CO 

IN     05 

' — 

O 

o 

s 

o 

O 

o 

3 

CN 

CN 

CN 

IN 

CN 

CM 

CN 

IN     CO 

■SP 

a 

_ 

CO 

CO 

CN 

in 

l> 

05 

,_, 

CD 

05      rH 

3 

o 

a> 

CO 

CO 

■* 

■<* 

in 

1?- 

00 

X 

05 

00 

05     ■* 
rH     CD 

03 

CJ 

CJ 

es 

05 

05 

05 

35 

C75 

05 

-a 

c 

*"* 

i— 1 

iH 

1-1 

*" * 

I — 

rH 

1—1 

iri 

' — 

CO 

_* 

X 

5 

in 

T-H 

!> 

CO 

» 

in 

CN 

IN 

in 

CO 

CO      CD 

a 

"3 

CO 

iri 

C-^ 

iri 

00 

rH 

t^ 

d 

d 

i> 

CO 

CO      '"■ 

*0     xi 

00 

X 

00 

00 

3! 

X 

/. 

*    I 

XI 

1-1 

1-1 

1-1 

7-1 

^ 

^H 

^H 

rH     t> 

-O     JA 

fc 

m      tx 

JW 

t~- 

03 

c- 

-# 

CO 

-* 

00 

CM 

00     ->* 

■"      C 

> 

t> 

o 
r. 

■* 

in 

00 

iri 

00 

CO 

CN 

d 

c> 

CO 
CO 

CD 

r- 

i  - 

c^ 

c- 

C~ 

c~ 

r 

t> 

t> 

CM 

rH      05 

-1-3 

t*H 

O 

CO 

CO 

CO 

CO 

in 

in 

05 

I> 

00 

CO 

00 

00 

Tf      CO 

CO 
03 

CD 

CN 

■* 

iri 

i> 

CO 

X 

t> 

^ 

CO 

d 

CO 
CO 

5§ 

•y. 

rH 

CO 

CO 

CO 

CO 

CO 

l> 

CD 

■- 

03 

— 

'""l 

rt 

rt 

^ 

1-1 

^ 

^ 

^ 

CM 

rH       05 

T3 

V 

CO 
CO 

^" 

CO 

■* 

Oj 

CO 

CD 

CO 

c- 

CO 

CN 

O) 

00 

05     CO 

a 

-t-j 

m 

in 

rH 

CO 

00 

'- 

CO 

t-^ 

iri 

ai 

t> 

iri 

in 

CO 

in    oo 

rH      CD 

£ 

£ 

CN 

i- 

m 

1- 

>- 

m 

m 

IT. 

m 

m 

o 
o 

3 
o 

o 

CN 

l> 

■a 
5 

£ 

00 

CN 

r— 

C~ 

CO 

a- 

Oi 

in 

"* 

00 

<J3 

^h 

r-t 

00 

I>     CO 

■* 

•* 

CM 

d 

rH 

rH 

CO 

rH 

CO 

CO 

iri 

d 

CO 

05 

g  § 

Xi 

CN 

-r 

•* 

re 

-T 

— 

-t 

-r 

-t 

■* 

-T 

— 

-T 

S 

=*H 

rH      CO 

cj 

o 

' — ' 

>. 

CO 

u 

05 

CJ) 

t-; 

CN 

CN 

rH 

O) 

in 

CN 

CO 

IM 

Tf 

05    in 

CO 

o 
ho 

00 

c— 

o 

t-' 

d 

00 

CO 

00 

00 

d 

[> 

CO 

d 

CO 

CO 

d 

CN 

05 

CO     2 

-c 

Oi 

CO 

<N 

r- 

CI 

CI 

CN 

CN 

:: 

CN 

?! 

N 

CO 

IN 

CO         . 

rH       OJ 

a> 

rt 

cm 

o 

CO 

CO 

rH     00 

"* 

00 

rH 

IN 

"* 

■* 

05 

Tf 

[^ 

00 

I> 

c- 

Tf 

en    in 

CO 

CM 

CN     CN 

in 

CO 

CN 

CO 

CO 

— 

CN 

CN 

~ 

iri 

- 

yi 

CN 
CN 

ri  - 

ft 

c 

-c 

CU 

CN 

rH      CO 

r~ 

O     05 

in 

in 

CO 

(N 

CD 

00 

CN 

CO 

00 

in 

CO 

CM 

C^ 

1           1 

-a 

c 

rn 

03  d 

CO 

CN 

rH 

CN 

(N 

CO 

CO 

05 

d 

rH 

CO 

CN 

CN 

1           1 

CJ 

'3 

CO 

05 

05 

05 

Ol 

05 

05 

05 

05 

05 

05 

(35 

OS 

t 

o 
CO 
XI 
O 

CO 

X 

CU 
CO 

l> 

i-j     t- 

00 

CO 

CO 

CN 

in 

CO 

i-i 

05 

CO 

00 

in 

CO 

X 

|           | 

-a 
c 

"cO 
3 

o 

iri   iri 

t> 

t> 

I> 

00 

00 

t> 

r> 

d 

■* 

d 

iri 

d 

CO 
r-i 

CO 

CO    CO 

in 

CO 

CO 

CO 

CO 

CO 

CO 

CD 

CD 

CO 

CO 

X 

CO 

T> 

-a 

'> 

cj 

■3 

~ 

"oj     O     CN 

03 

o 

CO 

CO 

CO 

m 

O) 

CO 

rH 

CN 

rH 

a 

4J     rH     CN 

■* 

OJ 

CO 

CN 

CN 

rH 

CO 

3 

HH 

o 

o 

3 

CJ 

T3 

CO 

CJ 

Ji 

CJ 
CO 

CD 

CD 

cm 

"cO 
o 

-g    cm    CO 

CO 
CQ 

Tf 

in 

CO 

t- 

00 

05 

= 

rH 

CI 

in 

C- 

c 

CO 

CJ 

- 

CO 

>  e 

c    « 

j3  S  3q 

o 

CQ 

"~" 

< 

164 


Fishery  Bulletin  102(1) 


Table  4 

Likelihood  ratio  tests 

comparing  estimates  of  von 

Bertalanffy  parameters 

for  males  (noted  as 

1 1  and  females 

(noted 

as  2)  for 

C.  signatus 

in  the  linear  constraints. 

Hypothesis 

Linear  constraints 

Residual  SS 

X2r 

df 

P 

HQ 

none 

60536.4 

Hwl 

£=oi  =  Lx2 

10511 

0.049 

1 

0.996 

Hto2 

A,  =  K2 

10524.3 

0.047 

1 

0.996 

Hto3 

'01  =  '02 

10205.6 

0.122 

1 

0.999 

HvA 

Same  L^,  A.  and  t0 

24301.2 

0.164 

3 

0.973 

useful  as  a  method  of  verification  ( Cailliet  et 
al.,  1983).  Although  no  specimens  younger 
than  2-years-old  were  caught  (perhaps  due 
to  the  gear  selection  bias),  the  presumed 
size  at  birth  was  about  60-65  cm  ( Compag- 
no,  1984),  which  is  similar  to  the  estimated 
size  in  the  present  study  (66.8  cm).  Also,  the 
estimated  Lr  value  (270  cm),  derived  from 
the  back-calculated  or  observed  VBGF  is 
close  to  the  maximum  size  of  276  cm  men- 
tioned by  Bigelow  and  Schroeder  ( 1948),  280 
cm  off  Cuba  (Compagno,  1984),  and  275  cm 
byGarrick(1985). 

Mean  observed  length-at-age  is  gener- 
ally higher  than  back-calculated  mean 
length-at-age  (Bonfil  et  al.,  1993;  Lessa  and 
Santana.  1998),  leading  to  lower  values  of 
La  and  higher  values  of  K.  However,  in  the 
present  study,  although  mean  observed 
length-at-age  is  higher  than  mean  back-cal- 
culated lengths,  parameters  derived  from  back-calculation 
provided  a  lower  Lx  and  a  higher  A' value.  Inconsistency  of 
the  observed  length-at-age  set  is  attributed  to  the  missing 
values  in  for  ages  0,  1,  13,  14,  and  16.  This  led  to  a  VBGF 
which  provided  an  unrealistic  birth  size  of  90  cm  and  which 
present  a  flatter  shape  than  the  back-calculated  curve. 

Von  Bertalanffy  growth  parameters  generated  from  both 
back-calculation  and  by  the  Fabens  method  were  all  consid- 
ered suitable  and  were  of  the  same  magnitude.  However, 
taking  into  account  1 )  parameters  close  to  those  derived 
for  length-frequency  analysis,  and  2)  the  best  statistical  fit, 
the  back-calculated  VBGF  was  chosen  as  best  representing 
growth  in  the  species. 

Comparisions  of  biological  features  such  as  maturity 
size  and  maximum  sizes  have  been  used  for  inferences  in 
growth  and  to  explain  differences  between  sexes  (Natanson 
et  al.,  1995;  Natanson  and  Kohler,  1996;  Lessa  et  al.,  2000). 
Tin'  studied  species  shows  a  disparity  of -15  cm  in  matu- 
rity sizes  between  sexes  (Hazin,  et  al.,  2000),  corresponding 
to  ~2  years.  In  addition,  the  largest  specimen,  for  which 
six  was  determined,  was  a  252-cm  female  and  the  largest 
male  was  248  cm.  These  disparities,  however,  did  not  bring 
about  differences  in  growth  between  sexes,  as  indicated  by 
results  of  both  tests  used.  Such  a  result  can  be  explained  by 
the  number  of  juveniles  used  for  age  determination  (-83'  I 


300 

250 

n    =1055 

>.    200 
0 

c 

S     150 

O" 

1 

"-     100 

■     | 

50 
0 

.ll 

Mil.. 

<1             3             5              7              9             11            13           15          >17 

Age  (years) 

Figure  7 

Age  composition  for  the  night  shark  (C.  signatus)  collected  off  northeastern 

Brazil. 

Thus,  the  number  of  adults  was  not  high  enough  to  bring 
about  any  differences  in  the  growth  equation  although 
differences  frequently  occur  after  maturity,  caused  by  dif- 
ferent growth  rates  between  sexes  (Natanson  et  al.,  1995; 
Sminkey  and  Musick,  1995). 

Assuming  that  the  time  elapsed  between  birth  and  the 
band  corresponding  to  age  1  is  one  year,  the  species  grows 
38%  of  its  birth  length  during  the  first  year.  This  growth 
rate  is  close  to  that  (50%)  generally  assumed  (Branstetter 
1990;  Cortes,  2000).  Furthermore,  the  estimated  K  value 
falls  within  the  range  suggested  by  the  first  author,  and 
according  to  him,  the  night  shark  is  a  relatively  fast  grow- 
ing species,  presenting  a  life  strategy  similar  to  that  of  C. 
falciformis,  and  apparently  depending  on  rapid  growth  for 
adequate  neonate  survival  due  to  vulnerability  to  preda- 
tion  from  large  sharks. 

In  summary,  considering  the  increasing  fishing  effort 
on  the  night  shark  as  a  targeted  species  and  that  catches 
are  mainly  composed  by  juveniles  (representing  74.7'  i  of 
specimens  in  landings),  we  believe  that  the  A'-selected 
characteristics  of  the  species  (including  late  maturity, 
long  gestation  period,  and  low  fertility  1  should  be  taken 
into  account  in  determining  the  management  of  this 
resource.  Demographic  analyses  will  be  required  for  the 
examination  of  consequences  of  current  levels  of  exploi- 


Santana  and  Lessa:  Age  and  growth  of  Carcharhtnus  signatus  off  the  northeastern  Brazilian  coast 


165 


c 


Q. 

C 


3 

2 


pq 

LT> 

c 

u 

a) 

.Q 

» 

,(B 

CO 

be 
3 


O 


bo 


tj, 


bf  no 


■"  cm 


2     I 


2   I 


S  2? 


oi  >5 


CO  «o 


CO    !N 


I  I  I         I         I  I  I         I         I  I  I         I         I         I         I  I         I         I  I  I         I         I         I  I 


I         I         I         I  I         I         I         I  I  I  I         I         I  I         I         I         I         I  I  I         I         I  I  I         I         I         I 


I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I  8    I 


I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I 


I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I 


I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    I    IS   I    I    I 


I    I    !    I    I    I    I    I    I    II    I    I    I    I    I    I    12    I8S    I    M    I    I 


I    I    I    I    I    I    I    I    I    I    I    I    I    I 


I    I    I    I    I    I    I    I    I    I    I    I    I    I 


I      CO     CO     [-     o 
1         '      CO     -*     CO     ^ 

^H       CO 


I   I   I   I   I   I 


CO     iO  lO     ■*  CO 

«-;    ^  oj    _j  ~-; 

00     t—  tH  CO 

CO  t>  CO 


I   I   I   I   I 


I   I   I   I   I   I   I   I   I   I 


I     co      I     co    co    t>    in    o 
1     •*    co'    ~    oi    "= 


I      I      I       I 


h   m    -*    co 


I      I      I       I 


I         I         I         I  I  I       t>     m     tH     CO     CO     C-        I       1> 

I      I      I      I       I       I    «    h    ™    ^    n    h      !     cb 


I      I      I      I      I      I      I      I 


H     CO     *  —I 


llllllliquaiaoqwco-*.   co«|||||||||| 

— i  cm    co    co    ■*    m    co 


I       I       I       I 


|       M10CB^^CDt-C5 
CONHCtirtT|'dH 

h    •-<    co    in    in    co    cm 


I     coocomcoin^-i^j- 
'     co'^co'cNcodcdiri 

CO      CO  CD  CM  CM      ^H 


I    I    I    I    I    I    I    I    I    I 


I    I    I    I    I    I    I    I    I    I 


I    I 


co  in  t> 


o  —  m 
10  cri  oci 


o  o  t-~  in 
2  2  «o  <n" 

"   "   CO   l-H 


I      I 


S  i    i 


I   I   I   I   I   I   I   I   I   I   I   I   I   I   I   I   I 


I   I   I   I   I   I   I   I   I   I   I   I   I   I   I   I   I 


COH!flCOCnmHCO!Mfflt*tOffiH05(NCO(Nt'0)NN 

HH-fCOCO(N(N(N  T-H  i-H 


HH^lNcoco^'<tlom^otD^^cocoo)a)OOHH(N(Ncoco't 


166 


Fishery  Bulletin  102(1) 


tation  to  ensure  the  sustainability  of  the  night  shark  in 
northeastern  Brazil. 


Acknowledgments 

The  present  research  was  funded  by  Ministerio  do  Meio 
Ambiente-MMA,  Secretaria  da  Comissao  Interministerial 
para  os  Recursos  do  Mar-SECIRM  in  the  scope  of  Programa 
Nacional  de  Avaliacao  do  Potencial  Sustentavel  de  Recur- 
sos Vivos-REVIZEE.  We  are  grateful  to  Norte  Pesca  S.  A. 
and  to  Conselho  Nacional  de  Desenvolvimento  Cientifico 
e  Tecnologico-CNPq  for  scholarships  and  research  grants 
iProcs:  301048/83,  38.0726/96-3  and  820652/87-3). 


Literature  cited 

Aniorim,  A.  F.,  C.  A.  Arfelli,  and  L.  Fagundes. 

1998.     Pelagic  elasmobranchs  caught  by  longliners  off  South- 
ern Brazil  during  1974-97:  an  overview.     Mar.  Freshw. 
Res..  49:621-632. 
Batista,  V.  S.,  and  T.  C.  Silva. 

1995.    Age  and  growth  of  juveniles  of  junteiro  shark  Car- 
charhinus  porosus  in  the  coast  of  Maranhao,  Brazil.     Rev. 
Bras,  de  Biol.  55  (suppl.  I):25-32. 
Bartoo,  N.  V.,  and  K.  E.  Parker. 

1983.  Reduction  of  bias  generated  by  age-frequency  esti- 
mation using  the  von  Bertalanffy  growth  equation.  In 
Proceedings  of  the  international  workshop  on  age  determi- 
nation of  oceanic  pelagic  fishes:  tunas,  billfishes.  and  sharks 
(E,  D.  Prince  and  L.  M.  Pulos,  eds. ),  p.  25-26.  NOAA  Tech- 
nical Report  NMFS  8. 
Beamish,  R.  J.,  and  D.  A.  Fournier. 

1981.    A  method  for  comparing  the  precision  of  a  set  of  age 
determinations.     Can.  J.  Fish.  Aquat.  Sci.,  38:  982-983. 
Beamish.  R.  J.,  and  G.  A.  McFarlane. 

1983.  Validation  of  age  determination  estimates:  The  for- 
gotten requirement.  /;;  Proceedings  of  the  international 
workshop  on  age  determination  of  oceanic  pelagic  fishes: 
tunas,  billfishes,  and  sharks  (E.  D.  Prince  and  L.  M.  Pulos, 
eds. ),  p.  29-33.  NOAA  Technical  Report  NMFS  circular  8. 
Bigelow,  H.  B.,  and  W.  C.  Schroeder. 

1948.     Fishes  of  the  western  North  Atlantic,  lancelets, 
cyclostomes,  sharks,  576  p.     Memoirs  of  Sears  Foundation 
for  Marine  Research  [. 
Bonfil,  R.,  R.  Mena,  and  D.  Anda. 

1993.     Biological  Parameters  of  Commercially  Exploited  Silky 
Shark,  Carcharhinus  falciformis,  from  the  Campeche  Bank, 
Mexico.     NOAA  Technical  Report  NMFS  1 15:73-76. 
Branstetter,  S. 

1990.  Early  life-history  implications  of  selected  carcharhi- 
noid  and  lamnoid  sharks  of  the  northwest  Atlantic.  //; 
Elasmobranchs  as  living  resources:  advances  in  the  biol- 
ogy, ecology,  systematics,  and  the  status  of  the  fisheries 
(H.  Pratt,  S.  Gruber,  and  T.  Taniuchi,  eds.),  17-28.  NOAA 
Technical  Report  NMFS  90. 
Branstetter,  S.,  J.  A.  Musick,  and  J.  A.  Colvocoresses. 

1987.     A  comparison  of  the  age  and  growth  of  the  tiger  shark. 
Galcorcrdo  nivicri,  from  "II  Virginia  and  from  the  north 
western  gulf  of  Mexico.     Fish.  Bull.  85:269-279. 
Branstetter,  S.,  and  R.  Stiles. 

L987.  Age  and  growth  estimates  of  the  hull  shark,  Car- 
charhinus  leucas,  from  Northern  Gulf  of  Mexico.  Environ. 
Biol.  Fish.  20(31:169-181. 


Brothers,  E.  B. 

1983.  Summary  of  round  table  discussions  on  age 
validation.  In  Proceedings  of  the  international  workshop 
on  age  determination  of  oceanic  pelagic  fishes:  tunas,  bill- 
fishes, and  sharks  (E.  D.  Prince,  L.  M.  Pulos,  eds.),  p.  35-44. 
NOAA  Technical  Report  NMFS  8. 

Cadenat  J.,  and  J.  Blache. 

1981.  Requins  de  Mediterranee  et  d  Atlantique.  Faune  trop- 
icale,  330  p.  ORSTOM  i  Office  de  Recherche  Scientifique  et 
Technique  d'Outre-Mer)  21. 

Cailliet,  G. 

1990.  Elasmobranch  age  determination  and  verification:  an 
updated  review.  In  Elasmobranchs  as  living  resources: 
advances  in  biology,  ecology,  systematics  and  the  status  of 
the  fisheries  (H.  L.  Pratt  Jr.,  S.  H.  Gruber,  and  T  Taniuchi. 
eds.  I.  p.  157-165.     NOAA  Technical  Report  NMFS  90. 

Caillet,  G,  K.  L.  Martin,  D.  Kusher,  P.  Wolf,  and  B.  A.  Welden. 

1983.  Techniques  for  enhancing  vertebral  bands  in  age 
estimation  of  California  elasmobranchs.  In  Proceedings 
of  the  international  workshop  on  age  determination  of 
oceanic  pelagic  fishes:  tunas,  billfishes,  and  sharks  lE.  D. 
Prince,  L.  M.  Pulos,  eds.  I.  p.  157-165.  NOAA  Technical 
Report  NMFS  8. 

Carlson,  J.  K.,  E.  Cortes,  and  A.  G.  Johnson. 

1999.  Age  and  growth  of  the  blacknose  shark  Carcharhinus 
acronotus  in  the  eastern  Gulf  of  Mexico.  Copeia  1999(31: 
684-691. 

Cerrato,  R.  M. 

1990.     Interpretable  statistical  tests  for  growth  comparisons 

using  parameters  in  the  von  Bertalanffy  equation.     Can.  J. 

Fish.  Aquat.  Sci.  47(71:1416-1426. 
Compagno,  L.  J.  V. 

1984.  FAO  species  catalogue.  Vol.  4:  Sharks  of  the  world. 
Part  2:  Carcharhiniformes.  FAO  Fisheries  Synopsis  125: 
251-655.     FAO.  Rome. 

1988.     Sharks  of  the  order  Carcharhiniformes.  486  p.     Prince- 
ton Univ.  Press,  Princeton,  NJ. 
Cortes,  E. 

2000.  Life  history  and  correlations  in  sharks.  Rev.  Fish. 
Sci.  8(41:299-344. 

Crabtree,  R.  E.,  and  L.  H.  Bullock. 

1998.     Age,  growth,  and  reproduction  of  black  grouper. 
Mycteroperca  bonaci,  in  Florida  waters.     Fish.  Bull.  96: 
735-753. 
Fabens,  A.  J. 

1965.     Properties  and  fitting  of  the  von  Bertalanffy  growth 
curve.    Growth  29:265-289. 
Gayanilo,  F.  C.  Jr.,  P.  Sparre,  and  D.  Pauly. 

1996.     FAO   ICLARM   stock   assessment   tools.     User's 
manual.     FAO  computerized  information  series  (Fisher- 
ies) 8,  126  p  and  3  diskettes.     FAO,  Rome. 
Garrick,  J.  A.  F. 

1982.  Sharks  of  the  genus  Carcharhinus.  NOAA  Tech.  Rep. 
NMFS  circular.  194  p. 

1985.  Additions  to  a  revision  of  the  shark  genus  Carcharhi- 
nus: Synonymy  of  the  Aprionodon  and  Hypoprion  and 
descriptions  of  as  new  species  of  Carcharhinus  (Car- 
charhinidaei.     NOAA  Tech.  Rep.  NMFS  34,  26  p. 

Gruber,  S.,  and  R.  Stout. 

1983.  Biological  materials  for  the  study  of  age  and  growth  in 
tropical  marine  elasmobranch.  the  lemon  shark, Negaprion 
brevirostris  (Poey).  In  Proceedings  of  the  international 
workshop  on  age  determination  of  oceanic  pelagic  fishes: 
tunas,  billfishes,  and  sharks  (E.  D.  Prince,  and  L.  M.  Pulos, 
eds.),  193-205.     NOAA  Technical  Report  NMFS  8. 


Santana  and  Lessa:  Age  and  growth  of  Carcharhinus  signatus  off  the  northeastern  Brazilian  coast  167 


Guitart  Manday,  D. 

1975.     Las  pesquerias  pelago-oceanicas  de  corto  radio  de 
accion  en  la  region  noroccidental  de  Cuba,     sen  Oceanol. 
Acad.  Cienc.  Cuba  31,  26  p. 
Hazin,  F.  H.  V.,  J.  R.  Zagaglia,  M.  Broadhurst,  P.  Travassos,  and 
T.  R.  Q.  Bezerra. 

1998.     Review  of  a  small-scale  pelagic  longline  fishery  off 
northeastern  Brazil.     Mar.  Fish.  Rev.  60(3):l-8. 
Hazin,  F.  H.,  F  Lucena,  T.  S.  L.  Souza,  C.  Boeckman, 
M.  Broadhurst,  and  R.  Menni. 

2000.     Maturation  of  the  night  shark,  Carcharhinus  signatus, 
in  the  south-western  equatorial  Atlantic  Ocean.     Bull.  Mar. 
Sci.  66(1):173-185. 
Kappenman,  R.  F 

1981.     A  method  for  growth  curve  comparisons.     Fish.  Bull. 
79:95-101. 
Killam,  K.  A.,  and  G.  R.  Parsons. 

1989.    Age  and  growth  of  the  blacktip  shark,  Carcharhinus 
limbatus,  near  Tampa  Bay,  Florida.    Fish.  Bull.  87:845-857. 
Krefft,  G. 

1968.     Neue  und  erstmalig  nachgewiesene  Knorpelfische 
aus  dem  Archibenthal  des  Sudwestatlantiks,  einsliesslich 
einer  Diskussion  einiger  Etmopterus-Arten  Siidlicher 
Meere.    Arch.  Fischereiwiss.  19(11:1-72. 
Lessa,  R.,  and  F.  M.  Santana. 

1998.     Age  determination  and  growth  of  the  smalltail  shark 
Carcharhinus porosus  from  northern  Brazil.     Mar.  Freshw. 
Res.  49:705-711. 
Lessa,  R.,  V.  Batista,  and  Z.  Almeida. 

1999a.     Occurrence  and  biology  of  the  daggernose  shark 
Isogomphodon  oxyrhynchus  (Chondrichthyes:  Carcharh- 
inidae)  off  the  Maranhao  Coast  (Brazil).     Bull.  Mar.  Sci. 
64(11:115-128. 
Lessa,  R.,  F.  Santana,  R.  Menni,  and  Z.  Almeida. 

1999b.     Population  structure  and  reproductive  biology  of 
the  smalltail  shark  (Carcharhinus  porosus)  off  Maranhao. 
Mar.  Freshw.  Res.  50:383-388. 
Lessa,  R.,  F  M.  Santana,  and  R.  Paglerani. 

1999c.     Age,  growth  and  stock  structure  of  the  oceanic 
whitetip  shark,  Carcharhinus  longimanus,  from  the  south- 
western Equatorial  Atlantic.     Fish.  Res.  42:21-30. 
Lessa,  R.,  F.  M.  Santana,  V.  Batista,  and  Z.  Almeida. 

2000.     Age  and  growth  of  the  daggernose  shark,  Isogompho- 
don oxyrhynchus,  from  northern  Brazil.     Mar.  Freshw.  Res. 
51:339-347. 
Marin,  Y.,  F.  Brum,  L.  F  Barea,  and  J.  F  Chocca. 

1998.     Incidental  catch  associated  with  swordfish  longline 
fisheries  on  the  South-West  Atlantic  Ocean.     Mar.  Freshw. 
Res.  49:633-639. 
Menni,  R.  C,  F  H.  V.  Hazin,  and  R.  P.  Lessa. 

1995.  Occurrence  of  the  night  shark  Carcharhinus  signa- 
tus, and  the  pelagic  stingray  Dasyatis  violacea  off  north- 
eastern Brazil.     Neotropica  41(105-1061:105-110. 


Natanson,  L.  J.,  J.  G.  Casey,  and  N.  E.  Kohler. 

1995.  Age  and  growth  estimates  for  the  dusky  shark,  Car- 
charhinus obscurus,  in  the  western  North  Atlantic  Ocean. 
Fish.  Bull.  93:116-126. 

Natanson,  L.  J.,  and  N.  E.  Kohler. 

1996.  A  preliminary  estimate  of  age  and  growth  of  the  dusky 
shark  Carcharhinus  obscurus  from  the  South-West  Indian 
Ocean,  with  comparisons  to  the  Western  North  Atlantic 
population.     S.  Afr.  J.  Mar.  Sci.  17:  217-224. 

Prager,  M.  M.,  S.  B.  Saila,  and  C.  M.  Recksiek. 

1987.  FISHPARM:  a  microcomputer  program  for  parameter 
estimationof  nonlinear  models  in  fishery  science.  Techni- 
cal Report.  870:1-37.  Dep.  Oceanography,  Old  Dominion 
Univ.,  Norfolk,  VA. 

Ricker,W  E. 

1969.  Effects  of  size-selective  mortality  and  sampling  bias 
on  estimates  of  groowth,  mortality,  production  and  yield.  J. 
Fish.  Res.  Board  Can.  26(31:479-541. 

Seki,  T,  T.  Taniuchi,  H.  Nakano,  and  M.  Shimizu. 

1998.  Age,  growth  and  reproduction  of  the  oceanic  whitetip 
shark,  Carcharhinus  longimanus  from  the  Pacific  Ocean. 
Fish.  Sci.  64(11:14-20. 

Shepherd,  J.  G. 

1987.  A  weakly  parametric  method  for  estimating  growth 
parameters  from  length  composition  data.  In  Length- 
based  methods  in  fisheries  research.  ICLARM  (Interna- 
tional Center  for  Living  Resources  Management)  conference 
proceedings  13  (D.  Pauly  and  G.  Morgan,  eds.),  p.  113-119. 
ICLARM,  Manila,  and  the  Kuwait  Institute  for  Scientific 
Research,  Safat,  Kuwait. 

Sminkey,  T.  R.,  and  J.  A.  Musick. 

1995.  Age  and  growth  of  the  sandbar  shark,  Carcharhinus 
plumbeus,  before  and  after  population  depletion.  Copeia 
1995(4):871-883. 

Sparre,  P.,  E.  Ursin,  and  S.  C.  Venema. 

1989.     Introduction  to  tropical  fish  stock.  Part  1.  Manual. 
FAO  Fisheries  Tech.  Paper  306. 1,  337  p.     FAO,  Rome. 
Suzuki,  Z..  Y.  Warashima,  and  M.  Kishida. 

1977.    The  comparison  of  catches  by  regular  and  deep  tuna 
longline  gear  in  the  western  and  central  equatorial  Pacific. 
Bull.  Far  Seas  Fish.  Res.  Lab.  15:51-89. 
von  Bertalanffy,  L. 

1938.    A  quantitative  theory  of  organic  growth.     Hum.  Biol. 
10:181-213. 
Wintner,  S.,  and  G.  Cliff. 

1996.  Age  and  growth  determination  of  the  blacktip  shark, 
Carcharhinus  limbatus,  from  the  east  coast  of  South 
Africa.    Fish.  Bull.  94:135-144. 

Zar,  J.  H. 

1996.  Biostatistical  analysis,  3rd  ed.,  662  p.  Prentice  Hall, 
Upper  Saddle  River,  NJ. 


168 


Abstract— The  prowfish  (Zaprora  sile- 
nus) is  an  infrequent  component  of 
bottom  trawl  catches  collected  on  stock 
assessment  surveys.  Based  on  pres- 
ence or  absence  in  over  40,000  trawl 
catches  taken  throughout  Alaskan 
waters  southward  to  southern  Cali- 
fornia, prowfish  are  most  frequently 
encountered  in  the  Gulf  of  Alaska  and 
the  Aleutian  Islands  at  the  edge  of  the 
continental  shelf  Based  on  data  from 
two  trawl  surveys,  relative  abundance 
indicated  by  catch  per  swept  area 
reaches  a  maximum  between  100  m 
and  200  m  depth  and  is  much  higher 
in  the  Aleutian  Islands  than  in  the 
Gulf  of  Alaska.  Females  weigh  3.7% 
more  than  males  of  the  same  length. 
Weight-length  functions  are  W  (gl  = 
0.0164  L292  (males)  and  W  =  0.0170 
L292  (females).  Length  at  age  does  not 
differ  between  sexes  and  is  described 


by  L  =  89.3(1  -  e 


-0  181i;+0554l 


),  where 


L  is  total  length  in  cm  and  t  is  age  in 
years.  Females  reached  50r/f  maturity 
at  a  length  of  57.0  cm  and  an  age  of  5.1 
years.  Prowfish  diet  is  almost  entirely 
composed  of  gelatinous  zooplankton, 
primarily  scyphozoa  and  salps. 


Manuscript  approved  for  publication 
20  September  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Mull.  102:168-178  (2004). 


Distribution  and  biology  of  prowfish 
(Zaprora  silenus)  in  the  northeast  Pacific 

Keith  R  Smith 
David  A.  Somerton 
Mei-Sun  Yang 
Daniel  G.  Nichol 

Alaska  Fisheries  Science  Center 

National  Marine  Fisheries  Service 

7600  Sand  Point  Way  NE 

Seattle,  Washington  98115 

E-mail  address  (for  K.  R.  Smith,  contact  author)  Keith.Smith@noaa.gov 


Current  taxonomy  distinguishes  the 
prowfish  (Zaprora  silenus,  Fig.  1)  as 
the  only  species  and  the  only  genus  of 
the  family  Zaproridae.  Other  families  in 
the  encompassing  suborder  Zoarcoidei 
include  Bathymasteridae  (ronquils), 
Cryptacanthodidae  (wrymouths),  Pholi- 
dae  (gunnels),  and  Stichaeidae  (prick- 
lebacks).  Systematics  of  most  families 
within  Zoarcoidei,  and  of  the  suborder 
itself  within  the  order  perciforms,  is 
uncertain  (Nelson,  1994).  Prowfish 
(adult)  physical  features  include  an 
elongate,  laterally  compressed  body;  a 
high  convex  brow  and  interorbital  area 
ending  with  a  short  blunt  snout;  and 
a  distinctive  protruding  area  below  a 
slightly  upturned  terminal  mouth. 
Fins  consist  of:  a  long,  moderately  high 
dorsal  fin;  a  moderately  long  anal  fin; 
a  discrete  truncate  caudal  fin  with  a 
short,  broad  peduncle;  and  moderately 
large,  rounded  pectoral  fins  (pelvic  fins 
are  absent).  Teeth  are  small,  sharp,  and 
close-set  in  a  single  row  attached  only 
to  the  jaws.  Scales  are  ctenoid.  Numer- 
ous distinctive  large  round  pores  occur 
on  the  sides  and  top  of  the  head.  Color 
is  olive-gray  to  brown  dorsally,  shading 
lighter  below,  suffused  on  the  sides 
and  back  with  many  small  dark  spots 
(Clemens  and  Wilby,  1961;  Eschmeyer 
et  al,  1983;  Hart,  1973;  Kessler,  1985). 
The  maximum  length  reported  is  more 
than  1  m  (Tokranov,  1999). 

Since  its  original  description  (Jordan, 
1897),  the  prowfish  has  been  observed 
infrequently  despite  numerous  and 
extensive  bottom  trawl  surveys  com- 
prising thousands  of  net  deployments 
off  Alaska  and  the  west  coast  of  North 


America.  It  is  not  clear  whether  this 
lack  of  documentation  indicates  a  spe- 
cies of  low  abundance  or  a  preference 
by  prowfish  for  a  habitat,  such  as  rough 
rock  substrate  or  steep  bottom  gradi- 
ents, that  is  poorly  sampled  by  bottom 
trawl  surveys.  Nevertheless,  the  spe- 
cies is  common  enough  to  be  considered 
representative  of  the  ichthyofauna  of 
certain  benthic  biotopes  within  its 
range  (Allen  and  Smith,  1988;  Tokranov, 
1999).  It  has  been  encountered  at  loca- 
tions along  the  outer  continental  shelf 
and  upper  slope  ranging  in  a  long  arc 
from  San  Miguel  Island,  California, 
north  through  the  Gulf  of  Alaska,  west 
through  the  Bering  Sea  and  Aleutian  Is- 
lands to  the  Asiatic  shelf,  thence  south 
to  Hokkaido,  at  depths  of  10-675  m 
(Allen  and  Smith,  1988;  Hart,  1973).  In 
addition  to  occurring  in  the  catches  on 
biological  surveys,  prowfish  have  been 
taken  incidentally,  and  occasionally 
processed,  in  commercial  fishing  op- 
erations on  the  outer  continental  shelf 
(Smith,  pers.  obs.;  Berger1). 

Prowfish  are  known  to  be  pelagic 
as  pre-adults  (Hart,  1973;  Doyle  et  al, 
2002).  After  larval  transformation  at 
30  mm  (Matarese  et  al.,  1989),  juve- 
niles maintain  close  proximity  to  the 
medusae  of  pelagic  cnidarians  (Schef- 
fer,  1940).  Brodeur  (1998)  observed 
juveniles  swimming  near  the  bells  of 
scyphomedusae  Cyanea  capillata  and 
Chrysaora  melanaster  and  retreating 


1  Berger,  J.  2002.  Personal  commun. 
Alaska  Fisheries  Science  Center,  National 
Marine  Fisheries  Service,  7600  Sand  Point 
Way  NE,  Bldg  4,  Seattle,  WA  98115-0070. 


Smith  et  al.:  Distribution  and  biology  of  Zaprora  silenus 


169 


Figure  1 

Aquarium  prowfish  specimen,  National  Marine  Fisheries  Service,  Kodiak  Laboratory, 
Kodiak,  AK.  Photograph  by  Jan  Haaga. 


behind  the  tentacles  or  within  the  bells  of  these  jellyfish 
when  approached  by  a  remotely  operated  vehicle,  appar- 
ently as  a  means  of  protection  from  predators.  Prowfish  are 
also  believed  to  later  become  demersal  and  have  a  prefer- 
ence for  rocky  areas  (Tokranov.  1999). 

The  association  with  scyphomedusae  and  other  large  ge- 
latinous zooplankton  exhibited  by  juveniles  may  continue 
throughout  their  lives,  because  such  prey  are  reported  to 
constitute  a  considerable  portion  of  the  prowfish  diet  (Car- 
olio  and  Rankin,  1998).  In  the  stomachs  of  16  juveniles 
of  5-13.3  cm  total  length  captured  at  midwater  depths  in 
Prince  William  Sound  in  1995,  Sturdevant2  found  prey 
biomass  was  composed  principally  of  hyperiid  amphipods 
but  also  found  unquantifiable  gelatinous  matter  which  was 
thought  to  be  the  remains  of  jellyfish  tentacles. 

Little  is  known  regarding  possible  predators  of  prow- 
fish, the  relative  frequency  of  prowfish  among  prey  items, 
or  the  sizes  of  prowfish  consumed.  Prowfish  have  been 
found  in  the  diets  of  diving  seabirds  and  have  comprised 
25%  of  food  biomass  delivered  to  tufted  puffin  {Lunda  cir- 
rhata )  chicks  (Sturdevant2).  Yang  ( 1993)  found  prowfish  in 
only  0.3%  of  467  stomachs  of  Pacific  halibut  (Hippoglossus 
stenolepis)  taken  by  bottom  trawl  in  the  Gulf  of  Alaska  in 
1990,  accounting  for  0.03%  (by  weight)  of  total  food  pres- 


2  Sturdevant,  M.  V.  1999.  Forage  fish  diet  overlap,  1994-1996. 
Exxon  Valdez  oil  spill  restoration  project  final  report  (restoration 
project  97163C),  184  p.  Alaska  Fish.  Sci.  Cent.,  Auke  Bay  Labo- 
ratory, Natl.  Mar.  Fish.  Serv.,  NOAA,  Juneau,  AK.  [Available  by 
order  no.  PB2000- 100700  from  Natl.  Tech.  Info.  Serv,  5285  Port 
Royal  Rd.,  Springfield,  Virginia  22161.] 


ent.  Orlov  (1998)  found  prowfish  in  0.13%  of  stomachs  of 
white-blotched  skate  (Bathyraja  metadata)  caught  by  bot- 
tom trawl  off  the  Northern  Kuril  Islands  and  Southeastern 
Kamchatka  in  1996.  In  comparisons  of  proximate  composi- 
tion among  17  taxa  of  forage-size  fish  from  the  northeast- 
ern Pacific  (Van  Pelt  et  al,  1997;  Payne  et  al,  1999),  juvenile 
prowfish  averaged  highest  in  moisture  content  (86-88%  by 
weight)  and  relatively  low  in  lipids  (10. 8±1.3%,  dry  weight 
analysis). 

In  this  study  we  examined  information  on  this  little- 
known  species,  investigating  spatial  and  depth  distribu- 
tions, size  frequency,  growth,  reproduction,  and  diet  in  the 
waters  off  Alaska. 


Materials  and  methods 

Data  and  sample  collection 

Data  used  in  this  investigation  were  collected  during 
bottom  trawl  surveys  for  groundfish  and  invertebrate 
stocks  conducted  by  the  Resource  Assessment  and  Con- 
servation Engineering  (RACE)  Division  of  the  Alaska  Fish- 
eries Science  Center  (APSC),  National  Marine  Fisheries 
Service.  Areas  surveyed  were  the  continental  shelf  and 
upper  slope  of  the  eastern  Bering  Sea,  Aleutian  Islands 
region  (Al),  Gulf  of  Alaska  (GOA),  and  west  coast  of  North 
America  from  Washington  to  California.  Trawl  catches 
were  sorted  to  species,  weighed,  and  individuals  were 
counted,  following  procedures  described  in  Wakabayashi 
etal.  (1985). 


170 


Fishery  Bulletin  102(1) 


To  characterize  prowfish  distribution  we  obtained  catch 
data  from  42,601  bottom  trawl  deployments  (hauls)  exe- 
cuted from  1953  through  2000  using  a  variety  of  net  de- 
signs. We  used  these  data  to  determine  presence  or  absence 
of  prowfish  at  each  haul  location.  Previous  observations 
have  indicated  that  prowfish  tend  to  be  pelagic  as  larvae 
and  become  demersal  as  adults  (Matarese  et  al.,  1989; 
Hart,  1973).  A  full  accounting  of  prowfish  distribution  by 
life  stage  is  beyond  the  scope  of  this  investigation,  which 
focuses  on  adults.  Therefore  we  confined  our  observations 
to  haul  catches  taken  on  bottom,  as  opposed  to  in  mid-water 
or  at  the  surface. 

On  two  of  the  bottom  trawl  surveys,  one  in  the  Gulf  of 
Alaska  from  22  May  to  30  July  1996,  and  the  other  in  the 
Aleutian  Islands  from  10  June  to  11  August  1997,  addi- 
tional prowfish  data  were  collected.  Consistency  between 
these  surveys  in  sampling  procedures  and  equipment 
(Martin,  1997,  and  Stark3)  facilitated  subsequent  data 
comparisons. 

Density  of  prowfish  at  each  sampling  location  was 
estimated  as  the  number  caught  divided  by  the  km2  of 
area  swept  by  the  trawl  (catch  per  unit  of  effort,  or  CPUE). 
Research  vessels  on  both  surveys  employed  the  standard 
RACE  Division  model  poly-Nor'eastern  high-opening  bottom 
trawl  net  with  roller  gear,  and  hauls  were  made  during  day- 
light. Net  configuration  and  bottom  contact  during  trawling 
were  monitored  by  Scanmar  instrumentation.  Data  were  ob- 
tained from  807  hauls  in  the  Gulf  of  Alaska  and  408  hauls  in 
the  Aleutian  Islands.  The  average  area  swept  per  haul  was 
0.025  km2  in  the  GOA  and  0.024  km2  in  the  AI. 

All  prowfish  were  sorted  to  sex  by  examination  of  the 
gonads  and  then  length  (total  length;  cm)  was  measured. 
Sample  sizes  were  84  males  and  90  females  for  the  Gulf 
of  Alaska;  396  males  and  431  females  for  the  Aleutian 
Islands.  Whole-body  weights  (g)  of  83  male  and  88  female 
prowfish  from  the  Gulf  of  Alaska  were  measured  and  the 
sagittal  otoliths  were  removed  and  stored  in  50%  ethanol. 
Whole  ovaries  from  a  representative  subsample  of  39  of 
the  females  were  removed,  frozen,  and  later  stored  in  10% 
buffered  formalin  solution. 

Diet  composition  was  examined  from  stomach  contents  of 
76  individuals  (18  from  the  Gulf  of  Alaska  and  58  from  the 
Aleutian  Islands).  Stomachs  containing  food  and  with  no 
signs  of  regurgitation  or  net-feeding  (e.g.  the  stomach  was 
in  an  inverted  or  flaccid  state  or  there  was  the  presence  of 
prey  in  the  mouth  or  around  the  gills)  were  removed  and 
preserved  in  10%  buffered  formalin. 

Laboratory  procedures 

Standard  otolith-prcparation  techniques  for  age  determi- 
nation were  modified  to  accommodate  the  relatively  small 
size  of  prowfish  otoliths  (usually  <5  mm  long).  An  anterior 
portion  of  each  otolith  was  removed  by  a  transverse  cut  with 
scalpel  perpendicular  to  the  sagittal  axis  and  anterior  to  the 


:!  Stark,  J.  1998.  Report  to  industry:  fishing  log  for  the  1997 
bottom  trawl  survey  of  the  Aleutian  Islands.  AFSC  Proc.  Rep. 
98-06, 96  p.  Alaska  Fish.  Sci.  Cent.,  Natl.  Mar.  Fish.  Serv.,  NOAA. 
7600  Sand  Point  Way  NE,  Bldg.  4,  Seattle,  WA  98115-0070. 


nucleus.  The  remainder,  which  contained  the  nucleus,  was 
baked  at  300-475°C  for  up  to  17  min  or  heated  over  an  alco- 
hol flame  to  enhance  visibility  of  annuli.  The  otoliths  were 
then  individually  mounted  on  slides  by  completely  embed- 
ding them  in  clear  thermoplastic  posterior  end  down.  On 
hardening,  each  mount  was  wet-sanded  on  increasingly  fine 
grades  of  sandpaper  (400-2000  grit),  parallel  to  the  slide, 
until  the  surface  intersected  the  otolith  nucleus  (trans- 
verse section).  Preparing  readable  mounts  was  a  delicate 
procedure;  besides  cutting  and  polishing  the  small  otoliths 
precisely  without  fracturing  them,  precise  heating  tem- 
perature and  time  were  especially  critical  to  expose  annuli 
without  again  causing  fractures  or  burning  the  otolith. 
Our  method  had  advantages  over  the  standard  "break  and 
burn"  method  of  simply  coating  the  surface  of  a  temporarily 
mounted  specimen  with  oil  to  enhance  visibility  of  annuli, 
in  lieu  of  polishing.  It  allowed  a  more  precise  intersection  of 
the  nucleus  by  the  viewed  surface  and  eliminated  the  need 
to  remove  oil  from  specimens  intended  for  further  viewing 
in  order  to  prevent  blurring  of  annuli.  After  preparation, 
slides  were  placed  in  sufficient  water  to  cover  the  surface 
scratches  and  were  examined  under  a  dissecting  microscope 
with  reflected  light.  Age  in  years  was  determined  by  count- 
ing the  annuli  or  hyaline  bands  according  to  the  criteria 
described  in  Chilton  and  Beamish  (1982). 

Prowfish  ovaries  were  prepared  for  histological  examina- 
tion by  removing  a  small  portion  from  the  middle  of  each 
ovary,  which  was  then  embedded  in  paraffin,  sectioned  at 
6  jim,  and  stained  with  hematoxylin  and  eosin.  The  histo- 
logical slides  were  examined  under  a  compound  microscope 
and  donor  females  were  classified  as  either  sexually  imma- 
ture or  mature  based  on  the  presence  of  yolk  in  the  oocytes 
(i.e.  vitellogenesis). 

Prowfish  stomachs  were  processed  by  first  neutralizing 
the  10%  formalin  used  for  initial  fixation  and  then  by  im- 
mediately transferring  the  stomachs  into  70%  ethanol.  The 
food  was  removed,  blotted  with  a  paper  towel,  and  exam- 
ined with  a  dissecting  microscope.  Prey  items  were  sorted 
to  the  lowest  practical  taxonomic  level  and  then  weighed 
to  the  nearest  0.1  gm.  The  percentage  of  total  prey  weight 
which  each  taxon  comprised,  as  well  as  the  percentage  of 
stomachs  containing  each  taxon,  was  calculated  for  each 
haul  sample  and  then  estimated  for  each  of  the  two  regions 
as  the  average  of  the  per-haul  percentages. 

Analysis  of  data 

The  distribution  of  prowfish  density  over  depth  in  the 
Gulf  of  Alaska  and  the  Aleutian  Islands  was  determined 
by  calculating  the  mean  CPUE  for  each  20-m  depth  inter- 
val from  20  m  to  480  m.  Both  surveys  utilized  a  stratified 
sampling  design  in  which  sampling  density  (hauls  per 
unit  area)  varied  by  geographical  subarea  (Martin,  1997; 
Stark1).  To  compensate  for  this  variation,  the  CPUE  of  each 
haul  was  weighted  by  the  inverse  of  the  sampling  density 
in  that  geographic  stratum.  The  mean  bottom  depth  as 
weighted  by  prowfish  density  was  calculated  for  each  of 
the  two  regions  as  the  weighted  average  of  the  midpoints 
of  the  depth  intervals,  where  the  weighting  factors  were 
the  interval-mean  CPUE  values. 


Smith  et  al.:  Distribution  and  biology  of  Zaprora  silenus 


171 


Frequency  distributions  of  prowfish  total  length,  sepa- 
rated by  sex  and  region,  were  calculated  as  the  weighted 
percent  of  measurements  within  10-cm  length  intervals. 
The  weighting  factors  were  calculated  for  each  fish  mea- 
surement as  the  inverse  of  geographic-stratum  sampling 
(i.e.  haul)  density,  multiplied  by  the  inverse  of  the  indi- 
vidual haul  area  that  was  swept .  Also,  differences  in  mean 
length  between  sexes  or  regions  were  examined  by  using 
analysis  of  variance  (ANOVA)4  to  test  the  significance  of 
statistical  differences  based  on  the  weighted  lengths.  Be- 
cause potential  grouping  of  prowfish  by  size  could  affect 
within-haul  variance,  source  haul  (i.e.  that  in  which  each 
measured  fish  was  caught)  was  included  in  analyses  as 
a  possible  random  variable  affecting  length.  Variances  in 
length  between  regions  and  sexes  were  each  tested  for  sig- 
nificance against  variance  among  hauls.  The  significance 
of  the  haul  variable  was  also  checked  by  testing  variance 
among  hauls  against  that  among  measurements. 

The  relationship  of  body  weight  (g;  W)  to  total  length 
(cm;  L)  was  assumed  to  be  an  exponential  function: 

W  =  e"U\ 

for  which  the  parameters  a  and  ft  were  estimated  from 
the  data  by  first  log-transforming  both  variables  and  then 
calculating  the  intercept  and  slope  of  the  least  squares 
linear  regression: 

\n(W)  =  a  +  p\n(L). 

To  determine  whether  the  relationship  differed  by  sex, 
analysis  of  covariance  ( ANCOVA;  Statgraphics  Plus,  Manu- 
gistics.  Inc.,  Rockville,  MD)  was  used  to  compare  the  fit  of 
a  model  with  two  regression  lines,  each  with  a  sex-specific 
intercept  and  slope,  to  the  fit  of  a  two-line  model  with  sex- 
specific  intercepts  and  a  common  slope  (null  hypothesis). 
If  no  significant  difference  was  observed,  then  a  second 
test  was  performed  by  testing  the  latter  model  against  the 
null  hypothesis  of  a  common  regression  line  with  single 
intercept  and  slope  for  both  sexes  combined.  The  relation- 
ships in  the  best-fit  model  were  then  transformed  back  to 
exponential  form. 

Prowfish  growth  was  described  by  fitting  the  von  Berta- 
lanffy  function  to  length  (L)  and  age  (year;  t )  data  by  using 
nonlinear  least  squares.  The  function  is 


L=L  (1 


-«'i/-(,ii) 


where  L^  =  asymptotic  maximum  length; 

k    -  a  constant  (per  year)  affecting  model  early 
growth  rate;  and 


U 


hypothetical  age  at  0  length. 


To  determine  whether  parameters  differed  between  prow- 
fish sexes,  we  fitted  the  function  separately  to  the  data  from 
each  sex  as  well  as  to  the  data  for  both  sexes  combined.  A 


4  Unless  otherwise  specified,  ANOVA,  log-likelihood,  and  nonlin- 
ear regression  analyses  were  accomplished  by  using  Systat  10 
software  (Systat  10  Statistics  I,  SPSS  Inc. .Chicago,  ID. 


likelihood  ratio  test  was  then  used  to  determine  whether 
the  separate-sex  model  fitted  the  data  significantly  better 
than  the  combined-sex  model  (Kimura,  1980).  Significance 
of  the  likelihood  ratio  was  based  on  the  chi-squared  sta- 
tistic with  degrees  of  freedom  equal  to  the  difference  in 
number  of  parameters  between  the  two  models. 

The  proportion  of  prowfish  females  that  were  mature 
(Pmo()  at  a  given  length  or  age  was  described  with  logistic 
functions  of  the  formPma,  =  1/(1  +  e"+'iV),  where  X  is  either 
length  (L)  or  age  in  years  (f),  and  a  and  /3  are  function  pa- 
rameters. The  models  were  fitted  to  the  data  by  using  maxi- 
mum likelihood.  After  the  relationships  were  estimated, 
the  length  and  age  at  which  50%  of  females  were  mature 
were  estimated  by  setting  Pmat  =  0.5  in  each  function  and 
solving  forX  The  959r  confidence  interval  for  each  estimate 
was  calculated  by  using  the  delta  method  (Seber,  1973). 


Results 

Geographic  distribution 

Prowfish  distribution  in  the  waters  off  Alaska,  as  indicated 
by  their  presence  at  1528  out  of  a  total  of  35,159  histori- 
cal bottom  trawl  locations,  is  shown  in  Figure  2.  The  total 
count  of  individuals  in  catches  was  11,401.  Distribution 
south  of  approximately  50°N  latitude  off  Vancouver  Island 
is  not  shown  because  here  6  of  7442  bottom  trawl  hauls 
caught  a  total  of  8  prowfish.  The  southernmost  occurrence 
was  at  34°13.4'N  latitude  near  San  Miguel  Island,  southern 
California.  Prowfish  were  taken  at  depths  ranging  from  24 
m  to  801  m  but  most  frequently  appeared  in  catches  close 
to  the  break  between  the  continental  shelf  and  upper  con- 
tinental slope  near  200  m  depth. 

Prowfish  CPUE  was  greater  than  zero  at  64  of  807  haul 
locations  in  the  Gulf  of  Alaska  in  1996  and  at  48  of  408 
locations  in  the  Aleutian  Islands  in  1997.  Over  all  areas  at 
the  depths  fished  the  range  of  per-haul  CPUE  was  0-547.5 
prowfish/km2  (average=6.7  prowfish/km2 )  in  the  GOA  and 
0-5220.1  prowfish/km-  (average=65.1  prowfish/km2)  in 
the  AI.  The  average  CPUE  within  20-m  bottom  depth  in- 
tervals in  each  region  indicated  that  fish  tend  to  be  most 
concentrated  at  intermediate  depths  (Fig.  3).  Depth  at 
trawl  locations  ranged  from  20  to  479  m  for  the  GOA  and 
from  22  to  474  m  for  the  AI,  and  prowfish  were  collected 
at  34-252  m  (GOA)  and  89-258  m  (AI),  respectively.  The 
CPUE-weighted  average  bottom  depth  was  163.8  m  for  the 
GOA  and  150.3  m  for  the  AI. 

The  CPUE  values  within  20-m  depth  intervals  (Fig.  3) 
indicated  that  the  regional  difference  in  mean  density  was 
largely  due  to  differences  at  the  same  depth  rather  than 
differences  between  regions  in  the  amount  of  area  available 
at  a  given  bottom  depth. 

Length  distribution 

Length-frequency  histograms  by  region  and  sex  for  prowfish 
from  the  Gulf  of  Alaska  (84  males,  90  females)  and  Aleutian 
Islands  (396  males,  431  females)  are  shown  in  Figure  4. 
Analysis  of  variance  tests  for  a  difference  in  mean  length 


172 


Fishery  Bulletin  102(1) 


65  °N 


HI    \ 


55  °N 


65  °N 


-  60°N 


-  ^  \ 


170°E 


65°W 


Ii.ii  \\ 


16(1   \\ 

155°W 

150°W                145°W 

40°W 

135°W 

130°W 

65°N- 

B 

Alaska 

60  °N- 

"V^ 

^ 

^ 

t0r 

:''  *UK5i8Si^ 

55°N- 

y^tBs 

N 

A 

Bathymetry  (m)                    Noprowf 
20Q                ♦            Prowfish 

sh  caught 
caught 

Ml    \  - 

h 

1 

1 

1 — 

1 1 1 

65°N 


hi  \ 


^  \ 


50  \ 


K.ll  W 


155°W 


150  \\ 


145  \\ 


140  \\ 


135°W 


130  \\ 


Figure  2 

Lm;  il  inns  i  if  Alaska  Fisheries  Science  Center  groundfish  survey  I  mt  turn  trawls  I  prior  to  year 
2001)  in  (Ai  the  eastern  Bering  Sea  and  the  Aleutian  Islands  region  and  (B)  the  Gulf  of 
Alaska,  indicating  trawls  in  which  prowfish  occurred. 


Smith  et  al.:  Distribution  and  biology  of  Zaprora  silenus 


173 


30  70         110        150        190        230       270        310 

Bottom  depth  (m) 


350       390       430       470 


Figure  3 

Average  prowfish  catch  per  unit  of  effort  (CPUE;  no/km2)  within  20-m  intervals  over  the 
range  of  trawl  depths  of  20-479  m  in  the  Gulf  of  Alaska  (GOA)  in  1996  (A)  and  22-474  m 
in  the  Aleutian  Islands  (AD  in  1997  (B).  The  mean  of  interval  midpoints,  each  weighted 
by  interval  average  CPUE,  was  163.8  m  for  the  GOA  and  150.3  m  for  the  AI. 


between  sexes  were  not  significant  for  either  the  GOA 
(P=0.83)  or  the  AI  (P=0.76).  Although  the  weighted  mean 
for  both  sexes  combined  was  61.0  cm  (range:  11-90  cm)  in 
the  GOA  and  51.9  cm  (range:  25-87  cm)  in  the  AI,  the  differ- 
ence in  length  between  regions  was  not  significant  (P=0.1 1 ). 
Grouping  of  prowfish  of  similar  size  within  hauls  was  highly 
significant  in  both  the  GOA  and  the  AI  (P«0.01 ). 

Weight-length  relationship 

In  the  between-sex  ANCOVA  comparison  of  the  linearized 
(i.e.  log-transformed)  weight-to-length  relationships  based 
on  prowfish  caught  in  the  Gulf  of  Alaska,  the  slopes  were 
not  significantly  different  between  sexes  (P=0.38).  How- 
ever, the  difference  in  intercepts  was  significant  (P=0.044 ). 
Thus  the  best  fitting  model  varied  by  sex  with  two  regres- 
sion lines  of  equal  slope  but  with  sex-specific  intercepts. 
The  equivalent  functions  in  terms  of  the  untransformed 
variables  (Fig.  5)  were 


and 


Wmales  =  0.0164  xL2922; 


W/mwles  =  0.017x7^22. 


The  model  indicated  that  adult  females  are,  on  average, 
3.7%  heavier  than  males  of  the  same  length. 


Age  and  growth 

Readable  otolith  specimens  were  produced  for  138  prowfish 
(71  males,  67  females)  of  the  172  from  which  samples  were 
collected.  Production  of  readable  specimens  did  not  appear 
related  to  fish  size  or  age.  The  likelihood  ratio  test  for  a 
difference  between  males  and  females  in  the  relationship 
of  length  to  age  was  not  significant  (P=0.53),  indicating 
that  there  was  no  difference  in  growth  between  sexes.  The 
best-fit  von  Bertalanffy  function  (Fig.  6)  had  the  following 
parameters  (with  95%  confidence  intervals):  La  =  89.33 
±6.5  cm;  k  =  0.18  ±0.05/year;  and  t()  =  -0.55  ±0.12  year. 

Female  maturity 

The  proportions  of  females  that  were  mature  were  highly 
significant  logistic  functions  of  length  and  age  (P<0.005; 
Fig.  7).  The  fitted  functions  of  length  and  age  were 


and 


Pmo,=  l/(l+e37114-6-51L); 


1/(1  +e9.66-1.90(). 


The  theoretical  length  and  age  at  which  50%'  of  females 
were  mature,  with  respective  95%  confidence  limits,  were 
57.0  ±0.4  cm  and  5.1  ±0.7  years. 


174 


Fishery  Bulletin  102(1) 


40  50 

Length  (cm) 

Figure  4 

Frequency  distribution  of  total  lengths  of  (A)  84  male  and  90  female  prowfish  from  the 
Gulf  of  Alaska  (GOA;  1996)  and  (B)  396  males  and  431  females  from  the  Aleutian  Islands 
(AI;  1997).  Gray  bars  show  the  percentage  of  male  lengths  and  black  bars  the  percentage 
of  female  lengths  within  continuous  10-cm  intervals  (e.g.  25-35  cm). 


9000  -I 

8000  ■ 

f 
O  Males                                                                                       °rJy> 

7000  ■ 

~     6000  ■ 

£      5000  ■ 
g> 

|j      4000  ■ 

□  Females                                                                       UfkJ/^B 

wmales  =  o.oi  64  x  u**                                   'jijr      ° 

—  w,emates  =  o.oi  79  *  l*  «*                      afljllo       ° 

3000  ■ 

2000  J 

r2  (combined  model)  =  0.95                      _^ffl^ 

1000- 

^«e**^ 

0                            20                           40                           60                           80                          100 

Length  (cm) 

Figure  5 

Prowfish  body  weight  (W)  fitted  by  an  exponential  function  of  fish  total  length  (/..land  sex. 

Data  for  males  (/i=83)  are  shown  by  diamonds  and  the  fitted  model  by  a  dashed  line.  Data 

for  females  (n=88)  are  shown  by  squares  and  the  fitted  model  by  a  solid  line. 

Food  habits 

Fish  used  for  diet  study  averaged  63.8  cm  in  total  length 
(range:  49-87  cm)  in  the  Gulf  of  Alaska  and  56.9  cm  (range: 


30-79  cm)  in  the  Aleutian  Islands.  The  contents  of  18 
prowfish  stomachs  from  the  Gulf  of  Alaska  and  58  from 
the  Aleutian  Islands  showed  that  jellyfish  (999r  and  31% 


Smith  et  al.:  Distribution  and  biology  of  Zaprora  silenus 


175 


100 
90 

80 

70 

E     60 

£     50  \ 

S     40 

30 

20 

10 

0 


L  =  89  33(1-e-°,8'<'-<-°554») 
r  =  0.752 


10  15 

Age  (years) 


20 


25 


Figure  6 

Prowfish  total  length  \L)  fitted  by  a  von  Bertalanffy  function  of  age  (f).  Data  for  males 
(«=71)  are  shown  by  diamonds;  data  for  females  (rc=67)  are  shown  by  squares.  The  fitted 
model  is  shown  by  a  solid  line. 


Table  1 

Mean  percent  weight  (%W)  and  mean  percent  freque 

icy  of  occurrence  (%FO)  of  the  prey  items  from  18  prowfish  stomachs  collected 

in  the  Gulf  of  Alaska  (GOA;  1996; 

total  prev 

weight= 

=299  g 

and  58  stomachs  from  the  Aleutian 

Islands 

area  (Al;  1997;  total  prey 

weight=1446.6  g).  Sample  prowfish  had  an  average  total  length  of  63.8  cm 

(range 

49 

-87  cm)  from  the  GOA  and  and  56.9  cm  (range: 

30-79  cm)  from  the  AI. 

Prey  name 

GOA(n  = 

18) 

AI(n 

=58) 

%W 

%FO 

%W 

%FO 

Scyphozoa  (jellyfish) 

98.84 

100 

30.45 

29.88 

Ctenophora  (comb  jelly) 

0.09 

1.23 

Polychaeta  (worm) 

0.03 

5.8 

Calanoida  (copepod) 

0.26 

28.13 

0.04 

29.14 

Thysanoessa  raschii  (euphausiid) 

0.05 

6.67 

Mysidacea  Mysida  (mysid) 

0.01 

3.13 

Hyperiidea  (amphipod) 

0.19 

33.46 

Gammaridea  (amphipod) 

0.12 

30.49 

Themisto  sp.  (amphipod) 

0.32 

28.57 

0.14 

36.91 

Salpa  sp.  (pelagic  salp) 

34.06 

46.79 

Larvacea  (pelagic  tunicate  1 

0.13 

12.5 

Sebastes  sp.  (rockfish)  larvae,  5-8 

mm  long 

0.43 

42.86 

Microsomus  paeifieus  (Dover  sole) 

eggs 

0.01 

3.13 

Unidentified  organic  material 

34.84 

32.59 

by  weight  of  total  food  in  the  two  regions,  respectively) 
and  gelatinous  pelagic  tunicates  (Salpa  spp.;  34%  in  the 
Aleutian  Islands  area  only)  were  the  most  important  food 
(Table  1).  Although  calanoid  copepods  and  Themisto  sp. 
(amphipod)  were  both  often  present  in  GOA  specimens 
(28.13%  and  28.57%  of  stomachs,  respectively),  they  were 
not  important  food  in  terms  of  weight.  The  same  was  true 
in  the  AI  for  calanoid  copepods,  Themisto  sp.,  gammaridean 


amphipods,  and  hyperiidean  amphipods  (29.14%,  36.91%, 
30.49%,  and  33.46%  respectively).  Mysids  and  larvaceans 
from  GOA  specimens  as  well  as  ctenophors,  polychaetes,  and 
euphasiids  from  AI  specimens  occurred  in  trace  amounts. 
Sebastes  larvae  (5-8  mm  standard  length),  the  only  fish 
species  found,  were  found  in  43%  of  Gulf  of  Alaska  stomachs 
but  made  up  only  0.43%  of  prey  weight.  Some  Dover  sole 
[Microstomas paeifieus)  eggs  had  also  been  consumed. 


176 


Fishery  Bulletin  102(1) 


0.8  H 

0.6 

0.4- 
0.2- 


0 


t* 


000000 


2 

-e- 


311      2  12 

000   000 


Pmat  =  1/(1+e3"'»-"'M 


00  i    0 


30         35         40         45  50         55         60         65         70         75  80  85         90 

Length  (cm) 


5        4 

1        3 

1         1 

0.8  - 

0.6  - 

/  3 

Jo 

0.4  - 

P 

1/(1  +e9«-,9°') 

0.2  - 

0- 

1^2 

6  0    i 

10  15 

Age  (years) 


20 


25 


Figure  7 

Proportion  of  female  prowfish  mature  (Pmat)  as  logistic  functions  of  length  (L)  and  age 
it).  Data  points  based  on  39  maturity-at-length  and  27  maturity-at-age  observations  are 
shown  by  diamonds,  and  numbers  of  females  of  each  cm-length  and  year-age  class  are 
shown  next  to  the  corresponding  symbol.  The  fitted  logistic  models  are  shown  by  solid 
lines.  The  length  and  age  at  which  Pmat  =  0.5  with  95^  confidence  limits  are  57.0  ±0.4  cm 
and  5.1  ±0.7  years. 


Discussion 

Geographic  distribution 

Historically  occurring  in  the  catch  in  AFSC  bottom  trawl 
surveys  in  areas  of  the  eastern  Bering  Sea,  Aleutian 
Islands,  and  Gulf  of  Alaska  regions,  prowfish  were  also 
observed  more  rarely  farther  south  along  the  West  Coast 
as  far  as  the  vicinity  of  San  Miguel  Island,  California. 
This  is  the  apparent  southern  limit  of  their  range  in  the 
northeastern  Pacific  (Allen  and  Smith,  1988).  They  were 
most  often  encountered  in  the  vicinity  of  the  edge  of  the 
continental  slope  near  200  m  depth  (Fig.  2),  although  our 
data  increase  the  maximum  known  depth  of  occurrence 
from  675  m  (Allen  and  Smith,  1988)  to  801  m.  As  indicated 
by  survey  CPUE,  prowfish  density  was  greatest  between 
the  depths  of  100  m  and  240  m  (Fig.  3).  Our  distribution 
data  show  similarities  with  those  of  Tokranov  ( 1999),  who 


studied  >300  bottom  trawls  executed  in  1995-97  on  the 
shelf  and  slope  off  the  southern  Kamchatka  Peninsula  and 
northern  Kuril  Islands,  in  which  adult  prowfish  were  taken 
at  100-480  m.  Tokranov  often  found  fish  concentrated  in 
areas  of  high-relief,  rocky  bottom — a  common  feature  of 
the  shelf  edge  in  the  Gulf  of  Alaska  and  Aleutian  Islands 
regions.  Such  areas  near  the  shelf  break  may  be  important 
prowfish  habitat.  Underwater  videos  taken  in  the  north- 
east Gulf  of  Alaska  by  the  Alaska  Department  of  Fish  and 
Game  (Brylinsky5)  show  numerous  adult  fish  resting  on  or 
just  above  this  type  of  substrate. 

Density  was  greater  in  the  AI  than  in  the  GOA,  over  all 
bottom  depths  combined  and  in  most  cases  by  individual 
depth  interval  (Fig.  3).  One  reason  may  be  that  preferred 
habitat  comprises  a  larger  proportion  of  the  Aleutian  Is- 


5  Brylinsky,  C.     2000.     Pers.  commun.    Alaska  Department  of 
Fish  and  Game,  304  Lake  Street,  Sitka,  AK  99835. 


Smith  et  al.:  Distribution  and  biology  of  Zaprora  stlenus 


177 


lands  area.  Because  of  the  lack  of  a  relatively  broad  shelf 
in  the  region,  a  larger  proportion  of  trawls  are  in  or  near 
areas  of  steep  seafloor  gradient  and  therefore  likely  over 
rough  bottom  (Fig.  2). 

Length  distribution 

In  both  the  Gulf  of  Alaska  and  the  Aleutian  Islands,  few 
prowfish  <40  cm  in  length  were  captured  (Fig.  4).  This 
paucity  of  small  prowfish  is  not  due  to  size  selection  by 
the  trawl  net  mesh  because  the  codend  is  lined  with  small 
mesh  (1.3  cm  stretched  measure)  webbing  that  retains 
small  individuals  of  other  species.  A  different  explanation, 
based  on  the  observations  of  Brodeur  (1998)  and  Scheffer 
(1940),  is  that  pre-adult  prowfish  are  pelagic,  remaining 
in  proximity  with  large  coelenterates  and  thus  avoiding 
bottom  trawls.  Thus,  the  minimum  capture  length  may 
indicate  the  length  at  which  prowfish  recruit  to  a  demer- 
sal habitat.  Our  data  showed  no  statistically  significant 
length  difference  between  sexes,  in  contrast  with  the  data 
of  Tokranov  (1999)  who  suggested  a  length  dimorphism 
where  females  are  generally  longer  than  males. 

Weight-length  relationship 

The  best-fitting  model  of  weight  versus  length  predicts 
that  for  any  length,  female  prowfish  are,  on  average,  3.7% 
heavier  than  males  (Fig.  5).  It  seems  unlikely  that  this 
relationship  exists  over  all  developmental  stages  because 
our  samples  were  almost  all  adults  and  such  a  (relative) 
difference  might  not  remain  constant  during  all  ontoge- 
netic sexual  divergence.  What  is  more  certain  is  simply 
the  existence  of  some  small  degree  of  length-weight 
dimorphism  (females  slightly  heavier  at  a  given  length). 
Also,  this  dimorphism  is  not  likely  to  stem  primarily  from 
a  sexual  difference  in  gonad  weight  because  the  maximum 
proportion  of  total  female  body  weight  composed  of  ovarian 
tissue  was  only  2.7%.  Thus  the  difference  is  due  to  other 
morphological  or  behavioral  differences. 

Growth 

There  was  no  significant  difference  between  sexes  in  length 
versus  age.  The  predicted  length  of  a  prowfish  of  given  age 
based  on  our  samples  was  higher  than  that  indicated  by 
Tokranov  (1999).  In  our  study  5-year-old  and  9-year-old 
fish  averaged  56.6  cm  and  73.5  cm  in  length,  respectively. 
Tokranov  (1999)  considered  that  prowfish  growth  deter- 
mined from  otoliths  of  102  specimens  from  the  Northwest 
Pacific  indicated  a  comparatively  fast-growing  species 
reaching  an  average  length  of  44.6  cm  by  5  years  of  age 
and  67  cm  after  9  years.  These  data  suggest  prowfish  are 
indeed  relatively  fast  growing  and  that  growth  rates  for  the 
Gulf  of  Alaska  are  faster  than  those  for  off  southeastern 
Kamchatka  and  the  northern  Kuril  Islands.  Alternatively, 
size-dependent  mortality  from  such  elements  as  incidental 
capture  by  commercial  fishing  may  affect  the  two  popula- 
tions differently. 

Historically,  two  other  prowfish  have  been  aged  from 
otoliths:  a  male  84  cm  long  taken  near  Eureka,  CA  (Fitch 


and  Lavenberg,  1971),  and  a  female  50.1  cm  long  ( standard 
length)  from  off  Monterey  (Cailliet  and  Anderson,  1975). 
The  ages  estimated  were  12  and  3  years,  respectively.  Af- 
ter converting  the  standard  length  record  to  an  estimate 
of  total  length  for  the  second  specimen  of  58  cm  by  using 
a  ratio  described  by  Baxter,6  both  lengths  are  slightly 
greater  than  our  predictions  for  the  same  ages,  albeit  near 
the  limits  of  our  data  range.  This  finding  contrasts  with 
the  predictions  of  lesser  length  at  a  given  age  presented 
by  Tokranov  (1999). 

Maturity 

Little  previous  data  exist  with  which  to  compare  our  obser- 
vations of  female  prowfish  rate  of  maturation.  Cailliet  and 
Anderson  (1975)  examined  the  ovaries  of  their  50.1-cm  3- 
year-old  female  specimen  for  vitellogenesis  and  predicted 
an  age  at  first  spawning  of  4  years,  slightly  less  than  the 
lower  95%  confidence  limit  of  4.4  years  for  our  expected 
average  age  at  50%  maturity. 

Food  habits 

Our  observation  that  gelatinous  zooplankton  was  the 
largest  constituent  in  the  contents  of  prowfish  stomachs 
(Table  1)  is  supported  by  Tokranov  ( 1999),  who  found  that 
the  two  most  common  prey  taxa  among  the  contents  of  102 
stomachs  of  adult  specimens  from  the  northwestern  Pacific 
were  Scyphozoa  (59.6-62.0%  of  stomachs)  and  Ctenophora 
(6.0-15.4%  of  stomachs).  Anecdotal  observations  have  also 
been  made  of  the  feeding  behavior  of  an  aquarium  specimen 
over  an  approximate  2-year  period  (Carollo  and  Rankin, 
1998).  When  first  obtained,  the  fish  ate  only  various  jel- 
lyfish species,  rejecting  other  food,  including  a  variety  of 
live  invertebrates.  In  our  food  samples,  we  observed  other 
taxa,  such  as  invertebrates  and  small  fish,  but  these  were  a 
minor  part,  possibly  first  captured  by  jellyfish  and  then  sec- 
ondarily ingested  by  prowfish.  Carollo  and  Rankin  (1998) 
found  that  the  aquarium  specimen  would  ingest  such  items 
when  eating  the  bells  of  Chrysaora  melanaster  in  which 
such  food  had  previously  been  placed,  indicating  the  pos- 
sibility of  this  occurring  naturally.  Possibly  more  reflective 
of  the  unnatural  circumstances,  the  specimen  later  began 
accepting  such  items  outside  the  bells  of  jellyfish. 

Apparent  adaptations  of  the  prowfish  to  a  diet  of  ge- 
latinous zooplankton  include  the  small,  sharp,  close-set 
teeth  in  a  single  row  attached  only  to  the  jaws,  which  are 
capable  of  a  180-degree  gape,  and  the  large  rough-scaled 
lips  (Clemens  and  Wilby,  1961;  Hart,  1973;  Carollo  and 
Rankin,  1998). 


Acknowledgments 

We  are  grateful  for  the  expert  advice  given  by  Alaska  Fish- 
eries Science  Center  colleagues  Delsa  Anderl  and  Nancy 


6  Baxter,  R.  1990.  Unpubl.  manuscript.  Annotated  key  to  the 
fishes  of  Alaska,  803  p.  [Available  from  Sera  Baxter,  Box  182, 
Seldovia,  AK  99663.1 


178 


Fishery  Bulletin  102(1) 


Roberson  regarding  age-reading  of  prowfish  otoliths, 
and  by  AFSC  colleagues  Kathy  Mier  and  Susan  Piquelle 
regarding  statistical  analyses  of  data. 

Literature  cited 

Allen,  James  M.,  and  Gary  B.  Smith. 

1988.     Atlas  and  zoogeography  of  common  fishes  in  the 
Bering  Sea  and  Northeastern  Pacific.     NOAA  Tech.  Rep. 
NMFS  66,  151  p. 
Brodeur,  R.  D. 

1998.     In  situ  observations  of  the  association  between  juve- 
nile fishes  and  scyphomedusae  in  the  Bering  Sea.     Mar. 
Ecol.  Prog.  Ser.  163:11-20. 
Cailliet,  G.  M.,  and  M.  E.  Anderson. 

1975.     Occurrence  of  prowfish  Zaprora  silenus  Jordan,  1896  in 
Monterey  Bay,  California.     Calif.  Fish  Game  61(l):60-62. 
Carollo,  M.,  and  P.  Rankin. 

1998.    The  care  and  display  of  the  prowfish,  Zaprora  silenus. 
Drum  and  Croaker  29:3-6. 
Clemens,  W.  A.,  and  G.  V.  Wilby. 

1961.    Fishes  of  the  Pacific  coast  of  Canada.    Fish.  Res. 
Board  Can.  Bull.  68,  2nd  ed.,  443  p. 
Chilton,  D.  E.,  and  R.  J.  Beamish. 

1982.  Age  determination  methods  for  fishes  studied  by  the 
groundfish  program  at  the  Pacific  Biological  Station.  Can. 
Spec.  Pub.  Fish.  Aquat.  Sci.  60, 102  p. 

Doyle,  M.  J.,  K.  L.  Meir,  M.  S.  Busby,  and  R.  D.  Brodeur. 

2002.     Regional  variation  in  springtime  ichthyoplankton 

assemblages  in  the  northeast  Pacific  Ocean.     Progress  in 

Oceanography  53  (2-4):247-281 
Eschmeyer,  W.  N,  E.  S.  Herald,  and  H.  Hammann. 

1983.  A  field  guide  to  Pacific  Coast  fishes  of  North  America, 
336  p.  Peterson  field  guide  ser.  Houghton  Mifflin,  Boston, 
MA. 

Fitch,  J.  E.,  and  R.  J.  Lavenberg. 

1971.     Marine  food  and  game  fishes  of  California,  179  p. 
U.C.  Press,  Berkeley,  CA. 
Hart,  J.  L. 

1973.     Pacific  fishes  of  Canada.     Fish.  Res.  Board  Can.  Bull. 
180,  740  p. 
Jordan,  D.  S. 

1897.     Notes  on  fishes  little  known  or  new  to  science.     Proc. 
Calif.  Acad.  Sci.,  2nd  sen,  6:203-205. 
Kessler,  D.  W. 

1985.  Alaska's  saltwater  fishes  and  other  sea  life:  a  field 
guide,  358  p.    Alaska  Northwest,  Anchorage,  AK 


Kimura,  D.  K. 

1980.     Likelihood  methods  for  the  von  Bertalanffy  growth 
curve.    Fish.  Bull.  77:765-776. 
Martin,  M.  H. 

1997.  Data  report:  1996  Gulf  of  Alaska  bottom  trawl 
survey.     NOAA  Tech.  Memo.  NMFS-AFSC-82,  235  p. 

Matarese,  A.  C,  A.  W.  Kendall,  D.  M.  Blood,  and  B.  M.  Vinter. 

1989.     Laboratory  guide  to  early  life  history  stages  of  north- 
east pacific  fishes.    NOAA  Tech.  Rep.  NMFS  80,  652  p. 
Nelson,  J.  S. 

1994.     Fishes  of  the  world,  3rd  ed.,  600  p.     John  Wiley,  New 
York,  NY.. 
Orlov,  A.  M. 

1998.  On  feeding  of  mass  species  of  deep-sea  skates  (Bathy- 
raja  spp.,  Rajidae)  from  the  Pacific  waters  of  the  Northern 
Kurds  and  Southeastern  Kamchatka.  J.  Icthyol.  38(8): 
635-644. 

Payne,  S.  A.,  B.  A.  Johnson,  and  R.  S.  Otto. 

1999.  Proximate  composition  of  some  north-eastern  Pacific 
forage  fish  species.     Fish.  Oceanogr.  8(3):  159-177. 

Scheffer,  V.  B. 

1940.    Two  recent  records  of  Zaprora  silenus  Jordan  from 
the  Aleutian  Islands.     Copeia  1940(31:203. 
Seber,  G.  A.  F 

1973.    The  estimation  of  animal  abundance,  506  p.     Hafner 
Press,  New  York,  NY. 
Tokranov,  A.  M. 

1999.     Some  features  of  biology  of  the  prowfish  Zaprora 
silenus  (Zaproridae)  in  the  Pacific  waters  of  the  Northern 
Kuril  Islands  and  Southeastern  Kamchatka.     J.  Ichthyol. 
39(61:475-478. 
Van  Pelt,  T  I.,  J.  F  Piatt,  B.  K.  Lance,  and  D.  D.  Roby. 

1997.     Proximate  composition  and  energy  density  of  some 
North  Pacific  forage  fishes.     Comp.  Biochem.  Physiol. 
118A(4 1:1393-1398. 
Wakabayashi,  K,  R.  G.  Bakkala  and  M.  S.  Alton. 

1985.  Methods  of  the  U.S.-Japan  demersal  trawl  surveys. 
In  Results  of  cooperative  U.S.-Japan  groundfish  investiga- 
tions in  the  Bering  Sea  during  May-August  1979  (R.  G. 
Bakkala  and  K.  Wakabayashi,  eds.).  p.  7-26.  Int.  North 
Pac.  Fish.  Comm.  Bull.  44. 
Yang,  Mei-Sun. 

1993.  Food  habits  of  the  commercially  important  ground- 
fishes  in  the  Gulf  of  Alaska  in  1990.  NOAA  Tech.  Memo. 
NMFS-AFSC-22,  150  p. 


179 


Abstract— Our  analyses  of  observer 
records  reveal  that  abundance  esti- 
mates are  strongly  influenced  by  the 
timing  of  longline  operations  in  rela- 
tion to  dawn  and  dusk  and  soak  time — 
the  amount  of  time  that  baited  hooks 
are  available  in  the  water.  Catch  data 
will  underestimate  the  total  mortal- 
ity of  several  species  because  hooked 
animals  are  "lost  at  sea."  They  fall  off, 
are  removed,  or  escape  from  the  hook 
before  the  longline  is  retrieved.  For 
example,  longline  segments  with  soak 
times  of  20  hours  were  retrieved  with 
fewer  skipjack  tuna  and  seabirds  than 
segments  with  soak  times  of  5  hours. 
The  mortality  of  some  seabird  species 
is  up  to  45%  higher  than  previously 
estimated. 

The  effects  of  soak  time  and  timing 
vary  considerably  between  species. 
Soak  time  and  exposure  to  dusk  periods 
have  strong  positive  effects  on  the  catch 
rates  of  many  species.  In  particular,  the 
catch  rates  of  most  shark  and  billfish 
species  increase  with  soak  time.  At  the 
end  of  longline  retrieval,  for  example, 
expected  catch  rates  for  broadbill 
swordfish  are  four  times  those  at  the 
beginning  of  retrieval. 

Survival  of  the  animal  while  it  is 
hooked  on  the  longline  appears  to  be  an 
important  factor  determining  whether 
it  is  eventually  brought  on  board  the 
vessel.  Catch  rates  of  species  that 
survive  being  hooked  (e.g.  blue  shark) 
increase  with  soak  time.  In  contrast, 
skipjack  tuna  and  seabirds  are  usu- 
ally dead  at  the  time  of  retrieval.  Their 
catch  rates  decline  with  time,  perhaps 
because  scavengers  can  easily  remove 
hooked  animals  that  are  dead. 

The  results  of  our  study  have  impor- 
tant implications  for  fishery  manage- 
ment and  assessments  that  rely  on 
longline  catch  data.  A  reduction  in  soak 
time  since  longlining  commenced  in  the 
1950s  has  introduced  a  systematic  bias 
in  estimates  of  mortality  levels  and 
abundance.  The  abundance  of  species 
like  seabirds  has  been  over-estimated 
in  recent  years.  Simple  modifications 
to  procedures  for  data  collection,  such 
as  recording  the  number  of  hooks 
retrieved  without  baits,  would  greatly 
improve  mortality  estimates. 


Manuscript  approved  for  publication 
22  September  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:179-195  (2004). 


Fish  lost  at  sea:  the  effect  of  soak  time 
on  pelagic  longline  catches 

Peter  Ward 

Ransom  A.  Myers 

Department  of  Biology 

Dalhousie  University 

Halifax,  B3H  4JI  Canada 

E-mail  address  (for  P  Ward)  wardiSmathstat.dal  ca 

Wade  Blanchard 

Department  of  Mathematics  and  Statistics 
Dalhousie  University 
Halifax,  B3H  44  Canada 


Our  knowledge  of  large  pelagic  fish  in 
the  open  ocean  comes  primarily  from 
tagging  and  tracking  experiments  and 
from  data  collected  from  longline  fish- 
ing vessels  since  the  1950s.  Abundance 
indices  for  pelagic  stocks  are  often 
derived  from  analyses  that  model  catch 
as  a  function  of  factors  such  as  year, 
area,  and  season.  However,  the  amount 
of  time  that  baited  hooks  are  available 
to  fish  is  likely  to  be  another  important 
factor  influencing  catch  rates  (Deriso 
and  Parma,  1987). 

The  activity  of  many  pelagic  animals 
and  their  prey  vary  with  the  time  of 
day.  Broadbill  swordfish,  for  example, 
feed  near  the  sea  surface  at  night.  They 
move  to  depths  of  500  m  or  more  during 
the  day  (Carey,  1990).  Other  species  may 
be  more  active  in  surface  waters  during 
the  day  (e.g.  striped  marlin)  or  at  dawn 
and  dusk  (e.g.  oilfish).  Longline  fishing 
crews  take  a  keen  interest  in  the  tim- 
ing of  their  fishing  operations  and  soak 
time  (the  total  time  that  a  baited  hook 
is  available  in  the  water).  However,  as- 
sessments have  not  accounted  for  those 
factors  in  estimating  the  abundance 
or  mortality  levels  of  target  species  or 
nontarget  species. 

In  many  assessments  that  use  pelagic 
longline  catch  rates,  fishing  effort  is  as- 
sumed to  be  proportional  to  the  number 
of  hooks  deployed.  The  effects  of  soak 
time  and  timing  may  have  been  omit- 
ted because  a  clear  demonstration  of 
their  effects  on  pelagic  longline  catch 
rates  is  not  available.  The  few  pub- 
lished accounts  on  soak  time  in  pelagic 
longline  fisheries  have  been  based  on 


limited  data  and  a  few  target  species. 
For  example,  in  analyzing  95  longline 
operations  or  "sets"  by  research  vessels 
Sivasubramaniam  ( 1961)  reported  that 
the  catch  rates  of  bigeye  tuna  increased 
with  soak  time,  whereas  yellowfin  tuna 
catch  rates  were  highest  in  longline  seg- 
ments with  intermediate  soak  times. 

In  contrast  to  the  limited  progress  in 
empirical  studies,  theoretical  approach- 
es are  well  developed  for  modeling  fac- 
tors that  may  influence  longline  catch 
rates.  Soon  after  large-scale  longlining 
commenced.  Murphy  (1960)  published 
"catch  equations"  for  adjusting  catch 
rates  for  soak  time,  bait  loss,  escape, 
hooking  rates,  and  gear  saturation.  He 
suggested  that  escape  rates  could  be  es- 
timated from  counts  of  missing  hooks 
and  hooks  retrieved  without  baits. 
Unfortunately,  such  data  are  rarely  col- 
lected from  pelagic  longline  operations. 

More  recently,  hook-timers  attached 
to  longline  branchlines  have  begun  to 
provide  information  on  the  time  when 
each  animal  is  hooked  and  also  whether 
animals  are  subsequently  lost,  e.g. 
Boggs  (1992),  Campbell  et  al.1-2  Such 
data  are  particularly  useful  to  under- 


1  Campbell,  R.,  W.  Whitelaw,  and  G.  Mc- 
Pherson.  1997.  Domestic  longline  fish- 
ing methods  and  the  catch  of  tunas  and 
non-target  species  off  north-eastern 
Queensland  (1st  survey:  October-Decem- 
ber 1995).  Report  to  the  Eastern  Tuna 
and  Billfish  Fishery  MAC.  71  p.  Aus- 
tralian Fisheries  Management  Authority, 
PO  Box  7051,  Canberra  Business  Centre, 
ACT  26 10,  Australia. 

2  See  next  page. 


180 


Fishery  Bulletin  102(1) 


standing  the  processes  affecting  the  probability  of  capture 
and  escape. 

The  purpose  of  our  study  is  to  determine  whether  varia- 
tions in  the  duration  and  timing  of  operations  bias  abun- 
dance and  mortality  estimates  derived  from  longline  catch 
rates.  We  present  a  theoretical  model  that  is  then  related  to 
empirical  observations  of  the  effects  of  soak  time  on  catch 
rates.  The  strength  in  our  approach  is  in  applying  a  random 
effects  model  to  large  data  sets  for  over  60  target  and  non- 
target  species  in  six  distinct  fisheries.  We  also  investigate 
the  survival  of  each  species  while  hooked  because  prelimi- 
nary analyses  suggested  that  the  effects  of  soak  time  on 
catch  rates  might  be  linked  to  mortality  caused  by  hooking 
(referred  to  as  "hooking  mortality"). 


Factors  affecting  catch  rates 

To  aid  interpretation  of  our  statistical  analysis  of  soak  time 
effects,  we  first  developed  a  simple  model  to  illustrate  how 
the  probability  of  catching  an  animal  may  vary  with  soak 
time. 

The  probability  of  an  animal  being  on  a  hook  when  the 
branchline  is  retrieved  is  a  product  of  two  probability 
density  functions:  first  the  probability  of  being  hooked 
and  then  the  probability  of  being  lost  from  the  hook.3  In- 
fluencing the  probability  of  being  hooked  are  the  species' 
local  abundance,  vulnerability  to  the  fishing  gear,  and  the 
availability  of  the  gear.  Catches  will  deplete  the  abundance 
of  animals  within  the  gear's  area  of  action,  particularly  for 
species  that  have  low  rates  of  movement.  Movement  will 
also  result  in  variations  in  exposure  of  animals  to  the  gear 
over  time — for  instance,  as  they  move  vertically  through 
the  water  column  in  search  of  prey  (Deriso  and  Parma, 
1987). 

Other  processes  that  will  reduce  the  probability  of  be- 
ing hooked  include  bait  loss  and  reduced  sensitivity  to  the 
bait  (Ferno  and  Huse,  1983).  Longline  baits  may  fall  off 
hooks  during  deployment,  deteriorate  over  time  and  fall 
off  or  they  may  lose  their  attractant  qualities.  They  may  be 
removed  by  target  species,  nontarget  species,  or  other  ma- 
rine life,  such  as  squids.  Hooked  animals  may  also  escape 
by  severing  the  branchline  or  breaking  the  hook.  Sections 
of  the  longline  may  become  saturated  when  animals  are 
hooked,  reducing  the  number  of  available  baits  (Murphy. 
1960;  Somerton  and  Kikkawa,  1995).  After  an  animal  has 
been  hooked,  it  may  escape,  fall  off  the  hook,  be  removed  by 
scavengers,  or  it  may  remain  hooked  until  the  branchline 
is  retrieved. 

Some  of  the  processes  affecting  the  probability  of  an  ani- 
mal being  on  a  hook  when  the  the  branchline  is  retrieved 


2  Campbell,  R.,  W.  Whitelaw,  and  G.  McPherson.  1997.  Do- 
mestic longline  fishing  methods  and  the  catch  of  tunas  and  non- 
target  species  off  north-eastern  Queensland  (2nd  survey:  May- 
August  1996).  Report  to  the  Eastern  Tuna  and  Billfish  Fishery 
MAC,  48  p.  Australian  Fisheries  Management  Authority,  PC) 
Box  7051,  Canberra  Business  Centre,  ACT  2610,  Australia. 
In  discussing  continuous  variables  we  use  the  terms  "proba- 
bility" and  "probability  density  function"  interchangeably. 


are  species-specific,  whereas  other  processes  may  affect  all 
species.  For  example,  bait  loss  during  longline  deployment 
will  reduce  the  catch  rates  of  all  species.  In  contrast,  the 
probability  of  a  hooked  animal  escaping  may  be  species-de- 
pendent; some  species  are  able  to  free  themselves  from  the 
hook  whereas  other  species  are  rarely  able  to  do  this. 

Our  simple  model  of  the  probability  of  an  animal  being 
on  a  hook  is  based  on  a  convolution  of  the  two  time-related 
processes  described  above:  1)  the  decay  in  the  probability 
of  capture  with  the  decline  in  the  number  of  baits  that  are 
available;  and  2)  gains  due  to  the  increased  exposure  of 
baits  to  animals  and  losses  due  to  animals  escaping,  falling 
off,  or  being  removed  by  scavengers. 

The  probability  of  an  animal  being  on  a  hook  when  the 
branchline  is  retrieved  is  the  integral  of  the  probability 
density  functions  of  capture  and  retention: 


rtT)  =  J  P(t)PrlT-t)dt, 


(1) 


where  jriT)  =  the  "catch  rate"  or  probability  of  an  animal 

being  on  a  hook  when  the  branchline  is 

retrieved  at  time  T  (T  is  the  total  soak  time 

of  the  hook); 

P(it)  =  the  probability  density  function  of  an  animal 

being  captured  at  time  t;  and 
Pr(t)  =  the  probability  density  function  of  a  cap- 
tured animal  being  retained  on  the  hook  for 
a  length  of  time  f. 

The  probability  density  function  of  capture  can  be  approxi- 
mated with  an  exponential  function: 


Pit)  =  P0e-", 


(2) 


where  P0  =  the  probability  of  capture  when  the  hook  is 
deployed  (r=0);  and 
o  =  a  parameter  determining  the  rate  of  change  in 
capture  probability  over  time. 

After  the  animal  is  hooked,  the  probability  density  function 
of  an  animal  being  retained  after  capture  can  be  approxi- 
mated as 


PIt)  =  e-pw, 


(3) 


where  /I  =  the  "loss  rate,"  a  parameter  determining  the 
rate  of  change  in  the  probability  of  an  animal 
being  retained  after  it  has  been  captured. 

Substituting  approximations  2  and  3  into  Equation  1 
gives 


7l(T)=  \P0e  '"e  <"'  '  dt 


(4) 


/?-«' 


Ward  et  al.:  The  effect  of  soak  time  on  pelagic  longlme  catches 


181 


Seabirds 
(-0.06) 


p 
c\i 


(XX) 


Skipjack 

(-0.04) 

() 

() 

(>           <> 

°  ()     o            () 

0 

O 

5  10         15         20         25 


0  5  10         15         20         25 


<) 

<) 

O    O 

o 

C)             °<»' 

j                     (X) 

o 

o 

o  o° 

Lancetfish 

(-0.02) 

in 

Swordfis 

h 

(M 

(+0.09) 

o  . 

<>°o 

<M 

o 

o 

m 

o 

() 

o 

o 

() 

T— 

( 

) 
() 

in 

o 

o 

o 

o' 

o 

o 

0  5  10         15         20         25  0  5  10         15         20         25 

Soak  time  (h) 

Figure  1 

Mean  catch  rates  plotted  against  soak  time  for  skipjack  tuna,  long-nosed  lancetfish.  and 
swordfish  in  the  South  Pacific  yellowfin  tuna  fishery  and  for  "other  seabirds"  in  the  South 
Pacific  bluefin  tuna  fishery.  To  reduce  variability,  the  estimates  were  limited  to  longline 
segments  with  more  than  25  hooks  and  soak  times  of  5-20  hours.  Vertical  bars  are  95% 
confidence  intervals  for  the  mean  hourly  catch  rate.  In  parentheses  are  the  soak-time  coef- 
ficients from  random  effects  models  ( note  that  the  soak-time  coefficient  is  not  the  same  as 
the  slope  coefficient  of  a  regression  of  the  data  presented  in  this  graph). 


Our  model  is  similar  to  the  parabolic  catch  model  exam- 
ined by  Zhou  and  Shirley  (1997).  It  is  simpler  than  catch 
equations  developed  by  other  authors  because  it  does  not 
include  specific  terms  for  the  loss  of  baits,  for  fish  competi- 
tion, and  gear  saturation. 

Preliminary  plots  of  observer  data  indicated  a  variety  of 
patterns  in  the  relationship  between  catch  rates  and  soak 
time  (e.g.  Fig.  1).  By  varying  the  values  of  P0  (probability 
of  capture),  a  (capture  rate),  and  p  (loss  rate),  our  simple 
catch  equation  (Eq.  4)  can  mimic  the  observed  patterns 
(Fig.  2).  However,  estimates  of  P0,  a  ,  and  /3  are  not  avail- 
able. Instead,  we  used  the  empirical  approach  described 


in  the  following  section  to  model  the  effect  of  soak  time 
on  catch  rates.  The  relationship  of  soak  time  to  catch  rate 
represents  the  product  of  the  probability  of  capture  and  the 
probability  of  being  retained. 

One  approach  to  investigating  the  effects  of  soak  time 
on  catch  rates  is  to  fit  linear  regressions  to  aggregated 
data  like  those  presented  in  Figure  1.  Such  an  approach, 
however,  would  violate  assumptions  of  independence 
(within  each  longline  operation,  catch  rates  in  consecutive 
segments  will  be  related),  normality  (these  are  binomial 
data),  and  homogeneity  of  variance  (for  binomial  data,  the 
variance  is  dependent  on  the  mean). 


182 


Fishery  Bulletin  102(1) 


hooks  retrieved  after  1 5  hours 
have  many  swordfish 


soak  times 
not  observed 


hooks  retrieved  before  6  hours 
have  few  swordfish 


10 


15 


Soak  time  (h) 


No  captures  after  deployment 
Soak  time  coefficient  <0 
e.g.  seabirds 


Captures  exceed  losses 
Soak  time  coefficient  >0 
e.g.  swordfish 


Losses  eventually  exceed  captures 
Soak  time  coefficient  <0 
e.g.  skipjack 


Captures  balance  losses 
Soak  time  coefficient  -0 
eg  lancetfish 


20 


Figure  2 

Illustration  of  different  patterns  in  the  theoretical  relationship  between  longline  catch  rates  and  soak  time.  The 
probability  of  an  animal  being  on  a  hook  when  a  branchline  is  retrieved  (the  "catch  rate")  is  estimated  from 
Equation  4  by  using  soak  times  iT)  ranging  from  0  to  20  hours  and  three  different  combinations  of  values  forPn 
(probability  of  capture),  «  (capture  rate),  and  /3  (loss  rate).  For  seabirds,  the  probabilities  were  estimated  from 
Equation  6.  The  probabilities  are  not  on  the  same  scale  for  all  species. 


Another  approach  might  be  to  fit  separate  logistic  regres- 
sions to  each  operation  and  then  to  combine  the  parameter 
estimates.  This  would  overcome  the  problems  of  normality 
and  homogeneity  of  variance.  However,  the  separate  re- 
gressions would  not  incorporate  information  that  is  com- 
mon to  all  operations. 

Instead,  we  used  a  logistic  regression  with  random  ef- 
fects. The  key  advantage  in  using  random-effects  models 
in  this  situation  is  that  they  carry  information  on  the  cor- 
relation between  longline  segments  that  is  derived  from 
the  entire  data  set  of  operations. 


Data  and  methods 

Fisheries 

We  analyzed  observer  data  from  six  different  fisheries  in 
the  Pacific  Ocean  to  determine  the  effects  of  soak  time 
and  timing  on  longline  catch  rates  (Table  1,  Fig.  3).  These 
fisheries  involve  two  different  types  of  longline  fishing 
operation:  1 )  distant-water  longlining  involves  trips  of 
three  months  or  longer  and  the  catch  is  frozen  on  board 


the  vessel;  and  2)  fresh-chilled  longlining,  which  involves 
small  vessels  (15-25  m)  undertaking  trips  of  less  than  four 
weeks  duration,  and  the  catch  is  kept  in  ice,  ice  slurries,  or 
in  spray  brine  systems.  The  fresh-chilled  longliners  deploy 
shorter  longlines  with  fewer  hooks  (-1000  hooks)  than 
the  distant-water  longliners  (-3000  hooks  per  operation) 
(Ward,  1996;  Ward  and  Elscot,  2000). 

The  six  fisheries  share  many  operational  similarities, 
such  as  the  types  of  bait  used  and  soak  time.  However, 
they  are  quite  different  in  terms  of  targeting,  which  is 
determined  by  fishing  practices,  e.g.  the  depth  profile  of 
the  longline,  timing  of  operations  and  the  area  and  season 
of  activity.  South  Pacific  bluefin  tuna  longliners  operate  in 
cold  waters  ( 10-16°C)  in  winter  to  catch  southern  bluefin 
tuna.  In  the  South  Pacific  yellowfin  tuna  longliners  tar- 
get tropical  species,  such  as  yellowfin  and  bigeye  tuna,  in 
warmer  waters  (19-22°C)  (Ward,  1996).  To  target  bigeye 
tuna,  longlines  in  the  Central  Pacific  bigeye  fishery  are 
deployed  in  the  early  morning  with  hook  depths  ranging 
down  to  about  450  m.  The  depths  of  the  deepest  hook  are 
much  shallower  (-150  m)  in  the  North  Pacific  swordfish 
fishery  where  the  longlines  are  deployed  late  in  the  after- 
noon and  retrieved  early  in  the  morning  (Boggs,  1992). 


Ward  et  al.:  The  effect  of  soak  time  on  pelagic  longline  catches 


183 


Observer  data 

National  authorities  and  regional  organizations  placed 
independent  observers  on  many  longliners  operating  in  the 
six  fisheries  during  the  1990s.  The  observer  data  consisted 
of  records  of  the  species  and  the  time  when  each  animal 
was  brought  on  board.  We  restricted  analyses  to  operations 
where  the  last  hook  that  had  been  deployed  was  retrieved 
first  ("counter- retrieved"),  where  there  was  no  evidence  of 
stoppages  due  to  line  breaks  or  mechanical  failure,  and 
where  there  was  continuous  monitoring  by  an  observer. 
Combined  with  records  of  the  number  of  hooks  deployed 
and  start  and  finish  times  of  deployment  and  retrieval,  the 
observer  data  allowed  calculation  of  soak  time  and  catch 
rates  of  longline  segments.  We  aggregated  catches  and  the 
number  of  hooks  into  hourly  segments.  The  soak  time  was 
estimated  for  the  midpoint  of  each  hourly  segment. 

The  Central  Pacific  bigeye  tuna  and  North  Pacific  sword- 
fish  fisheries  differed  from  the  other  four  fisheries  in  the 
species  that  were  recorded  and  the  method  of  recording 
the  time  when  each  animal  was  brought  on  board.  Observ- 
ers reported  catches  according  to  a  float  identifier  in  the 
Central  and  North  Pacific  fisheries.  Therefore  we  estimated 
soak  times  for  each  longline  segment  from  the  time  when 
each  float  was  retrieved.  For  those  fisheries,  observers  re- 
ported the  float  identifier  only  for  tuna,  billfish,  and  shark 
(Table  2).  Data  are  available  for  protected  species,  such 
as  seals,  turtles,  and  seabirds  but  were  not  sought  for  the 
present  study. 

We  assumed  a  constant  rate  of  longline  retrieval 
throughout  each  operation.  The  number  of  hooks  retrieved 
during  each  hourly  segment  was  the  total  number  of  hooks 
divided  by  the  duration  of  monitoring  (decimal  hours).  For 
each  species  we  analyzed  only  the  operations  where  at  least 
one  individual  of  that  species  was  caught. 

Longline  segments  that  involved  a  full  hour  of  monitor- 
ing had  several  hundred  hooks.  Segments  at  either  end 
of  the  longline  involved  less  than  an  hour  of  monitoring 
and  had  fewer  hooks.  Catch  rates  may  become  inflated  in 
segments  with  very  small  numbers  of  hooks.  Therefore  we 
arbitrarily  excluded  segments  where  the  observer  moni- 
tored less  than  25  hooks. 

For  four  of  the  fisheries,  data  were  available  on  survival 
rates,  allowing  the  investigation  of  the  relationship  be- 
tween soak  time  and  hooking  mortality.  For  the  Western 
Pacific  and  South  Pacific  fisheries,  observers  reported 
whether  the  animal  was  alive  or  dead  when  it  was  brought 
on  board.  We  calculated  survival  rate  (the  number  alive 
divided  by  the  total  number  reported  dead  or  alive)  for  spe- 
cies where  data  were  available  on  the  life  status  of  more 
than  ten  individuals. 

Generalized  linear  mixed  model 

Logit  model  We  applied  a  generalized  linear  mixed 
model  to  the  observer  data.  The  model  is  based  on  a  logis- 
tic regression,  with  the  catch  (y)  on  each  hook  assumed 
to  have  a  binomial  distribution  with  y  ~  b(ra,  n).  n  is  the 
expected  value  of  the  distribution  for  a  specified  soak  time. 
We  refer  to  it  as  the  probability  of  catching  an  animal  or 


•s  ■$ 

CD                       CO 

c 

"2  ° 

CJ 

■2  2 

•-  °  5.  o 

J!      Or'* 

B   a  -S  o 

^H 

CO 

CO 

o 

o 

CO 

CJ 

8.  " 

lO 

CM 

CM 

CO 

^f 

CM 

Cfi 

a  "in  o 

^ 

o          2 

O 

o 

X 

-    ^ 

CJ 

o 

g 

co  >-* 
o    ° 

C 
O 

<3J      tG 

•*-»    ■ — 
t_    - • 

, — 1 

05 

, — I 

CN 

CM 

CM 

t- 

a.  a> 

CM 

,—1 

CM 

CM 

CM 

CM 

o 

»  "3 

=3 

CO 

.—1 

cd    & 

Q 

>> 

u      CO 

-- 

■m-h 

o 

O     C 

c 

-C 

„     o 

co            O 

en 

3  6 

£  S 

M    u  '43 

CM 

lO 

o 

t> 

O 

CO 

cc 

OQ}C3 

rH 

CD 

<N 

■* 

CO 

00 

cd 

O      Q,     '- 

CO 

00 

CD 

CO 

, — 1 

o 

c 

x  S 

X         g. 

rH 

rH 

CM 

CO 

CO 

3 

o 

O 

© 

■+*      - 

>, 

C8     2 

CD 

3   2 

bp 

cd    a> 

CO 

en 

CO 

lO 

■* 

a> 

CD 

IS 

i  -C 

'.2 

CO 

■* 

rH 

CO 

,— i 

CD 

£  +* 

6    es 
P. 

lO 

CM 

Ol 

CM 

■* 

CD 

cfl 

to  ,o 

~ 

CO 

^ 

"cfl 

O       QJ 

o 

u 

"-J3   b 

C 

nj     ™ 

CM 

CM 

^H 

,H 

a; 

a»    S 

O 

O 

O 

O 

a  £ 

t3 

o 

o 

o 

o 

r- 

t> 

cu 

_c 

c 

o     CC 

°  x 

o 
"C 

CM 

4 

CM 

4 

CM 

1 

o 

1 
O 

<3i 

1 

CM 

03 

1 

CM 

Ph 

O) 

en 

ai 

o> 

05 

OS 

o 

c    « 
.2  to 

en 

o> 

o> 

Ol 

cn 

ai 

CO 

+J      o 

CO     „, 

CtJ 

u   2 

cO 

c 

CO 

c 

CO 

C 

o 

en 

3 

3 

3 

c 

C    M" 

a  .2 

+j 

-M 

cu 
■5 

c 

c 

CD 

CO 

CD      Lh 

* 

* 

>> 

c 

-P 

a    cd 
c  X 

-      CO 

C   cC 

CO 

0) 

"o 
<n 
a 

co 

o 

CJ 
CO 

x 

o 

"CD 

"CD 

CD 

.SP 

X 

3 
c 

CC 

J3 

cn 

o 

.2   o 

"cO 

>, 

>> 

CO 

a 

CD 

_3 

o 

CO" 

CO" 

CO" 

3 

m 

Table 

er  oper 
?rn  Pac 

CD 

i-. 
H 

X 

CO 

cC 

•a 

u 
O 

c 

3 
CJ 
CD 

d 

3 

CD 

CU 

c 

3 
*-> 

CD 
>> 
CD 

4^ 

c 

cC 
o 

X 

c 
u 

CD 
X 
+a 

3 

V 

r- 

CD 

-C 

o,  .^ 

CU3 

be 

bo 

w 

en 

CD 

O 

cC 

CO      (Tj 

CO 

X 

X 

X 

>> 

CO 

■g* 

CO 

he  mean  number  of  ho 
perations.  For  the  two 
and  SP  =  South  Pacific. 

CO 

CO 

-c 

X 

t- 

u 

CO 

CO 

CO 

1 

CO 

CD 

3 

o 

S3 

'cj 

CO 

cu 

X 
+J 

S- 
O 

2: 

u 

S3 

CO 

"3 
a 
o 

CJ 

S3 

o 

CO 
Oh 
0 
0) 
CO 

1 

CJ 
CC 
'u 

CO 
Oh 

C 

CD 
*J 
CO 

1 

3 
< 

C 

u 

CD 

-4-> 

CO 

CO 
CD 

X 
-tJ 
u 
o 
c 

3 

<c 

C 
u 

CD 
+J 
CO 

CO 

CD 
X 
+j 
3 
o 

CO 

EC 

'o 
aj 
CU 

-C 

r-C 

o 

_C 

CO 

+J    o    -  - 

CJ 

C1C    CD    CC 

„ 

1 

howin 
onglin 
n  Paci 

CO 

CO 

bo 

d 

60 
CO 

CO 

CO 

-3 

SB 

CO 

o 

yzed,  s 
ier  of  1 
Wester 

CO 

3 
o 

3 
o 

u 
CD 

CD 

CO 

cd 

T3 
CD 

■a 

_cu 

CO 

cO 

> 

CO 

5 

CO 

o 

tj  -" 

> 

c 

CD 

2 

cj 

X 
o 

-a 
J2 

Cr 

CD 
CO 

c 

CO 

c 

CO 
±3 

fisheries 
the  total  : 
1  Pacific;  \ 

X 

CO 

CD 

<r« 

X 

CO 

a) 

eft 

X 

CJ 

X 

c 

CO 

CO 

•a 
c 

CO 

CO 

■3 
c 

CO 

cu 

-O 

CO 

o 
o 

rC 

V- 

o 

13 

CO 

t3 

CO 

01 

■ft 

CO 

■3 

a 

CO 
•-5 

a 

cO 
<-3 

2         co 

V- 

ST)    h 

a> 

a  -e 

CO 

£2 

ary  of  the 
er  data,  a: 
;  CP  =  Cen 

s 
a 

c 

X 
to 

cC 

■a 
s- 
o 

"cO 
C 
3 

cu 
>, 

CO 

a 

3 

CU 
>> 
CD 

bo 
X 

Ph 

+3 

c 

CO 

c 

3 

C 
CC 

•s 

o 

CO 

a 

3 

+3 

c 

cC 
CD 

C 
cu 

CO   _; 

Summ 
observ 
Pacific 

Ir 

0) 
XI 
en 

E 

CO 

Oh 
2 

X 

cm 

O 

■3 

"CD 

>> 

Oh 

cn 

_3 
X 

Oh 

CO 

3  .2 

184 


Fishery  Bulletin  102(1) 


North  Pacific  Swordfish 


o 


o 
eg 


Western  Pacific  Bigeye 


o 

CM 

I 


o 
I 


Western  Pacific 
Bigeye 


Western  Pacific  Distant 


South  Pacific 
Yellowfin 


South  Pacific  Bluefin 


— I — 
140 


T 


T 


160 


220 


180  200 

Longitude  (degrees) 

Figure  3 

Geographical  distribution  of  the  observer  data  analyzed  for  each  fishery. 


240 


the  expected  number  of  animals  per  hook.  For  each  longline 
segment  (j)  within  each  operation  (£),  we  link  jr  to  a  linear 
predictor  ( ?;(  )  through  the  equation 

rjj  is  then  modeled  as  a  function  of  soak  time: 

r?y  =  ft+ATy,  (5) 

where  TtJ  =  the  hook's  soak  time  (decimal  hours)  in  long- 
line  segment  j; 

P0    =  the  intercept;  and 

/3j  =  the  slope  coefficient,  which  we  term  the  "soak 
time  coefficient." 

Modeling  the  probability  of  a  catch  on  each  individual 
hook  would  result  in  large  numbers  of  zero  observations 
and  thus  test  the  limits  of  current  computer  performance. 
Therefore  we  aggregated  hooks  and  catches  into  hourly 
segments  for  each  longline  operation. 

We  assumed  that  each  longline  segment  had  the  same 
configuration  and  that  the  probability  of  capture  was  the 
same  for  each  segment  within  a  longline  operation.  The 
assumption  may  be  violated  where  segments  pass  through 
different  water  masses  or  where  they  differ  in  depth  profile 
or  baits.  Saturation  of  segments  with  animals  will  also  al- 
ter the  capture  probability  between  segments.  The  effects 


of  water  masses,  depth  profiles,  baits,  and  gear  saturation 
were  not  analyzed  in  the  present  study. 

Capture  probability  may  also  vary  through  the  differen- 
tial exposure  of  segments  to  the  diurnal  cycle  of  night  and 
day.  The  addition  of  dawn  and  dusk  as  fixed  effects  allowed 
us  to  model  diurnal  influences  on  catch  rates. 

Fixed  effects  To  explore  factors  that  might  affect  the  rela- 
tionship between  soak  time  and  catch  rate,  we  added  four 
fixed  effects  to  the  logit  model:  year,  season,  and,  as  men- 
tioned above,  whether  the  segment  was  available  at  dawn 
or  dusk.  A  full  model  without  interaction  terms  would  be 

iu  =  A.  +  /Wj  +  AA>  +  PAi  +  PAj +  P*Yu +  °- 

where  71,     =  the  hook's  soak  time  (decimal  hours)  in  long- 
line  segment  j; 

At     =  an  indicator  of  whether  the  hook  was  exposed 
to  a  dawn  period; 

P     =  an  indicator  of  whether  the  hook  was  exposed 
to  a  dusk  period; 

S: ,  =  the  season  (winter  or  summer); 

Y-  =  the  year; 

Oi  =  the  random  effect  for  operation  that  we  mod- 
eled as  an  independent  and  normally  distrib- 
uted variable  (see  "Random  effects"  section); 
and 
)30-/34  are  parameters  (fixed  effects)  to  be  estimated. 
We  refer  to  fix  as  the  "soak  time  coefficient." 


Ward  et  al.:  The  effect  of  soak  time  on  pelagic  longline  catches 


185 


Table  2 

List  of  common  and  scientific 

names  of  the  species  analyzed.  Also  shown  is 

the  number  of  individuals  of  each  species 

analyzed  in 

each  fishery.  A  dash  indicates  that  the  species  was  not  analyzec 

in  the  present  study 

it  does  not 

necessarily  mean  that  the  spe- 

cies  was  not  taken  in  the  fishery.  In  particular,  observer  data  on 

the  time  of  capture  were  not  aval 

lable  for 

'other  bony  fish"  in  the 

North  Pacific  swordfish  and  Centra]  Pacific  bigeye  tuna  fisheries 

.  NP  =  North  Pacific; 

CP  =  Central  Pacific 

WP  =  Western  Pacific; 

SP  =  South  Pacific;  LN  =  long- 

nosed;  and  SN  =  short-nosed. 

Fishery 

CP 

WP 

SP 

SP 

NP 

bigeye 

bigeye 

WP 

yellowfin 

Bluefin 

Common  name 

Species 

swordfish 

tuna 

tuna 

distant 

tuna 

tuna 

Tuna  and  tuna-like  species 

Albacore 

Thunnus  alalunga 

9707 

23,128 

14,194 

11,976 

21,550 

1399 

Bigeye  tuna 

77?  annus  obesus 

5409 

45,476 

9814 

2581 

1846 

- 

Butterfly  mackerel 

Gasterochisma  melumpus 

— 

— 

— 

— 

— 

533 

Skipjack  tuna 

Katsuwonus  pelamis 

546 

13,882 

1456 

445 

691 

— 

Slender  tuna 

Allothunnus  fallai 

— 

— 

— 

— 

— 

28 

Southern  bluefin 

Thunnus  maccoyii 

— 

— 

— 

— 

1030 

10.537 

Yellowfin  tuna 

Thunnus  albacares 

2811 

21,654 

16,029 

4689 

12,454 

— 

Wahoo 

Acanthocybium  solandri 

383 

5508 

1345 

— 

308 

— 

Billfish 

Black  marlin 

Makaira  indica 

25 

41 

353 

226 

160 

— 

Blue  marlin 

Makaira  nigricans 

981 

2379 

1467 

529 

179 

— 

Sailfish 

Istiophorus  platypterus 

49 

193 

706 

399 

151 

— 

Shortbill  spearfish 

Tetrapturus  angustirostris 

543 

5467 

529 

398 

654 

— 

Striped  marlin 

Tetrapturus  audax 

1963 

8332 

681 

182 

724 

— 

Swordfish 

Xiphias  gladius 

22,457 

1680 

1472 

287 

1173 

92 

Other  bony  fish 

Barracouta 

Thyrsites  atun 

— 

— 

— 

— 

53 

— 

Barracudas 

Sphyraena  spp. 

— 

— 

707 

153 

— 

— 

Escolar 

Lepidocybium  flavubrunneum 

1208 

3983 

1343 

878 

1726 

84 

Great  barracuda 

Sphyraena  barracuda 

32 

743 

303 

442 

92 

— 

Lancetfish  (LN) 

Alepisaurus  ferox 

2788 

30,136 

325 

419 

2868 

610 

Lancetfish(SN) 

Alepisaurus  brevirostris 

— 

— 

155 

84 

257 

59 

Lancetfishes 

Alepisaurus  spp. 

— 

— 

1431 

98 

— 

— 

Long-finned  bream 

Taractichthys  longipinnis 

— 

— 

— 

— 

— 

292 

Mahi  mahi 

Coryphaena  hippurus 

17,463 

19,090 

1436 

211 

447 

— 

Oilfish 

Ruvettus  pretiosus 

555 

1091 

420 

456 

653 

900 

Opah 

Lampris  guttatus 

68 

4724 

527 

129 

80 

213 

Pomfrets 

Family  Bramidae 

— 

— 

623 

60 

— 

40 

Ray's  bream 

Brama  brama 

— 

— 

— 

— 

1074 

10,547 

Ribbonfishes 

Family  Trachipteridae 

— 

— 

— 

— 

— 

22 

Rudderfish 

Centrolophus  niger 

— 

— 

— 

— 

— 

90 

Sickle  pomfret 

Taractichthys  steindachneri 

— 

— 

122 

21 

— 

— 

Slender  barracuda 

Sphyraena  jello 

— 

— 

— 

— 

121 

— 

Snake  mackerel 

Gempylus  serpens 

1971 

9881 

256 

44 

— 

— 

Snake  mackerels 

Family  Gempylidae 

— 

— 

456 

10 

— 

— 

Southern  Ray's  bream 

Brama  spp. 

— 

— 

— 

— 

— 

28 

Sunfish 

Mola  ramsayi 

— 

— 

— 

— 

249 

99 

Sharks  and  rays 

Bigeye  thresher  shark 

Alopias  superciliosus 

149 

1930 

145 

61 

— 

— 

Blacktip  shark 

Carcharhinus  limbatus 

— 

— 

445 

125 

— 

— 

Blue  shark 

Prionace  glauca 

31,503 

31,413 

5601 

1628 

1689 

12.797 

Bronze  whaler 

Carcharhinus  brachyurus 

— 

— 

— 

— 

116 

— 

Crocodile  shark 

Pseudocarcharias  kamoharai 

153 

73 

continued 

186 


Fishery  Bulletin  102(1) 


Table  2  (continued) 

Fishery 

CP 

WP 

SP 

SP 

NP 

bigeye 

bigeye 

WP 

yellowfin 

Bluefin 

Common  name 

Species 

swordfish 

tuna 

tuna 

distant 

tuna 

tuna 

Sharks  and  rays  (continued) 

Dog  fishes 

Family  Squalidae 

— 

— 

— 

— 

— 

60 

Dusky  shark 

Carcharhinus  obscurus 

— 

112 

— 

— 

20 

— 

Grey  reef  shark 

Carcharhinus  amblyrhynchos 

— 

— 

282 

64 

— 

— 

Hammerhead  shark 

Sphyrna  spp. 

— 

— 

142 

191 

22 

— 

Long  finned  mako 

Isurus  paucus 

— 

83 

108 

15 

— 

— 

Oceanic  whitetip  shark 

Carcharhinus  longimanus 

568 

2373 

2376 

384 

142 

— 

Porbeagle 

Lamna  nasus 

— 

— 

— 

— 

27 

1011 

Pelagic  stingray 

Dasyatis  violacea 

2374 

2849 

1212 

248 

534 

109 

Pelagic  thresher  shark 

Alopias  pelagicus 

— 

— 

77 

34 

— 

— 

School  shark 

Galeorhinus  galeus 

— 

— 

— 

— 

— 

143 

Short  finned  mako 

Isurus  oxyrinchus 

476 

685 

408 

169 

432 

128 

Silky  shark 

Carcharhinus  falciformis 

25 

1433 

5396 

2406 

8 

— 

Silvertip  shark 

Carcharhinus  albimarginatus 

— 

— 

168 

74 

— 

— 

Thintail  thresher  shark 

Alopias  vulpinus 

— 

74 

— 

— 

— 

31 

Thresher  shark 

Alopias  superciliosus 

— 

— 

415 

— 

93 

18 

Tiger  shark 

Galeocerdo  cuvier 

— 

— 

56 

18 

38 

— 

Velvet  dogfish 

Zameus  squamulosus 

— 

— 

— 

— 

— 

156 

Whip  stingray 

Dasyatis  akajei 

— 

— 

78 

15 

— 

— 

Seabirds 

Albatrosses 

Family  Diomedeidae 

— 

— 

— 

— 

— 

88 

Petrels 

Family  Procellariidae 

— 

— 

— 

— 

— 

29 

Other  seabirds 

Family  Procellariidae 

— 

— 

— 

— 

38 

200 

All  operations 

104,054 

238,340 

73,212 

30,222 

51,699 

40,343 

To  maintain  a  focus  on  the  effects  of  soak  time,  the  models 
were  limited  to  simple  combinations  of  fixed  effects  and 
interaction  terms.  Dawn  and  dusk  were  added  to  various 
models  of  each  species  in  each  fishery.  To  reduce  complex- 
ity, year  and  season  were  limited  to  models  of  seven  spe- 
cies (bigeye  tuna,  oilfish,  swordfish,  blue  shark,  albacore, 
southern  bluefin  tuna,  long-nosed  lancetfish)  in  the  two 
South  Pacific  fisheries.  The  seven  species  represented 
four  taxonomic  groups  and  the  full  range  of  responses 
observed  in  preliminary  analyses  of  the  soak-time-catch- 
rate  relationship. 

Random  effects  We  added  random  effects  to  all  models  to 
allow  catch  rates  of  segments  within  each  longline  opera- 
tion to  be  related.  The  random  effects  model  assumes  that 
there  is  an  underlying  distribution  from  which  the  true 
values  of  the  probability  of  catching  the  species,  jt,  are 
drawn.  The  distribution  is  the  among-operation  varia- 
tion or  "random  effects  distribution."  The  operations  are 
assumed  to  be  drawn  from  a  random  sample  of  all  opera- 
tions, so  that  the  random  effects  (0()  in  the  relationship 
between  catch  rate  and  soak  time  for  each  operation  (i)  are 


independent  and  normally  distributed  with  0~N(Q,  a2). 
The  random  effects  and  various  combinations  of  the  fixed 
effects  were  added  to  the  linear  predictor  presented  in 
Equation  5. 

For  each  species  in  the  South  Pacific  yellowfin  tuna 
data  set  we  compared  the  performance  of  models  under 
an  equal  correlation  structure  with  that  of  models  under 
an  autoregressive  correlation  structure.  Under  an  au- 
toregressive  structure,  catch  rates  in  the  different  hourly 
segments  within  the  operations  are  not  equally  correlated. 
For  example,  the  correlation  between  segments  might  be 
expected  to  decline  with  increased  time  between  seg- 
ments. However,  we  used  an  equal  correlation  structure 
for  all  models  because  the  Akaike's  information  criterion 
(AIC)  and  Sawa's  Bayesian  information  criterion  (BIO 
indicated  that  there  was  no  clear  advantage  in  using  the 
autoregressive  structure  rather  than  an  equal  correlation 
structure. 

Implementation  We  implemented  the  models  in  SAS 
(version  8.0)  using  GLIMMIX,  a  SAS  macro  that  uses 
iteratively  reweighted  likelihoods  to  fit  generalized  linear 


Ward  et  al.:  The  effect  of  soak  time  on  pelagic  longline  catches 


187 


Seabirds 


Other  fish 


Tuna 


Billfish 


Sharks 


i 
-0.2 


SP  Bluefin 

— « — Other  seabirds  (1  07) 
-Albatross  (0-99) 

Petrel  (1.17) 

— Lancetfish(SN)(1) 
"Opah(1 .03) 

Pomfret(1  16) 

-Lancetfish(LN)(1.02) 

Southern  Ray's  bream  (0.96) 
"© — Long  finned  bream  (111) 
^Ray's  bream  (2.47) 

— © Sunfish  (0.97) 

© Ribbonfish  (0.93) 


Seabirds 


Other  fish 


e Rudderfish  (0.89) 

© Escolar  (0.66) 

-e-Oilfish  (0.98) 
-e-Albacore  (0.94) 

Slender  tuna  (0.9) 

"-Butterfly  mackerel  (0.93) 
eSouthern  bluefin  (1.4) 
—©—Swordfish  (0  9) 
— Thintail  thresher  shark  (0.88) 
— Mako  (0.93) 
©Blue  shark  (1  87) 
-©"Porbeagle  (0.92) 
-© Ray  (0.89) 


Tuna 


Billfish  - 


Sharks 


— I — 
0.0 


"I 


0.2 


SP  Yellowfin 

© — ; Other  seabirds  (1.26) 

— © iBarracouta  (0.99) 

— © — 'Slender  barracuda  (0.98) 

©-?—  Opah  (0.99) 

-©iancetfish  (LN)  (1  14) 

-s-Mahi  mahi  (1.09) 

— «r-Lancetfish  (SN)  (0.96) 

© Great  barracuda  (0.95) 

:-e-Ray,s  bream  (1.71) 
■ — e — Sunfish  (0.99) 
;  — e— Oilfish  (1.23) 

-©"Escolar  (1.33) 
-©"Skipjack  (1 .06) 
©Yellowfin  (2.33) 
—f3— Southern  bluefin  (2.2) 
©Albacore(2.12) 
-r6— Wahoo  (0.96) 

-®-Bigeye(1  16) 
— © — Sailfish  (1  03) 
— 9 — Blue  marlin  (0.88) 

r-e-Shortbill  spearfish  (0.99) 
:— e — Black  marlin  (0.92) 
■  -©-Striped  marlin  (0.94) 
-©"Swordfish  (0.85) 

©i Porbeagle  (0.87) 

j-e Silky  shark  (0.86) 


Tiger  shark  (0.87) 

"Mako  (1.06) 

-Ray  (0  99) 

> — Bronze  whaler  (0.95) 

© — Oceanic  whitetip  (0.99) 

©-Blue  shark  (0.99) 

© Hammerhead  (0.93) 

© Dusky  shark  (0.85) 


-0.2 


Soak  time  coefficient 


0.0  0.2 

Figure  continued  on  next  page. 


Figure  4 

Coefficients  for  the  effect  of  soak  time  on  the  catch  rates  of  the  most  abundant  species  in  each  fishery.  The  coefficients  are  from 
random  effects  models  where  soak  time  is  the  only  factor.  Horizontal  bars  are  95%  confidence  intervals  for  the  estimated  coefficient. 
The  dispersion  parameter  is  shown  in  parentheses  (it  is  1.00  for  species  that  are  distributed  as  predicted  by  the  model,  but  may  be 
higher  for  species  that  have  a  more  clumped  distribution  along  the  longline). 


mixed  models  (Wolfinger  and  O'Connell,  1993).  To  judge 
the  performance  of  the  various  model  formulations,  we 
checked  statistics,  such  as  deviance  and  dispersion,  and 
examined  scatter  plots  of  chi-square  residuals  against  the 
linear  predictor  I  rj)  and  QQ  plots  of  chi-square  residuals. 
We  used  the  AIC  and  BIC  to  compare  the  performance  of 
the  various  model  formulations. 

Variance  in  the  binomial  model  depends  on  only  one  pa- 
rameter, P.  A  dispersion  parameter  is  therefore  necessary 
to  allow  the  variance  in  the  data  to  be  modeled.  In  effect, 
the  dispersion  parameter  scales  the  estimate  of  binomial 
variance  for  the  amount  of  variance  in  the  data.  The  disper- 
sion parameter  will  be  near  one  when  the  variance  in  the 
data  matches  that  of  the  binomial  model.  Values  greater 
than  one  ("over-dispersion")  imply  that  the  species  may 
have  a  clumped  distribution  along  the  longline. 


Results 


Soak  time 


For  most  species,  soak  time  had  a  positive  effect  on  catch 
rates  (Fig.  4).  In  addition  to  being  statistically  significant, 
the  effect  of  soak  time  made  a  large  difference  to  catch 
rates  at  opposite  ends  of  the  longline.  In  the  South  Pacific 
yellowfin  tuna  fishery,  for  example,  the  expected  catch  rates 
of  swordfish  can  vary  from  0.6  (CI  ±0.1)  per  1000  hooks 
(5  hours)  to  1.9  (CI  ±0.3)  per  1000  hooks  (20  hours) 
(Table  3).  A  soak  time  of  5  hours  and  3500  hooks  (if  that 
were  possible)  would  result  in  a  total  catch  of  about 
two  swordfish.  In  contrast,  almost  seven  swordfish  are 
expected  from  a  longline  operation  of  the  same  number  of 
hooks  with  20  hours  of  soak  time. 


188 


Fishery  Bulletin  102(1) 


WP  Bigeye 


WP  Distant 


Other  fish 


Tuna 


Billfish  - 


Sharks 


-0.2 


Great  barracuda  (0.94) 
-Mahi  mahi(1.15) 
-«— Lancetfish  (LN)  (0.99) 
Lancetfish  (SN)  (0.97) 


■e — Snake  mackerel  (1.06) 
-Barracudas  (0.96) 
*-  Opah  (11) 
■9-Escolar  (0.97) 

-9 Sickle  pomfret  (1 .2) 

-e-Pomfret  (0.99) 
- &—  Escolars  (1.07) 
— ° — OHfish  (1.12) 
-^Skipjack  (1  12) 
■«-Wahoo(1) 
?Yellowfin  (1.61) 
:eAlbacore(1.45) 
!eBigeye(1.18) 
tShortbill  spearfish  (0.98) 
-^Swordfish  (1.04) 
_e_Stnped  marlin  (1.23) 
— e—  Sailfish  (1.51) 
— e — Black  marlin  (1.32) 
-o-Blue  marlin  (1.14) 
Hammerhead  (0.87) 
Grey  reef  shark  (1 .49) 

1 Pelagic  thresher  shark  (1.16) 

■®       Bigeye  thresher  shark  ( 1 .06) 
-° Tiger  shark  (1) 


Other  fish  - 


Pomfret  (1.43) 
-Mahi  mahi  (1.45) 
Lancetfish  (LN)  (1.02) 
e—  Great  barracuda  (0.84) 

— e Opah  (1.24) 

e Barracudas  (0.97) 

— e —  — Sickle  pomfret  (2.13) 

Escolar  (1.15) 


-Lancetfish  (SN)  (0  84) 
Snake  mackerel  (3.56) 


-Ollfish  (1.17) 


Tuna 


Billfish 


-e Sllvertip  shark  (1.17) 

■Q-Silky  shark  (115) 
-6— Thresher  sharks  (0.88) 
-s— Short  finned  mako  (0.91 ) 
"^Pelagic  stingray  (0.92) 

"Long  finned  mako  (0  96) 


Sharks  ■ 


-° — Blacktip  shark  (1.3) 
9Blue  shark  (0.93) 
-^Oceanic  whitetip  (0.91) 

e Whip  stingray  (1 .37) 

e Crocodile  shark  (1 .2) 


0.0 


I 

0.2 


Skipjack  (0.91) 

Wahoo  (0.97) 
e-Yellowfin  (2.02) 
&Albacore(1.51) 
-e"Bigeye(1.32) 
■Black  marlin  (0.89) 
Striped  marlin  (1.19) 


— Shortbill  spearfish  (1.17) 
-°—  Sailfish  (1 .04) 
-° — Swordfish  (0.89) 
— s— Blue  marlin  (0  98) 
Tiger  shark  (1.2) 
— Crocodile  shark  (0.95) 
Hammerhead  (0  88) 

Whip  stingray  (1.01) 

Sllvertip  shark  (1 .46) 

■Blue  shark  (0.95) 

Blacktip  shark  (0.91) 

-e-Silky  shark  (1.5) 

e      Pelagic  stingray  (1) 
e     Oceanic  whitetip  (1 .06) 

"Pelagic  thresher  shark  (1.81) 


-Bigeye  thresher  shark  (1.17) 
"Short  finned  mako  (0.91) 


-0.2 


0.0 


0.2 


Soak  time  coefficient 


Figure  4  (continued) 


For  some  species  (e.g.  seabirds,  skipjack  tuna,  and  mahi 
mahi),  soak  time  had  a  negative  effect  on  catch  rates  that 
was  often  statistically  significant  (Fig.  4).  For  skipjack 
tuna  in  the  Western  Pacific  distant  fishery,  for  example. 
catch  rates  decreased  from  1.3  (CI  ±0.2)  per  1000  hooks 
for  a  soak  time  of  5  hours  to  1.0  (CI  ±0.1)  per  1000  hooks 
(20  hours).  Soak  time  had  a  small  or  statistically  insignifi- 
cant effect  on  catch  rates  for  several  species,  such  as  yel- 
lowfin  tuna  and  shortbill  spearfish. 

Fixed  effects 

Exposure  to  dusk  had  a  positive  effect  on  the  catch  rates 
for  most  species  (Fig.  5).  Dusk  often  had  a  negative  effect 
on  the  catch  rates  of  billfish,  such  as  striped  marlin  and 
sailfish.  For  most  species,  however,  the  effect  of  dawn  was 
weaker,  and  it  influenced  the  catch  rates  of  fewer  species. 
Like  soak  time,  timing  made  a  substantial  difference  to 
catch  rates  (Table  4).  For  a  soak  time  of  12  hours  in  the 
South  Pacific  yellowfin  fishery,  for  example,  longlinc  seg- 


ments exposed  to  both  dawn  and  dusk  have  a  catch  rate 
of  2.0  (CI  ±0.5)  escolar  per  1000  hooks.  The  catch  rate  is 
0.8  (CI  ±0.1)  per  1000  hooks  for  segments  that  were  not 
exposed  to  dawn  or  dusk. 

The  effects  of  timing  on  catch  rates  were  most  pro- 
nounced in  the  South  Pacific  bluefin  tuna  fishery.  The 
fishery  also  showed  the  greatest  range  in  soak  time  coef- 
ficients, and  the  coefficients  tended  to  be  larger  than  those 
of  other  fisheries  (Fig.  4). 

Separately,  the  fixed  effects  often  had  statistically  signifi- 
cant relationships  with  catch  rates  of  the  seven  species  that 
we  investigated  in  detail.  However,  the  interaction  between 
soak  time  and  each  fixed  effect  was  less  frequently  signifi- 
cant. Season  was  significant,  for  example,  in  none  of  the 
six  models  that  included  a  soak-time-season  interaction 
term.  By  comparison,  season  was  significant  in  six  of  the 
18  models  that  included  season  as  a  factor  but  not  with  a 
soak-time-season  interaction  term.  The  effect  of  soak  time 
was  not  significant  for  southern  bluefin  tuna  in  any  model 
for  the  South  Pacific  bluefin  tuna  fishery.  It  was  significant 


Ward  et  al.:  The  effect  of  soak  time  on  pelagic  longline  catches 


189 


Tuna 


NP  Swordfish 

hBlgeye  (0  94) 

3 Pacific  bluefin  (0.95) 


CP  Bigeye 


Billfish 


Sharks 


Tuna 


*— Skipjack  (0-84) 
-s-Yellowfin  (0  87) 
^Albacore  (1.03) 
Slue  marlin  (0.91) 
— Shortbill  spearfish  (0.93) 
"Swordfish  (0  96) 
-°-Striped  marlin  (0.88) 
e Sailfish  (0  96) 


Billfish 


-Salmon  shark  (0.96) 


"° — Oceanic  whitetip  (0.98) 

"e Crocodile  shark  (0  89) 

-e — Short  finned  mako  (0  95) 
eBlue  shark  (0.96) 
e Bigeye  thresher  shark  (0.83) 


Sharks 


-0.2 


0.0 


— I 
0.2 


-0.2 


Soak  time  coefficient 


Figure  4  (continued) 


^Skipjack  (0.85) 
eAlbacore  (0.81) 
eYellowfin  (0.93) 
bigeye  (0.88) 

Black  marlin  (0.89) 

eStnped  marlin  (0  86) 
■^Shortbill  spearfish  (0  89) 
"^Blue  marlin  (0.86) 

— e Sailfish  (1.05) 

"^Swordfish  (0.9) 


— Sandbar  shark  (1.24) 

e Bignose  shark  (1.19) 

_e—  Short  finned  mako  (0.94) 

eBlue  shark  (0  81) 
-e-Silky  shark  (0.93) 
-° — Pelagic  thresher  shark  (0.88) 

"^Oceanic  whitetip  (1.01) 
"^Bigeye  thresher  shark  (0.86) 

e Long  finned  mako  (0.86) 

e Thintail  thresher  shark  (0.9) 


0.0 


"Dusky  shark  (1.05) 
-°—  Crocodile  shark  (0.89) 
I 
0.2 


in  36  of  the  48  models  for  the  other  six  species.  We  con- 
cluded that  the  fixed  effects  modified  the  intercept  of  the 
soak-time-catch-rate  relationship,  but  they  rarely  altered 
the  slope  of  the  relationship. 

Akaike's  information  criterion  (AIC)  and  Sawa's  Bayes- 
ian  information  criterion  (BIC )  both  indicated  that  models 
with  soak  time  as  the  only  variable  were  the  most  or  second 
most  parsimonious  model.  This  was  the  case  for  all  models, 
except  for  several  models  of  albacore  and  long-nosed  lan- 
cetfish.  Therefore  the  following  discussion  concentrates  on 
the  effects  of  soak  time  and  timing  on  catch  rates. 


Discussion 

In  considering  results  of  the  random  effects  models,  we 
examined  patterns  in  the  effects  of  soak  time  and  timing 
among  taxonomic  groups,  the  mechanisms  that  may  cause 
the  patterns,  and  their  implications.  First,  however,  we 
investigated  whether  the  effects  were  consistent  for  the 
same  species  between  fisheries. 


Comparison  of  fisheries 

The  effect  of  soak  time  was  consistent  for  several  spe- 
cies between  the  fisheries,  despite  significant  differences 
in  fishing  practices  and  area  and  season  of  activity.  For 
example,  the  soak  time  coefficients  for  species  in  the  South 
Pacific  yellowfin  tuna  fishery  were  very  similar  to  those  of 
the  same  species  in  the  Central  Pacific  bigeye  tuna  fishery 
(r=0.79)  (Fig.  6). 

Several  species  had  a  narrow  range  of  soak  time  coef- 
ficients over  all  the  fisheries  analyzed.  Estimates  of  the 
coefficient  of  yellowfin  tuna,  for  example,  ranged  from  0.00 
(CI  ±0.01)  in  the  South  Pacific  yellowfin  fishery  to  0.04 
( CI  ±0.0 1 )  in  the  North  Pacific  swordfish  fishery.  A  coefficient 
of  0.04  is  equivalent  to  a  difference  of  1.3  yellowfin  tuna  per 
1000  hooks  between  longline  segments  with  soak  times  of 
5  and  20  hours.  The  range  in  coefficients  is  also  small  for 
other  abundant  and  widely  distributed  species,  such  as  al- 
bacore (r=0. 00-0.05)  and  blue  shark  (r=0.01-0.05). 

For  many  species,  however,  the  correlation  between  soak- 
time  coefficients  from  different  fisheries  was  poor  (Fig.  6). 


190 


Fishery  Bulletin  102(1) 


Table  3 

Examples  of  the  effect  of 

soak  time  on  expected  catch 

rates  of  species  in  the  South  Pacific  yellowfin  tuna 

ishery. 

The  expected  catch  rates 

i  number  per  1000  hooks  I  are 

predicted  from  the  soak-time  coefficient  for  each 

species 

for  longline  segments  exposed  to  a  dusk  period  with 

a  soak 

time  of  5  or  20  hours.  Figu 

re  4  shows  the  95%  confidence 

intervals  for  soak-time  coe 

fficients  used  to  calculate  the 

expected  catch  rates.  LN  = 

ong-nosed;  SN  =  short-nosed. 

Species 

Soak  time 

h) 

5 

20 

Tuna  and  tuna-like  species 

Albacore 

15.5 

13.4 

Bigeye  tuna 

1.1 

2.3 

Skipjack  tuna 

1.3 

1.0 

Southern  bluefin  tuna 

5.2 

5.5 

Yellowfin  tuna 

8.4 

7.7 

Billfish 

Black  marlin 

0.4 

1.6 

Blue  marlin 

1.2 

0.4 

Sailfish 

0.8 

1.0 

Shortbill  spearfish 

1.0 

1.6 

Striped  marlin 

0.8 

1.0 

Swordfish 

0.6 

1.9 

Other  bony  fish 

Barracouta 

0.8 

0.7 

Escolar 

0.8 

3.1 

Great  barracuda 

0.9 

1.1 

Lancetfish  (LN) 

2.7 

2.4 

Lancetfish  (SN) 

1.6 

1.4 

Mahi  mahi 

1.0 

0.9 

Oilfish 

0.8 

2.2 

Opah 

0.7 

0.5 

Ray's  bream 

1.8 

2.0 

Slender  barracuda 

1.7 

1.6 

Sunfish 

0.6 

1.3 

Wahoo 

1.0 

1.1 

Sharks  and  rays 

Blue  shark 

1.1 

2.0 

Bronze  whaler 

0.7 

0.8 

Dusky  shark 

0.4 

0.8 

Hammerhead 

0.2 

1.8 

Mako 

0.6 

0.8 

Oceanic  whitetip 

0.5 

0.9 

Porbeagle 

1.2 

1.1 

Pelagic  stingray 

0.9 

1.2 

Thresher  shark 

0.6 

1.0 

Tiger  shark 

0.5 

0.5 

Table  4 

Examples  of  the  effect  of  timing  on  expected  catch  rates 
of  species  in  the  South  Pacific  yellowfin  tuna  fishery.  The 
expected  catch  rates  (number  per  1000  hooks  I  are  pre- 
dicted from  the  soak-time  coefficient  for  each  species  for  a 
longline  operation  with  a  soak  time  of  12  hours.  The  differ- 
ent catch  rates  are  for  longline  segments  exposed  to  nei- 
ther the  dawn  or  dusk  period,  for  dawn  only,  and  for  dawn 
and  dusk  periods.  LN  =  long-nosed;  SN  =  short-nosed. 


Period 

Neither 

Dawn 

Dawn 

Species 

period 

only 

+  dusk 

Tuna  and  tuna-like  species 

Albacore 

12.3 

14.0 

16.5 

Bigeye  tuna 

0.9 

1.2 

2.1 

Skipjack  tuna 

1.4 

1.2 

1.0 

Southern  bluefin  tuna 

3.8 

2.9 

4.1 

Yellowfin  tuna 

7.7 

7.6 

8.0 

Billfish 

Black  marlin 

1.2 

0.6 

0.4 

Blue  marlin 

0.4 

1.0 

1.4 

Sailfish 

0.8 

0.7 

0.7 

Shortbill  spearfish 

1.3 

0.9 

0.9 

Striped  marlin 

0.8 

0.9 

0.9 

Swordfish 

0.5 

0.7 

1.3 

Other  bony  fish 

Barracouta 

1.1 

1.2 

0.7 

Escolar 

0.8 

1.0 

2.0 

Great  barracuda 

1.0 

0.8 

0.8 

Lancetfish  (LN) 

2.8 

2.7 

2.5 

Lancetfish  (SN) 

1.2 

1.1 

1.3 

Mahi  mahi 

1.2 

1.3 

1.1 

Oilfish 

0.8 

1.1 

1.8 

Opah 

0.5 

0.5 

0.6 

Ray's  bream 

0.8 

0.7 

1.6 

Slender  barracuda 

2.0 

1.5 

1.2 

Sunfish 

0.8 

0.6 

0.7 

Wahoo 

1.2 

1.3 

1.1 

Sharks  and  rays 

Blue  shark 

1.3 

1.4 

1.4 

Bronze  whaler 

0.6 

0.9 

1.0 

Dusky  shark 

0.1 

0.1 

0.6 

Hammerhead 

0.4 

0.2 

0.3 

Mako 

0.7 

0.8 

0.8 

Oceanic  whitetip 

0.7 

0.8 

0.7 

Porbeagle 

1.0 

0.6 

0.6 

Pelagic  stingray 

0.9 

0.9 

1.1 

Thresher  shark 

0.6 

0.6 

0.7 

Tiger  shark 

0.4 

0.5 

0.7 

For  a  few  species  (e.g.  tiger  shark)  the  poor  correlation  may 
have  been  a  function  of  small  sample  sizes  and  the  wide 
confidence  intervals  of  the  estimates.  For  other  species  the 
estimates  were  well  determined,  yet  poorly  correlated, 
e.g.  the  coefficient  for  short-nosed  lancetfish  was  0.09 


(CI  ±0.05)  in  the  Western  Pacific  distant  fishery  compared 
to  0.01  (CI  ±0.04)  in  the  Western  Pacific  bigeye  tuna  fishery. 
Therefore,  we  urge  caution  in  applying  our  estimates  to  the 
same  species  in  longline  fisheries  in  other  areas. 


Ward  et  a!.:  The  effect  of  soak  time  on  pelagic  longlme  catches 


191 


SP  Yellowfin                                                                     WP  Bigeye 

o 

dusk  preference 

dawn  &  dusk                                    ^  - 

dusk  preference 

dawn  &  du>k 

• 
Raj 's  bream 

• 

•  Swordfish 

C3 

•                                                                o 

Oilfish         . 
Hammerhead      4                            Blue  martin 

Oiltish 

• 
Tiger  shark      * 

sk  coefficient 

0.0 

• 
•    •  Tiger  shark 

;•  L ' 

•  • 

• 
*          •                   Swordfish 

Q 

•  • 

l 

• 

3 

• 
Black  marlin 

9  ' 

• 
Spnped  marlin 

• 

•      • 

Black  marlin 

Hammerhead  • 

c 

nol  dawn  or  dusk                          dawn  preference                                    _ 

not  dawn  or  dusk                            dawn  preference 

~~ 

-1.0           -0.5            0.0            0.5             10                                           -1.0           -0.5            0.0            0.5             10 

Dawn  coefficient 

Figure  5 

Pair-wise  comparison  of  coefficients  for  the  effects  of  dawn  and  dusk  on  catch  rates  for  two 

fisheries.  The  shading  of  each  symbol  represents  the  sum  of  the  standard  errors  of  the  dawn 

and  dusk  estimates  (heavy  shading  for  the  lowest  standard  errors;  light  shading  for  large 

standard  errors).  Not  all  species  names  are  shown. 

Underlying  mechanisms 

The  broad  taxonomic  groups  taken  by  longlme  each  rep- 
resent a  wide  range  of  life  history  strategies  and  feeding 
behaviors.  Nevertheless,  the  results  show  a  tendency  for 
soak  time  to  have  a  positive  effect  on  catch  rates  of  most 
shark  species  (Fig.  4).  It  also  had  a  positive  effect  on  catch 
rates  of  many  billfish  species,  including  striped  marlin, 
black  marlin,  and  swordfish.  There  is  no  clear  pattern  in 
the  effect  of  soak  time  on  catch  rates  of  tuna  or  other  bony 
fish.  It  had  a  negative  effect  on  the  four  seabird  groups. 

The  results  imply  that  the  ability  of  a  species  to  stay 
alive  and  to  escape  or  avoid  scavengers  while  hooked  is 
important  in  determining  the  catch  that  is  actually  brought 
on  board.  The  effect  of  soak  time  is  significantly  correlated 
with  the  ability  of  a  species  to  survive  while  hooked  on  the 
longline  in  the  four  fisheries  with  data  on  survival  (Fig.  7). 
Soak  time  has  a  strong,  positive  effect  on  catch  rates  of  spe- 
cies like  blue  shark,  which  are  almost  always  alive  when 
branchlines  are  retrieved.  Species  like  skipjack  tuna  and 
seabirds  are  usually  dead.  Soak  time  had  a  negative  effect 
on  their  catch  rates.  The  opposite  trend  would  be  expected 
if  escape  is  a  significant  process  that  affects  catch  rates;  if 
escape  is  important,  soak  time  should  have  a  negative  af- 
fect on  the  catch  rates  of  the  most  active  species.  Therefore 
removal  by  scavengers  is  likely  to  be  more  important  than 
escape  in  determining  catch  rates  for  many  species. 


Longline  branchlines  are  usually  20-30  m  in  length,  al- 
lowing considerable  room  for  a  live,  hooked  animal  to  evade 
predators  or  scavengers.  Or,  scavengers  may  be  attracted 
by  immobile  and  dead  animals.  The  scavenger  avoidance 
hypothesis  is  attractive,  but  it  is  difficult  to  test  with  ob- 
server data.  Data  from  hook-timer  experiments  may  help 
to  estimate  the  total  number  of  animals  that  are  lost  or 
removed  from  the  longline.  Data  presented  by  Boggs  ( 1992 ) 
showed  a  large  number  of  hook-timers  that  were  triggered 
but  which  did  not  hold  an  animal  when  the  branchline  was 
retrieved,  e.g.  his  data  show  that  2-4"7r  of  hook-timers  on 
10,236  branchlines  that  had  "settled"  were  activated  but 
did  not  have  an  animal.  It  is  unclear  whether  the  trigger- 
ing of  hook-timers  was  due  to  equipment  malfunction  or 
whether  it  represents  high  loss  rates.  Of  particular  signifi- 
cance to  the  question  of  loss  rates  is  the  fact  that  current 
hook-timer  technology  does  not  identify  the  species  that 
were  lost  and  whether  they  were  alive  or  dead. 

We  noticed  that  soak-time  coefficients  tended  to  be  poorly 
correlated  between  fisheries  and  that  the  effects  of  soak 
time  on  catch  rates  were  most  pronounced  in  the  South  Pa- 
cific bluefin  tuna  fishery.  Our  scavenging  hypothesis  might 
explain  those  observations  as  evidence  that  the  activities  of 
scavengers  vary  between  fisheries.  For  example,  blue  shark 
might  be  an  important  scavenger.  They  are  most  abundant 
in  temperate  areas  (Last  and  Stevens,  1994).  Our  analyses 
showed  a  predominance  of  negative  soak-time  coefficients 


192 


Fishery  Bulletin  102(1) 


CD 

a. 


=0.10* 


•>•' 

y 


-0  1         ( I  0  2 

WP  Distant  coefficient 


m 
o. 


/-0.65* 


-o  I         o  o         ii  I         ii  2 
SP  Yellowfin  coefficient 


Q. 


,-=oo,s 


-ol         I  0.2 

CP  Bigeye  coefficient 


r=0.15* 


o 
u 

•ii 

<u 

•  • 

0 

=    - 

• 

fl 

en 

• 

n 

• 

C/3 

-  . 

-0  1 

0 

Ml 

SP  Yellowfin 

coefficient 

o 

S 
w 

0_ 


CL 


r=0. 12ns 


«*• 


-ii  1  0.0  0  I  0  2 

CP  Bigeye  coefficient 


r=0.79* 


3= 
o 


S      - 


•  • 


-i'  |  0.0  ii  |  (12 

CP  Bigeye  coefficient 


Figure  6 

Pair-wise  comparison  of  soak  time  coefficients  for  species  that  were  common  to  fisheries.  The  coefficients  are  from  ran- 
dom effects  models  where  soak  time  is  the  only  factor.  The  shading  of  each  symbol  represents  the  size  of  the  standard 
error  of  the  estimate.  V  is  the  correlation  coefficient  of  a  linear  regression  of  coefficients  ( *  indicates  that  the  regres- 
sion slope  is  significantly  different  from  one  at  the  95%  level,  whereas  "ns"  indicates  that  the  null  hypothesis,  that  the 
regression  slope  equals  one,  cannot  be  rejected). 


in  the  South  Pacific  bluefin  tuna  fishery — perhaps  indicat- 
ing that  loss  rates  may  be  particularly  high  where  blue 
shark  are  abundant. 

Nevertheless,  there  are  other  plausible  explanations  for 
the  differences  in  soak-time  effects  between  fisheries.  The 
movement  of  branchlines  caused  by  wave  action  will  cause 
animals  to  fall  off  hooks,  especially  when  branchlines  are 
near  the  sea  surface.  Rough  seas  are  frequently  experi- 
enced in  the  North  Pacific  swordfish  and  South  Pacific 
bluefin  tuna  fisheries  where  the  soak-time  effects  were 
most  pronounced. 

Another  source  of  loss  might  be  the  breakage  of  longline 
branchlines.  The  animal's  teeth  or  rostrum  might  abrade 
the  branchline  causing  the  branchline  to  fail  and  allow- 
ing the  animal  to  escape.  In  this  regard  it  is  noteworthy 
that  Central  Pacific  bigeye  tuna  longliners  often  use  wire 
for  the  end  of  branchlines  or  "leader"  whereas  North  Pa- 
cific wwordfish  longliners  use  monofilament  nylon  leaders 
(Ito4). 


Mortality  estimates 

The  results  of  our  study  show  that  longline  catch  rates  that 
are  not  adjusted  for  the  effects  of  soak  time  will  under- 
estimate the  level  of  mortality  of  several  species  because 
they  are  lost  after  being  hooked.  The  soak  time  effect  was 
negative  for  albatrosses  and  other  seabirds.  This  finding 
agrees  with  field  observations  (e.g.  Brothers,  1991)  that 
most  seabirds  are  taken  during  longline  deployment  in 
the  brief  period  after  the  bait  is  cast  from  the  vessel  until 
the  bait  sinks  beyond  the  depth  that  seabirds  can  dive  to. 
Those  observations  indicate  that  counts  of  seabirds  when 
they  are  brought  on  board  do  not  cover  the  total  number 
hooked  because  many  fall  off  or  are  removed  by  scavengers 
or  are  lost  during  the  operation. 


1  It...  K.  2002.  Personal  commun.  National  Marine  Fisheries 
Service  (NOAA),  2570  Dole  Street,  Honolulu  Hawaii  96822- 
2396. 


Ward  et  al.:  The  effect  of  soak  time  on  pelagic  longlme  catches 


193 


WP  Bigeye 

CM    . 

d 

d 

Whip  stingray 

o 
o 

m 
• 

■a 
•  •         • 

• 
•       • 

•           m 

r.~    • 
•• 

Skipjack                  • 

• 

ci 

r=0.42* 

WP  Distant 

Grej  reef  shark 

• 

• 

• 

• 
•  • 

• 

• 

• 

• 
• 

• 

• 

•"     •         • 

•  • 
• 

• 
Skipjack 

r=0.46* 

0  2"  4(1  <M  80         100 


SP  Bluefin 


0  20         -40         no         80        100 


SP  Yellowfin 


•  Oiltish 

Escolar 


"...      1  %  « 


Other  seabirds 


=0.32ns 


Escnlai 


:  •:    • 


Skipjack 


f": 


=0.54 


20         40         60         80         100 


0  20         40         60         80        100 


Proportion  alive  (%) 

Figure  7 

Soak-time  coefficients  plotted  against  the  proportion  of  each  species  reported  to  be  alive 
when  brought  on  board.  Not  included  are  species  where  less  than  ten  individuals  for  the 
fishery  had  a  record  of  life  status.  The  coefficients  are  from  random  effects  models  where 
soak  time  is  the  only  factor.  The  shading  of  each  symbol  represents  the  size  of  the  standard 
error  of  the  estimate.  The  proportion  alive  is  assumed  to  be  measured  without  error.  V  is 
the  correlation  coefficient  of  a  linear  regression  of  coefficients  (*  indicates  that  the  regres- 
sion slope  is  significantly  different  from  zero  at  the  95<7c  level i. 


Seabirds  provide  a  unique  case  for  estimating  loss  rates 
because  they  are  only  caught  when  the  longline  is  deployed 
(Brothers.  1991).  Within  minutes  of  the  branchline  being 
deployed,  the  capture  rate  ( a  in  Eq.  4 )  falls  to  zero  whereas 
the  loss  rate  /3  might  be  constant  or  it  might  vary.  There- 
fore, the  probability  of  a  seabird  being  on  a  hook  when  the 
branchline  is  retrieved  is 


n\T) 


-0T 


(6) 


We  estimated  a  soak-time  coefficient  of -0.0302  (CI 
±0.0462)  for  albatrosses  in  the  South  Pacific  bluefin  tuna 


fishery.  Substituting  0.0302  for  ft  in  Equation  6  and  10.4 
hours  for  T  ( the  average  soak  time  of  hooks  deployed  by 
the  longliners),  we  estimated  that  27%  of  albatrosses  are 
lost  after  being  hooked  but  before  the  branchlines  are 
retrieved.  The  loss  rate  is  about  12%  for  petrels  1/3=0.0123) 
and  45%  for  other  seabirds  (0=0.0582).  It  is  about  26%  for 
other  seabirds  in  the  South  Pacific  yellowfin  tuna  fishery 
(j6=0.0307,  T=10.0  hours). 

For  fishes  and  sharks,  we  do  not  know  how  the  probabil- 
ity of  capture,  or  capture  rate,  or  loss  rate  varies  during 
a  longline  operation.  However,  hook-timer  experiments 


194 


Fishery  Bulletin  102(1) 


and  observer  programs  may  provide  estimates  of  those  pa- 
rameters. Broad  limits  for  the  probability  of  capture  may 
also  be  obtained  if  observers  were  to  report  the  number  of 
branchlines  that  are  retrieved  with  missing  baits  or  miss- 
ing hooks. 

For  most  species,  capture  rates  must  balance  or  outweigh 
loss  rates.  In  this  case,  captures  result  from  the  increased 
exposure  of  animals  to  the  longline  as  a  result  of  movement 
and,  perhaps,  the  dispersal  of  chemical  attractants  during 
the  operation.  However,  we  must  stress  that  losses  are  also 
likely  to  be  occurring  for  the  species  that  have  positive  co- 
efficients. The  analyses  indicate  the  relative  levels  of  loss 
between  longline  segments  of  varying  soak  time.  Other 
than  those  for  seabirds,  we  cannot  estimate  the  levels  of 
catch  that  are  lost. 

Adding  to  the  uncertainty  over  loss  rates  is  the  unknown 
fate  of  lost  animals.  For  seabirds  it  is  known  that  most 
drown  soon  after  being  hooked.  The  few  seabirds  that  sur- 
vive while  hooked  eventually  drown  during  longline  re- 
trieval (Brothers,  1991).  However,  it  is  not  known  whether 
other  lost  animals  are  dead  or  alive. 

Results  of  our  analyses  may  also  be  useful  for  monitoring 
programs.  Observers  are  increasingly  being  placed  on  long- 
liners  to  collect  data  on  bycatch  and  to  independently  verify 
data  reported  in  logbooks.  A  sampling  approach  is  neces- 
sary in  some  fisheries  because  observers  are  often  unable 
to  monitor  the  entire  longline  retrieval.  Indications  that 
catch  rates  of  some  species  at  the  end  of  the  retrieval  are 
double  those  at  the  beginning  necessitate  care  in  designing 
observer  monitoring  protocols  and  in  the  interpretation  of 
the  data.  Observers  could  also  collect  information  on  the 
number  of  hooks  retrieved  without  baits.  Such  data  would 
greatly  improve  the  estimates  of  a  and  fi  required  for  the 
theoretical  model.  For  the  empirical  model,  catch  rate  data 
from  research  surveys  where  longline  segments  have  very 
short  (<4  hour)  soak  times  would  improve  estimates  of 
soak-time  coefficients. 

Historical  changes 

The  interaction  of  year  and  soak  time  was  rarely  significant 
for  the  random  effects  models  of  the  seven  species  exam- 
ined in  detail.  This  might  suggest  that  soak-time-catch- 
rate  relationships  are  stable  over  time.  However,  the  range 
of  years  that  we  analyzed  was  limited  to  1992-97.  Over 
larger  time  scales  there  have  been  large  variations  in  the 
abundance  of  individual  species  and  the  mix  of  species 
comprising  the  pelagic  ecosystem.  We  cannot  predict  how 
soak-time-catch-rate  relationships  would  change  with 
those  long-term  variations. 

Our  original  motivation  for  examining  the  effects  of 
soak  time  was  the  hypothesis  that  the  number  of  hooks 
per  operation  and  soak  time  have  increased  since  longlin- 
ing  commenced  and  that  this  may  have  resulted  in  an 
overestimation  of  billfish  catch  rates  in  early  years.  Wardr' 
presented  information  on  temporal  trends  in  soak  time 


5  Ward,  P.  2002.  Historical  changes  and  variations  in  pela- 
gic longline  fishing  operations,  http://fish.dal.ca/-myers/pdf 
papers.html.     (Accessed  22  February  2003.1 


and  timing  for  several  longline  fleets.  Although  there  is 
uncertainty  over  the  early  operations,  the  available  infor- 
mation indicates  significant  historical  changes  in  Japan's 
distant-water  longline  operations.  Average  soak  time 
shows  a  decline  from  over  11.5  hours  before  1980  to  10.0 
hours  in  the  1990s.  For  species  with  a  negative  soak-time 
coefficient,  this  apparently  modest  reduction  in  soak  time 
would  inflate  catch  rate  estimates  for  recent  years.  It  would 
result  in  reduced  catch-rate  estimates  for  species  with  posi- 
tive coefficients.  For  example,  the  expected  catch  rate  for 
swordfish  is  0.94  (CI  ±0.06)  per  1000  hooks  for  a  soak  time 
of  11.5  hours  compared  to  0.82  (CI  ±0.06)  per  1000  hooks 
for  10.0  hours. 

More  significant  may  be  changes  in  the  timing  of  op- 
erations. During  1960-80  most  baits  used  with  Japan's 
distant-water  longliners  were  available  to  fish  at  dawn 
whereas  about  50%  were  also  available  at  dusk.  Longlines 
were  deployed  and  retrieved  at  later  times  in  the  1990s  so 
that  about  30%  of  baits  were  available  at  dawn  and  about 
70%  available  at  dusk.  In  the  case  of  swordfish,  the  changes 
in  timing  would  moderate  the  effects  of  reduced  soak  time. 
The  expected  catch  rate  for  swordfish  is  0.89  per  1000 
hooks  in  the  early  operations  compared  to  0.83  per  1000 
hooks  in  the  later  operations. 


Conclusions 

The  results  have  important  implications  for  fishery  man- 
agement and  assessments  that  rely  on  longline  catch 
data.  Modifications  to  data  collection,  such  as  recording 
the  number  of  hooks  with  missing  baits  during  longline 
retrieval,  would  greatly  improve  mortality  estimates.  The 
mortality  of  species  like  seabirds  is  significantly  higher 
than  previously  estimated.  Such  underestimation  may  be 
particularly  critical  for  the  assessment  and  protection  of 
threatened  species  of  seabirds.  Furthermore,  the  changes 
in  timing  and  reduction  in  soak  time  have  resulted  in  a 
systematic  bias  in  estimates  of  mortality  levels  and  abun- 
dance indices  for  many  species.  For  species  like  swordfish, 
where  soak  time  has  a  positive  effect  on  catch  rates,  the 
stocks  might  be  in  better  shape  than  predicted  by  current 
assessments  ( if  assessments  were  solely  based  on  catch  and 
effort  data).  The  opposite  situation  would  occur  for  species 
with  negative  soak-time  coefficients:  assessments  that  use 
long  time-series  of  longline  catch  data  will  over-estimate 
the  species'  abundance  so  that  population  declines  are 
more  severe  than  previously  believed. 


Acknowledgments 

Grants  from  the  Pew  Charitable  Trust,  Pelagic  Fisheries 
Research  Program,  and  the  Killam  Foundation  provided 
financial  support  for  this  work.  Peter  Williams  (Secretariat 
of  the  Pacific  Community).  U.S.  National  Marine  Fisher- 
ies Service  staff  (Kurt  Kawamoto,  Brent  Miyamoto,  Tom 
Swenarton,  and  Russell  Ito )  and  Thim  Skousen  (Australian 
Fisheries  Management  Authority)  provided  observer  data 
and  operational  information  on  the  fisheries.  We  are  espe- 


Ward  et  al.:  The  effect  of  soak  time  on  pelagic  longline  catches 


195 


cially  grateful  to  the  observers  who  collected  the  data  used 
in  this  study  and  thank  the  masters,  crew  members,  and 
owners  of  longliners  for  their  cooperation  with  the  observer 
programs.  Pierre  Kleiber,  Ian  Jonsen,  Julia  Baum,  Boris 
Worm  and  an  anonymous  referee  provided  many  useful 
comments  on  the  manuscript. 


Literature  cited 

Boggs.  C.  H. 

1992.     Depth,  capture  time,  and  hooked  longevity  of  longline- 
caught  pelagic  fish:  timing  bites  of  fish  with  chips.     Fish. 
Bull.  90:642-658. 
Brothers,  N. 

1991.     Albatross  mortality  and  associated  bait  loss  in  the 
Japanese  longline  fishery  in  the  Southern  Ocean.     Biol. 
Conserv.  55:255-268. 
Carey,  F.  G. 

1990.     Further  acoustic  telemetry  observations  of  swordfish. 
Mar.  Recr.  Fish.  13:103-22.     Sportfishing  Institute,  Wash- 
ington D.C. 
Deriso,  R.  B.,  and  A.  M.  Parma. 

1987.     On  the  odds  of  catching  fish  with  angling  gear.    Trans. 
Am.  Fish.  Soc.  116:244-256. 
Ferno,  A.,  and  I.  Huse. 

1983.     The  effect  of  experience  on  the  behaviour  of  cod  iGadus 


morhua  L.)  towards  a  baited  hook.     Fish.  Res.  (Amst.)  2: 
19-28. 
Last,  P.  R.,  and  J.  D.  Stevens. 

1994.  Sharks  and  rays  of  Australia.  CSIRO  Australia. 
513  pages  +  color  plates. 

Murphy,  G.  I. 

1960.  Estimating  abundance  from  longline  catches.  J.  Fish. 
Res.  Board  Can.  17:33-40. 

Sivasubramaniam,  K. 

1961.  Relation  between  soaking  time  and  catch  of  tunas  in 
longline  fisheries.    Bull.  Jpn.  Soc.  Sci.  Fish.  27:835-845. 

Somerton,  D.  A.,  and  B.  S.  Kikkawa. 

1995.  A  stock  survey  technique  using  the  time  to  capture 
individual  fish  on  longlines.  Can.  J.  Fish.  Aquat.  Sci.  52: 
260-267. 

Ward,  P.  J. 

1996.  Japanese  longlining  in  eastern  Australian  waters, 
1962-90,  249  p.     Bureau  of  Resource  Sciences,  Canberra. 

Ward,  P.,  and  S.  Elscot. 

2000.     Broadbill  swordfish:  status  of  world  fisheries,  242  p. 
Bureau  of  Rural  Sciences,  Canberra. 
Wolfinger,  R.  and  M.  O'Connell. 

1993.     Generalized  linear  mixed  models:  a  pseudo-likelihood 
approach.     J.  Stat.  Comput.  Simul.  48:233-243. 
Zhou,  S.,  and  T.  C.  Shirley. 

1997.  A  model  expressing  the  relationship  between  catch 
and  soak  time  for  trap  fisheries.  N.  Am.  J.  Fish.  Manag. 
17:  482-487. 


196 


Abstract— Annual  mean  fork  length 
(FL)  of  the  Pacific  stock  of  chub  mack- 
erel {Scomber  japonicus )  was  examined 
for  the  period  of  1970-97.  Fork  length  at 
age  0  (6  months  old)  was  negatively  cor- 
related with  year-class  strength  which 
fluctuated  between  0.2  and  14  billion  in 
number  for  age-0  fish.  Total  stock  bio- 
mass  was  correlated  with  FL  at  age  but 
was  not  a  significant  factor.  Sea  surface 
temperature  (SST)  between  38-40°N 
and  141-143°E  during  April-June 
was  also  negatively  correlated  with  FL 
at  age  0.  A  modified  von  Bertalanffy 
growth  model  that  incorporated  the 
effects  of  population  density  and  SST  on 
growth  was  well  fitted  to  the  observed 
FL  at  ages.  The  relative  FL  at  age  0  for 
any  given  year  class  was  maintained 
throughout  the  life  span.  The  variabil- 
ity in  size  at  age  in  the  Pacific  stock  of 
chub  mackerel  is  largely  attributable  to 
growth  during  the  first  six  months  after 
hatching. 


Effects  of  density-dependence  and 

sea  surface  temperature  on  interannual  variation 

in  length-at-age  of  chub  mackerel 

(Scomber  japonicus)  in  the 

Kuroshio-Oyashio  area  during  1970-1997 

Chikako  Watanabe 
Akihiko  Yatsu 

National  Research  Institute  of  Fisheries  Science 

Fisheries  Research  Agency 

2-12-4  Fukuura,  Kanazawa 

Yokohama  236-8648.  Japan 

E-mail  address  (for  C  Watanabe):  falconer  a  affrc  go  ip 


Manuscript  approved  for  publication 
22  September  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:196-206(2004). 


Variability  in  growth  of  marine  fishes 
has  been  attributed  to  the  effects  of 
density-dependence  or  environmental 
factors  such  as  water  temperature,  or 
to  the  effects  of  both  factors  (e.g.  Moyle 
and  Cech,  2002).  Size-at-age  data  are 
crucial  because  they  are  necessary 
for  stock  assessment  methods  such 
as  virtual  population  analysis,  yield 
per  recruit,  and  spawning-per-recruit 
analyses  (Pauly,  1987;  Mace  and  Sissen- 
wine,  1993;  Haddon,  2001)  and  are  pos- 
sibly useful  for  detecting  regime  shifts 
as  well  (Yatsu  and  Kidoroko,  2002). 
Around  Japan,  the  effects  of  population 
density  and  sea  water  temperature  on 
fish  growth  have  been  shown  for  the 
Pacific  stock  of  chub  mackerel  (Scomber 
japonicus)  (Iizuka,  1974),  Japanese 
Spanish  mackerel  {Scomberomorus 
niphonius)  (Kishida,  1990),  the  Pacific 
and  Tsushima  Current  stocks  of  Japa- 
nese sardine  (Sardinops  melanostietus) 
(Hiyama  et  al,  1995;  Wada  et  al.,  1995), 
and  Japanese  common  squid  ( Todarodes 
pacificus)  (Kidokoro,  2001). 

The  Pacific  stock  of  chub  mackerel  is 
one  of  the  most  important  commercially 
exploited  fish  populations  in  Japan  and 
has  been  managed  by  the  total  al- 
lowable catch  (TAC)  system  in  Japan 
since  1997.  Chub  mackerel  seasonally 
migrate  along  the  Pacific  coast  of  Japan 
from  Kyushu  to  Hokkaido.  They  spawn 
in  the  coastal  waters  around  Izu  Islands 
and  off  southwestern  Japan  between 
February  and  June  (Fig.  1,  Watanabe, 
1970;  Usami,  1973;  Murayama  et  al., 
1995;  Watanabe  et  al.,  1999).  Adult  fish 
(after  spawning)  and  their  offspring 
migrate  eastward  along  the  Pacific 


coast  with  the  Kuroshio  Current.  Ju- 
venile mackerel  of  about  6  months  old 
usually  recruit  to  the  purse-seine  and 
set-net  fisheries  off  the  coast  of  north- 
eastern Japan  at  the  end  of  summer 
(Fig.  1,  Odate,  1961;  Kawasaki,  1966; 
Watanabe,  1970;  Iizuka,  1974).  The 
total  catch  of  the  Pacific  stock  of  chub 
mackerel  increased  during  the  1960s 
and  1970s,  peaked  at  1.5  million  metric 
tons  in  1978,  and  then  declined  to  2.3 
thousand  tons  in  1990  (Fig.  2).  The  es- 
timated total  biomass  increased  in  the 
1970s  from  2.8  million  tons  in  1970  to 
5.9  million  tons  in  1977,  and  the  consec- 
utive occurrences  of  large  year  classes 
exceeded  7  billion  age-0  (6-month-old) 
fish  in  the  early  and  mid  1970s.  In  1990, 
the  biomass  was  reduced  to  a  minimum 
of  0.2  million  tons  in  1990  (Table  1,  Fig. 
2;  Yatsu  et  al.1).  Relatively  large  year 
classes  occurred  in  1992  (2.8  billion 
fish)  and  1996  (4.5  billion  fish),  and 
the  total  biomass  increased  in  the  mid 
1990s,  but  it  remained  at  about  10% 
of  the  level  attained  in  the  mid  1970s 
(Yatsu  et  al.1). 

On  the  basis  of  year-class  strength 
and  variations  in  fork  length  (FL)  at 
ages  0-2  for  the  11  year  classes  present 
from  1963  to  1973,  Iizuka  (1974)  sug- 
gested an  effect  of  density-dependent 
growth  on  young  chub  mackerels.  In 


1  Yatsu,  A.,  C.  Watanabe,  and  H.  Nishida. 
2001.  Stock  assessment  of  the  Pacific 
stock  of  chub  mackerel  in  fiscal  2000 
year.  /;;  Stock  assessment  report,  p. 
64-87.  |In  Japanese.  Available  from  Fish- 
eries Research  Agency,  2-12-4  Fukuura, 
Kanazawa.  Yokohama  236-8648.  Japan.] 


Watanabe  and  Yatsu:  Interannual  variation  in  length  at  age  of  Scomber  japonicus 


197 


45°N 


40CN 


35°N 


30°N  - 


Kuroshio-Oyashio  transition  zone 


izu  Islands 


— i 

130°E 


140°E 


150°E 


160°E 


Figure  1 

Distribution  of  the  Pacific  stock  of  chub  mackerel  (Scomber  japonicus) 
and  major  oceanographic  features  around  Japan.  The  hatched  areas  show 
spawning  grounds.  The  dotted  areas  show  feeding  grounds.  Major  purse-seine 
fishing  grounds  are  surrounded  by  dashed  lines.  The  fishing  grounds  around 
the  eastern  coast  of  Hokkaido  failed  in  1977  with  the  decline  in  biomass 
(Hirai,  1991). 


6,000    -, 

-  20 

uT 

CO 

o 
o 

i      i  Stock  number  at  age  0 

o 
o 
I    4,000    - 

£1 

Total  biomass 

—-  Catch 

k  number  a 

/ 

03 
CJ 
"O 

c 
ro 

nj 

' 

|-| 

CD 
CO 

-io  S 

to 
in 

CD 

E 

.2     2,000    - 
.o 

rn 

rn 

o 

5' 

Q. 

ro 
o 

H 

^ 

. 

• 

• 

\ 

•. 

-5      a 

v 

jO\           y 

♦J 

rv^^D 

nr^rt 

rn 

70                  75                  80                  85                  90                  95 

Year 

Figure  2 

Total  biomass,  total  catch,  and  year-class  strength  for  the  Pacific  stock  of  chub  mackerel 

{Scomber  japonicus)  (Yatsu,  et  al.1).  Year-class  strength  was  represented  by  the  stock  in 

number  at  age  0,  estimated  by  virtual  population  analysis  (Yatsu,  et  al.1). 

this  study  we  describe  the  variation  in  FL  at  age  of  the 
Pacific  stock  of  chub  mackerel  in  the  Kuroshio-Oyashio 
area,  using  data  from  1970  to  1997  when  the  stock  biomass 


fluctuated  between  0.2  and  5.9  million  metric  tons.  We  use 
these  data  to  evaluate  the  effects  of  population  density  and 
sea  surface  temperature  on  FL  at  age. 


198 


Fishery  Bulletin  102(1) 


Materials  and  methods 

Biological  data 

Biological  data  have  been  compiled  since  1964  for  purse- 
seine,  set-net,  dip-net  catches,  and  other  catches  by  national 
fisheries  research  institutes  and  local  government  fisher- 
ies experimental  stations  in  Japan.  Fork  length  (FL)  was 
measured  for  one  thousand  to  100  thousand  fish  per  year 
and  body  weight  (BW)  and  gonad  weight  were  measured  for 
10-100Tf  of  these  fish.  The  monthly  FL  compositions  and 
the  relationships  of  FL  to  BW  were  established  for  each 
year  with  this  data  set.  Year-specific  age-length  keys  from 
1970  to  1994  were  adopted  from  the  reports  of  cooperative 
research  on  Pacific  mackerel  by  local  government  agencies 
in  Chiba,  Kanagawa,  Shizuoka,  and  Tokyo.2  Between  1995 
and  1997,  age-length  keys  were  developed  by  national 
fisheries  research  institutes  and  local  government  fisher- 
ies experimental  stations. 

For  calculating  the  mean  FL  for  ages  0, 1,  2,  3,  4,  5.  and  6 
years  and  older,  we  used  data  from  the  purse-seine  fishery 
of  northeastern  Japan  during  September-December  for 
28  years,  from  1970  to  1997.  The  catch  of  this  fishery  in 
these  four  months  constituted  26-80%  ( the  28-year  mean  is 
about  639c )  of  the  total  annual  catch  of  the  Pacific  stock  of 
chub  mackerel.  Catch  in  number  at  FL  class  i  (cm)  of  each 
month  were  calculated  by 


n.,=C. 


(1) 


I* 


where  na  i  =  catch  in  number  at  FL  class  i  (cm)  (=  1,  .  .  .  , 
k,  .  .  50)  of  month  a  {=  Sep.,  Oct.,  Nov.,  Dec.]; 
da  t  =  frequency  at  FL  class  i  of  month  a; 
w a  (  =  a  mean  weight  of  each  FL  class  derived  from 
the  FL-BW  relationship;  and 
Ca  =  a  total  catch  of  month  a. 

We  then  summed  na  ,  of  4  months  to  derive  the  annual 
catch  in  number  at  FL  class  i: 


where  ntj  =  the  annual  catch  in  number  at  FL  class  i  at 
age  J;  and 
r:j  =  the  proportion  of  agej  at  FL  class  i  ( rt  0  +  r,  1  + 

••••  +  '-,*=  1»- 

From  nlt,  we  calculated  the  mean  and  variance  of  FL  at 
age./': 


I"    » 
l.,=^\ 


(4) 


and 


I",X,-// 


Va /•(/,)  =  - 


2Xi 


(5) 


where  L  =  mean  FL  at  age./';  and 

/,■  i  =  mean  FL  at  FL  class  i  at  age  j. 

Sea  surface  temperature 

Time-series  data  for  sea  surface  temperature  (SST  tem- 
peratures averaged  over  10  days  for  1°  latitude  x  1°  longi- 
tude blocks  over  the  northwestern  North  Pacific  between 
0-53°N  and  110-180°E  since  1950)  were  provided  by  the 
Oceanographical  Division  of  the  Japan  Meteorological 
Agency.  The  SST  data  for  each  block  was  averaged  for 
periods  of  three  months  (i.e.  January-March,  April^June, 
July-September,  and  October-December).  The  relationship 
between  the  SST  of  each  block  and  FL  at  age  0  were  exam- 
ined from  1970  to  1997. 

Autocorrelation 

For  correlation  analysis,  effective  sample  sizes  («')  were 
calculated  for  all  time  series  data  to  take  autocorrelation 
into  account,  n*  was  computed  by  the  formula  (Pyper  and 
Peterman,  1998): 


"i  =  X "°" 


(2) 


a   Sep 


where  /;,  =  the  annual  catch  in  number  at  FL  class  i. 

Using  the  age-length  key,  we  converted  nt  to  catch  at  FL 
class  i  at  age  j: 


(3) 


J_-i     1 


5/jaC/XrO"). 


(6) 


where  rx>Xj)  and  r^Xj)  are  the  autocorrelations  of  X  and  Y 
at  lagj,  defined  here  with  the  additional  weighting  factor 
proposed  by  Pyper  and  Peterman  ( 1998): 


^(Xt-XKXl+j-X) 


r„U)  = 


"-./ 


£«■,-*) 


(7) 


2  Age-length  keys.  In  Kanto  Kinkai  no  Masaba  ni  tuite,  Ap- 
pendix 1,  vol.  30,  30  p.  [In  Japanese.  Available  from  Kanagawa 
Prefectual  Fisheries  Research  Institute.  Jyogashima,  Misaki. 
Miura.  Kanagawa  238-0237.  .Iapan.1 


Growth  model 

We  used  the  modified  von  Bertalanffy  growth  model  to 
incorporate  the  effects  of  population  density  and  sea  sur- 


Watanabe  and  Yatsu:  Interannual  variation  in  length  at  age  of  Scomber  /aponicus 


199 


face  temperature  according  to  Millar 
and  Myers,3  who  nvestigated  three 
formulations  of  the  modified  von  Berta- 
lanffv  equations:  1)  a  reversible  effect 
on  the  growth  constant  k;  2)  a  revers- 
ible effect  on  the  asymptotic  length  Lr; 
and  3)  an  irreversible  effect  on  Lx  or  k. 
We  tested  two  of  the  models,  1  and  2, 
to  investigate  the  effect  of  population 
density  and  SST.  We  did  not  test  model 
3  because  we  did  not  consider  that  the 
environmental  effects  on  growth  were 
permanent.  Mean  length  at  age  i  of 
year-class  y  was  estimated  with  the  fol- 
lowing formulas: 

Model  1:  reversible  environmental 
effect  on  k 

L,v=L„(l-e"*°'-'°)  (8) 

4v  =  A-lv  + ( L~  ~  A-i.v  X 1  -  e~*  ■  )     ( 9 ) 
klv=k  +  PlT,fv+P2D:v.  110) 

Model  2:  reversible  environmental 
effect  on  L 


L0,,  =  L^v(l-e-*'») 


Llv  =  L,_ly+(L^-L,_lvn-e- 


Year 


Figure  3 

Interannual  fluctuations  in  mean  fork  length  (FLl  at  age  0,  age  1,  and  age  2  for 
chub  mackerel  iScomberjaponicus)  in  1970-97.  Horizontal  lines  show  the  28  year 
mean  FL  at  age  0,  1,  and  2,  respectively.  Vertical  bars  show  standard  deviations. 


(Ill 
(12) 
(13) 


We  ran  the  models  with  all  possible  combinations  of 
explanatory  variables  (T,  D,  T,  and  D),  and  compared  AIC 
with  that  obtained  with  the  base  parameters  (L,,  r0,  k). 


Results 


where  tr, 


XL,  V 


D. 


=  the  age  at  length  0  (year); 
=  the  asymptotic  length;  and 
=  the  growth  coefficient; 
=  L,  at  age  i  of  year-classy; 
=  k  at  age  ;'  of  year-class  y; 
=  the  sea  surface  temperature  in  year  i+y;  and 
=  a  population  density  presented  by  the  number 
of  stock  at  age  i  of  year-class  y. 


These  variables  were  z-score  standardized.  The  model 
parameters  ax  and  /32  were  estimated  to  represent  the 
effects  of  Tl+v  and  DI  v  on  k  or  Lv. 

The  parameters  were  estimated  by  maximizing  the  like- 
lihood function  which  is  represented  by 


and 


L(i,y)  =  L:  v  +£,, 
f,  -MO.cr), 


UL,,k,t0,pvP2,o'f)  = 

{L(/y)-L,v}2 


nnM'-p 


2a; 


(14) 
(15) 


(16) 


3  Millar,  R.  B.,  and  R.  A.  Myers.  1990.  Modeling  environmen- 
tally induced  change  in  growth  for  Atlantic  Canada  cod  stock. 
ICES  CM  1990/G:24. 


Fork  length  at  age 

Mean  FL  at  age  0  varied  substantially  over  the  time  series 
examined.  For  example,  it  ranged  from  16.9  (Sd  ±3.0)  cm 
in  1975  to  25.9  (Sd  ±1.0)  cm  in  1989.  The  mean  FL  for  the 
28  years  period  was  21.7  (±2.1)  cm  (coefficient  of  variation: 
CV=9.8%,  Table  1,  Fig.  3).  The  FL-at-age-0  values  were 
smaller  than  the  28-year  mean  FL  for  the  1970s,  varied 
around  the  mean  in  the  early  and  mid  1980s,  reached  a 
maximum  in  1989,  and  were  at  about  22-24  cm  in  the 
1990s  (Fig.  3). 

Mean  FL  at  age  1  was  similarly  variable;  it  ranged  from 
24.3  (±1.9)  cm  in  1976  to  31.6  (±1.4)  cm  in  1995.  The  28- 
year  mean  FL  was  27.7  (±1.6)  cm  (CV=5.6%,Table  1).  The 
trend  in  interannual  variability  was  similar  to  that  in  age 
0,  i.e.  it  was  smaller  in  the  1970s  and  larger  in  the  1990s 
(Fig.  3).  In  age-2  fish  the  28-year  minimum  FL  of  29.1  (±1.8) 
cm  was  observed  in  1986  and  the  maximum  of  34.5  (±1.3) 
cm  was  observed  in  1990  (the  28-year  mean  FL=31.1  (±1.5) 
cm,  CV=4.7%,  Table  1,  Fig.  3). 

In  fish  age  3  and  older,  mean  FL  varied  year-to-year  in  a 
manner  similar  to  that  found  in  the  younger  ages  ( Table  1 ). 
Annual  mean  FLs  for  3-,  4-,  and  5-year-olds  were  33.7 
(±1.3)  cm  (3.8%),  36.2  (CI  ±1.4)  cm  (CV=4.0%),  and  38.5 
(CI  ±1.5)  cm  (CV=3.8%),  respectively  (Table  1).  The  mean 
FLs  for  ages  0-5  of  each  year  were  significantly  different 
among  different  years  (one-way  ANOVA,  P<0.01 ). 


200 


Fishery  Bulletin  102(1 


Table  1 

Total  biomass,  year  class  strength  I  stock  number  at  age 
1970  to  1997.  Blanks  show  the  lack  of  data. 

0;Yats 

u,  et  al. 

),  SST,  and  mean  fork  length  (FL>  of  Scomber  japonicus  from 

Year 

Total 

Biomass 

(103t) 

Stock  number 

at  age  0 

(106  individuals  I 

SST 
(°C)' 

Mean  FL  (SD)  cm 

0 

1 

2 

3 

4 

5 

1970 

2833 

10,199 

11.5 

19.2 

12.6) 

26.3 

(1.8) 

30.5 

2.4) 

34.2 

(1.7) 

37.7 

(1.6) 

40.5 

1.4) 

1971 

3781 

14.138 

10.9 

20.2 

(2.3) 

26.8 

(1.9) 

31.4 

1.5) 

34.3 

(1.6) 

37.7 

(1.6) 

40.4 

1.3) 

1972 

4860 

8342 

13.2 

19.3 

il. 0) 

27.2 

(1.4) 

31.1 

1.6) 

34.3 

(1.5) 

37.3 

(1.7) 

40.0 

1.5) 

1973 

4650 

7154 

11.1 

22.2 

U.4) 

27.9 

(1.5) 

29.4 

1.6) 

31.2 

(1.8) 

33.1 

(2.0) 

36.1 

1.9) 

1974 

4048 

7854 

10.5 

19.7 

(1.4) 

27.7 

(2.5) 

30.4 

1.4) 

31.9 

(1.7) 

33.9 

(1.8) 

37.6 

1.7) 

1975 

3558 

10,353 

12.3 

16.9 

(3.0) 

25.4 

(1.8) 

30.3 

2.6) 

32.7 

(1.6) 

33.8 

(1.6) 

35.5 

1.7) 

1976 

3896 

14,402 

11.5 

19.7 

(2.0) 

24.3 

(1.9) 

29.4 

2.4) 

33.7 

(1.9) 

35.3 

(1.8) 

38.1 

1.8) 

1977 

5868 

11.701 

10.9 

21.4 

(1.3) 

26.2 

(1.8) 

30.1 

2.8) 

33.5 

(2.2) 

35.7 

(1.7) 

37.4 

1.4) 

1978 

5285 

6249 

10.0 

21.5 

(1.1) 

28.5 

(1.7) 

29.8 

1.6) 

32.1 

(2.3) 

34.5 

(2.D 

36.1 

1.9) 

1979 

3250 

2931 

12.3 

19.5 

(1.1) 

27.1 

(2.0) 

30.2 

2.0) 

33.0 

(1.7) 

35.2 

(1.6) 

37.2 

1.3) 

1980 

1898 

2952 

11.3 

20.7 

(1.1) 

25.8 

(2.6) 

30.3 

2.2) 

32.4 

(1.8) 

33.9 

(1.8) 

35.6 

1.6) 

1981 

1683 

3374 

9.4 

22.7 

(1.3) 

27.2 

(1.7) 

30.5 

1.5) 

33.1 

(2.1) 

36.5 

(1.8) 

38.0 

1.5) 

1982 

1567 

2883 

10.8 

22.5 

(1.8) 

27.9 

(1.6) 

29.3 

1.8) 

33.6 

(2.2) 

36.6 

(1.6) 

38.3 

1.4) 

1983 

1516 

3175 

11.5 

19.6 

(1.2) 

26.7 

(2.2) 

30.8 

1.6) 

33.6 

(1.5) 

35.5 

(2.0) 

37.8 

1.2) 

1984 

1759 

3605 

9.3 

22.7 

(1.3) 

27.0 

(2.4) 

31.0 

1.8) 

34.8 

(1.9) 

36.6 

(1.8) 

38.2 

2.0) 

1985 

1565 

4998 

11.4 

20.1 

(2.2) 

27.3 

(2.11 

30.9 

1.9) 

33.3 

(1.9) 

37.4 

(1.7) 

39.0 

1.8) 

1986 

1373 

1833 

9.7 

21.5 

(1.7) 

26.4 

(1.4) 

29.1 

1.8) 

32.5 

(2.4) 

35.9 

(2.1) 

38.9 

1.9) 

1987 

812 

583 

10.9 

20.5 

(2.1) 

27.6 

(1.7) 

30.2 

1.3) 

32.8 

(1.6) 

36.4 

(2.3) 

39.2 

0.8) 

1988 

555 

236 

11.4 

24.9 

(1.4) 

28.1 

(1.5) 

30.5 

1.4) 

32.8 

(1.7) 

36.8 

(1.6) 

40.1 

1.2) 

1989 

289 

219 

9.8 

25.9 

(1.0) 

29.7 

(2.3) 

32.2 

1.4) 

34.6 

(1.5) 

35.7 

(1.5) 

39.2 

1.5) 

1990 

185 

356 

11.7 

24.4 

(1.3) 

30.3 

(2.6) 

34.5 

1.3) 

35.8 

(1.5) 

38.2 

(1.1) 

39.7 

0.8) 

1991 

338 

1017 

12.2 

24.1 

(1.6) 

28.9 

(1.8) 

33.5 

1.9) 

35.5 

(1.2) 

36.7 

(1.9) 

39.0 

1.8) 

1992 

724 

2839 

9.7 

24.0 

(1.6) 

29.0 

(1.7) 

32.1 

1.4) 

34.1 

(1.5) 

37.5 

(1.6) 

40.5 

1.6) 

1993 

685 

589 

10.7 

23.9 

(0.9) 

29.3 

(1.3) 

31.7 

l.D 

33.2 

(0.5) 

1994 

343 

547 

11.3 

23.7 

(1.7) 

28.8 

(2.5) 

32.8 

1.0) 

34.6 

(0.8) 

35.9 

(0.7) 

39.1 

1.0) 

1995 

351 

1183 

11.3 

22.0 

(1.3) 

31.6 

(1.4) 

32.9 

1.8) 

35.5 

(1.8) 

38.0 

(1.3) 

39.2 

0.8) 

1996 

726 

4452 

9.9 

22.5 

(1.1) 

28.7 

(2.5) 

34.1 

1.2) 

36.1 

(1.1) 

37.8 

(0.9) 

39.7 

0.7) 

1997 

682 

529 

9.9 

23.6 

(1.4) 

29.0 

(1.5) 

33.0 

1.3) 

35.4 

(1.7) 

37.6 

(0.7) 

38.6 

0.5) 

28-year 

mean  of  FLs  at  ages 

21.7 

(2.1) 

27.7 

(1.6) 

31.1 

1.5) 

33.7 

(1.3) 

36.2 

(1.4) 

38.5 

1.5) 

'  SST  during  ApriUJ 

une  in  the  waters  bounded  by  38^tO°N  ar 

d  141- 

143°E. 

Mean  growth  increments  I  of  each  year  class  from  age  0 
(6  months  old)  to  ages  1-5  (/0_,)  showed  significantly  nega- 
tive correlations  (Table  2).  Correlations  between  the  two 
variables  tended  to  increase  with  age:  -0.69  for  /,, _,,  -0.71 
for  I„_.,,  -0.80  for  /„_.,,  and  -0.77  for  I0^. 

The  relative  FL  at  age  0  for  any  given  year  class  was 
maintained  throughout  the  life  span.  A  correlation  be- 
tween the  mean  FL  at  age  0  and  age  1  within  each  year 
class  (1970  to  1996  year  class)  was  positive  and  statisti- 
cally significant  (P<0.05,  Fig.  4).  Similarly,  the  positive  cor- 
relations between  the  mean  FL  at  age  0  and  age  3  ( 1970 
to  1994  year  class,  P<0.01,  Fig.  4),  and  age  0  and  age  4 
(1970  to  1993  year  class.  P<0.05,  Fig.  4)  were  significant 
(P<0.05.  Fig.  4). 


Correlation  between  FL  and  population  density 

Population  densities  represented  by  stock  in  number  at  age 
0  and  total  biomass  were  negatively  correlated  to  FL  at  age. 
Negative  correlations  between  the  logarithm  of  abundance 
of  age  0  (ln/V0)  and  FL  at  ages  were  relatively  high  in  age  0 
to  3  (-0.69  to  -0.83,  Table  3)  and  low  in  age  4  and  5  (-0.63 
and  -0.64,  Table  3).  Correlations  were  statistically  signifi- 
cant for  ages  0,  2,  and  3  (Table  3).  Negative  correlations 
between  the  logarithm  of  total  biomass  and  FL  at  ages 
were  relatively  high  at  ages  0  to  2  (-0.73  to  -0.75)  and 
moderate  for  age  3  to  5  (-0.50  to  -0.52,  Table  4).  However, 
the  relationships  were  not  statistically  significant  for  all 
ages  (Table  4). 


Watanabe  and  Yatsu:  Interannual  variation  in  length  at  age  of  Scomber  /aponicus 


201 


32 


E 
3   29 


26  -- 


23 


15 


H h 


20  25 

FL  at  age  0  (cm) 


38 


36 


S.    34 

IB 


32  -■ 


30 


B 


15 


H — H 


H — i- 


20  25 

FL  at  age  0  (cm) 


20  25 

FL  at  age  0  (cm) 


Figure  4 

Scatter  plots  of  FL  at  age  1  (A),  age  3  (Bl  and  age  4  (C)  on  FL  at  age  0  for  chub  mackerel  (Scomber japonicus).  Correlations  between 
FL  at  age  1  with  age  0  (r=0.83.  n=28,  actual  sample  size  n'=8,  df=6),  age  3  with  age  0  (r=0.62,  ;i=26.  n  =11.  df  =9)  and  age  4  with 
age  0  (r=0.67,  u=24,  n'=10)  were  all  significant  at  P  <  0.05. 


Table  2 

Correlation  of  FL  at  age  0  and  growth  increment  after  age 
0.  n  =  actual  sample  size,  n*  and  degree  of  freedom  (df) 
show  the  effective  ;?  and  df  when  the  data  were  corrected 
for  autocorrelation  (Pyper  and  Peterman,  1998).  Signifi- 
cance level:  **,  P<0.01. 


Growth  increment 


Ages  0-1 
Ages  0-2 
Ages  0-3 
Ages  0-4 
Ages  0-5 


df 


0.69** 

0.48 

27 

21 

19 

0.71** 

0.51 

26 

25 

23 

0.80** 

0.64 

25 

23 

21 

0.77** 

0.59 

23 

24 

22 

0.78** 

0.61 

22 

22 

20 

Table  3 

Correlation  between  the  natural  logarithm  of  the  abun- 
dance of  age  0  and  mean  FL  for  each  age.  n  =  actual  sample 
number,  n*  and  degree  of  freedom  (df)  show  the  significant 
n  and  df  when  autocorrelation  was  considered  (Pyper  and 
Peterman,  1998).  Significance  levels:  *,  P  <  0.05. 

Age 

r 

r2 

n 

n              df 

0 

-0.75* 

0.57 

28 

8               6 

1 

-0.69 

0.48 

27 

7               5 

2 

-0.83* 

0.69 

26 

6               4 

3 

-0.71* 

0.51 

25 

9               7 

4 

-0.63 

0.40 

23 

8               6 

5 

-0.64 

0.40 

22 

6               4 

Correlation  between  FL  and  SST 

Growth  in  the  first  six  months  of  life  was  correlated  with 
SST.  We  detected  significant  negative  correlation  between 
FL-at-age  0  and  SST  between  April  and  June  in  the  waters 
bounded  by  38-40°N  and  141-143°E  U--0.45,  r2=0.20, 
n=28,  n  =27,  df=25,  P<0.05,  Fig.  5).  The  SST  between  July 
and  September  of  this  area  was  also  negatively  correlated 
with  FL  at  age  0  although  the  correlation  coefficient  was 
not  significant  at  5%  level. 

Growth  analysis 

Model  1  that  incorporated  SST  ( T)  and  population  density 
(D)  gave  a  minimum  Akaike's  information  criterion  (AIC) 
of  457.68  (Table  5)  and  the  model  was  expressed  by 


L-    =  43.98  1  -  exp(  -2.585  )exp 


-5X 


:0. 271-0. 008T     -0.2LD, 


(17) 


(18) 


Table  4 

Correlation  between  natural  logarithm  of  total  biomass 
and  mean  FL  for  each  age.  n  =  actual  sample  size,  n*  and 
dgree  of  freedom  ( df )  show  the  effective  n  and  df  when  the 
data  were  corrected  for  autocorrelation  (Pyper  and  Peter- 
man, 1998).  No  correlations  were  significant  (P>0.05). 


Age 


df 


0.74 

0.38 

27 

6 

4 

0.73 

0.32 

27 

6 

4 

0.75 

0.36 

27 

5 

3 

0.52 

0.26 

27 

11 

9 

0.51 

0.26 

26 

9 

7 

0.50 

0.22 

26 

7 

5 

This  model  estimated  the  FL  at  ages  0-5  well  (Fig.  6). 
The  AIC  of  model  1  incorporating  T  and  D  was  smaller  than 
the  AIC  of  model  2;  therefore  the  environmental  factors  had 
an  affect  on  k  rather  than  LT. 


202 


Fishery  Bulletin  102(1) 


45  N 


40  N 


35  N 


30  N 


140N 


145  N     150 


B 


25  -- 


8>  20  + 

CD 


r  =o.20 


15  -I — l — l — l — i — I — i — i — i — i — I — i — i — i — i — I 
8        10       12       14 
Mean  SST 


Figure  5 

(A)  Map  to  show  correlation  between  sea  surface  temperatures  (SST)  and  mean  fork  length  (FLl 
at  age  0  for  chub  mackerel  (Scomber  japonicus).  The  dotted  area  indicates  the  negative  correla- 
tion coefficient  r  above  0.4.  The  contour  interval  is  0.1  of  the  correlation  coefficient  and  positive 
contours  are  shown  as  dashes.  (B)  Relationship  between  mean  SST  for  the  area  38°-40°N  and 
141-143°E  from  April  to  June  and  mean  FL  at  age  0.  Correlation  was  significant  at  the  5^  level 
(r=-0.45,  n=28,  n  "=27,  df=25). 


To  investigate  the  effect  of  T  and  D,  we  calculated  the 
total  effect  on  k  for  year-class  v  according  to  Sinclair  et  al. 
(2002): 


lA^ 


I  ft  A, 


for  T,  and 


for  D. 


Discussion 

Estimated  population  abundance  of  age-0  fish  and  total 
biomass  may  explain  density-dependent  growth.  FL  at 
age  0,  2,  and  3  of  the  Pacific  stock  of  chub  mackerel  were 
negatively  correlated  with  the  number  of  age-0  recruits. 
Correlations  between  biomass  and  FL  at  ages  0-5  were  low 
and  not  significant.  Therefore,  year-class  strength  is  indi- 
cated to  have  a  greater  negative  influence  on  the  growth 
of  the  Pacific  stock  of  chub  mackerel  than  total  biomass, 
as  reported  for  the  Atlantic  mackerel  (Scomber  scombrus) 
(Agnalt,  1989;  Overholtz,  1989;  Neja,  1995)  and  Atlantic 
herring  (.CI upea  harengus)  (Toresen,  1990). 

Density-dependent  growth  in  fish  populations  seems  to 
be  a  common  phenomenon  for  pelagic  fishes  found  in  the 
temperate  waters  of  Japan.  The  FL  at  age  0  of  the  1963-69 
year  classes  ranged  from  16  to  20  cm,  and  were  smaller  than 
those  of  the  1970s,  possibly  indicating  density-dependent 
growth  ( Iizuka,  1974 ).  According  to  Honma  et  al.  ( 1987 ),  the 
stock  abundance  of  the  Pacific  stock  of  chub  mackerel  from 
1963  to  1969  was  larger  than  it  was  in  the  1970s.  Wada  et 
al.  (1995)  and  Hiyama  et  al.  ( 1995)  found  negative  relation- 
ships between  total  biomass  and  body  length  in  the  Pa- 


Table  5 

Summary  of  statistics  from  the  estimation  of  growth  for 

chub  mackerel  (Scomber  japonicus).  AIC 

=  Akaikf 

's  infor- 

mation  criterion. 

No.  of 

Log 

unknown 

likeli- 

Model          Variables 

parameters 

hood 

AIC 

1           L„,  k,  tn,  (jj  .  .  .a5 

9 

-280.20 

578.40 

L,.k,t0,a1  ...a5, /3, 

10 

-270.10 

560.20 

Lr,  k,  t0,  CTj  .  .  .a5,  p2 

10 

-222.38 

464.77 

L„,  k,  t0,  a,  . .  .cts,/S] 

ft    11 

-217.84 

457.68 

2          La,k,t0,a1...as 

9 

-280.20 

578.40 

L,,k.t„.ax  .  .  -cr5,  jSj 

10 

-268.63 

557.25 

Lr,  k,  t0,  at . .  .ii-,  />., 

10 

-224.01 

468.02 

Lx,  k,  tQ,   (Jj    .    .    .Or,,  fix 

ft    11 

-220.81 

463.62 

cific  and  Tsushima  Current  stock  of  the  Japanese  sardine 
(Sarclinops  melanostictus).  Kishida  (1990)  demonstrated  a 
density-dependent  relationship  between  the  growth  and 
total  stock  density  (CPUE)  of  Japanese  Spanish  mackerel 
(Scomberomortis  nipkonius). 

Our  results  do  not  agree  with  the  positive  effect  of  sea  wa- 
ter temperature  on  somatic  growth  that  has  been  shown  for 
several  species,  including  Japanese  common  squid  (Kidokoro, 
2001).  Atlantic  herring  ( Moores  and  Winters,  1981;  Toresen, 
1990).  and  Atlantic  cod  (Gadus  morhua )  (Brander,  1995;  Du- 
til  et  al,  1999;  Ratz  et  al.  1999;  Otterson  et  al.,  2002). 


Watanabe  and  Yatsu:  Interannual  variation  in  length  at  age  of  Scomber  /aponicus 


203 


15  I i mini 20  I i 

70   75   80   85   90   95        70   75   80   85   90   95 


35 


30 


Age  2 


35 


30  -- 


Age  3 


25  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I  II  I  I  I  I  I  I  I  I  I    25  I  I  I  I  I  I  I  I  I  I  I  I  II  I  I  I  I  I  I  I  I  I  I  I  I  I  I  I 
70   75   80   85   90   95        70   75   80   85   90   95 


40 


35 


Age  4 


..  Age  5 


40  -- 


35 


30  I  I  I  II  I  I  II  I  I  I  I  I  I  II  II  I  I  I  I  I  I  II  I  I   30  I  I  I  I  I  I  II  I  I  I  I  I  I  III  I  I  I  I  I  I  I  I  I  I  I  I 
70   75   80   85   90   95       70   75   80   85   90   95 

Year 

Figure  6 

Time  series  of  observed  (open  circles)  and  modeled  (solid  line)  values  of  mean  fork 
length  (FL)  at  ages  0-5  during  1970-97  for  chub  mackerel  iScomber  japonieus). 


There  was  a  positive  correlation  between  FL  at  age  0  and 
l°xl°  block  SST  in  the  waters  of  32-34°N  and  144-149°E, 
located  south  of  the  Kuroshio  Extension  flowing  eastward 
at  the  latitude  of  35-37°N  from  April  to  June  (Figs.  1  and 
5A).  But  the  correlation  coefficient  was  not  significant,  and 
this  area  was  not  considered  to  be  inhabited  by  juvenile 
mackerel  (Watanabe,  1970).  Thus,  we  considered  that  the 
SST  in  the  waters  of  32-34°N  and  144-149°E  was  not  a 
significant  factor  on  the  variation  of  FL  at  age  0. 

The  low  SST  in  the  waters  bounded  by  38-40°N  and 
141-143°E  is  indicative  of  a  large  inflow  of  Oyashio  Cur- 
rent waters  (Hirai  and  Yasuda,  1988),  which  is  a  cold 
water  current  and  has  high  productivity  (Odate,  1994), 
into  the  Kuroshio-Oyashio  transition  zone,  where  is  one 
of  the  main  feeding  grounds  of  mackerels  (Odate,  1961; 
Watanabe,  1970;  Watanabe  and  Nishida,  2002;  Fig.  1). 
Thus,  we  hypothesized  that  the  large  inflow  of  Oyashio 
current  waters  into  the  Kuroshio-Oyashio  transition  zone 
improved  the  feeding  condition  and  accelerated  the  growth 
of  juvenile  mackerel.  Jobling  ( 1988)  suggested  a  parabolic 
relationship  between  water  temperature  and  fish  growth. 
The  range  of  SST  in  this  area,  which  was  negatively  cor- 


related with  FL  at  age  0  of  mackerel,  was  9-13°C  (Table  1 ). 
This  temperature  range  is  near  the  lowest  nonstressful 
temperatures  for  mackerel  ( 10-12°C,  Schaefer,  1986).  Thus, 
we  do  not  consider  that  the  negative  relationship  between 
growth  and  SST  was  the  result  of  suppressed  growth  by 
the  high  ambient  temperature. 

In  mackerel,  maximum  egg  production  appears  to  have 
shifted  to  later  in  spring  during  the  1990s,  as  compared  to 
the  late  1970s  and  1980s,  resulting  in  a  shorter  period  of 
growth  and  thus  smaller  fish  (Fig.  8,  Mori  et  al.4;  Kikuchi 
and  Konishi5;  Ishida  and  Kikuchi6;  Zenitani  et  al.7;  Kubota 
et  al.8).  In  the  early  1970s,  the  main  spawning  period  was 


4  Mori,  K.,  K.  Kuroda,  and  Y.  Konishi.  1988.  Monthly  egg 
production  of  the  Japanese  sardine,  anchovy,  and  mackerels  off 
the  southern  coast  of  Japan  by  egg  censuses.  Datum  Collect. 
Tokai  Reg.  Fish.  Res.  Lab.  12:1-321.  [In  Japanese.  Available 
from  National  Research  Institute  of  Fisheries  Science,  2-12-4 
Fukuura,  Kanazawa,  Yokohama  236-8648,  Japan.] 

5  See  next  page. 

6  See  next  page. 

7  See  next  page. 

8  See  next  page. 


204 


Fishery  Bulletin  102(1) 


0.015  T 


-0.015 


0.015 


i  i  i i i i i i i i i  ii  n  i  i 


70 


75 


80 


85 


90 


95 


-0.015    I  I  I  I  I  I  I  I  I  I  I  I  I  I 
70         75         80 
Year 


Figure  7 

The  total  effect  of  I  A)  mean  SST  for  the  area  of  38-40°N  and  141-143"E  from  April  to 
June,  and  (Bl  population  density  on  k  for  each  year  class  of  chub  mackerel  i Scomber 
japonicus). 


also  in  April  (Kuroda9).  Delayed  spawning  in  the  1990s 
should  have  resulted  in  a  reduction  in  the  mean  FL  at  ages 
during  September-December  in  the  1990s  compared  to  the 
1970s  and  1980s;  however  the  present  study  showed  the  op- 
posite result  (Table  1 ).  We  hypothesize  that  the  effect  of  the 
shift  of  spawning  period  on  the  FL  at  ages  may  have  been 
overwhelmed  by  the  effect  of  population  density  (Fig.7). 


■"'  Kikuchi,  H.,andY.  Konishi.  1990.  Monthly  egg  production  of 
the  Japanese  sardine,  anchovy,  and  mackerels  off  the  southern 
coast  of  Japan  by  egg  censuses:  January,  1987  through  December, 
1988,  72  p.  National  Research  Institute  of  Fisheries  Science. 
Tokyo.  [In  Japanese.  Available  from  National  Research  Insti- 
tute of  Fisheries  Science,  2-12-4  Fukuura,  Kanazawa.  Yokohama 
236-8648,  Japan.] 

6  Ishida,  M.  and  H  Kikuchi.  1992.  Monthly  egg  production  of 
the  Japanese  sardine,  anchovy,  and  mackerels  off  the  southern 
coast  of  Japan  by  egg  censuses:  January,  1989  through  December, 
1990,  86  p.  National  Research  Institute  of  Fisheries  Science, 
Tokyo.  [In  Japanese.  Available  from  National  Research  Insti- 
tute of  Fisheries  Science,  2-12-4  Fukuura,  Kanazawa,  Yokohama 
236-8648,  Japan.]. 

7  Zenitani,  H.,  M.  Ishida,  Y  Konishi,  T  Goto,  Y.  Watanabe,  and  R. 
Kimura.  1995.  Distributions  of  eggs  and  larvae  of  Japanese 
sardine,  Japanese  anchovy,  mackerels,  round  herring,  jack  mack- 
erel and  Japanese  common  squid  in  the  waters  around  Japan. 
1991  through  1993.  Resources  Management  Research  Report 
Series  A-2, 368  p.  National  Research  Institute.  Japan  Fisheries 
Agency,  Tokyo.  [In  Japanese.  Available  from  National  Research 
Institute  of  Fisheries  Science,  2-12-4  Fukuura,  Kanazawa,  Yoko- 
hama, 236-8648  Japan] 

s  Kubota,  H.,Y  Oozeki,  M.  Ishida,  Y  Konishi,  T  Goto,  H.  Zenitani, 
and  R.  Kimura.  1999.  Distributions  of  eggs  and  larvae  of 
Japanese  sardine,  Japanese  anchovy,  mackerels,  round  her- 
ring, jack  mackerel  and  Japanese  common  squid  in  the  waters 
around  Japan,  1994  through  1996,  352  p.  Resources  Manage- 
ment Research  Report  Series  A-2.,  National  Research  Institute, 
Japan  Fisheries  Agency,  Tokyo.  [In  Japanese.  Available  from 
National  Research  Institute  of  Fisheries  Science,  2-12-4  Fuku- 
ura, Kanazawa,  Yokohama  236-8648,  Japan.] 

9  Kuroda,  K.  2002.  Personal  commun.  1-1-3-406.  Kasumi. 
Narashino.  Chiba  275-0022,  Japan. 


Jun  6 


May  5  -  - 


Apr  4 


Mar  3 


H — l — I — I — I — I— 
78       80       82       84 


86      88 
Year 


—l — l — l — l — l — i — i 
90       92       94       96 


Figure  8 

Interannual  variation  in  the  peak  period  (weighted 
monthly  means)  of  egg  production  for  the  Pacific  stock 
of  chub  mackerel  (Scomber  japonicus),  which  includes  a 
small  portion  from  the  eggs  of  spotted  mackerel  (Scomber 
australasicus)  (Mori  et  al.4:  Kikuchi  and  Konishi'';  Ishida 
and  Kikuchi1';  Zenitani  et  al.7;  Kubota  et  al.8). 


The  estimated  FL  at  age  from  our  growth  model,  with 
the  use  of  AIC,  fitted  well  to  the  observed  FL  at  age 
(Fig.  6).  Mean  growth  increments  /  of  each  year  class 
from  age  0  (6  months  old)  to  ages  1-5  (/„_,)  were  signifi- 
cant and  negatively  correlated  with  FL  at  age  0  (Table  2), 
indicating  that  the  growth  rate  of  mackerel  had  changed 
from  year  to  year  for  a  given  year  class.  This  negative 
correlation  indicated  that  the  effects  of  population  density 
and  SST  was  temporal,  and  influenced  k  rather  than  L r. 
The  negative  correlation  between  FL  at  age  0  and  growth 
increments  also  suggested  that  the  FL  at  age  of  mack- 
erel approximated  the  asymptotic  length.  Thus,  mackerel 
growth  was  best  fitted  to  the  modified  von  Bertalanffy 
growth  model  with  the  temporal  environmental  effect  on 
k  (Table  5). 


Watanabe  and  Yatsu:  Interannual  variation  in  length  at  age  of  Scomber  /aponicus 


205 


The  effect  of  population  density  on  growth  of  mackerel 
was  higher  than  the  effect  of  SST  (Fig.  7,  Table  6).  Our 
result  agreed  with  the  results  for  Japanese  sardine  ( Wada 
et  al.,  1995 )  and  Atlantic  cod  ( Sinclair  et  al.,  2002 ).  Particu- 
larly, the  effect  of  population  density  was  significant  in  the 
late  1980s,  which  resulted  in  a  remarkable  increase  in  FL 
at  age  0  (Figs.  3  and  7). 

The  relative  size  at  age  0  was  carried  over  to  older  ages 
(Fig.  4),  indicating  that  the  cohorts  that  were  small  at  age 
0  could  not  compensate  for  this  early  small  size.  Iizuka 
(1974)  reported  that  the  trend  of  growth  established  at 
age  0  for  chub  mackerel  was  maintained  until  age  2  for 
the  1963-73  year  classes.  Toresen  (1990)  demonstrated 
from  length  data  that  a  trend  in  rate  of  growth  for  a  given 
year  class  of  Norwegian  herring  was  determined  at  the  im- 
mature stage  and  was  consistent  after  maturation.  Total 
length  of  Hokkaido-Sakhalin  herrings  iClupea  pallasii)  at 
age  5  and  older  was  positively  correlated  with  the  length 
at  age  4  (Watanabe  et  al.,  2002).  Because  fish  first  mature 
at  age  4.  this  implied  that  the  trend  in  total  length  of  each 
year  class  was  determined  by  the  age  at  maturity.  From 
these  results  we  hypothesize  that  the  variability  in  size  at 
age  in  the  Pacific  stock  of  chub  mackerel  is  largely  attribut- 
able to  growth  before  maturity,  especially  during  the  first 
6  months  after  hatching. 


Acknowledgments 

We  would  like  to  thank  K.  Meguro  of  Chiba  Prefecture 
Governmental  Office  and  K.  Kobayashi  of  Shizuoka  Pre- 
fecture Governmental  Office  for  providing  insights  into 
chub  mackerel's  growth  and  into  age  determination.  We 
also  thank  T.  Akamine,  M.  Suda,  and  N.  Yamashita  of  the 
National  Research  Institute  of  Fisheries  Science  for  advice 
on  the  statistical  analysis.  We  also  thank  Y.  Watanabe  and 
C.  B.  Clarke  of  the  Ocean  Research  Institute,  University  of 
Tokyo,  for  their  constructive  comments  on  this  manuscript. 


Literature  cited 

Agnalt,  A.  -L. 

1989.     Long-term  changes  in  growth  and  age  at  maturity  of 
mackerel.  Scomber  scombrus  L..  from  the  North  Sea.     J. 
Fish  Biol.  35  (suppl.  A):305-311. 
Brander.  K.  M. 

1995.    The  effect  of  temperature  on  growth  of  Atlantic  cod 
iGadus  morhua  L.).     ICES  J.  Mar.  Sci.  52:1-10. 
Dutil.  J.  -D.,  M.  Castonguay,  D.  Gilbert,  and  D.  Gascon. 

1999.  Growth,  condition,  and  environmental  relationships 
in  Atlantic  cod  iGadus  morhua)  the  northern  Gulf  of  St. 
Lawrence  and  implications  for  management  strategies 
in  the  Northwest  Atlantic.  Can.  J.  Fish.  Aquat.  Sci.  56: 
1818-1831. 
Haddon,  M. 

2001.     Modelling  and  quantitative  methods  in  fisheries, 
406  p.     Chapman&Hall/CRC,  New  York,  NY. 
Hirai,  M. 

1991.     Fisheries  oceanographic  study  on  purse-seine  fish- 
ing-grounds for  chub  mackerel  in  the  Sanriku  coastal 


waters.     Bull.  Tohoku  Natl.  Fish.  Res.  Inst.  53:59-147.     [In 

Japanese.] 
Hirai,  M.,  and  I.  Yasuda. 

1988.     Interannual  variability  of  the  temperature  field  at  100 

m  depth  near  the  east  coast  of  Japan.     Bull.  Otsuchi  Ocean 

Res.  Center.  Univ.  Tokyo.  14:184-186. 
Hiyama,  Y  ,  H.  Nishida.  and  T  Goto. 

1995.     Interannual  fluctuations  in  recruitment  and  growth  of 

the  sardine,  Sardinops  melanostictus,  in  the  Sea  of  Japan 

and  adjacent  waters.     Res.  Popul.  Ecol.  37(21:177-183. 
Honma,  M.,  Y  Sato,  and  S.  Usami. 

1987.  Estimation  of  the  population  size  of  the  Pacific  mack- 
erel by  the  cohort  analysis.  Bull.  Tokai  Reg.  Fish.  Res.  Lab. 
121:1-11.     [In  Japanese.) 

Iizuka,  K. 

1974.    The  ecology  of  young  mackerel  in  the  north-eastern 

sea  of  Japan  FY  Estimation  of  the  population  size  of  the 

0-age  group  and  the  tendencies  of  growth  patterns  on  0.  I. 

and  II  age  groups.     Bull.  Tohoku  Reg.  Fish.  Res.  Lab.  34: 

1-16.     [In  Japanese.] 
Jobling,  M. 

1988.  A  review  of  the  physiological  and  nutritional  energet- 
ics of  Cod,  Gadus  morpha  L.,  with  particular  reference  to 
growth  under  farmed  conditions.    Aquaeulture,  70:1-19. 

Kawasaki,  T. 

1966.     Structure  of  the  Pacific  population  of  the  mackerel. 
Bull.  Tokai  Reg.  Fish.  Res.  Lab.  47:1-34.     [In  Japanese.] 
Kidokoro,  H. 

2001.  Fluctuations  in  body  size  and  abundance  of  Japanese 
common  squid  {Todarodes  pacificus)  in  the  Sea  of  Japan. 
GLOBEC  report  15:42. 

Kishida.  T 

1990.  Relationship  between  growth  and  population  density 
of  Japanese  Spanish  mackerel  in  the  central  and  western 
waters  of  the  Seto  Inland  Sea.  Bull.  Nansei  Natl.  Fish.  Res. 
Inst.  23:35-41.     |In  Japanese.] 

Mace,  P.  M.,  and  M.  O.  Sissenwine. 

1993.  How  much  spawning  per  recruit  is  enough?  Can. 
Spec.  Publ.  Fish.  Aquat.  Sci.  120:101-118 

Moores,  J.  A.,  and  G.  H.  Winters. 

1981.     Growth  patterns  in  a  Newfoundland  Atlantic  herring 

iClupea  harengus  harengus)  stock.     Can.  J.  Fish.  Aquat. 

Sci.  39:454-461. 
Moyle,  P.  B.,  and  J.  J.  Cech  Jr. 

2002.  Fishes:  An  introduction  to  ichthyology,  4th  ed.,  612  p. 
Prentice  Hall,  Englewood  Cliffs.  NJ. 

Murayama.T.,  I.  Mitani,  and  I.  Aoki. 

1995.  Estimation  of  the  spawning  period  of  the  Pacific  mack- 
erel Scomber  japonicus  based  on  the  changes  in  gonad  index 
and  the  ovarian  histology.  Bull.  Jpn.  Soc.  Fish.  Oceanogr. 
59:11-17.     [In  Japanese.] 

Neja,  Z. 

1995.  The  stock  size  and  changes  in  the  growth  rate  of  the 
northwest  Atlantic  mackerel  [Scomber  scombrus  L.)  in 
1971-1983.    Acta  Ichthyologica  et  Piscatoria  25:113-121. 

Odate,  K. 

1994.  Zooplankton  biomass  and  its  long-term  variation  in 
the  western  north  Pacific  ocean,  Tohoku  sea  area,  Japan. 
Bull.  Tohoku  Natl.  Fish.  Res.  Inst.  56, 115-173.     [In  Japanese.] 

Odate,  S. 

1961.     Study  on  the  larvae  of  the  fishes  in  the  north-eastern 
sea  area  along  the  Pacific  coast  of  Japan.  Part  1.  Mackerel. 
Pneumatophorus japonicus  ( Houttuyn  I.     Bull.  Tohoku  Reg. 
Fish.  Res.  Lab.  19:98-108.     [In  Japanese.] 
Otterson,  G,  K.  Helle,  and  B.  Bogstad. 

2002.     Do  abiotic  mechanisms  determine  interannual  vari- 


206 


Fishery  Bulletin  102(1) 


ability  in  length-at-age  of  juvenile  Arcto-Norwegian  cod? 
Can.  J.  Fish.  Aquat.  Sci.  59:57-65. 
Overholtz,  W.  J. 

1989.  Density-dependent  growth  in  the  northwest  Atlan- 
tic stock  of  Atlantic  mackerel  (Scomber  scombrus).  J. 
Northw.  Atl.  Fish.  Sci.  9:115-121. 

Pauly,  D. 

1987.    Application  of  information  on  age  and  growth  of  fish 

to  fishery  management.    In  Age  and  growth  of  fish  (R.  C. 

Summerfelt  and  G.  E.  Hall,  eds.)  p.495-506.    Iowa  State 

Univ  Press,  Ames,  IA. 
Pyper,  B.  J.,  and  R.  M.  Peterman. 

1998.  Comparison  of  methods  to  account  for  autocorrelation 
in  correlation  analyses  of  fish  data.  Can.  J.  Fish.  Aquat. 
Sci.  55:2127-2140. 

Ratz,  H.  -J.,  M.  Stein,  and  J.  Lloret. 

1999.  Variation  in  growth  and  recruitment  of  Atlantic  cod 
(Gadus  morhua  I  off  Greenland  during  the  second  half  of  the 
twentieth  century.     J.  Northw.  Atl.  Fish.  Sci.  25:161-170. 

Schaefer,  K.  M. 

1986.     Lethal  temperature  and  the  effect  of  temperature 
change  on  volitional  swimming  speeds  of  chub  mackerel. 
Scomber  japonicus.     Copeia  1986:37-44. 
Sinclair,  A.  F..  D.  P.  Swain,  and  J.  M.  Hanson. 

2002.     Disentangling  the  effects  of  size-selective  mortality, 
density,  and  temperature  on  length-at-age.     Can.  J.  Fish. 
Aquat.  Sci.  59:372:382. 
Toresen,  R. 

1990.  Long-term  changes  in  growth  of  Norwegian  spring- 
spawning  herring.    J.  Cons.  Int.  Explor.  Mer  47:48-56. 

Usami,  S. 

1973.    Ecological  studies  of  life  pattern  of  the  Japanese 


mackerel,  Scomber  japonicus  Houttuyn:  on  the  adult  of  the 
Pacific  subpopulation.     Bull.  Tokai  Reg.  Fish.  Res.  Lab.  76: 
71-178.     |In  Japanese.] 
Wada,  T.,  Y.  Matsubara,  Y.  Matsumiya,  and  N.  Koizumi. 

1995.     Influence  of  environment  on  stock  fluctuations  of 
Japanese  sardine,  Sardinops  melanostictus.     Can.  Spec. 
Publ.  Fish.  Aquat.  Sci.  121:387-394. 
Watanabe,  C,  and  H.  Nishida. 

2002.     Development  of  assessment  techniques  for  pelagic 
fish  stocks:  applications  of  daily  egg  production  method  and 
pelagic  trawl  in  the  northwestern  Pacific  ocean.     Fisheries 
Science.  68:97-100. 
Watanabe,  C,  T.  Hanai,  K.  Meguro,  R.  Ogino,  Y  Kubota,  and 
R.  Kimura. 

1999.     Spawning  biomass  estimates  of  chub  mackerel 
Scomber  japonicus  of  Pacific  subpopulation  off  central 
Japan  by  a  daily  egg  production  method.     Nippon  Suisan 
Gakkaishi  651 4 1:695-702.     [In  Japanese.] 
Watanabe,  T. 

1970.     Morphology  and  ecology  of  early  stages  of  life  in  Japa- 
nese common  mackerel.  Scomber  japonicus  Houttuyn,  with 
special  reference  to  fluctuation  of  population.     Bull.  Tokai 
Reg.  Fish.  Res.  Lab.  62:1-283.     [In  Japanese.] 
Watanabe,  Y,  Y  Hiyama.  C.  Watanabe,  and  S.  Takayanagi. 

2002.     Inter-decadal  fluctuations  in  length-at-age  of  Hok- 
kaido-Sakhalin herring  and  Japanese  sardine  in  the  Sea  of 
Japan.     PICES  Scientific  Report  20:63-67. 
Yatsu,  A.,  and  H.  Kidokoro. 

2002.  Coherent  low  frequency  variability  in  biomass  and  in 
body  size  of  Japanese  common  squid,  Tadorodes  padificus. 
during  1964-2002,  89  p.  Abstracts  of  PICES  11th  annual 
meeting. 


207 


Latitudinal  and  seasonal  egg-size  variation  of  the 
anchoveta  (Engraulis  ringens)  off  the  Chilean  coast 


Alejandra  Llanos-Rivera 

Leonardo  R.  Castro 

Laboratorio  de  Oceanografia  Pesquera  y  Ecologia  Larval 

Departamento  de  Oceanografia 

Universidad  de  Concepcion 

Casilla  160-C,  Concepcion,  Chile 

E-mail  address  (for  L.  R  Castro,  contact  author)  lecastro@udeccl 


occur  among  populations  of  E.  ringens 
along  its  distribution.  In  this  study,  we 
1 )  report  changes  in  egg  size  through- 
out the  anchoveta  spawning  season 
as  well  as  for  the  peak  months  of  the 
spawning  season,  2)  evaluate  whether 
egg  size  varies  with  respect  to  latitude, 
and  3 )  evaluate  whether  differences  in 
larval  length  and  yolksac  volume  occur 
in  hatching  larvae  from  the  two  major 
spawning  stocks  along  Chile  (central 
and  southern  stocks). 


The  anchoveta  Engraulis  ringens  is 
widely  distributed  along  the  eastern 
South  Pacific  (from  4°  to  42°S;  Serra  et 
al.,  1979)  and  it  has  also  supported  one 
of  the  largest  fisheries  of  the  world  over 
the  last  four  decades.  However,  there 
are  few  interpopulation  comparisons 
for  either  the  adult  or  the  younger 
stages.  Reproductive  traits,  such  as 
fecundity  or  spawning  season  length, 
are  known  to  vary  with  latitude  for 
some  fish  species  (Blaxter  and  Hunter, 
1982;  Conover,  1990;  Fleming  and 
Gross,  1990;  Castro  and  Cowen,  1991). 
and  latitudinal  trends  for  some  early 
life  history  traits,  such  as  egg  size  and 
larval  growth  rates,  have  been  reported 
for  others  clupeiforms  and  other  fishes 
(Blaxter  and  Hempel,  1963;  Ciechom- 
ski.  1973;  Imai  and  Tanaka,  1987, 
Conover  1990,  Houde  1989).  However, 
there  is  no  published  information  on 
potential  latitudinal  trends  during  the 
adult  or  the  early  life  history  of  the 
anchoveta,  even  though  this  type  of 
information  may  help  in  understand- 
ing recruitment  variability,  especially 
during  recurring  large  scale  events 
( such  as  El  Nino  or  La  Nina)  that  affect 
the  entire  species  range. 

Egg  volume  has  been  found  to  vary 
widely  among  species  and  among  popu- 
lations of  the  same  species.  For  fish  that 
broadcast  planktonic  or  benthic  eggs, 
egg  size  often  varies  as  the  spawning 
season  progresses  (Bagenal,  1971),  and 
the  magnitude  of  this  variation  depends 
on  the  species.  For  instance,  the  egg  vol- 
ume of  the  pelagic  spawners  Engraulis 
anchoita  and  Solea  solea  decreases  23% 
and  38%.  respectively,  throughout  the 
spawning  season  (Ciechomski,  1973; 
Rijnsdorp  and  Vingerhoed,  1994).  Ma- 
ternal and  environmental  factors  may 
also  affect  egg  volume  (Bagenal,  1971; 


Thresher,  1984;  Rijnsdorp  and  Vinger- 
hoed, 1994;  Chambers  and  Waiwood. 
1996;  Chambers,  1997).  Variations  in 
size  of  the  spawning  females  and  shifts 
in  energy  allocation  from  reproduction 
to  growth  as  the  spawning  season  pro- 
gresses may  influence  the  egg  volume 
(Wootton,  1990).  Alternatively,  seasonal 
variations  in  photoperiod,  seawater 
temperature,  and  food  supply  during 
the  spawning  season  may  affect  the 
reproductive  output  (Wootton,  1990). 

Scarce  information  exists  on  the 
variability  of  egg  sizes  for  fishes  in  the 
Humboldt  Current.  In  this  extensive 
area,  the  heavily  exploited  anchoveta 
Engraulis  ringens  is  the  dominant 
small  pelagic  species.  Throughout  this 
range,  three  major  stocks  are  recog- 
nized: the  northern  stock  off  northern 
Peru  ( the  largest );  the  central  stock  off 
southern  Peru  and  northern  Chile  (mid- 
size), and  the  southern  stock  off  central 
Chile  (the  smallest  of  the  three).  For 
the  entire  distribution  of  anchoveta, 
the  main  spawning  season  is  from  July 
through  September,  but  may  extend  to 
December  or  January  (Cubillos  et  al., 
1999).  The  wide  latitudinal  range  and 
prolonged  spawning  period  suggest  the 
possibility  of  egg-size  variation,  as  ob- 
served in  other  clupeifoms  (Blaxter  and 
Hempel,  1963;  Ciechomski,  1973;  Imai 
and  Tanaka,  1987).  Egg  size  correlates 
with  larval  characteristics  such  as  lar- 
val length  at  hatching,  the  time  to  first 
feeding,  and  time  before  irreversible 
starvation  (Shirota,  1970;  Ware,  1975; 
Hunter,  1981;  Marteinsdottir  and  Able, 
1992).  To  explore  whether  differences 
in  potential  early-life-stage  survival 
would  exist  among  populations  and  (or 
seasons ),  the  objective  of  our  research 
was  to  determine  whether  variations 
in  some  early-life-stage  characteristics 


Materials  and  methods 

We  collected  anchovy  eggs  from  four 
locations  along  the  coastal  zone  (<20 
nmi  offshore )  off  northern  and  central 
Chile  during  the  austral  winter  and 
spring  spawning  seasons  1995-97 
(Fig.  1).  Eggs  were  collected  with  a 
Calvet  net  (150  urn  mesh)  in  Iquique 
and  Antofagasta  (northern  Chile), 
with  a  standard  conical  net  (330  um) 
in  Valparaiso  and  with  either  a  Tucker 
trawl  (250  ;<m  mesh)  or  a  standard 
bongo  net  ( 500  um)  in  Talcahuano.  The 
shorter  axis  of  the  anchovy  eggs  varied 
from  0.563  mm  (SD=0.032)  in  Iquique 
to  0.657  mm  (SD=0.027 )  in  Talcahuano. 
Consequently,  egg  extrusion  from  the 
nets  was  ruled  out  as  a  potential  source 
of  variation  in  our  collections.  Egg  size 
(length  and  width)  was  measured  with 
an  ocular  micrometer  on  a  dissecting 
microscope  at  25x  magnification.  Upon 
collection,  all  eggs  were  preserved  in 
5%  buffered  formalin.  Previous  studies 
on  anchoveta  eggs  have  reported  no  egg 
size  shrinkage  or  shape  changes  with 
formalin  preservation  (Fisher,  1958). 
Similarly,  reports  on  this  species  and 
other  anchovies  show  that  egg-size 
variations  throughout  their  develop- 
ment do  not  occur  iEngr-aulis  ringens, 
Fisher.  1958;  Engraulis  japonica,  Imai 
and  Tanaka,  1987).  We  tested  this 
hypothesis  using  eggs  that  we  col- 
lected in  northern  Chile  and  found  no 
size  differences  among  different  egg 
stages  (ANOVA.  n=535,  P=0.1176). 


Manuscript  approved  for  publication 
12  August  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:207-212  (2004). 


208 


Fishery  Bulletin  102(1) 


Egg  volume  was  calculated  considering  the  anchovy  egg 
as  an  ellipsoid  (  V=4jt  x  a  x  b  x  c/3,  where  a,  b  and  c  are  the 
ellipse  radii). 

Statistical  analyses  included  one-way  ANOVA  tests  for 
volume  differences  among  eggs  spawned  in  three  subpe- 
riods  during  the  spawning  season  (initial,  middle,  and 
final)  in  Valparaiso  and  Talcahuano  (1996-97).  We  did 
not  have  samples  from  the  start  of  the  spawning  season 
in  1996  for  Valparaiso;  thus,  we  used  samples  collected  in 
late  June  1995  for  this  subperiod  (Table  1).  Variations  in 
the  egg-size  frequency  distributions  between  contiguous 
subperiods  were  also  tested  with  nonparametric  tests 
(Kolmogorov-Smirnov  test).  Changes  in  egg  volume  with 
latitude  were  tested  by  using  egg  sizes  measured  at  four 
localities  (20°,  23°,  33°,  and  36°S)  during  the  peak  spawn- 
ing months  (initial  subperiod).  The  same  statistical  tests 
were  used  as  in  the  previous  objective  (ANOVA,  Kolmogo- 
rov-Smirnov test). 

To  evaluate  the  relationship  between  egg  size  and  lar- 
val length  at  hatching,  live  eggs  from  ichthyoplankton 
samples  from  the  field  in  August  2000  were  transported  to 
the  laboratory  and  reared  at  the  normal  mixed-layer  tem- 
perature off  Antofagasta  and  Talcahuano  (15°C  and  12°C, 
respectively)  (Escribano  et  al.,  1995;  Castro  et  al.,  2000). 
A  subsample  of  the  incubated  egg  batch  was  preserved  in 
5%  formalin  and  measured  at  the  beginning  of  the  experi- 
ments. The  rest  of  the  eggs  were  placed  in  1-L  flasks  and 
incubated  until  hatching  under  12h/12h  photoperiod.  This 
procedure  was  repeated  twice  (4  days  apart)  in  each  zone. 
From  each  rearing  experiment,  30  just-hatched  larvae 
(max.  30  min  from  hatch)  were  anesthetized  and  measured 
(notochord  length)  under  a  dissecting  microscope  with  the 
aid  of  an  ocular  micrometer.  Additionally,  yolksac  sizes  of 
recently  hatched  larvae  were  measured  and  volume  esti- 
mated as  one  half  of  an  ellipsoid  by  using  the  algorithms 
given  above. 


Results 

A  total  of  7196  anchovy  eggs  were  measured.  The  egg  size 
tended  to  decrease  as  the  spawning  season  progressed 
(Fig.  2).  From  late  June  through  January  the  mean 
volume  decreased  by  about  20%  in  Valparaiso  and  by 
about  10"%  in  Talcahuano  (ANOVA,  P<0.05)( Table  1).  The 
size-frequency  distribution  between  consecutive  subperi- 
ods (initial,  middle,  and  final)  also  differed  in  both  areas 
(Kolmogorov-Smirnov  test,  P<0.05).  The  mean  size  of  the 
eggs  in  Valparaiso  was  smaller  than  in  Talcahuano  during 
all  subperiods  (Table  1 ),  and  the  largest  difference  between 
areas  was  at  the  end  of  the  spawning  season  ( 15% ). 

During  the  spawning  peak  (spawning  commencement ). 
when  the  eggs  were  larger,  the  mean  anchoveta  egg  size  in- 
creased with  latitude  (ANOVA,  P<0.01;  Fig.  3).  At  Iquique 
(20"S),  the  mean  egg  volume  was  21%  smaller  than  at  An- 
tofagasta (23°S),  49%  smaller  than  at  Valparaiso  (33°S), 
and  5695  smaller  than  the  eggs  from  Talcahuano  (36°S), 
the  southernmost  location  (Table  2).  The  egg-size  fre- 
quency distribution  differed  between  adjacent  areas  (Kol- 
mogorov-Smirnov test.  P<0.05).  Interestingly,  the  smallest 


Pacific  Ocean 


> 


73 
Longitude  West 

Figure  1 

Areas  where  anchoveta  eggs  were  collected  to 
determine  egg-size  variations  along  the  Chil- 
ean coast.  Arrows  show  the  locations  depicted 
in  Table  2. 


egg  sizes  measured  in  Iquique  (<0.19  mm3)  did  not  occur 
in  Talcahuano.  Similarly,  the  largest  sizes  determined  in 
Talcahuano  (>0.30  mm3)  did  not  occur  in  Iquique.  at  the 
lowest  latitude. 

Larval  length  at  hatching  determined  in  the  rearing 
experiments  at  normal  field  temperatures  was  greater  for 
the  southernmost  population  (Talcahuano)  (Table  3).  The 
mean  larval  size  for  the  southern  location  (2.70  mm  noto- 
chord length)  was  8.2%  greater  than  the  larvae  hatched 
from  eggs  collected  at  the  northern  experimental  location 
(Antofagasta,  2.50  mm).  Furthermore,  the  yolksac  volume 
in  the  recently  hatched  larvae  in  Talcahuano  (0.130  mm3) 


Note     Llanos-Rivera  and  Castro:  Egg-size  variation  of  Engraul/s  ringens 


209 


80 
60 
40 
20 
0 
„  80 

&  60 

c 

<D 

=   40 

(D 

^       20 

0 
80 

60 

40 

20 

0 


Valparaiso  (33°) 


middle 


0  15-0.19  020-0.24  0.25-0.29  0.30-0.34  0  35-0  39  0.40-0.44 


80 

60  - 

40 

20 

0 

80 
60  - 
40 
20  - 


Talcahuano  (36  ) 


o 

80 
60 
40 
20 
0 


0.15-0.19   020-0.24   0.25-0.29   0.30-0  34    0.35-0  39   0.40-0.44 


Size  interval  (mm3) 

Figure  2 

Seasonal  variation  in  egg  size  of  the  anchoveta  (£.  ringens)  off  Valparaiso  and  Talcahuano.  y-axis  is  frequency  over 
the  total  number  of  eggs  measured  at  each  locality.  Initial,  middle,  and  final  are  subperiods  within  the  spawning 
season  (see  Table  2). 


Table  1 

Mean  volume  of  Engraulis  ringens  eggs  at  the  beginning,  middle,  and  end  of  the  spawning  season  off  Valparaiso 
central  Chile.  SD  =  standard  deviations,  n  =  number  of  eggs  measured. 

and  Talcahuano, 

Valparaiso 

Talcahuano 

June  1995 

October  1996 

December  1996 

August  1996 

October  1996 

January  1997 

Volume  (mm3)                      0.298 
SD                                           0.026 
n                                         62 

0.281 
0.029 
759 

0.247 
0.022 
630 

0.312 
0.030 
1833 

0.308 
0.034 
718 

0.286 
0.029 
1099 

was  much  larger  (33%  larger)  than  the  yolk  volume  of  the 
larvae  hatched  in  Antofagasta  (0.098  mm3). 


Discussion 

The  results  of  this  study  identified  several  trends  that 
are  related  to  egg-size  variation.  First,  egg  size  tends  to 
decrease  with  the  progression  of  the  spawning  season. 
Second,  egg  size  increases  with  latitude  during  the  peak 
spawning  period.  Third,  larval  size  at  hatching  is  smaller 
in  the  northern  latitude  populations.  Fourth,  the  yolk  sac 
of  recently  hatched  larvae  is  much  larger  than  expected 


(based  on  the  larval  size  at  hatching)  in  the  southern 
population. 

A  number  of  hypotheses  have  been  proposed  to  ex- 
plain egg-size  variations  in  fish  that  spawn  at  multiple 
times  as  the  reproductive  season  progresses.  It  has  been 
proposed  that  in  clupeiforms,  the  decrease  in  egg  size 
may  result  from  maternal  reduction  of  energy  reserves 
over  the  spawning  season,  a  switch  in  the  stored  energy 
from  reproduction  to  growth,  seasonal  changes  in  the  age 
structure  of  the  spawning  population,  or  changes  during 
ovogenesis  that  are  correlated  with  temperature  (Blaxter 
and  Hunter,  1982;  Chambers,  1997,  for  a  recent  review).  In 
the  anchoveta  E.  ringens,  published  data  suggest  that  some 


210 


Fishery  Bulletin  102(1) 


80 
60 
40 
20 
0 


80  -p 
60  -- 
40 
20 
0 


80 
60 
40 
20 

0 

80 


20-- 


20  S 


n 


.I     ii 


23  S 


-- 

33°S 

-- 

+ 

,1     1, 

H 1 1 

36 >S 


-+- 


0  10-0  14      0  15-0  19     0  20-0  24     0.25-0.29     0.30-0.34     0.35-0.39     0.40-0.44 

Size  interval  (mm3) 

Figure  3 

Latitudinal  variation  in  egg  size  of  the  anchoveta  (£.  ringens )  along  northern 
and  central  Chile  during  the  peak  months  of  the  spawning  season.  Y-axis  is 
frequency  over  the  total  number  of  eggs  measured  at  each  locality. 


of  these  factors  co-occur.  For  instance,  changes  in  growth 
rates  for  yearly  cohorts  during  the  spawning  season  (low  at 
the  beginning,  fast  at  the  end)  have  been  documented  for 
the  southernmost  population  (Cubillos  et  al.,  2001).  Alter- 
natively, variations  in  the  population  age  structure  during 
the  spawning  season  have  also  been  reported  as  the  1.5 
year-old  new  recruits  begin  to  spawn  in  early  summer  (late 
December-January,  Cubillos  et  al.,  1999,  2001 ).  Changes  in 
environmental  factors  affecting  the  spawning  adults  also 
correlate  with  the  egg-size  variations.  The  photoperiod  and 
nearsurface  temperatures  increase  as  the  spawning  season 
progresses  from  mid-winter  to  late  spring. 

Larger  egg  size  at  the  beginning  of  the  spawning  season 
in  winter  may  be  advantageous  for  these  offspring  because 
the  chances  of  survival  increase  with  the  larger  sizes  of  the 
hatching  larvae.  According  to  Cushing  (1967),  larger  size 
larvae  should  be  favored  over  smaller  larvae  in  seasons 
with  variable  environmental  conditions.  In  theTalcahuano 
area,  strong  fluctuations  in  the  hydrographic  regime  occur 
during  winter  as  strong  north  wind  storms  alternate  with 


short  periods  of  south  winds,  and  also  because  of  the  in- 
creased river  flow  to  the  coastal  zone  (Castro  et  al.,  2000). 
Larval  food,  although  variable,  seems  to  be  sufficient  to 
support  most  of  the  larval  growth  demands  for  larger 
exogenous  feeding  larvae  during  winter  (Hernandez  and 
Castro,  2000).  For  recently  hatched  larvae,  however,  the 
picture  might  be  slightly  different  because,  in  addition  to 
food  supply  variability,  the  strong  turbulent  environmental 
conditions  may  jeopardize  first  feeding  success.  In  these 
highly  variable  areas,  therefore,  larger  larval  size  at  hatch- 
ing and  larger  yolk  reserves  may  be  even  more  important 
than  in  other  less  hydrographically  variable  areas  and 
seasons. 

A  remarkable  increase  in  egg  size  at  the  peak  spawn- 
ing season  occurred  with  respect  to  latitude.  Egg  from 
the  northernmost  (20°S)  latitude  were  at  a  maximum 
559c  larger  than  eggs  from  the  southernmost  (36°S)  lati- 
tude. Latitudinal  variations  in  egg  size  have  been  previ- 
ously reported  for  other  anchovies  (i.e.  Engraulis  anchoita; 
Ciechomski,  1973).  However,  egg-size  variations  for  fishes 


Note     Llanos-Rivera  and  Castro:  Egg-size  variation  of  Engraulis  nngens 


211 


Table  2 

Width,  length,  and  volume  of  anchoveta  eggs  collected  at  different  latitudes  along  the  Chilean 

coast  during  the  peak  months  of  the 

spawning  season.  SD  = 

-  standard  deviations,  n  =  number  of 

eggs  measured. 

Latitude  and  area 

Width  (mm) 

Length  (mm) 

Volume  (mm3] 

mean               SD 

mean              SD 

mean           SD 

?! 

20°  Iquique 

0.563             0.032 

1.201            0.076 

0.201          0.031 

1670 

23°  Antofagasta 

0.597             0.030 

1.293            0.083 

0.243         0.034 

425 

33°  Valparaiso 

0.643            0.023 

1.373            0.064 

0.298         0.026 

62 

36°  Talcahuano 

0.657            0.027 

1.377            0.063 

0.312         0.030 

1833 

Table  3 

Morphological  characteristics  of 

recently  hatched  Engraulis  ringens 

larvae  from  rearing  experiments  at  normal  field  tempera- 

tures  in  Antofagasta  (15°C)  and  Talcahuano  (12°C). 

SD 

=  standard  deviations.  N  =  number  of  eggs 

measured 

Exp.  =  expe 

riment. 

Egg 

volume 

Larval  length 

Yolksac 

size 

( mm3 

) 

at  hatching  (mm) 

at  hatching 

(mm3) 

Exp.  1 

Exp.  2 

Exp.  1 

Exp.  2 

Exp.  1 

Exp.  2 

Antofagasta 

15°C 

Mean 

0.264 

0.260 

2.49 

2.50 

0.099 

0.096 

SD 

(0.023) 

(0.023) 

(0.170) 

(0.1041 

(0.012) 

(0.0121 

n 

358 

325 

30 

30 

30 

30 

Talcahuano 

12°C 

Mean 

0.302 

0.292 

2.71 

2.69 

0.126 

0.134 

SD 

(0.023) 

(0.026) 

(0.111) 

(0.103) 

(0.016) 

(0.017) 

n 

254 

66 

30 

30 

30 

30 

are  not  necessarily  always  associated  with  latitude  (i.e. 
north  Atlantic  herring  stocks)  because  local  environmen- 
tal conditions  that  trigger  spawning  (i.e.  specific  tempera- 
ture or  others)  may  have  a  stronger  effect  in  some  species 
(Chambers,  1997).  Because  of  the  extremely  wide  distri- 
bution range  of  the  anchoveta  (4-42°S)  and  its  residence 
along  an  almost  linear  coast  oriented  exactly  north-south, 
we  proposed  that  any  potential  differences  in  egg  size  due 
to  specific  local  conditions  is  probably  over-driven  by  the 
larger  scale  changes  in  environmental  conditions  associ- 
ated with  latitude. 

The  strong  latitudinal  gradient  in  egg  size  of  the  ancho- 
veta may  be  an  adaptive  measure  if  different  egg  sizes  are 
favored  at  different  latitudes  or  if  there  is  a  correlation 
between  egg  size  and  adult  life  history  traits  that  maximize 
net  reproductive  output.  Unfortunately,  an  analysis  of  the 
anchoveta  in  which  fecundity,  age  of  first  reproduction, 
longevity,  or  other  adult  traits  are  compared  in  relation 
to  latitude  has  not  yet  been  carried  out.  The  timing  and 
length  of  the  spawning  season  seem  to  be  similar  for  the 
northern  (Iquique,  20°S)  and  southern  (Talcahuano,  36°S) 
stocks  along  Chile,  despite  the  different  temperatures  at 
which  anchoveta  spawn  (Castro  et  al.,  2001 ).  The  decrease 
in  egg  size  coincides  with  known  temperature  effects  on 
physiological  rates  (Houde,  1989)  and  on  ecological  factors 


related  to  the  need  of  anchoveta  at  early  life  stages  to  re- 
main in  nearshore  environments  (Bakun.  1996).  At  lower 
latitudes,  the  sea  temperature  is  higher  and  the  seaward 
surface  Ekman  transport  is  stronger  and  therefore  eggs 
and  larvae  in  such  conditions  would  likely  develop  rapidly. 
Alternatively,  anchovy  egg  and  larvae  at  higher  latitudes 
are  retained  nearshore  in  winter  (because  the  Ekman 
transport  is  negative,  Castro  et  al.,  2000)  but  are  exposed 
to  lower  temperatures  and  to  strong  turbulence  that  may 
not  facilitate  the  first  feeding  of  recently  hatched  larvae 
and  subsequent  rapid  larval  development.  Larger  eggs, 
larger  larvae  at  hatching,  and  more  energy  reserves  may 
be  the  favored  early  life  history  strategy  in  southern  popu- 
lations. How  the  latitudinal  variations  in  environmental 
characteristics  affect  the  rest  of  the  life  history  traits  of  the 
different  populations  of  Engraulis  ringens,  one  of  the  most 
important  fish  species  in  the  world  in  terms  of  catches, 
remains  to  be  assessed. 


Acknowledgments 

We  acknowledge  help  from  R.  Escribano  (U.  Antofagasta), 
G.  Claramunt  (U.  Arturo  Prat),  and  F.  Balbontin  (U.  of 
Valparaiso)  who  facilitated  ichthyoplankton  collections. 


212 


Fishery  Bulletin  102(1) 


H.  Moyano  (U.  of  Concepcion)  allowed  the  use  of  his  labora- 
tory and  optical  material.  This  study  was  financed  by  the 
project  FONDECYT  1990470  to  L.  R.  Castro.  E.  Tarifeno, 
and  R.  Escribano.  Alejandra  Llanos-Rivera  was  also  par- 
tially supported  by  the  Graduate  School  of  the  Universidad 
de  Concepcion. 


Literature  cited 

Bagenal,  T.B. 

1971.     The  interrelation  of  the  size  of  fish  eggs,  the  date 

of  spawning  and  the  production  cycle.     J.  Fish  Biol.  3: 

207-219 
Bakun,  A. 

1996.  Patterns  in  the  ocean.  Ocean  processes  and  marine 
population  dynamics,  233  p.  California  Sea  Grant  College 
Publ.,  NOAA  in  Cooperation  with  the  Centro  de  Investiga- 
ciones  Biologicas  del  Noreste,  La  Paz,  BCS.  Mexico. 

Blaxter,  J.,  and  G.  Hempel. 

1963.    The  influence  of  egg  size  on  herring  larvae  iClupea 
harengus  L.).     J.  Con.  Perm.  Int.  Explor.  Mer  28:  211-240. 
Blaxter,  J.  H.  S.,  and  J.  R.  Hunter 

1982.    The  biology  of  clupeoid  fishes.    Adv.  Mar.  Biol.  20: 
3-194. 
Castro,  L.  R.,  and  R.  K.  Cowen 

1991.     Environmental  factors  affecting  the  early  life  history 
of  bay  anchovy  (Anchoa  mitchilli)  in  Great  South  Bay,  New 
York.    Mar.  Ecol.  Prog.  Ser.  76:235-247. 
Castro,  L.  R.,  G.  R  Salinas,  and  E.  H.  Hernandez. 

2000.  Environmental  influences  on  winter  spawning  of  the 
anchoveta,  Engruulis  ringens,  off  Central  Chile.  Mar.  Ecol. 
Prog.  Ser.  197:247-258. 

Castro.  L.  R.,  A.  Llanos,  J.  L.  Blanco,  E.  Tarifeno,  R.  Escribano, 
and  M.  Landaeta. 

2001.  Latitudinal  variations  in  spawning  habitat  charac- 
teristics: influence  on  the  early  life  history  traits  of  the 
anchoveta,  Engraulis  ringens,  off  northern  and  central 
Chile.     GLOBEC  Report  16:42-45. 

Chambers.  R.  C. 

1997.  Environmental  influences  on  egg  and  propagule  sizes 
in  marine  fishes.  In  Early  life  history  and  recruitment  in 
fish  populations  (R.  C.  Chambers  and  E.  A.  Trippel,  eds.), 
p.  63-102.     Chapman  &  Hall,  London. 

Chambers,  R.  C,  and  K.  Waiwood. 

1996.     Maternal  and  seasonal  differences  in  egg  sizes  and 
spawning  characteristics  of  captive  Atlantic  cod,  Gadus 
morhua.     Can.  J.  Fish.  Aquat.  Sci.  53:1986-2003. 
Ciechomski,  J.  de 

1973.     The  size  of  the  eggs  of  the  Argentine  anchovy, 
Engraulis  anchoita  (Hubbs  &  Marinil  in  relation  to  the 
season  of  the  year  and  the  area  of  spawning.     J.  Fish  Biol. 
5:393-398. 
Conover,  D. 

1990.    The  relation  between  capacity  for  growth  and  length  of 
the  growing  season:  evidence  for  and  implications  of  coun- 
tergradient  variation.     Trans.  Am.  Soc.  119:416-430. 
Cubillos,  L.  A.,  D.  Arcos,  D.  Bucarey,  and  M.  Canales. 

2001.  Seasonal  growth  of  small  pelagic  fish  off  Talcahuano, 
Chile  (37°S,  73°W):  a  consequence  of  their  reproductive 


strategy  to  seasonal  upwelling?    Aquat.  Living  Resour. 
14:1-10. 
Cubillos,  L.  A.,  M.  Canales,  D.  Bucarey,  A.  Rojas.  and  R.  Alarcon. 

1999.  Reproductive  period  and  mean  size  at  first  maturity 
for  Strangomera  bentincki  and  Engraulis  i-ingens  from  1993 
to  1997,  off  central-southern  Chile.  Invest.  Mar.  27:73-85. 
(In  Spanish.] 

Cushing,  D.  H. 

1967.    The  grouping  of  herring  populations.     J.  Mar.  Biol. 
Assoc.  U.K.  47:193-208 
Escribano,  R.,  L.  Rodriguez,  and  C.  Irribarren 

1995.     Temporal  variability  of  sea  temperature  in  bay  of 
Antofagasta,  Northern  Chile  ( 199 1-1995 1.    Estud.  Oceanol. 
14:39-47.  [In  Spanish.] 
Fisher.  W. 

1958.     Huevos,  crias  y  primeras  prelarvas  de  la  anchoveta 
Engraulis  ringens  Jenyns.     Rev.  Biol.  Mar.  8  (1-2-31:111- 
124. 
Fleming.  I.  A.,  and  M.  Gross. 

1990.     Latitudinal  clines:  a  trade-off  between  egg  number 
and  size  in  pacific  salmon.     Ecology  71(11:1-11. 
Hernandez,  E.  H..  and  L.  R.  Castro. 

2000.  Larval  growth  of  the  anchoveta,  Engraulis  ringens, 
during  the  winter  spawning  season  off  central  Chile.  Fish. 
Bull.  98:704-710. 

Houde,  E.  D. 

1989.  Comparative  growth,  mortality  and  energetics  of 
marine  fish  larvae:  temperature  and  implied  latitudinal 
effects.     Fish.  Bull.  87:471^195. 

Hunter,  J.  R. 

1981.     Feeding  ecology  and  predation  of  marine  fish  larvae. 
In  Marine  fish  larvae  (R.  Lasker.  ed.l.  p.  34-77.     Univ. 
Washington  Press,  Seattle,  WA. 
Imai,  Ch.,  and  S.  Tanaka. 

1987.     Effect  of  sea  water  temperature  on  egg  size  of  Jap- 
anese anchovy.     Nippon  Suisan  Gakkaishi  53(12):  2169- 
2178. 
Marteinsdottir,  G.,  and  K.  Able. 

1992.     Influence  of  egg  size  on  embryos  and  larvae  ofFundu- 
lus  heteroclitus  (L.l.    J.  Fish  Biol.  41:883-896. 
Rijnsdorp,  A.,  and  B.  Vingerhoed. 

1994.     The  ecological  significance  of  geographical  and  sea- 
sonal differences  in  egg  size  in  sole  Solea  solea  (L.).     Neth. 
J.  Sea  Res.  32(3/41:255-270. 
Serra  J.,  O.  Rojas,  M.  Aguayo.  F.  Inostroza,  and  J.  Canon. 

1979.    Anchoveta  {Engraulis  ringens).    In  Estado  actual  de  las 
principales  pesquerias  nacionales.  Bases  para  un  desarrollo 
pesquero.  36  p.     Corporacibn  de  fomento  de  la  production.     In- 
stituto  de  Fomento  Pesquero.  Santiago,  Chile. 
Shirota,  A. 

1970.     Studies  on  the  mouth  size  offish  in  the  larval  and  fry 
stages.    Bull.  Jap.  Soc.  Sci.  Fish.  36:353-368. 
Thresher,  R.  E. 

1984.     Reproduction  in  reef  fishes,  399  p.    T.F.H.     Publica- 
tions. Inc.  Ltd. .The  British  Crown  Colony  of  Hong  Kong. 
Ware,  D.  M. 

1975.     Relation  between  egg  size,  growth,  and  natural  mortal- 
ity of  larval  fish.     J.  Fish.  Res.  Board  Can.  32:  2503-2512. 
Wootton.  R. 

1990.  Ecology  of  teleost  fishes,  404  p.  Chapman  &  Hall. 
London. 


213 


Molecular  methods  for  the  genetic  identification 
of  salmonid  prey  from  Pacific  harbor  seal 
(Phoca  vitulina  richardsi)  scat 


Maureen  Purcell 

Greg  Mackey 

Eric  LaHood 

Conservation  Biology  Molecular  Genetics  Laboratory 
Northwest  Fisheries  Science  Center 
National  Marine  Fisheries  Service.  NOAA 
2725  Montlake  Blvd.  E. 
Seattle,  Washington  98112-2097 


Harriet  Huber 

National  Marine  Mammal  Laboratory 
Alaska  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
7600  Sand  Point  Way  NE 
Seattle,  Washington  98115 


Linda  Park 

Conservation  Biology  Molecular  Genetics  Laboratory 

Northwest  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

2725  Montlake  Blvd.  E. 

Seattle,  Washington  98112-2097 

E-mail  address  (for  L.  Park,  contact  author):  linda  parkig'noaa  gov 


Twenty-six  stocks  of  Pacific  salmon 
and  trout  [Oncorhynchus  spp.),  rep- 
resenting evolutionary  significant 
units  (ESU),  are  listed  as  threatened 
or  endangered  under  the  Endangered 
Species  Act  (ESA)  and  six  more  stocks 
are  currently  being  evaluated  for 
listing.1  The  ecological  and  economic 
consequences  of  these  listings  are 
large;  therefore  considerable  effort  has 
been  made  to  understand  and  respond 
to  these  declining  populations.  Until 
recently.  Pacific  harbor  seals  (Phoca 
vitulina  richardsi)  on  the  west  coast 
increased  an  average  of  5%  to  1%  per 
year  as  a  result  of  the  Marine  Mammal 
Protection  Act  of  1972  (Brown  and 
Kohlman2).  Pacific  salmon  are  season- 
ally important  prey  for  harbor  seals 
(Roffe  and  Mate,  1984;  Olesiuk,  1993); 
therefore  quantifying  and  understand- 
ing the  interaction  between  these  two 
protected  species  is  important  for 
biologically  sound  management  strat- 
egies. Because  some  Pacific  salmonid 
species  in  a  given  area  may  be  threat- 


ened or  endangered,  while  others  are 
relatively  abundant,  it  is  important 
to  distinguish  the  species  of  salmonid 
upon  which  the  harbor  seals  are  prey- 
ing. This  study  takes  the  first  step  in 
understanding  these  interactions  by 
using  molecular  genetic  tools  for  spe- 
cies-level identification  of  salmonid 
skeletal  remains  recovered  from  Pacific 
harbor  seal  scats. 

Most  studies  of  harbor  seal  food  hab- 
its rely  on  morphological  identification 
of  indigestible  parts  (e.g.  otoliths  and 
bones)  from  scat.  Otoliths  can  be  used 
to  identify  fish  species  (Ochoa-Acuna 
and  Francis,  1995)  but  are  not  always 
present  in  scats,  which  can  result  in  an 
underestimate  of  the  number  of  species 
and  the  number  offish  consumed  (Har- 
vey, 1989).  Skeletal  remains  in  scat  are 
much  more  common  and  generally 
bones  can  be  identified  to  the  species 
level  (Cottrell  et  al.,  1996).  Morpho- 
logical identification  is  possible  to  the 
family  level  only  with  Pacific  salmonid 
bones;  however,  genetic  markers  have 


the  ability  to  discriminate  between 
species,  and  the  feasibility  of  extracting 
DNA  from  bones  has  been  clearly  dem- 
onstrated (Hochmeister  et  al.,  1991). 

Mitochondrial  DNA  (mtDNA)  has 
been  widely  employed  in  systematic 
studies  (reviewed  by  Avise,  1994)  mak- 
ing it  ideal  for  animal  species  identifi- 
cation. In  this  study,  we  explored  three 
regions  of  the  mitochondrial  genome 
that  have  been  previously  character- 
ized in  Pacific  salmonids  (Shedlock 
et  al.,  1992;  Domanico  and  Phillips, 
1995:  Parker  and  Kornfield,  1996). 
DNA  sequencing  of  these  regions 
provided  an  unambiguous  way  to  de- 
termine species  identity.  Because  high 
throughput  sequencing  can  be  prohibi- 
tively expensive  for  laboratories  with 
limited  facilities,  restriction  fragment 
length  polymorphism  (RFLP)  analysis 
was  also  explored  as  an  alternative  for 
species  identification.  A  previous  study 
had  established  a  species-specific  poly- 
merase chain  reaction  (PCR)  test  for 
Pacific  Northwest  salmon  and  coastal 
trout  species  (McKay  et  al.,  1997).  The 
PCR  test  is  based  on  the  initial  ampli- 
fication of  an  approximately  1000-bp 
fragment  of  the  nuclear  growth  hor- 
mone 2  gene.  The  degraded  state  of  the 
DNA  isolated  from  bones  recovered 
from  scat  has  generally  limited  suc- 
cessful PCR  to  amplicons  of  300  bp  or 
less  (data  not  shown).  Furthermore, 
the  amount  of  DNA  isolated  from  bone 
fragments  can  be  quite  small;  mtDNA 
is  present  in  higher  copy  number  per 
cell  than  is  nuclear  DNA.  Thus,  we 
considered  mtDNA  it  to  be  a  more 


1  http://www.nwr.noaa.gov/lsalmon/salmesay 
specprof.htm.     [Accessed  June  17,  2003.] 

-  Brown,  R.  F.  and  S.  G.  Kohlman.  1998. 
Trends  in  abundance  and  current  status 
of  the  Pacific  harbor  seal  tPhoca  vitulina 
richardsi)  in  Oregon:  1977-1998.  ODFW 
(Oregon  Department  of  Fish  and  Wildlife  i 
Wildlife  Diversity  Program  Technical 
Report,  98-6-01.  16  p.  [Available  from 
ODFW,  7118  NE  Vandenberg  Ave.  Corval- 
lis,  OR  97333.] 


Manuscript  approved  for  publication 
9  October  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:213-220  (2004). 


214 


Fishery  Bulletin  102(1) 


appropriate  target  for  our  assay.  We  chose  to  explore 
smaller  regions  of  the  mitochondrial  genome,  including  the 
d-loop  (Shedlock  et  al.,  1992),  a  portion  of  the  16s  ribosomal 
gene  (Parker  and  Kornfield,  1996),  and  a  region  spanning 
the  cytochrome  oxidase  III,  t-RNA  glycine,  and  ND3  genes 
(hereafter,  referred  to  as  COIII/ND3)  (Domanico  and  Phil- 
lips, 1995 ).  Significant  interspecific  variation  but  not  intra- 
specific  variation  was  observed  in  the  COIII/ND3  region 
among  salmonid  species  in  previous  studies,  making  it  a 
particularly  good  candidate  region  for  the  development  of 
diagnostic  markers  (Domanico  and  Phillips,  1995). 

In  the  first  phase  of  the  study,  we  developed  and  vali- 
dated the  genetic  tools  for  species  identification  by  using 
frozen  or  ethanol-preserved  tissues  collected  from  known 
species  and  populations.  In  the  second  phase,  we  applied 
these  tools  to  the  identification  of  bone  remains  from  har- 
bor seal  scats  collected  at  the  Umpqua  River  (Oregon). 
A  number  of  Pacific  salmonid  species  are  present  in  the 
Umpqua  River  but  of  particular  concern  were  the  sea- 
run  cutthroat  (Oncorhynchus  clarki)  that  were  listed  as 
endangered  under  the  ESA  during  1996  (Johnson  et  al., 
1999).  Here  we  report  the  method  associated  with  these 
two  phases  of  the  project.  The  salmonid  bones  that  were 
identified  genetically  were  incorporated  into  a  larger  study 
of  the  harbor  seal  diet  and  are  reported  in  a  companion 
paper  (Orr  et  al.,  2004). 


Materials  and  methods 

Salmonid  tissue  samples  of  known  species  have  been 
collected  over  the  past  decade  by  geneticists  from  the 
Conservation  Biology  Molecular  Genetics  Laboratory 
(NOAA/NMFS/NWFSC)  or  generously  donated  by  others 
(see  "Acknowledgments"  section)  and  maintained  either 
frozen  at  -80°C  or  preserved  in  95%  ethanol.  Reference 
populations  were  chosen  to  represent  the  geographic 
range  of  chinook  salmon  (O.  tshawytscha),  coho  salmon 
(O.  kisutch),  sockeye  salmon  (O.  nerka).  pink  salmon  (O. 
gorbuscha),  chum  salmon  (O.  keta),  steelhead  (O.  mykiss), 
coastal  cutthroat  trout  (O.  clarki  clarki),  and  Yellowstone 
cutthroat  trout  ( O.  clarki  bouvieri )  ( collection  information  is 
listed  in  Table  1 ).  Tissues  were  extracted  with  either  a  stan- 
dard phenol  and  chloroform  extraction  (Sambrook  et  al., 
1989)  or  by  using  the  DNAeasy  96-well  tissue  kit  (Qiagen, 
Valencia,  CA),  following  the  manufacturer's  instruction 
for  tissue  preparations.  PCR  primers  were  either  taken 
directly  from  the  published  studies  or  designed  from  the 
reported  sequences  (Table  2).  All  primers  were  cycled  with 
2.5  mM  MgCl2, 0.8  mM  dNTPs,  0.04  ,«M  primers,  0.25  units 
of  Taq  DNA  polymerase  (Promega,  Madison,  WI),  20-40  ng 
of  DNA,  and  cresol  red  loading  buffer  (final  concentration 
2' <  sucrose  and  0.005%  cresol  red)  for  35-45  cycles  of 
94°C  for  45  seconds,  55°C  for  45  seconds,  and  72°C  for 
1  minute. 

A  single  individual  of  each  salmonid  species  listed  in 
Table  1  was  sequenced  for  both  the  16s  rRNA  and  COIII/ 
ND3  regions.  For  DNA  sequencing,  the  PCR  products  were 
purified  with  an  Ultrafree  MC  column  (Millipore,  Beverly, 
MA  i  and  resuspended  in  20  ,uL  of  sterile  water.  The  puri- 


fied product  (1-10  uL  depending  on  band  intensity)  was 
manually  sequenced  by  using  the  USB  ThermoSeque- 
nase  cycle  sequencing  kit  (Cleveland.  OH),  following  the 
manufacturer's  instructions.  MACDNASIS  (Miraibio  Inc., 
Alameda.  CA)  and  SEQUENCHER  (Gene  Codes  Corp.,  Ann 
Arbor.  MI)  were  used  for  sequence  alignment  and  identifi- 
cation of  diagnostic  restriction  enzyme  cut  sites. 

RFLP  analysis  of  the  unpurified  COIII/ND3  PCR  product 
was  performed  in  the  presence  of  a  cresol  red  loading  buf- 
fer. Restriction  digests  were  incubated  for  6  to  12  hours  at 
37°C  for  Dpn  II,  Sau  961,  Fok  I,  Ase  I,  at  50°  for  Apo  I,  and 
at  60°C  for  Bst  NI  with  the  supplied  buffers  (NEB,  Beverly, 
MA)  and  1-5  units  of  enzyme.  Restricted  products  were 
electrophoresed  in  a  47c  3:1  high-resolution  and  medium- 
resolution  agarose  gel  (Continental  Laboratory  Products, 
San  Diego,  CA).  DNA  bands  on  the  agarose  gels  were 
visualized  with  SYBR  Gold,  following  the  manufacturer's 
instructions  (Molecular  Probes,  Eugene,  OR). 

Personnel  from  the  National  Marine  Mammal  Laboratory 
(NMML)  collected  and  processed  harbor  seal  scat  samples 
from  the  Umpqua  River  (Orr  et  al.,  2004).  NMML  research- 
ers identified  bone  remains  to  either  family  or  species  level 
by  using  morphological  characteristics  of  skeletal  remains 
(Orr  et  al.,  2004).  From  39  harbor  seal  scats,  116  bones  were 
identified  morphologically  to  the  genus  Oncorhynchus  and 
subjected  to  DNA  analysis  for  species  identification.  For  a 
positive  DNA  extraction  control,  we  simulated  digestion 
by  treating  coastal  cutthroat  bones  (collected  from  Cowlitz 
Trout  Hatchery,  Winlock,  WAi  in  a  mixture  of  laboratory- 
grade  trypsin  (a  digestive  enzyme),  baking  soda,  and  water 
for  1  to  2  days.  These  trypsin-treated  bones  from  a  coastal 
cutthroat  trout  were  used  as  positive  DNA  extraction  and 
amplification  control. 

To  prepare  samples  for  DNA  extraction,  bones  were 
soaked  in  107c  sodium  hypochlorite  for  10  minutes  to 
destroy  any  contaminating  DNA  that  may  have  adhered 
to  the  outside  of  the  bone  and  were  rinsed  twice  in  sterile 
water.  Bones  ranged  in  weight  from  0.1  to  105.6  mg  and 
included  teeth,  vertebrae,  gillrakers,  radials,  and  bone 
fragments  (hereafter,  all  bony  parts  and  teeth  will  be  re- 
ferred to  as  "bone").  The  bones  were  decalcified  overnight 
in  0.5M  EDTA  solution  (Hochmeister  et  al.,  1991);  fragile 
or  small  fragments  were  not  decalcified.  The  EDTA  was 
removed  and  the  decalcified  samples  were  extracted  with 
the  QIAamp  tissue  extraction  kit  (Qiagen.  Valencia.  CA) 
according  to  the  manufacturer's  instructions  with  the 
following  modifications:  1)  samples  were  proteinase  K 
digested  overnight  or  until  completely  digested;  2)  10 
mg/«L  yeast  t-RNA  carrier  was  added  to  the  extractant 
before  placement  on  the  QIAQuick  column;  and  3)  DNA 
was  eluted  in  a  reduced  volume  (50-100  «L)  of  buffer  AE. 
Negative  controls  containing  no  tissue  were  simultane- 
ously processed  to  verify  that  the  extraction  was  free  of 
contaminating  DNA.  The  trypsin-treated  coastal  cutthroat 
bones  were  used  as  positive  extraction  and  PCR  controls. 

Five  to  ten  microliters  of  the  extracted  DNA  were  used 
in  each  amplification  reaction.  Amplification  success  was 
determined  by  electrophoresis  through  a  27c  agarose  gel 
followed  by  staining  with  ethidium  bromide  or  the  more 
sensitive  SYBR  Gold  i Molecular  Probes).  Species  identifi- 


NOTE     Purcell  et  al.:  Genetic  identification  of  salmonid  prey  from  scat  of  Phoca  vitulina  nchardsi 


215 


Table  1 

Species,  locations,  and  sampl 

;  sizes  (n  1  examined  for  RFLP  analysis. 

Species 

Population 

Location 

71 

Chinook 

Walker  Creek 

Upper  Frasier  River.  British  Columbia 

10 

Grovers  Creek  Hatchery 

Puget  Sound,  Washington 

12 

Lookingglass  Hatchery 

Snake  River.  Oregon 

12 

Carson  Hatchery 

Columbia  River,  Washington 

12 

Abernathy  Hatchery. 

Columbia  River,  Washington 

11 

Upper  Sacramento  Mainstem 

Sacramento  River.  California 

10 

Coho 

Edison  Creek 

Oregon  Coast 

13 

Sandy  River 

Columbia  River,  Oregon 

15 

North  Fork  Moclips  River 

Washington  Coast 

15 

Minter  Creek  Hatchery 

Puget  Sound,  Washington 

15 

Yakoun  River 

Queen  Charlotte  Island,  British  Columbia 

7 

Sockeye 

Nehalem  Ponds 

Oregon  Coast 

4 

Redfish  Lake 

Snake  River,  Idaho 

4 

Alturas  Lake 

Snake  River,  Idaho 

2 

Ozette  Lake 

Washington  Coast 

14 

Lake  Wenatchee 

North  Cascades.Washington 

10 

Babine  Lake 

Central  British  Columbia 

2 

Kamchatka  River 

Kamchatka  Peninsula,  Russia 

9 

Chum 

Hamma  Hamma  River 

Hood  Canal.  Washington 

11 

Frosty  Creek 

Alaskan  Peninsula 

12 

Utka  River 

Chucotka  Peninsula,  Russia 

9 

Miomote  River 

West  Honshu.  Japan 

11 

Pink 

Nisqually  River 

South  Puget  Sound.  Washington 

6 

Snohomish  River  Even  Year 

North  Puget  Sound,  Washington 

12 

Skagit  River 

North  Puget  Sound,  Washington 

7 

Hood  Canal  Hatchery 

Hood  Canal,  Washington 

9 

Steelhead 

Gaviota  Creek 

South  California  Coast 

4 

Coquille  River 

Oregon  Coast 

8 

Upper  Tucannon  River 

Snake  River,  Washington 

12 

Finney  Creek 

Puget  Sound,  Washington 

12 

Quinault  Hatchery 

Washington  Coast 

12 

Tigil  River 

Kamchatka  Peninsula.  Russia 

12 

Cutthroat' 

Alsea  River 

Oregon  Coast 

2 

Alsea  Hatchery 

Oregon  Coast 

3 

Duwamish  River 

Puget  Sound  Washington 

12 

Yellowstone  River 

Yellowstone  River.  Montana 

5 

'  Cutthroat  trout  from  the  Yellowstone  River  are  a  different  subspecies  (O.  clarki  bouvieri)  from  the  Washington  and  Oregon  coastal  cutthroat 

trout 

(O.  clarki  clarki). 

cation  was  accomplished  by  sequencing  of  either  the  d-loop 
or  the  COIII/ND3  region.  RFLP  analysis  was  performed 
as  described  above  with  the  following  modifications:  Bst 
NI  was  excluded  because  it  is  redundant  with  Dpn  II, 
the  enzyme  amount  was  reduced  to  0.4-1.0  units  per 
reaction,  and  incubation  time  did  not  exceed  2  hours.  The 
COIII/ND3  primers  are  specific  to  the  family  Salmonidae. 
To  test  the  possibility  that  the  failure  to  obtain  amplifica- 
tion with  the  COIII/ND3  primers  was  due  to  morphologi- 
cal misidentification  of  an  Oncorhynchus  species  we  used 
the  16s  primers  that  are  conserved  across  a  broad  set  of 


taxa  from  Platyhelminthes  through  Chordata  ( Parker  and 
Kornfield,  1996). 


Results 

The  COIII/ND3  and  16s  sequences  were  confirmed  for  all 
seven  salmonid  naturally  present  in  the  Pacific  Northwest 
(Figs.  1  and  2)  and  deposited  in  Genbank  (COIII/ND3: 
AF294827-AF294833;  16S:  AF296341-AF296347).  Two 
chinook  salmon  were  sequenced  representing  two  Dpn  II 


216 


Fishery  Bulletin  102(1) 


Primer 

sequences 

size  of  amplified  product  in  base 

Table  2 

Dairs,  and  references  for  mitochondria]  loci  used  in  this  study. 

Locus 

Primer  sequences  (5'  to  3') 

Product  size 

Reference 

d-loop 

COIII/ND3 

16sV 

P2:  tgt  taa  ace  cct  aaa  cca  g 
P4:  gec  gaa  tgt  aaa  gca  tct  ggt 

F:  tta  caa  teg  ctg  acg  gcg 

R:  gaa  aga  gat  agt  ggc  tag  tac  tg 

F:  tac  ata  aca  cga  gaa  gac  c 
R:  gtg  att  gcg  ctg  tta  tec 

230 
368 

260 

Shedlocketal..  1992 
Domanico  and  Phillips 
Parker  and  Kornfield, 

1995 
1997 

Table  3 

Restriction  fragment 

length  polymorphisms  of  the  cytochrome  oxidase  III  a 

id  ND3  region  digested  w 

ith  six  restriction 

?nzymes. 

The  "A"  haplotype  does  not  cut  with  the 

enzyme,  "B"  cuts 

with  the 

enzyme, 

and  "C"  cuts  with  the  enzyme  but  at  a  different  site 

than  "B." 

Species 

Dpn  II 

Sau  961 

Fok  I 

Asel 

Apo  I 

Bst  NI 

Chinook 

A/B; 

B 

B 

A 

A 

A 

Coho 

A 

A 

B 

A 

A 

A 

Sockeye 

A 

A 

A 

A 

C 

B 

Chum 

A 

A 

A 

B 

C 

A 

Pink 

C 

A 

A 

B 

C 

C 

Steelhead 

A 

A 

A 

B 

B 

A 

Cutthroat 

A 

A 

A 

A 

A 

A 

1  Spring-running  chinook  from  the  Columbia  and  Snake  Rivers  were  polymorphic  foi 

the  Dpn  II  cut  site.    Spring 

chinook  from  Carson 

Hatchery 

(derived  from  the  upper  Columbia  River  spri 

ng-running  ESU  [evolutionary 

significam 

unit]  I  had  the  "A"  haplotype  at  a  frequency  of  0.91  ( 

n=12) and 

spring  chinook  from  Lookingglass  Hatchery  (Snake  River  spring-summer- 

■unning  ESU)  had  the  "A"  haplotype  at 

a  frequency  of  0.83  (n  =  12).  All 

other  chinook  samples 

from  Table  1  were  invariant  for  the  "B"  h£ 

plotvpe. 

haplotypes  (A  and  B)  and  their  sequences  are  presented 
in  Figure  1;  the  chinook  salmon  individuals  were  from  the 
Upper  Columbia  River  summer  and  fall  ESU  (Methow 
River,  WA).  A  second  intraspecific  polymorphism  in  chi- 
nook salmon  was  observed  at  position  341  between  our 
ND3  sequence  and  the  published  sequence  (Domanico 
and  Phillips,  1995)  (Fig.l).  Sufficient  nucleotide  varia- 
tion exists  in  the  d-loop  (Shedlock  et  al.,  1992)  and  in  the 
COIII/ND3  region  ( Fig.  1 )  to  distinguish  among  the  salmon 
species  by  sequencing;  both  regions  were  used  for  bone 
identification. 

Six  restriction  enzymes  were  selected  from  the  COIII/ 
ND3  sequence  that  appeared  to  distinguish  among  all  the 
species  (Dpn  II,  Sau  961,  Fok  I,  Ase  I,  Apo  I,  and  Bst  NI) 
(Fig.  1).  The  Dpn  II  and  Bst  NI  cut  patterns  are  redundant 
in  that  only  one  of  these  enzymes  is  required  for  species 
identification  when  used  in  conjunction  with  the  other  four 
enzymes  (however,  only  Dpn  II  exhibits  the  intraspecific 
chinook  polymorphism,  see  below).  Haplotype  patterns  for 
all  species  are  listed  in  Table  3.  The  haplotypes  were  scored 
with  a  simple  alphabetic  system:  "A"  was  uncut  (368  base- 
pair  (bp)  band)  and  "B"  was  cut  (the  size  differed  depending 


on  enzyme).  A  few  of  the  enzymes  had  an  alternative  cut 
site,  and  the  resulting  haplotype  we  labeled  "C."  The  "B" 
haplotype  produced  by  Apo  I  occurs  in  steelhead  and  the 
bands  migrate  at  300  and  68  bp,  whereas  the  bands  of  the 
"C"  haplotype  in  sockeye,  chum,  and  pink  salmon  migrate 
at  250  and  118  bp.  The  enzyme  Bst  NI  also  has  two  cut  pat- 
terns: the  sockeye  salmon  "B"  haplotype  bands  migrate  at 
282  and  87  bp  and  the  "C"  haplotype  bands  in  pink  salmon 
migrate  at  271  and  98  bp.  The  Dpn  II  "B"  haplotype  in 
chinook  salmon  creates  two  fragments,  290  and  80  bp;  the 
"C"  haplotype  in  pink  salmon  creates  three  fragments,  292, 
53,  and  24  bp. 

To  confirm  that  the  restriction  enzyme  polymorphisms 
were  diagnostic  within  each  species,  we  surveyed  all  seven 
Pacific  salmon  species  representing  multiple  populations 
spanning  a  large  geographic  range  (Table  II.  No  intra- 
specific polymorphisms  were  detected  among  populations 
with  the  exception  of  chinook  salmon  (Tables  1  and  3).  A 
single  intraspecific  polymorphism  was  found  with  the  Dpn 
II  enzyme  in  chinook  salmon  lineages  in  the  Columbia 
and  Snake  River  basins  (Tables  1  and  3).  Chinook  salmon 
from  the  Snake  River  spring-summer  run  (Lookingglass 


NOTE     Purcell  et  al.:  Genetic  identification  of  salmonid  prey  from  scat  of  Phoca  vitulina  nchardsi 


217 


Chinook  A 

Chinook  B 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook  A 

Chinook  B 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook  A 

Chinook  B 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook  A 

Chinook  B 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook  A 

Chinook  B 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook  A 

Chinook  B 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook  A 

Chinook  B 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


20     DpnII  40  60 

*  *  *  * 

TTACAATCGCTGACGGCGTGTACGGCTCTACTTTCTTTGTCGCCACCGGATTCCATGGCC 


. . . . A. 
.T. .A. 
.T. .A. 


DpnII  80 


100 


Apol/  Sau96I 


TACACGTGATTATTGGCTCAACCTTTCTAGCCGTTTGCCTTCTGCGACAGGTCCAATACC 
A 


.  .  .  . A. 
.  .  .  . A. 
.T. .A. 


:.  .G. 
.  .  .G. 
.T.G. 


.A. .T. 
. AA . T . 
. AA.T. 
. AA.T. 


Fokl        140  160  180 

********** 

ACTTTACATCCGAACATCATTTTGGCTTTGAAGCTGCTGCTTGATATTGACACTTTGTAG 


.T. . . . 
.T.  .  .  . 
.T. .G. 


200  220  start  tRNA  glycine 

--> 

ACGTTGTGTGACTCTTCCTATACGTCTCTATTTACTGATGAGGCTCATAATCTTTCTAGT 


.A. .G. 
. A.  .  .  . 


Asel 

****** 


260 


BSTNI 


280 


Start  ND3 
—  > 

ATTAACACGTATAAGTGACTTCCAATCACCCGGTCTTGGTTAAAATCCAAGGAAAGATAA 


.  .G 

.  TGA 

.  TTA 

.TTA. . .CG. 
.T 


Apol  DpnII  340  360 

******  **** 

TGAACTTAATTACAACAATCATCACTATTACCATCACATTRTCCGCAGTACTAGCCACTA 


.CG. 
.C.G. 


.CG. 
.  .  .A. 
.  .  .G. 


TTTCTTTC 


Figure  1 

Aligned  sequences  of  the  3'  region  of  the  cytochrome  oxidase  III  gene  (COM  I,  the  tRNA  glycine  gene, 
and  the  5'  region  of  the  ND3  gene  for  seven  species  of  the  genus  Oncorhynchus.  The  cutthroat  trout 
sequence  is  represented  by  the  coastal  cutthroat  subspecies  (O.  clarki  clarki).  Chinook  "A"  refers  to 
the  "A"  Dpn  II  haplotype;  chinook  "B"  refers  to  the  "B"  Dpn  II  haplotype.  Sequence  identity  relative 
to  the  chinook  salmon  "A"  sequence  is  denoted  by  dots;  nucleotide  substitutions  are  indicated.  The 
arrow  at  basepair  (bp)  230  is  the  start  of  the  tRNA  glycine  gene  and  the  arrow  at  bp  300  is  the  start 
of  the  ND3  gene.  Stars  above  the  sequence  correspond  to  restriction  enzyme  cut  sites  used  in  this 
study.  At  position  341  in  chinook.  the  R  represents  an  A  or  G. 


218 


Fishery  Bulletin  102(1) 


Chinook 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


Chinook 

Coho 

Sockeye 

Chum 

Pink 

Steelhead 

Cutthroat 


20                    40                    60 
GGAGCTTTAGACACCAGGCAGATCACGTCAAACAACCTTGAATTAACAAGTAAAAACGCAGT 
G 


80  100  120 

GACCCCTAGCCCATATGTCTTTGGTTGGGGCGACCGCGGGGGAAAATTAAGCCCCCATGTGG 


140                   160                   180 
ATGGGGGCATGCCCCCACAGCCAAGAGCCACAGCTCTAAGCACCAGAATATCTGACCAAAAA 
T T...A 


200  220 

TGATCCGGCAAACGCCGATCAACGGACCGAGTTACCCTAG. . . 


Figure  2 

Aligned  sequences  of  a  variable  portion  of  the  16s  gene  for  seven  species  of  the  genus  Oncorhynchus. 
Sequence  identity  in  relation  to  the  chinook  salmon  "A"  sequence  is  denoted  by  dots;  nucleotide 
substitutions  are  indicated. 


Hatchery)  and  hatchery  stocks  descended  from  the  Upper 
Columbia  River  spring  run  (Carson  Hatchery)  had  the  "A" 
(uncut)  haplotype  at  a  frequency  of  83%  and  91%,  respec- 
tively, whereas  those  from  the  Lower  Columbia  River  ESU 
were  invariant  for  the  "B"  (cut)  haplotype.  The  "B"  hap- 
lotype was  also  invariant  in  the  other  lineages  examined 
(Sacramento  River,  CA;  Puget  Sound,  WA;  and  the  Fraser 
River,  BC).  Despite  this  Dpn  II  polymorphism,  the  haplo- 
type patterns  were  still  chinook-specific. 

Extractions  from  the  trypsin-treated  cutthroat  trout 
bones,  used  as  positive  controls,  were  amplified  consis- 
tently, but  of  the  116  salmonid  bones  from  harbor  seal 
scats,  only  78  (67%)  were  amplified.  Failed  samples  were 
repeated  several  times  with  all  possible  primer  sets.  Be- 
cause each  scat  contained  multiple  bones,  we  were  able 
to  amplify  bones  representing  35  of  the  39  scats  (90%). 
The  smallest  bone  we  successfully  amplified  was  a  O.'2-mg 
tooth  and  the  largest  was  a  21.8-mg  vertebra.  There  did 
not  appear  to  be  a  relationship  between  bone  size  and  DNA 
extraction  success;  no  significant  difference  in  mean  bone 
size  was  detected  between  32  bones  that  either  amplified 
or  failed  (P=0.280;  unpaired  t-test;  SYSTAT  8.0  [Chicago, 
IL| ).  The  bone  samples  that  failed  to  amplify  repeatedly 
were  also  tested  by  using  the  evolutionarily  conserved 
16s  primers.  Some  samples  were  still  refractory  to  PCR, 
indicating  that  the  overall  DNA  quality  or  quantity  was 


insufficient  for  this  assay;  however,  those  samples  that  did 
amplify  were  identified  by  sequencing  as  salmon.  In  an  un- 
related study  using  river  otter  bones  (data  not  presented), 
one  bone  sample  morphologically  identified  as  salmonid 
yielded  a  sequence  with  100%  identity  to  the  published  16s 
sequence  available  for  Northern  squawfish  {Ptychocheilus 
oregonensis)  (Simons  and  Mayden,  1998). 

After  verifying  the  specificity  of  the  RFLP  analysis  for 
differentiating  the  Pacific  salmon  species,  the  assay  was 
applied  to  the  bone  samples.  Restriction  enzyme  digestion 
required  some  modification  when  applied  to  bone.  On  occa- 
sion, the  restriction  enzyme  protocol  developed  for  the  fresh 
tissue  resulted  in  degradation  of  the  amplified  bone  PCR 
product.  Enzyme  amount  and  digestion  times  were  scaled 
back  for  the  analysis  of  the  bone  samples.  The  Fok  I  enzyme 
proved  the  most  difficult  for  the  bone  samples,  which  was 
likely  due  to  nonspecific  restriction  that  occurs  when  the 
enzyme  is  present  at  a  high  concentration  in  relation  to  its 
target  or  if  the  reaction  is  allowed  to  digest  for  more  than 
two  hours.  In  some  cases,  only  very  weak  amplification  was 
achieved  with  the  bone  samples  and  it  was  difficult  to  get 
digestion  without  degradation.  Although  sequencing  was 
the  main  technique  used  for  bone  identification.  23  bones 
in  this  study  were  identified  by  using  the  RFLP  technique. 
Fourteen  of  these  23  bones  were  additionally  confirmed  by 
sequencing  and  the  two  techniques  gave  matching  results. 


NOTE     Purcell  et  al.:  Genetic  identification  of  salmonid  prey  from  scat  of  Phoca  vitulina  nchardsi 


219 


Discussion 

This  study  focused  on  the  development  of  tools  for  the 
genetic  identification  of  Pacific  salmon  skeletal  remains 
recovered  from  harbor  seal  scats.  These  tools  help  to  deter- 
mine the  diet  of  marine  mammals  and  can  also  be  used  to 
address  direct  management  questions  regarding  interspe- 
cific interactions  in  rivers  such  as  the  Umpqua  River  where 
salmonid  species  of  concern  (cutthroat  trout  (occur  with  pro- 
tected marine  mammal  species.  The  harbor  seal  diet  in  the 
Umpqua  River  consisted  of  nonsalmonid  fish  and  chinook. 
coho,  and  steelhead;  no  cutthroat  trout  were  observed  in  the 
scat  samples  (Orr  et  al.,  2004).  The  majority  of  salmonid 
species  identifications  were  possible  only  by  using  genetic 
methods  because  very  few  otoliths  were  recovered  in  the 
Umpqua  River  scats.  A  number  of  other  sites  exist  were  this 
technology  may  also  be  applicable.  In  Hood  Canal  ( WA)  the 
summer  chum  salmon  run  is  listed  as  threatened  under  the 
ESA.  A  report  of  seal  diets  in  Hood  Canal  determined  that 
2T7c  of  the  fish  consumed  by  harbor  seals  were  salmonids 
(Jeffries  et  al.3).  The  study  used  both  bones  and  otoliths, 
but  only  25%  of  the  samples  contained  otoliths  that  allowed 
species-level  identification.  In  the  Alsea  River  (OR),  coho 
salmon  are  listed  as  threatened.  A  report  by  Riemer  et  al.4 
indicated  that  69r  of  fish  consumed  by  pinnipeds  in  the 
Alsea  River  are  salmonids;  none  of  the  salmonid  remains 
were  morphologically  identifiable  to  species. 

Extraction  of  DNA  from  bones  can  be  done  with  a  com- 
mercially available  kit  with  minor  modifications.  In  our 
study,  only  67%  of  the  bone  DNA  extracts  could  be  ampli- 
fied by  PCR.  PCR  failure  could  be  due  to  DNA  degradation 
during  the  digestive  process  or  to  environmental  exposure 
after  defecation.  However,  multiple  bones  are  often  present 
in  scats  and  we  were  able  to  amplify  DNA  from  at  least  one 
bone  representative  from  35  out  of  the  39  scats  examined. 
Sequencing  or  RFLP  analyses  of  the  COIII/ND3  locus  are 
both  viable  methods  of  identifying  the  seven  common  On- 
corhynchus  species.  This  study  used  manual  sequencing 
with  radioactivity  and  we  did  have  better  results  using 
this  method  compared  to  the  RFLP  method.  A  recently 
published  study  also  identified  restriction  enzymes  in  the 
cytochrome  B  gene  that  distinguish  among  the  salmonid 
species  (Russell  et  al.,  2000).  The  study  reported  diagnostic 
RFLP  differences  among  these  species  but  did  not  confirm 
the  lack  of  intraspecific  variation  in  a  wide  geographic  sur- 
vey of  each  species.  The  goal  of  the  cytochrome  B  RFLP  as- 
say designed  by  Russell  et  al.  (2000)  was  to  identify  salmon 
species  found  in  processed  food  products  but  the  primers 


3  Jeffries,  S.  J.,  J.  M.  London,  and  M.  M.  Lance.  2000.  Obser- 
vations of  harbor  seal  predation  on  Hood  Canal  summer  chum 
salmon  run  1998-1999.  Annual  progress  report  to  Pacific 
States  Marine  Fisheries  Commission,  39  p.  [Available  from 
WDFW,  Marine  Mammals  Investigations,  7801  Phillips  Rd.  SW, 
Tacoma,  WA  98498.] 

4  Riemer,  S.  D.,  R.  F.  Brown,  B.  E.  Wright  and  M.  I.  Dhruv. 
1999.  Monitoring  pinniped  predation  on  salmonids  at  Alsea 
River  and  Rogue  River,  Oregon:  1997-1999.  Oregon  Depart- 
ment of  Fish  and  Wildlife,  Marine  Mammal  Research  Program, 
Corvallis,  OR,  36  p.  [Available  from  ODFW,  7118  NE  Vanden- 
berg  Ave.,  Corvallis,  OR  97333.] 


may  also  prove  useful  in  species  identification  of  bone  re- 
mains. The  16s  primer  set  is  also  valuable  for  bones  that 
are  morphologically  unidentifiable.  However  for  salmonid 
species  identification,  the  16s  region  contains  fewer  diag- 
nostic nucleotide  substitutions  in  relation  to  the  d-loop  and 
the  COIII/ND3  region.  Overall,  the  techniques  established 
here  would  be  useful  for  further  study  of  marine  mammal 
diets  and  may  have  the  potential  for  forensic  application. 


Acknowledgments 

The  authors  acknowledge  Robert  Delong  for  suggesting 
this  study.  Jon  Baker  at  the  Northwest  Fisheries  Science 
Center  and  Paul  Spruell  at  the  University  of  Montana 
kindly  provided  cutthroat  DNA.  James  Shaklee  at  the 
Washington  Department  of  Fish  and  Wildlife  kindly 
provided  pink  salmon  samples.  Sam  Wasser  and  Virginia 
Butler  provided  advice  on  the  recovery  of  DNA  from  scat 
and  bone  samples.  Gail  Bastrup  assisted  in  technical 
aspects  of  this  study. 


Literature  cited 

Avise,  J.  C. 

1994.  Molecular  markers,  natural  history,  and  evolution, 
p.  44-91.     Chapman  and  Hall,  New  York,  NY. 

Cottrell,  P.  E..  A.  W  Trites,  and  E.  H.  Miller. 

1996.  Assessing  the  use  of  hard  parts  in  faeces  to  identify 
harbour  seal  prey:  results  of  captive-feeding  trials.  Can. 
J.  Zool.  74:875-880. 

Domanico,  M.  J.,  and  R.  B.  Phillips. 

1995.  Phylogenetic  analysis  of  Pacific  salmon  (genus 
Oncorhynchus)  based  on  mitochondrial  DNA  sequence 
data.    Mol.  Phylogenet.  Evol.  4:366-371. 

Harvey.  J.  T. 

1989.    Assessment  of  errors  associated  with  harbour  seal 
{Phoca  vitulina)  faecal  sampling.     J.  Zool.  219:101-111. 
Hochmeister,  M.  N,  B.  Budowle.  U.  V.  Borer,  U.  Eggmann, 
C.  T.  Comey,  and  R.  Dirnhofer. 

1991.    Typing  of  deoxyribonucleic  acid  (DNA)  extracted  from 
compact  bone  from  human  remains.     J.  Forensic  Sci.  36: 
1649-1661. 
Johnson,  O.,  M.  Ruckelshaus,  W.  Grant.  F.  Waknitz,  A.  Garrett, 
G  Bryant,  K.  Neely,  and  J.  Hard. 

1999.     Status  review  of  coastal  cutthroat  trout  from  Washing- 
ton, Oregon,  and  California.     NOAA  Tech.  Memo.  NMFS- 
NWFSC-37,  p  138-139. 
McKay,  S.  J.,  M.  J.  Smith,  and  R.  H.  Devlin. 

1997.  Polymerase  chain  reaction-based  species  identifica- 
tion of  salmon  and  coastal  trout  in  British  Columbia.  Mol. 
Mar.  Biol.  Biotechnol.  6:131-140. 

Ochoa-Acuna,  H.,  and  J.  M.  Francis. 

1995.     Spring  and  summer  prey  of  the  Juan  Fernandez  fur 
seal,  Arctocephalus philippii.     Can.  J.  Zool.  73:1444-1452. 
Olesiuk,  P.  F. 

1993.    Annual  prey  consumption  by  harbor  seals  (Phoca 
vitulina)  in  the  Strait  of  Georgia,  British  Columbia.     Fish. 
Bull.  91:491-515. 
Orr,  A.,  A.  Banks,  S.  Mellman,  and  H.  Huber. 

2004.  Examination  of  Pacific  harbor  seal  (Phoca  vitulina 
richardsi)  foraging  habits  to  describe  their  use  of  the 


220 


Fishery  Bulletin  102(1) 


Umpqua  River,  Oregon,  and  their  predation  on  salmonids. 
Fish.  Bull.l02:108-117. 
Parker,  A.,  and  I.  Kornfield. 

1996.    An  improved  amplification  and  sequencing  strategy 
for  phylogenetic  studies  using  the  mitochondrial  large  sub- 
unit  rRNA  gene.     Genome  39:793-797. 
Roffe,  T.,  and  B.  Mate. 

1984.    Abundance  and  feeding  habits  of  pinnipeds  in  the 
Rogue  River,  OR.    J.  Wildl.  Manag.  48:1262-1274. 
Russell,  V.  J.,  G.  L.  Hold,  S.  E.  Pryde,  H.  Rehbein,  J.  Quinteiro. 
M.  Rey-Mendez,  C.  G.  Sotelo,  R.  Perez-Martin,  A.  T.  Santos, 
and  C.  Rosa. 

2000.  Use  of  restriction  fragment  length  polymorphism  to 
distinguish  between  salmon  species.  J.  Agricult.  Food 
Chem.  48:2184-2188. 


Sambrook,  J.,  E.  F  Fritsch,  and  T.  Maniatis. 

1989.     Molecular  cloning:  a  laboratory  manual,  p.  9.17-9.19. 
Cold  Spring  Harbor  Laboratory.  Cold  Spring  Harbor.  NY. 
Shedlock,  A.  M.,  J.  D.  Parker,  D.A.  Crispin,  T.  W.  Pietsch,  and 
G.  C.  Burmer. 

1992.     Evolution  of  the  salmonid  mitochondrial  control 
region     Mol.  Phylogenet.  Evol.  1:179-192. 
Simons,  A.  M.,  and  R.  L.  Mayden. 

1998.  Phylogenetic  relationships  of  the  western  North  Amer- 
ican phoxinins  (Actinopterygii:  Cyprinidae  I  as  inferred  from 
mitochondrial  12S  and  16S  ribosomal  RNA  sequences. 
Mol.  Phylogenet.  Evol.  9:308-329. 


221 


Diel  vertical  migration  of  the 

bigeye  thresher  shark  (Alopias  superci/iosus), 

a  species  possessing  orbital  retia  mirabilia 

Kevin  C.  Weng 

Barbara  A.  Block 

Tuna  Research  and  Conservation  Center 
Hopkins  Marine  Station  of  Stanford  University 
120  Oceanview  Boulevard 
Pacific  Grove,  California  93950 

E-mail  address  (for  K.  C.  Weng):  kevin  cm  wengia'stanford  edu 


The  bigeye  thresher  shark  {Alopias 
superciliosus,  Lowe  1841)  is  one  of 
three  sharks  in  the  family  Alopiidae, 
which  occupy  pelagic,  neritic,  and 
shallow  coastal  waters  throughout  the 
tropics  and  subtropics  (Gruber  and 
Compagno,  1981;  Castro,  1983).  All 
thresher  sharks  possess  an  elongated 
upper  caudal  lobe,  and  the  bigeye 
thresher  shark  is  distinguished  from 
the  other  alopiid  sharks  by  its  large 
upward-looking  eyes  and  grooves 
on  the  top  of  the  head  (Bigelow  and 
Schroeder,  1948).  Our  present  under- 
standing of  the  bigeye  thresher  shark 
is  primarily  based  upon  data  derived 
from  specimens  captured  in  fisheries, 
including  knowledge  of  its  morpho- 
logical features  (Fitch  and  Craig,  1964; 
Stillwell  and  Casey,  1976;  Thorpe, 
1997),  geographic  range  as  far  as  it 
overlaps  with  fisheries  (Springer,  1943; 
Fitch  and  Craig,  1964;  Stillwell  and 
Casey,  1976;  Gruber  and  Compagno, 
1981;  Thorpe,  1997),  age,  growth  and 
maturity  (Chen  et  al.,  1997;  Liu  et  al., 
1998),  and  aspects  of  its  reproductive 
biology  (Gilmore,  1983;  Moreno  and 
Moron,  1992;  Chen  et  al..  1997). 

Limited  information  on  the  move- 
ment patterns  of  bigeye  thresher 
sharks  has  been  obtained  from  mark- 
recapture  studies  by  using  conven- 
tional tags.  The  longest  straight-line 
movement  of  a  conventionally  tagged 
bigeye  thresher  shark  to  date  is  2767 
km  from  waters  off  New  York  to  the 
eastern  Gulf  of  Mexico  (Kohler  and 
Turner,  2001).  The  bigeye  thresher 
shark  has  been  captured  on  longlines 
set  near  the  surface  at  night  (0  m  to  65 
m,  Fitch  and  Craig,  1964;  Stillwell  and 


Casey,  1976;  Thorpe,  1997;  Buencuerpo 
et  al.,  1998)  and  at  400  m  to  600  m 
during  the  day  (Nakamura1).  There 
is  no  published  information  available 
regarding  its  habitat  and  behavior,  al- 
though Francis  Carey  tracked  a  bigeye 
thresher  with  an  acoustic  tag  for  six 
hours  (Carey2). 

Endothermy  is  a  rare  trait  in  fishes 
and  has  been  demonstrated  only  in 
tunas  (Thunnini),  billfishes  (Xiphiidae, 
Istiophoridae),  and  lamnid  sharks 
(Lamnidael  (Carey  and  Teal,  1969; 
Carey,  1971,  1982a;  Block,  1991).  In  all 
endothermic  fishes,  the  blood  supply 
to  aerobic  tissues  such  as  slow-twitch 
swimming  muscle,  visceral  organs, 
extraocular  muscles,  retina,  and 
brain  occurs  by  counter-current  heat 
exchangers  known  as  retia  mirabilia. 
The  vascular  supply  reduces  heat  loss 
to  the  environment  and  enables  heat 
conservation  in  metabolically  active 
tissues  (Carey,  1971).  Lamnid  sharks 
have  retia  mirabilia  in  the  circulatory 
anatomy  supplying  the  slow-oxidative 
swimming  muscles,  viscera,  brain,  and 
eyes  (Burne,  1924;  Block  and  Carey, 
1985;  Tubbesing  and  Block,  2000).  In 
many  lamnid  species,  tissue  tempera- 
tures significantly  above  ambient  have 
been  recorded  from  freshly  captured 
specimens  and  through  telemetry  stud- 
ies of  swimming  animals  (Carey,  1971; 
Carey  et  al.,  1981,  1982,  1985;  McCos- 
ker,  1987;  Goldman,  1997;  Tubbesing 
and  Block,  2000). 

The  anatomy  of  alopiid  sharks  sug- 
gests that  endothermy  may  occur  in 
this  family.  The  bigeye  thresher  and  the 
common  thresher  (Alopias  vulpinus) 
have  centrally  located  slow-oxidative 


muscle  and  primitive  retia  mirabilia 
supplying  blood  to  them  (Carey,  1982b: 
Bone  and  Chubb,  1983).  Burne  (1924) 
noted  a  coiling  of  the  pseudobranchial 
artery  supplying  the  orbit  and  cranial 
regions  in  the  common  thresher.  No 
internal  tissue  temperature  measure- 
ments have  been  taken  for  free-swim- 
ming thresher  sharks  to  ascertain 
whether  heat  is  conserved  in  oxidative 
tissues.  A  freshly  caught  bigeye  thresh- 
er shark  was  found  to  have  a  body-core 
thermal  excess  of  4°C  (Carey,  1971); 
thus  the  species  may  have  the  ability 
to  conserve  metabolic  heat. 

In  this  study  we  present  electronic 
tagging  data  on  the  movements,  div- 
ing behavior,  and  habitat  preferences 
of  the  bigeye  thresher  shark  based  on 
two  individuals  studied  with  pop-up 
satellite  archival  tags.  In  addition, 
we  provide  a  brief  description  of  the 
orbital  rete  mirabile  of  the  species. 
The  presence  of  this  highly  developed 
rete  mirabile  within  the  orbital  sinus 
suggests  a  physiological  mechanism 
to  buffer  the  eyes  and  brain  from  the 
large  temperature  changes  associated 
with  diel  vertical  migration,  potentially 
conferring  enhanced  physiological  per- 
formance. 


Materials  and  methods 

The  movements  of  two  bigeye  thresher 
sharks  were  monitored  with  pop-up 
satellite  archival  tags  (PAT  tag  version 
2.00,  Wildlife  Computers,  Redmond, 
WA;  Gunn  and  Block,  2001;  Marcinek 
et  al.,  2001).  The  first  shark  was  cap- 
tured on  a  longline  set  in  the  Gulf  of 
Mexico  at  26.5°N,  91.3°W  on  12  April 


1  Nakamura.  I.  2002.  Personal  commun. 
Institut  National  des  Sciences  et  Technolo- 
gies de  la  Mer.  28  rue  2  Mars  1934,  2025 
Salammbo.  Tunisia. 

2  Carey.  F.  G.  (deceased).  1990.  Personal 
commun.  Woods  Hole  Oceanographic 
Institution,  Woods  Hole,  MA  02543. 


Manuscript  approved  for  publication 
15  August  2003  by  Scientific  Editor. 

Manuscript  received  20  October  2003 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:221-229  (2004). 


222 


Fishery  Bulletin  102(1) 


2000  in  waters  with  a  surface  temperature  of  21.9°C.  The 
longline  set  contained  184  hooks  set  at  depths  between 
70  m  and  90  m  and  was  made  at  06:00  h  and  retrieved  at 
09:00  h.  Circle  hooks  (L2045  20/0  circle  hook,  Eagle  Claw, 
Denver,  CO)  were  used  to  avoid  hooking  of  the  gut,  and 
the  shark  in  this  study  was  hooked  in  the  corner  of  the 
jaw.  Hooks  were  baited  with  squid,  and  chemical  light 
sticks  were  attached  to  every  other  line.  The  mass  of  the 
shark  was  visually  estimated  at  170  kg  by  an  experienced 
commercial  longline  fisherman,  which  corresponds  to  a 
fork  length  of  229  cm,  and  a  total  length  of  377  cm,  based 
on  the  weight-length  relationship  of  Kohler  et  al.  (1995). 
According  to  this  size  estimation  and  the  published  size-at- 
maturity  data  (Chen  et  al.,  1997;  Liu  et  al.,  1998),  the  shark 
was  mature.  The  sex  of  the  shark  was  not  determined.  The 
second  shark  was  captured  by  hook-and-line  gear  near 
Hawaii  at  19.5°N,  156.0°W  on  13  May  2003  in  waters  with  a 
surface  temperature  of  25.5°C.  A  baited  circle  hook  set  at  a 
depth  of  40  m  was  taken  by  the  shark  at  02:00  h.  The  mass 
of  the  shark  was  estimated  at  200  kg  by  an  experienced 
sportfishing  captain,  which  corresponds  to  a  fork  length  of 
242  cm,  and  a  total  length  of  400  cm  (after  Kohler  et  al., 
1995).  Given  this  size,  the  shark  was  mature  (Chen  et  al., 
1997;  Liu  et  al.,  1998),  but  its  sex  was  not  determined. 

Each  pop-up  satellite  archival  tag  was  attached  to  a  tita- 
nium dart  (59  mm  x  13  mm)  with  a  17  cm  segment  of  136- 
kg  monofilament  line  ( 300-lb  test  extra-hard  Hi-Catch,  Mo- 
moi  Fishing  Net  Mfg.  Co.  Ltd.,  Ako  City,  Hyogo  prefecture, 
Japan).  The  dart  was  inserted  into  the  dorsal  musculature 
of  the  shark  at  the  base  of  the  first  dorsal  fin,  such  that  the 
tag  trailed  behind  the  fin.  Following  attachment  of  each 
tag,  the  fishing  line  was  cut  near  the  hook  and  both  sharks 
swam  away  vigorously.  Tagging  locations  were  recorded  by 
using  the  vessel's  global  positioning  system.  After  the  Gulf 
of  Mexico  shark  was  tagged,  a  depth-temperature  recorder 
(ABT-1,  Alec  Electronics,  Kobe,  Japan)  was  used  to  deter- 
mine the  temperature-depth  profile  of  the  upper  200  m  of 
the  ocean  at  the  release  site,  at  a  resolution  of  1  m. 

The  pop-up  satellite  archival  tag  deployed  in  the  Gulf 
of  Mexico  was  programmed  to  collect  pressure  and  tem- 
perature data  at  two-minute  intervals,  which  the  on-board 
software  (PAT  software  version  1.06,  Wildlife  Computers, 
Redmond,  WA)  summarized  into  six-hour  bins.  This  version 
of  PAT  software  did  not  permit  light-based  geolocation.  The 
summary  data  for  each  time  interval  comprised  percentage 
distributions  of  time-at-depth  and  time-at-temperature, 
and  profiles  of  temperature-at-depth.  Temperature-depth 
profiles  for  this  generation  of  software  were  recorded  at 
intervals  by  measuring  a  single  temperature  at  depths  of 
0,  25,  50,  75, 100,  125, 150,  200,  250,  300,  350,  and  400  me- 
ters for  the  deepest  dive.  A  mean  temperature-depth  profile 
was  obtained  by  calculating  the  mean  temperature  at  each 
specified  depth  for  all  profiles  taken  during  the  track.  The 
endpoint  position  of  the  shark's  track  was  obtained  from 
the  tag's  radio  transmissions  to  the  Argos  satellites.  The 
six-hour  bins  were  later  combined  into  12-hour  bins  repre- 
senting day  (06:00  to  17:59  h  local  time)  and  night  ( 18:00  to 
05:59  h  local  time).  At  the  time  and  place  of  tag  deployment, 
sunrise  occurred  at  05:45  h  and  sunset  at  18:28  h;  whereas 
at  the  popup  time  and  position,  sunrise  occurred  at  05:02  h 


and  sunset  at  18:55  h  (U.S.  Naval  Observatory),  such  that 
the  day  and  night  bin  cutoffs  were  always  within  one  hour 
of  true  sunrise  and  sunset. 

The  pop-up  satellite  archival  tag  deployed  off  Hawaii  col- 
lected data  at  30-second  intervals  and  summarized  them 
into  four-hour  bins  (PAT  software  version  2.08e,  Wildlife 
Computers,  Redmond,  WA).  The  data  were  later  combined 
into  day  and  night  bins  as  for  the  first  tag,  and  the  actual 
sunrise  and  sunset  times  were  within  one  hour  of  06:00  h 
and  18:00  h,  respectively  (U.S.  Naval  Observatory).  The  tag 
measured  the  minimum  and  maximum  temperature  at  the 
surface,  maximum  depth,  and  six  intermediate  depths,  for 
the  deepest  dive  in  each  time  interval.  Temperature-depth 
profiles  for  each  time  interval  were  later  constructed  by  us- 
ing the  maximum  temperature  at  each  depth  for  all  profiles 
taken  during  the  track,  and  a  curve  was  fitted  by  using  a 
LOWESS  (locally  weighted  regression  smoothing)  function 
(Cleveland,  1992).  Version  2.08e  PAT  software  collected 
light  data  for  geolocation;  however  the  diel  dive  pattern  of 
the  shark  prevented  the  calculation  of  accurate  positions. 

The  vascular  circulation  to  the  brain  and  eyes  was  exam- 
ined in  two  bigeye  thresher  sharks:  one  common  thresher 
shark  and  one  pelagic  thresher  shark  iAlopias  pelagicus). 
A  female  bigeye  thresher  (1.5  m  fork  length)  was  captured 
off  Cape  Hatteras,  North  Carolina,  and  a  male  (1.4  m  fork 
length)  was  captured  in  the  Gulf  of  Mexico.  The  circula- 
tory systems  of  the  bigeye  threshers  were  injected  with 
latex  to  aid  in  identifying  the  blood  vessels.  A  male  com- 
mon thresher  (1.3  m  fork  length)  was  captured  off  Cape 
Hatteras,  North  Carolina,  and  was  examined  without 
being  frozen  or  preserved.  An  immature  female  pelagic 
thresher  shark  (1.37  m  fork  length)  was  captured  in  the 
Indian  Ocean.  The  orbital  retia  mirabilia  were  prepared 
from  casts  of  the  vascular  circulation  that  were  removed 
from  the  orbit. 


Results 

One  bigeye  thresher  shark  was  tracked  in  the  Gulf  of 
Mexico  for  60  days,  and  another  in  the  Hawaiian  Archi- 
pelago for  27  days,  by  using  pop-up  satellite  archival  tags. 
Both  tags  released  from  the  sharks  as  programmed  and 
transmitted  summary  information  to  Argos  satellites.  The 
tag  deployed  in  the  Gulf  of  Mexico  popped  up  on  10  June 
2000  at  27.95°N,  89.54°W  (Fig.  1A).  The  shark  moved  a 
straight-line  distance  of  320  km  during  the  track,  start- 
ing from  the  central  Gulf  in  depths  exceeding  3000  m  and 
moving  to  waters  150  km  south  of  the  Mississippi  Delta 
where  depths  were  approximately  1000  m.  The  second 
shark  was  tagged  off  the  Kona  coast  of  Hawaii  and  the  tag 
released  on  9  June  2003  at  24.2°N,  165.6°W.  northeast  of 
French  Frigate  Shoals,  a  straight-line  distance  of  1125  km 
from  the  deployment  position  (Fig.  IB). 

The  depth  and  temperature  distributions  of  the  bigeye 
thresher  sharks  showed  a  strong  diel  movement  pattern 
(Fig.  2).  The  Gulf  of  Mexico  shark  spent  the  majority  of 
the  daytime  (84f*  (±2.39H.  mean  [±1  SE])  below  the  ther- 
mocline  between  300  m  and  500  m  and  the  majority  of 
nighttime  (809?  [±4.7%],  mean  (±1  SE] )  in  the  mixed  layer 


NOTE     Weng  and  Block:  Diel  vertical  migration  in  Alopias  superaliosus 


223 


30°  N 


82°W 


24°N 


166 


164 


162 


160 


158 


156°W 


Figure  1 

Deployment  (A)  and  end-point  (•)  positions  for  the  two  pop-up  satellite  archival  tags 
attached  to  bigeye  thresher  sharks.  Both  tags  surfaced  on  the  programmed  dates  and 
transmitted  data  to  Argos  satellites.  Pressure  sensors  in  the  tags  confirmed  that  the  tags 
remained  attached  to  the  sharks  for  the  duration  of  the  tracks.  (A)  In  the  Gulf  of  Mexico  a 
shark  was  tagged  and  released  on  12  April  2000  and  the  tag  surfaced  on  10  June  2000.  The 
shark  moved  a  straight-line  distance  of  .320  km  during  the  60-day  track.  (B)  In  the  Hawaiian 
Archipelago  a  shark  was  tagged  on  13  May  2003  off  Kona,  Hawaii,  and  the  tag  surfaced  on 
9  June  2003  northeast  of  French  Frigate  Shoals.  The  shark  moved  a  straight-line  distance 
of  1125  km  during  the  27-day  track. 


and  upper  thermocline  between  10  m  and  100  m  (Fig.  2A). 
The  shark  spent  most  of  the  daytime  in  deeper  waters  of 
6°C  to  12°C  (70%  [±4.4%],  mean  [±1  SE]),  and  most  of  the 
nighttime  in  shallower  waters  from  20°C  to  26°C  (70% 
[±2.7%],  mean  [±1  SE])  (Fig.  2B).  A  temperature-depth 


profile  taken  by  the  tag  during  the  first  day  of  the  shark's 
track  closely  matched  a  profile  taken  from  the  vessel  with  a 
bathythermograph  (Fig.  3A).  The  mean  temperature-depth 
profile  for  the  60-day  track  (Fig.  3B),  when  compared  with 
the  shark's  depth  preferences  (Fig.  2A),  indicated  that 


224 


Fishery  Bulletin  102(1) 


Percent  time 
00  75     50    25      0     25     50    75    100 


0-5 

5-10 

10-50 

50-100 

~     100-150 


g     150-200 


200-250 
250-300 
300-500 
500-700 
700-1000 


,r 


Percent  time 
50        25         0         25        50        75 


45 


25 


Percent  time 
5  25 


28-30 
26-28 
24-26 
22-24 
20-22 
18-20 
16-18 
14-16 
10-14 
10-12 
6-10 
<6 


_i i   .   i i_ 


45 

j i 


Percent  time 
50      25      0        25      50      75 


Figure  2 

Depth  and  temperature  distributions  of  two  bigeye  thresher  sharks  showing  diel  vertical  migration. 
The  tags  recorded  depth  and  temperature  at  two-minute  (A,  B>  or  30-second  (C,  Di  intervals;  data 
are  summarized  into  a  series  of  bins  for  the  full  duration  of  each  track.  (Al  Depth  distribution  for  the 
Gulf  of  Mexico  shark  is  shown  as  the  percentage  of  day  (□>  and  night  ■  spent  within  depth  bins 
ranging  from  the  surface  to  1000  m.  Error  bars  are  1  SE.  (B)  Temperature  distribution  for  the  Gulf 
of  Mexico  shark  is  shown  as  the  percentage  of  day  (□)  and  night  ■  spent  within  temperature  bins 
ranging  from  6°C  to  30°C.  The  shark  occupied  cool  waters  during  the  day  and  warm  waters  during 
the  night,  a  consequence  of  its  deep  daytime  and  shallow  nighttime  habitats.  Error  bars  are  1  SE. 
(Cl  Depth  distribution  for  the  Hawaii  shark  showing  diel  vertical  migration.  The  shark  spent  most 
of  the  daytime  at  the  base  of  the  thermocline  and  must  of  the  nighttime  in  the  mixed  layer  and  upper 
thermocline.  iD)  Temperature  distribution  for  the  Hawaii  shark  showing  cool  daytime  and  warm 
nighttime  water  temperatures. 


NOTE     Weng  and  Block:  Diel  vertical  migration  in  Alopias  superaliosus 


225 


10  15  20  25 


Temperature  (C) 
10  15  20  25 


5         10         15         20         25        30 


50 

100-1 

-jT    150 
n 

Q_ 

q    200 -I 
250 

300  -I  c 

350 

Figure  3 

Temperature-depth  profiles  characterizing  the  thermal  habitat  of  two  bigeye  thresher  sharks.  (Al  Profiles  of  the 

Gulf  of  Mexico  taken  with  a  bathythermograph  ( )  sampling  at  1-m  intervals  deployed  from  the  fishing  vessel 

after  the  tagging  event,  and  by  the  pop-up  satellite  archival  tag  (O)  during  the  first  day  it  was  attached  to  the 
bigeye  thresher  shark.  The  two  profiles  are  similar,  indicating  that  the  pop-up  satellite  archival  tag  is  capable 
of  characterizing  thermal  habitat.  (B)  Average  temperature-depth  profile  for  the  60-day  track  of  the  bigeye 
thresher  shark  in  the  Gulf  of  Mexico,  showing  a  mixed  layer  shallower  than  50  m  and  a  thermocline  extending 
beyond  400  m  where  waters  were  10°C.  The  curve  was  fitted  by  using  a  LOWESS  function  and  error  bars  are 
1  SD,  because  1  SE  bars  are  invisible  at  this  scale.  (C)  Average  temperature-depth  profile  for  the  27-day  track  of 
the  bigeye  thresher  shark  in  the  Hawaiian  Archipelago,  showing  a  shallow  mixed  layer  a  thermocline  extending 
to  approximately  600  m  where  waters  were  6°C.  Curve  was  fitted  by  using  a  LOWESS  function  and  error  bars  are 
1  SD,  because  1  SE  bars  are  invisible  at  this  scale. 


the  shark  spent  most  of  the  daytime  below  the  maximum 
gradient  of  the  thermocline  where  temperatures  were  ap- 
proximately 10°C.  On  25  April  and  25  May  2000  the  shark 
spent  two  hours  of  the  day  in  waters  between  4°C  and  6°C. 
The  Hawaii  shark  showed  a  similar  diel  vertical  migration, 
with  a  lesser  contrast  between  day  and  night  (Fig.  2,  C  and 
D).  The  shark's  modal  nighttime  depth  was  between  10  m 
and  50  m,  whereas  its  modal  daytime  depth  was  between 
400  m  and  500  m  (Fig.  2C).  The  temperature-depth  profile 
for  the  Hawaii  shark  ( Fig.  3C )  indicated  that  it  spent  night- 
time above  the  thermocline  and  daytime  below  it. 

The  bigeye  thresher  shark  possesses  a  large  arterial 
plexus  between  the  posterior  part  of  the  eye  and  the  wall 
of  the  orbital  sinus,  which  appears  to  be  a  rete  mirabile 
(Fig.  4).  The  orbital  rete  is  bathed  in  venous  blood  from  the 
orbital  sinus  and  its  anterior  surface  is  contoured  to  the 
posterior  surface  of  the  eye.  The  sources  of  venous  input 
to  the  orbital  sinus  remain  unknown  but  are  most  likely 
within  the  surrounding  extraocular  muscles,  which  are 
large  and  comprise  numerous  aerobic  muscle  fiber  types, 
and  the  retina.  The  rete  shown  in  Figure  4  measures  72 
mm  by  49  mm  by  19  mm.  A  reduced  structure  of  similar 
form  is  also  found  in  the  pelagic  thresher  shark,  but  is  not 
present  in  the  common  thresher.  The  orbital  rete  of  the 
bigeye  and  pelagic  threshers  is  larger  in  absolute  size  and 


occupies  a  greater  cross  sectional  proportion  of  the  orbital 
sinus  than  the  lamnid  orbital  rete  noted  by  Burne  (1924). 
The  arterial  vessels  form  a  finer  and  more  orderly  mesh- 
work  than  those  in  the  lamnid  sharks  (Block  and  Carey, 
1985;  Tubbesing  and  Block,  2000)  and  appear  similar  in 
physical  structure  to  the  mammalian  carotid  rete  used  for 
brain  cooling  (Baker,  1982). 


Discussion 

Observations  of  the  biological  features  of  the  bigeye 
thresher  shark  are  rare  and  our  knowledge  of  the  species 
is  based  primarily  on  incidental  catches  in  fisheries.  Using 
pop-up  satellite  archival  tags  we  were  able  to  record  behav- 
ior for  a  total  of  87  days,  and  for  individual  periods  up  to  60 
days  without  recapturing  or  following  the  study  animals. 
We  observed  a  pronounced  diel  alternation  between  warm 
shallow  waters  and  cool  deep  waters  and  a  rete  mirabile 
that  may  confer  physiological  benefits  during  deep  dives  by 
stabilizing  brain  and  eye  temperatures. 

The  depth  data  obtained  for  the  bigeye  thresher  shark 
shows  a  striking  pattern  of  diel  vertical  migration.  The  big- 
eye thresher  shark's  vertical  movement  pattern  is  distinct 
from  those  of  most  other  sharks  for  which  observations 


226 


Fishery  Bulletin  102(1) 


Figure  4 

Orbital  rete  of  a  bigeye  thresher  shark,  showing  the  highly  developed  arterial  network.  The 
rete  was  injected  with  latex  so  that  the  arterial  structure  (72  mm  by  49  mm  by  19  mm)  could  be 
photographed.  The  structure  of  the  rete  and  its  position  in  the  orbital  sinus  suggest  that  it  may 
be  a  heat  exchanging  vascular  plexus.  Retention  of  metabolic  heat  in  the  eyes  and  brain  would 
buffer  these  sensitive  organs  from  the  large  ambient  temperature  swings  that  occur  as  a  result 
of  the  bigeye  thresher  shark's  diel  vertical  migrations.  A  smaller  but  similar  structure  is  found 
in  A.  pelagicus  but  not  in  A.  vulpinus. 


exist.  In  satellite  or  acoustic  tracks,  diel  vertical  migra- 
tion was  not  observed  for  white  sharks  (Carcharodon  car- 
charias;  Carey  et  al.,  1982;  Goldman  and  Anderson,  1999; 
Boustany  et  al.,  2002),  salmon  sharks  (Lamna  ditropis; 
Block  et  al.3),  shortfin  mako  (Isurits  oxyrhynchus;  Carey, 
1982b;  Holts  and  Bedford,  1993),  blue  (Prionace  glauca, 
Carey,  1982b;  Carey  and  Scharold,  1990),  sixgill  (Hexan- 
chus  griseus;  Carey  and  Clark,  1995),  tiger  (Galeocerdo 
cuvier;  Tricas  et  al.,  1981;  Holland  et  al.,  1999),  Pacific 
angel  (Squatina  californica;  Standora  and  Nelson,  1977), 
whale  [Rhincodon  typus;  Gunn  et  al.,  1999),  or  scalloped 
hammerhead  sharks  (Sphyrna  lewini;  Klimley,  1993). 

Diel  vertical  migration  has  been  observed  in  the  sword- 
fish  (Xiphias  gladius;  Carey  and  Robison,  1981;  Carey4), 
the  megamouth  shark  (Megachasma  pelagios;  Nelson  et 
al.,  1997),  and  the  school  shark  iGaleorhinus  ga/eus;  West 


Block,  B.A.,  K.G.Goldman,  and  J.  A.  Musick.  1999.  Unpubl. 
data.  Hopkins  Marine  Station  of  Stanford  University.  120 
Oceanview  Boulevard,  Pacific  Grove,  CA  93950. 
Carey,  R  G.  1990.  Further  acoustic  telemetry  observations  of 
swordfish.  In  Planning  the  future  of  billfishes;  proceedings  of 
the  second  international  billfish  symposium,  1-5  August  1988, 
Kailua-Kona,  Hawaii  (R.  H.  Stroud,  ed.),  p.  103-122.  National 
Coalition  for  Marine  Conservation,  3  North  King  St.,  Leesburg, 
VA  20176. 


and  Stevens,  2001).  Carey  and  Robison  ( 1981)  and  Carey4 
studied  swordfish  in  both  the  Pacific  and  Atlantic  Oceans, 
acoustically  tracking  fish  that  moved  from  the  surface  at 
night  to  over  600  m  during  day.  A  megamouth  shark  showed 
a  strong  diel  vertical  migration  when  tracked  acoustically 
off  southern  California  (Nelson  et  al.,  1997)  with  shallow 
nighttime  and  deep  daytime  distribution  in  a  vertical  range 
of  20  m  to  160  m.  West  and  Stevens  (2001)  studied  school 
sharks  in  southern  Australia  using  archival  tags  and  noted 
that  they  ascended  in  the  water  column  at  night. 

The  ambient  temperature  at  the  modal  day-  and  night- 
time depths  of  the  two  bigeye  thresher  sharks  differed  by 
15°  to  16°C,  requiring  them  to  be  eurythermal.  The  sharks 
spent  most  of  the  nighttime  in  shallow  waters  warmer 
than  20°C  and  commonly  spent  8  or  more  hours  during 
the  daytime  in  deep  waters  cooler  than  10°C.  The  coolest 
waters  occupied  had  temperatures  between  4°C  and  6°C. 
The  bigeye  thresher  sharks  tracked  in  our  study  spent  a 
higher  proportion  of  their  time  in  waters  below  10°C  than 
did  white  sharks  (Carey  et  al.,  1982;  Boustany  et  al.,  2002) 
and  mako  sharks  (Carey  and  Scharold,  1990;  Klimley  et 
al.,2002). 

The  presence  of  a  rete  mirabile  in  the  cranial  region 
may  indicate  a  mechanism  for  heat  conservation.  Heat 
conservation  in  the  brain  and  eyes  would  enable  the  big- 


NOTE     Weng  and  Block:  Diel  vertical  migration  in  Alopias  superaliosus 


227 


eye  thresher  shark  to  prolong  its  foraging  time  beneath 
the  thermocline,  as  we  observed  for  both  of  the  sharks 
tagged  in  our  study.  The  retina  and  brain  are  extremely 
temperature  sensitive  in  most  vertebrates  and  the  large 
changes  in  depth  and  temperature  recorded  would  impose 
significant  effects  on  the  biochemical  processes  occurring  in 
these  tissues  (Block  and  Carey,  1985;  Block,  1994).  Delayed 
responses  to  retinal  stimulation  can  be  caused  by  cooling, 
whereas  increased  noise  and  random  firing  of  neurons  can 
be  caused  by  warming — both  responses  having  adverse 
affects  on  sensory  function  (Konishi  and  Hickman,  1964; 
Friedlander  et  al.,  1976;  Prosser  and  Nelson,  1981). 

Anatomical  and  physiological  adaptations  to  warm  the 
brain  and  eyes  have  evolved  independently  in  divergent 
pelagic  fish  lineages,  including  the  lamnid  sharks  (Block 
and  Carey,  1985),  billfishes  of  the  Xiphiidae  and  Istiophori- 
dae  (Carey,  1982a;  Block.  1983)  and  some  scombrid  fishes 
(Linthicum  and  Carey,  1972).  A  cranial  rete  mirabile  also 
has  been  identified  in  mobulids  (Schweitzer  and  Notarbar- 
tolo  di  Sciara,  1986)  and  is  thought  to  be  a  heat  exchanger 
(Alexander,  1995,  1996).  Although  it  is  premature  to  sug- 
gest that  the  orbital  rete  of  the  bigeye  thresher  shark  is  a 
heat  exchanger  without  direct  evidence  of  elevated  tissue 
temperatures  in  the  brain  and  eyes,  the  structure  is  larger 
than  the  rete  mirabile  of  lamnid  sharks,  for  which  elevat- 
ed brain  and  eye  temperatures  have  been  demonstrated 
(Block  and  Carey,  1985).  The  anatomical  arrangement  of 
an  arterial  plexus  in  an  orbital  sinus  is  correlated  with 
heat  conservation  strategies  in  other  vertebrates  (Baker, 
1982).  The  phylogenetic  relationships  of  the  alopiid  and 
lamnid  sharks  (Compagno,  1990;  Naylor  et  al.,  1997)  sug- 
gest that  endothermic  traits  evolved  independently  in  the 
two  families. 

This  note  presents  new  information  on  the  depth  and 
ambient  temperature  preferences  of  the  bigeye  thresher 
shark  based  on  observations  of  two  individuals,  as  well  as 
the  anatomy  of  the  orbital  rete  mirabile,  which  appears  to 
function  as  a  vascular  heat  exchanger.  Behavior  of  many 
organisms  varies  with  ontogeny,  season  and  location; 
therefore  the  present  study  should  be  considered  as  only 
the  beginning  of  an  understanding  of  the  bigeye  thresher 
shark's  physical  habitat  preferences  and  adaptations  to 
temperature  change.  Further  studies  on  individuals  of 
different  sizes  and  in  different  regions  will  enhance  our 
understanding  of  the  behavior,  and  morphological  and 
physiological  adaptations,  of  the  bigeye  thresher  shark  to 
variations  in  temperature. 


Acknowledgments 

This  research  was  supported  by  grants  from  the  National 
Marine  Fisheries  Service,  the  National  Fish  and  Wildlife 
Federation  and  the  Packard  Foundation.  The  authors  wish 
to  thank  Captain  David  Price  and  crew  of  the  FV  Allison, 
and  Captain  John  Bagwell  and  crew  of  the  FY  Silky.  Shana 
Beemer  provided  scientific  assistance  on  the  cruise  and 
Captain  McGrew  Rice  assisted  in  tagging  and  releasing  the 
Gulf  of  Mexico  shark.  This  research  was  conducted  under 
Scientific  Research  Permit  TUNA-SRP-2000-002,  issued 


by  the  Office  of  Sustainable  Fisheries,  National  Marine 
Fisheries  Service,  Silver  Spring,  MD  20910. 


Literature  cited 

Alexander,  R.  L. 

1995.  Evidence  of  counter-current  heat  exchanger  in  the 
ray,  Mobu la  tarapacana  (Chondrichthyes:  Elasmobranchii: 
Batoidea:  Myliobatiformes).    J.  Zool.  (Lond)  237:377-384. 

1996.  Evidence  of  brain-warming  in  the  mobulid  rays, 
Mobula  tarapacana  and  Manta  birostris  (Chondrichthyes: 
Elasmobranchii:  Batoidea:  Myliobatiformes).  Zool.  J.  Linn. 
Soc.  118:151-164. 

Baker,  M.  A. 

1982.  Brain  cooling  in  endotherms  in  heat  and  exercise. 
Annu.  Rev.  Physiol.  44:85-96. 

Bigelow,  H.  B..  and  W.  C.  Schroeder. 

1948.     Sharks.     In  Fishes  of  the  western  North  Atlantic,  part 

one  (A.  Parr  and  Y.  Olsen,  eds),  p.  59-546.     Sears  Found. 

Mar.  Res.,  Yale  Univ.,  Mem.  1. 
Block,  B.  A. 

1983.  Brain  heaters  in  the  billfish.     Am.  Zool.  23:936. 
1991.     Endothermy  in  fish:  thermogenesis  ecology  and 

evolution.  In  Biochemistry  and  molecular  biology  of 
fishes.  Volume  1:  Phylogenetic  and  biochemical  perspec- 
tives (P.  Hochachka  and  T.  Mommsen,  eds.),  p.  269-311. 
Elsevier,  Amsterdam. 
1994.  Thermogenesis  in  muscle.  Annu.  Rev.  Physiol.  56: 
535-577 
Block,  B.  A.,  and  F.  G.  Carey. 

1985.     Warm  brain  and  eye  temperatures  in  sharks.     J. 
Comp.  Physiol.  B.  Biochem.  Syst.  Environ.  Physiol.  156: 
229-236. 
Bone,  Q.,  and  A.  D.  Chubb. 

1983.     The  retial  system  of  the  locomotor  muscles  in  the 
thresher  shark  tAlopias  uulpinus).     J.  Mar.  Biol.  Assoc. 
U.K.  63:239-242. 
Boustany,  A.  M.,  S.  F.  Davis,  P.  Pyle,  S.  D.  Anderson, 
B.  J.  Le  Boeuf,  and  B.  A.  Block. 

2002.     Satellite  tagging:  expanded  niche  for  white  sharks. 
Nature  415:35-36. 
Buencuerpo,  V.,  S.  Rios,  and  J.  Moron. 

1998.     Pelagic  sharks  associated  with  the  swordfish,  Xiphias 
gladius,  fishery  in  the  eastern  North  Atlantic  Ocean  and 
the  strait  of  Gibraltar.     Fish.  Bull.  96:667-685. 
Burne,  R.  H. 

1924.     Some  peculiarities  of  the  blood  vascular  system  of 
the  porbeagle  shark  (Lamna  comubica).     Philos.  Trans. 
R.  Soc.  Lond.  B.  Biol.  Sci.  212B:209-257. 
Carey,  F.  G. 

1971.    Warm  bodied  fish.  Am.  Zool.  11:135-143. 

1982a.  A  brain  heater  in  the  swordfish  {Xiphias  gladius). 

Science  216:1327-1329. 
1982b.    Warm  fish.    In  A  companion  to  animal  physiology; 
5th  international  conference  on  comparative  physiology; 
Sandbjerg,  Denmark,  July  22-26,  1980  (C.  R.  Taylor.  K. 
Johansen  and  L.  Bolis,  eds.),  p.  216-234.     Cambridge  Univ. 
Press,  Cambridge,  England;  New  York,  NY. 
Carey,  F.  G,  J.  G.  Casey,  H.  L.  Pratt,  D.  Urquhart,  J.  E.  McCosker, 
J.  A.  Seigel,  and  C.  C.  Swift. 

1985.  Temperature,  heat  production  and  heat  exchange  in 
lamnid  sharks.  In  Biology  of  the  white  shark  (J.  A.  Seigel 
and  C.  C.  Swift,  eds. ),  p.  92-108.  Mem.  South.  Calif.  Acad. 
Sci.,  vol.  9,  Fullerton,  CA. 


228 


Fishery  Bulletin  102(1) 


Carey,  F.  G.,  and  E.  Clark. 

1995.     Depth  telemetry  from  the  sixgill  shark,  Hexanchus 
griseus,  at  Bermuda.     Environ.  Biol.  Fishes  42:7-14. 
Carey,  F.  G.,  J.  W.  Kanwisher,  O.  Brazier.  G.  Gabrielson, 
J.  G.  Casey,  and  H.  L.  J.  Pratt. 

1982.  Temperature  and  activities  of  a  white  shark,  Carcha- 
rodon  carcharias.     Copeia  1982:254-260. 

Carey,  F.  G.,  and  B.  H.  Robison. 

1981.     Daily  patterns  in  the  activities  of  wwordfish.Xip/i/as 
gladius,  observed  by  acoustic  telemetry.     Fish.  Bull.  79: 
277-292. 
Carey,  F.  G,  and  J.  V.  Scharold. 

1990.     Movements  of  blue  sharks  iPrionaee  glauca )  in  depth 
and  course.     Mar.  Biol.  106:329-342. 
Carey,  F.  G,  and  J.  M.  Teal. 

1969.     Mako  and  porbeagle:  warm-bodied  sharks.     Comp. 
Biochem.  Physiol.  28:199-204. 
Carey,  F.  G,  J.  M.  Teal,  and  J.  W.  Kanwisher. 

1981.     The  visceral   temperatures   of  mackerel   sharks 
l  Lamnidae  I.     Physiol.  Zool.  54:334-344. 
Castro,  J.  I. 

1983.  The  sharks  of  North  American  waters,  180  p.     Texas 
A&M  Univ.  Press,  College  Station,  TX. 

Chen,  C.-T..  K.-M.  Liu,  and  Y.-C.  Chang. 

1997.     Reproductive  biology  of  the  bigeye  thresher  shark, 
Alopias  superciliosus  (Lowe,  1839)(Chondrichthyes:  Alo- 
piidae),  in  the  northwestern  Pacific.     Ichthyol.  Res.  44: 
227-235. 
Cleveland,  W.  S.,  E.  Grosse,  and  W.  M.  Shyu. 

1992.     Local  regression  models.     Chapter  8  in  Statistical 
models  in  S  (J.  M.  Chambers  and  T.  J.  Hastie  eds.),  608  p. 
Wadsworth  &  Brooks/Cole,  Pacific  Grove,  CA. 
Compagno,  L.  J.  V. 

1990.  Relationships  of  the  megamouth  shark,  Megachasma 
pelagios  (Lamniformes:  Megachasmidae).  with  comments 
on  its  feeding  behavior.  In  Elasmobranchs  as  living 
resources:  advances  in  the  biology,  ecology,  systematics, 
and  the  status  of  the  fisheries  (H.  L.  J.  Pratt,  S.  H.  Gruber, 
and  T  Taniuchi,  eds.),  p.  357-379.  NOAA  Technical 
Report  90. 
Fitch,  J.  E.,  and  W.  L.  Craig. 

1964.     First  records  for  the  bigeye  thresher  (Alopias  super- 
ciliosus) and  slender  tuna  (Allothunnus  fallal)  from 
California,  with  notes  on  eastern  Pacific  scombrid  oto- 
liths.    Calif.  Fish  Game  50:195-206. 
Friedlander,  M.  J.,  N.  Kotchabhakdi,  and  C.  L.  Prosser. 

1976.     Effects  of  cold  and  heat  on  behavior  and  cerebellar 
function  in  goldfish.     J.  Comp.  Physiol.  A  Sens.  Neural. 
Behav.  Physiol.  112:19-45. 
Gilmore,  R.  G. 

1983.     Observations  on  the  embryos  of  the  longfin  mako, 
Isurus  paucus  and  the  bigeye  thresher,  Alopias  super- 
ciliosus.   Copeia  1983:375-382. 
Goldman.  K.  J. 

1997.     Regulation  of  body  temperature  in  the  white  shark. 
Carcharodon  carcharias.    J.  Comp.  Physiol.  B.  Biochem. 
Syst.  Environ.  Physiol.  167:423-429. 
Goldman,  K.  G,  and  S.  D.  Anderson. 

1999.     Space  utilization  and  swimming  depth  of  white 
sharks,  Carcharodon  can  lianas,  at  the  South  Farallon  Is- 
lands, central  California.     Environ.  Biol.  Fishes  56:351- 
364. 
Gruber,  S.  H.,  and  L.  J.  V.  Compagno. 

1981.    Taxonomic  status  and  biology  of  the  bigeye  thresher, 
Alopias  superciliosus.     Fish.  Bull.  79:617-640. 


Gunn,  J.  S.,  J.  D.  Stevens,  T.  L.  O.  Davis,  and  B.  M.  Norman. 

1999.  Observations  on  the  short-term  movements  and 
behaviour  of  whale  sharks  tRhincodon  typus)  at  Ningaloo 
Reef,  Western  Australia.     Mar.  Biol.  135:553-559. 

Gunn,  J.  S„  and  B.  A.  Block. 

2001.  Advances  in  acoustic,  archival  and  satellite  tagging 
of  tunas.  In  Tunas:  physiology,  ecology  and  evolution  (B. 
A.  Block  and  E.  D.  Stevens,  eds.),  p.  167-224.  Academic 
Press,  San  Diego,  CA. 

Holland,  K.  N,  B.  M.  Wetherbee,  C.  G.  Lowe,  and  C.  G.  Meyer. 

1999.     Movements  of  tiger  sharks  (Galeocerdo  cuvier)  in 
coastal  Hawaiian  waters.     Mar.  Biol.  134:665-673. 
Holts,  D.  B..  and  D.  W.  Bedford. 

1993.     Horizontal  and  vertical  movements  of  the  shortfin 
mako  shark,  Isurus  oxyrinchus,  in  the  southern  California 
bight.    Aust.  J.  Mar.  Freshw.  Res.  44:901-909. 
Klimley,  A.  P. 

1993.     Highly  directional  swimming  by  scalloped  hammer- 
head sharks,  Sphyrna  lewini,  and  subsurface  irradiance. 
temperature,  bathymetry,  and  geomagnetic  field.     Mar. 
Biol.  117:1-22. 
Klimley,  A.  P.,  S.  C.  Beavers,  T.  H.  Curtis,  and  S.  J.  Jorgensen. 

2002.  Movements  and  swimming  behavior  of  three  species 
of  sharks  in  La  Jolla  Canyon,  California.  Environ.  Biol. 
Fishes  63:117-135. 

Kohler,  N.  E.,  J.  G.  Casey,  and  P.  A.  Turner. 

1995.     Length-weight  relationships  for  13  species  of  sharks 
from  the  western  North  Atlantic.    Fish.  Bull.  93:412-418. 
Kohler,  N.  E.,  and  P.  A.  Turner. 

2001.     Shark  tagging:  A  review  of  conventional  methods  and 
studies.     Environ.  Biol.  Fishes  60:191-223. 
Konishi,  J.,  and  C.  P.  Hickman. 

1964.     Temperature  acclimation  in  the  central  nervous 
system  of  rainbow  trout  [Salmo  gairdnerii).     Comp.  Bio- 
chem. Physiol.  13:433-442. 
Linthicum,  D.  S.,  and  F.  G.  Carey. 

1972.     Regulation  of  brain  and  eye  temperatures  by  the 
bluefin  tuna.     Comp.  Biochem.  Physiol.  A  Comp.  Physiol. 
43:425-433. 
Liu.  K.-M..  P. -J.  Chiang,  and  C.-T  Chen. 

1998.    Age  and  growth  estimates  of  the  bigeye  thresher 
shark,  Alopias  superciliosus.  in  northeastern  Taiwan 
waters.     Fish.  Bull.  96:482-491. 
Marcinek,  D.  J.,  S.  B.  Blackwell,  H.  Dewar.  E.  V.  Freund. 
C.  Farwell,  D.  Dau,  A.  C.  Seitz.  and  B.  A.  Block. 

2001.     Depth  and  muscle  temperature  of  Pacific  bluefin  tuna 
examined  with  acoustic  and  pop-up  satellite  archival  tags. 
Mar.  Biol.  138:869-885. 
McCosker,  J.  E. 

1987.    The  white  shark,  Carcharodon  carcharias.  has  a  warm 
stomach.     Copeia  1987:195-197. 
Moreno,  J.  A.,  and  J.  Moron. 

1992.     Reproductive  biology  of  the  bigeye  thresher  shark 
Alopias  superciliosus  Lowe  1839.    Aust.  J.  Mar.  Freshat. 
Res.  43:77-86. 
Naylor,  G.  J.  P.,  A.  P.  Martin,  E.  G.  Mattison,  and  W.  M.  Brown. 
1997.     Interrelationships  of  lamniform  sharks:  Testing  phv- 
logenetic  hypotheses  with  sequence  data.     In  Molecular 
systematics  of  fishes  (T.  D.  Kocher  and  C.  A.  Stepien,  eds.), 
p.  199-218.    Academic  Press,  San  Diego,  CA. 
Nelson.  D.  R.,  J.  N.  MeKibben.  W.  R.  Strong  Jr.,  C.  G.  Lowe, 
J.  A.  Sisneros,  D.  M.  Schroeder.  and  R.  J.  Lavenberg. 

1997.  An  acoustic  tracking  of  a  megamouth  shark,  Mega- 
chasma pelagios:  A  crepuscular  vertical  migrator.  Envi- 
ron Biol.  Fishes  49:389-399. 


NOTE     Weng  and  Block:  Diel  vertical  migration  in  Alopias  superahosus 


229 


Prosser,  C.  L.,  and  D.  O.  Nelson. 

1981.     Role  of  nervous  systems  in  temperature  adaptation  of 
poikilotherms.    Annu.  Rev.  Physiol.  43:281-300. 
Schweitzer.  J.,  and  G.  Notarbartolo  di  Sciara. 

1986.     The  rete  nurabile  cranica  in  the  genus  Mobula:  a  com- 
parative study.    J.  Morphol.  188:167-178. 
Springer,  S. 

1943.     A  second  species  of  thresher  shark  from  Florida. 
Copeia  1943:54-55. 
Standora,  E.  A.,  and  D.  R.  Nelson. 

1977.    A  telemetric  study  of  the  behavior  of  free  swimming 
Pacific  angel  sharks  Squatina  californica.     Bull.  South. 
Calif.  Acad.  Sci.  76:193-201. 
Stillwell,  C.  E.,  and  J.  G.  Casey. 

1976.  Observations  on  the  bigeye  thresher  shark,  Alopias 
superciliosus,  in  the  western  North  Atlantic.  Fish.  Bull. 
74:221-225. 


Thorpe,  T. 

1997.     First  occurrence  and  new  length  record  for  the  bigeye 
thresher  shark  in  the  northeast  Atlantic.     J.  Fish  Biol.  50: 
222-224. 
Tricas,  T.  C.  L.  R.  Taylor,  and  G.  Naftel. 

1981.     Diel  behavior  of  the  tiger  shark  Galeocerdo  cuvier  at 
French  Frigate  Shoals  Hawaiian  Islands  USA.     Copeia 
1981:904-908. 
Tubbesing,  V.  A.,  and  B.  A.  Block. 

2000.  Orbital  rete  and  red  muscle  vein  anatomy  indicate 
a  high  degree  of  endothermy  in  the  brain  and  eye  of  the 
salmon  shark.    Acta  Zool.  (Stockh.)  81:49-56. 

West,  G.  J.,  and  J.  D.  Stevens. 

2001.  Archival  tagging  of  school  shark,  Galeorhinus  galeus. 
in  Australia:  Initial  results.  Environ.  Biol.  Fishes  60: 
283-298. 


Fishery  Bulletin  102(1) 


231 


Superintendent  of  Documents  Publications  Order  Form 

*5178 

I I  YrLo,  please  send  me  the  following  publicatic 


tions: 


Subscriptions  to  Fishery  Bulletin 
for  $55.00  per  year  ($68.75  foreign) 


The  total  cost  of  my  order  is  $ 


.  Prices  include  regular  domestic 


postage  and  handling  and  are  subject  to  change. 


(Company  or  Personal  Name) 


(Please  type  or  print) 


(Additional  address/attention  line) 


(Street  address) 


(City,  State,  ZIP  Code) 


(Daytime  phone  including  area  code) 


(Purchase  Order  No.  I 


Charge 
your 
order. 

IT'S 
EASY! 


Please  Choose  Method  of  Payment: 

]  Check  Payable  to  the  Superintendent  of  Documents 
]  GPO  Deposit  Account 


]  VISA  or  MasterCard  Account 


(Credit  card  expiration  date) 


-□ 


To  fax 

your  orders 

(202)  512-2250 


(Authorizing  Signature) 

Mail  To:    Superintendent  of  Documents 

P.O.  Box  371954,  Pittsburgh,  PA  15250-7954 


Thank  you  for 
your  order! 


UH    nxB    H 


Fishery  Bulletin 

Guidelines  for  contributors 


Content  of  papers 

Articles 

Articles  are  reports  of  10  to  30  pages  (double 
spaced)  that  describe  original  research  in  one  or 
a  combination  of  the  following  fields  of  marine 
science:  taxonomy,  biology,  genetics,  mathematics 
(including  modeling),  statistics,  engineering,  eco- 
nomics, and  ecology. 

Notes 

Notes  are  reports  of  5  to  10  pages  without  an 
abstract  that  describe  methods  and  results  not 
supported  by  a  large  body  of  data.  Although  all 
contributions  are  subject  to  peer  review,  responsi- 
bility for  the  contents  of  articles  and  notes  rests 
upon  the  authors  and  not  upon  the  editor  or  the 
publisher.  It  is  therefore  important  that  authors 
consider  the  contents  of  their  manuscripts  care- 
fully. Submission  of  an  article  is  un-derstood  to 
imply  that  the  article  is  original  and  is  not  being 
considered  for  publication  elsewhere.  Manuscripts 
must  be  written  in  English.  Authors  whose  native 
language  is  not  English  are  strongly  advised  to 
have  their  manuscripts  checked  for  fluency  by 
English-speaking  colleagues  prior  to  submission. 

Preparation  of  papers 
Text 

Title  page  should  include  authors'  full  names  and 
mailing  addresses  (street  address  required)  and 
the  senior  author's  telephone,  fax  number,  e-mail 
address,  as  well  as  a  list  of  key  words  to  describe  the 
contents  of  the  manuscript.  Abstract  must  be  less 
than  one  typed  page  (double  spaced)  and  must  not 
contain  any  citations.  It  should  state  the  main  scope 
of  the  research  but  emphasize  the  author's  con- 
clusions and  relevant  findings.  Because  abstracts 
are  circulated  by  abstracting  agencies,  it  is  impor- 
tant that  they  represent  the  research  clearly  and 
concisely.  General  text  must  be  typed  in  double- 
spaced  format.  A  brief  introduction  should  state  the 
broad  significance  of  the  paper;  the  remainder  of 
the  paper  should  be  divided  into  the  following  sec- 
tions: Materials  and  methods.  Results,  Discussion 
lor  Conclusions),  and  Acknowledgments.  Headings 
within  each  section  must  be  short,  reflect  a  logical 
sequence,  and  follow  the  rules  of  multiple  subdi- 
vision (i.e.  there  can  be  no  subdivision  without  at 
least  two  subheadings).  The  entire  text  should  be 
intelligible  to  interdisciplinary  readers;  therefore, 
all  acronyms  and  abbreviations  should  be  written 
out  and  all  lesser-known  technical  terms  should  be 
defined  the  first  time  they  are  mentioned.  The 
scientific  names  of  species  must  be  written  out  the 
first  time  they  are  mentioned;  subsequent  mention 
of  scientific  names  may  be  abbreviated.  Follow  Sci- 
entific style  and  format;  CBE  manual  for  authors, 
editors,  and  publishers  (6th  ed.)  for  editorial  style 
and  the  most  current  issue  of  the  American  Fish- 
enes  Society's  common  and  scientific  names  of 
fishes  from  the  United  States  and  Canada  for 
fish  nomenclature.  Dates  should  be  written  as  fol- 
lows: 11  November  1991.  Measurements  should  be 
expressed  in  metric  units,  e.g.  metric  tons  (t).  The 
numeral  one  ( 1)  should  be  typed  as  a  one,  not  as  a 
lower-case  el  1 1 1. 

Footnotes 

Use  footnotes  to  add  editorial  comments  regarding 
claims  made  in  the  text  and  to  document  unpub- 


lished works  or  works  with  local  circulation.  Foot- 
notes should  be  numbered  with  Arabic  numerals 
and  inserted  in  10-point  font  at  the  bottom  of  the 
first  page  on  which  they  are  cited.  Footnotes  should 
be  formatted  in  the  same  manner  as  citations. 
If  a  manuscript  is  unpublished,  in  the  process 
of  review,  or  if  the  information  provided  in  the 
footnote  has  been  conveyed  verbally,  please  state 
this  information  as  "unpubl.  data."  "manuscript 
in  review,"  and  "personal  commun.,"  respectively. 
Authors  are  advised  wherever  possible  to  avoid  ref- 
erences to  nonstandard  literature  (unpublished  lit- 
erature that  is  difficult  to  obtain,  such  as  internal 
reports,  processed  reports,  administrative  reports, 
ICES  council  minutes.  IWC  minutes  or  working 
papers,  any  "research"  or  "working"  documents, 
laboratory  reports,  contract  reports,  and  manu- 
scripts in  review)  If  these  references  are  used, 
please  indicate  whether  they  are  available  from 
NTIS  (National  Technical  Information  Service)  or 
from  some  other  public  depository.  Footnote  format: 
author  (last  name,  followed  by  first-name  initials); 
year;  title  of  report  or  manuscript;  type  of  report 
and  its  administrative  or  serial  number;  name  and 
address  of  agency  or  institution  where  the  report  is 
filed. 

Literature  cited 

The  literature  cited  section  comprises,  won 
have  been  published  and  those  accepted  for  pub- 
lication (works  in  press)  in  peer- reviewed  jour- 
nals and  books.  Follow  the  name  and  year 
for  citation  format.  In  the  text,  write  "Smith  and 
Jones  (1977)  reported"  but  if  the  citation  takes 
the  form  of  parenthetical  matter,  write  "(Smith 
and  Jones,  1977)."  In  the  literature  cited  section, 
list  citations  alphabetically  by  last  name  of  senior 
author:  For  example.  Alston,  1952;  Mannly,  1988; 
Smith,  1932;  Smith,  1947;  Stalinsky  and  Jones, 
1985.  Abbreviations  of  journals  should  conform 
to  the  abbreviations  given  in  the  Serial 
for  the  BIOSIS  previews  database.  Authors  are 
responsible  for  the  accuracy  and  completeness  of 
all  citations.  Literature  citation  format:  author 
(last  name,  followed  by  first -name  initials 
title  of  report  or  article;  abbreviated  title  of  the 
journal  in  which  the  article  was  published,  volume 
number,  page  numbers.  For  books,  please  provide 
publisher,  city,  and  state 

Tables 

Tables  should  not  be  excessive  in  size  and  must  be 
cited  in  numerical  order  in  the  text.  Headings  in 
tables  should  be  short  but  ample  enough  to  allow 
the  table  to  be  intelligible  on  its  own.  All  unusual 
symbols  must  be  explained  in  the  table  legend. 
Other  incidental  comments  may  be  footnoted  (use 
italic  arabic  numerals  for  footnote  marker 
asterisks  only  to  indicate  probability  in  statistical 
data.  Place  table  legends  on  the  same  page  as  the 
table  data.  We  accept  tables  saved  in  most  spread- 
sheet software  programs  (e.g.  Microsoft  Excel'. 
Please  note  the  following: 

•  Use  a  comma  in  numbers  of  five  digits  or  more 
(e.g.  13,000  but  3000). 

•  Use  zeros  before  all  decimal  points  for  values 
less  than  one  (e.g.  0.31). 

Figures 

Figures  include  line  illustrations,  computer-gener- 
ated line  graphs,  and  photographs  (or  slides).  They 


must  be  cited  in  numerical  order  in  the  text.  Line 
illustrations  are  best  submitted  as  original  draw- 
ings. Computer-generated  line  graphs  should  be 
printed  on  laser-quality  paper.  Photographs  should 
be  submitted  on  glossy  paper  with  good  contrast. 
All  figures  are  to  be  labeled  with  senior  author's 
name  and  the  number  of  the  figure  (e.g.  Smith, 
Fig.  4).  LTse  Helvetica  or  Arial  font  to  lain 
tomical  parts  (line  drawings)  or  variables  (graphs) 
within  figures:  use  Times  Roman  bold  font  to  label 
the  different  sections  of  a  figure  (e.g.  A,  B,  C>. 
Figure  legends  should  explain  all  symbols  and 
abbreviations  seen  within  the  figure  and  should  be 
typed  in  double-spaced  format  on  a  separate  page 
at  the  end  of  the  manuscript.  We  advise  authors  to 
peruse  a  recent  issue  of  Fishery  Bulletin  for  stan- 
dard formats.  Please  note  the  following: 

•Capitalize  the  first  letter  of  the  first  word  of 
axis  labels. 

•  Do  not  use  overly  large  font  sizes  to  label  axes 
or  parts  within  figures. 

•  Do  not  use  boldface  fonts  within  figures. 

•  Do  not  create  outline  rules  around  graphs. 
•Do  not  use  horizontal  lines  through  graphs. 

•  Do  not  use  large  font  sizes  to  label  degrees  of 
longitude  and  latitude  on  maps. 

•  Indicate  direction  of  degrees  longitude  and 
latitude  on  maps  (e.g.  170°E). 

•Avoid  placing  labels  on  a  vertical  plane 

(except  on  ) 
•Avoid  odd  (nonstandard)  patterns  to  mark 
I  ions  of  bar  graphs  and  pie  charts. 

Copyright  law 

Fishery  Bulletin,  a  U.S.  government  publication,  is 
not  subject  to  copyright  law.  [fan  author  wishes  to 
reproduce  any  part  of  Fishery  Bulletin  in  his  or  her 
work,  he  or  she  is  obliged,  however,  to  acknowledge 
the  source  of  the  extracted  literature. 

Submission  of  papers 

Send  four  printed  copies  (one  original  plus  three 
copies) — clipped,  not  stapled — to  the  Scientific  Edi- 
tor, at  the  address  shown  below.  Send  photocopies 
of  figures  with  initial  submission  of  manuscript. 
Original  figures  will  be  requested  later  when  the 
manuscript  has  been  accepted  for  publication. 
Do  not  send  your  manuscript  on  diskette  until 
requested  to  do  SO. 

Dr.  Norman  Bartoo 

National  Marine  Fisheries  Service,  NOAA 

8604  La  Jolla  Shores  Drive 

La  Jolla.CA  92037 

Once  the  manuscript  has  been  accepted  for  pub- 
lication, you  will  be  asked  to  submit  a  sofl 
copy  of  your  manuscript.  The  software  copy  should 
be  submitted  in  WordPerfect  or  Word  form 
Word,  save  as  Rich  Text  Format'.  Please  note  that 
we  do  not  accept  ASCII  text  files. 

Reprints 

Copies  of  published  articles  and  notes  are  avail- 
able free  of  charge  to  the  senior  author  (50  i 
and  to  his  or  her  laboratory  (50  copies).  Additional 
copies  may  be  purchased  in  lots  of  100  when  the 
author  receives  page  proofs. 


«>^v-»zr  >-^:  a  ii :  \l*tX>isr 


->>v-»sr 


Itv^^f  f*^^^  tv^2*f  f^^^Wp 

U.S.  Department  ■■#  I  _ 

of  Commerce  Mam  II  tf^  m^k  £  ^  B^  ^  ^ 

Volume  102  ■       1^1    1  ^T  ■       W 

Number  2  ^^^  H  H  A 

Bulletin 

►  ^i^ga  ^f^fel^  ^5^n )  ^^&J>  j®c 

WAY ^  _ 


U.S.  Department 
of  Commerce 

Donald  L  Evans 
Secretary 


National  Oceanic 
and  Atmospheric 
Administration 

Vice  Admiral 

Conrad  C.  Lautenbacher  Jr., 

USN  (ret.) 

Under  Secretary  for 
Oceans  and  Atmosphere 


National  Marine 
Fisheries  Service 

William  T.  Hogarth 

Assistant  Administrator 
for  Fisheries 


^tf"0^, 


srATES  0f 


The  Fishery  Bullet  in  (ISSN  0090-0656) 
is  published  quarterly  by  the  Scientific 
Publications  Office,  National  Marine  Fish- 
eries Service.  NOAA.  7600  Sand  Point  Way 
NE,  BIN  C15700,Seattle,WA98115-0070. 
icals  postage  is  paid  at  Seattle,  WA, 
and  at  additional  mailing  offices.  POST- 
MASTER: Send  address  changes  for  sub- 
scriptions to  Fishery  Bulletin,  Superin- 
tendent of  Documents,  Attn.:  Chief.  Mail 
List  Branch,  Mail  Stop  SSOM.  Washing- 
ton. DC  20402-9 

Although  the  contents  of  this  publica- 
tion have  not  been  copyrighted  and  may 
be  reprinted  entirely,  reference  to  source 
is  appreciated 

The  Secretary  of  Commerce  lias  deter- 
mined that   the  publication  of  this  pen 
ording  to  law  for 
insaction  of pubh  of  this 

Department.  Use  of  funds  for  printing  of 
this  periodical  has  been  approved  by  the 
ir  of  the  l  Mine  of  Management  and 
Budget. 

For  salt-  by  the  Superintendent  of 
Documents.  rnment    Printing 

Office.  Washington,  DC  20402.  Subscrip- 
tion price  per  year:  $55.00  domestic  and 
foreign.   Cost    per   aingle    issue: 
$28.00  domi  gn.  See 

back  for  order  form. 


Scientific  Editor 

Dr.  Norman  Bartoo 

Associate  Editor 

Sarah  Shoffler 

National  Marine  Fisheries  Service,  NOAA 
8604  La  Jolla  Shores  Drive 
La  Jolla,  California  92037 

Managing  Editor 

Sharyn  Matriotti 

National  Marine  Fisheries  Service 
Scientific  Publications  Office 
7600  Sand  Point  Way  NE,  BIN  C15700 
Seattle,  Washington  98115-0070 


Editorial  Committee 

Dr.  Harlyn  O.  Halvorson 
Dr.  Ronald  W.  Hardy 
Dr.  Richard  D.  Methot 
Dr.  Theodore  W.  Pietsch 
Dr.  Joseph  E.  Powers 
Dr.  Harald  Rosenthal 
Dr.  Fredric  M.  Serchuk 
Dr.  George  Waiters 


University  of  Massachusetts,  Boston 
University  of  Idaho,  Hagerman 
National  Marine  Fisheries  Service 
University  of  Washington,  Seattle 
National  Marine  Fisheries  Service 
Universitat  Kiel,  Germany 
National  Marine  Fisheries  Service 
National  Marine  Fisheries  Service 


Fishery  Bulletin  web  site:  www.fishbull.noaa.gov 


The  Fishery  Bulletin  carries  original  research  reports  and  technical  notes  on  investigations  in 
fishery  science,  engineering,  and  economics.  It  began  as  the  Bulletin  of  the  United  Stat. 
Commission  in  1881    i    be  one  the  Bulletin  of  the  Bureau  of  in  1904  and  the  Fishery 

Bulletin  of  the  Fish  and  Wildlife  Service  in  1941.  Separates  were  issued  as  documents  through 
volume  46;  the  last  document  was  No.  1103.  Beginning  with  volume  17  in  19:!1  and  continuing 
through  volume  62  in   1963,  each  separate'  appeared  as  a  numbered  bulletin.  A  new   system 
began  in    1963  with  volume  63  in  which  papers  arc  bound  together  in  a  single  issue  of  the 
bulletin.  Beginning  with  volume  70,  number  1.  January  1972,  the  Fishery  Bulletin  bee 
quarterly.  In  this  form,  it  is  available  by  subscription  from  the  Superinte 
of  Documents,  U.S.  <  kn  eminent  Printing  ( illice.  Washington.  DC  20402.  It  is  also  available 
limited  numbers  to  libraries,  research  institutions,  Stale  and  Federal  agencies,  and  in  exchange 
for  other  scientific  publicat  ii 


U.S.  Department 
of  Commerce 

Seattle,  Washington 

Volume  102 
Number  2 
April  2004 


Fishery 
Bulletin 


Contents 


The  conclusions  and  opinions  expressed 
in  Fishery  Bulletin  are  solely  those  of  the 
authors  and  do  not  represent  the  official 
position  of  the  National  Marine  Fisher- 
ies Service  (NOAAi  or  any  other  agency 
or  institution. 

The  National  Marine  Fisheries  Service 
iNMFS)  does  not  approve,  recommend,  or 
endorse  any  proprietary  product  or  pro- 
prietary material  mentioned  in  this  pub- 
lication. No  reference  shall  be  made  to 
NMFS,  or  to  this  publication  furnished  by 
NMFS.  in  any  advertising  or  sales  pro- 
motion which  would  indicate  or  imply 
that  NMFS  approves,  recommends,  or 
endorses  any  proprietary  product  or  pro- 
prietary material  mentioned  herein,  or 
which  has  as  its  purpose  an  intent  to 
cause  directly  or  indirectly  the  advertised 
product  to  be  used  or  purchased  because 
of  this  NMFS  publication. 


Articles 

233-244  Archer,  Frederick,  Tim  Gerrodette,  Susan  Chivers, 

and  Alan  Jackson 

Annual  estimates  of  the  unobserved  incidental  kill  of 
pantropical  spotted  dolphin  (Stenella  attenuata  attenuata) 
calves  in  the  tuna  purse-seme  fishery  of  the  eastern 
tropical  Pacific 

245-250  Chernova,  Natalia  V.,  and  David  L.  Stein 

A  remarkable  new  species  of  Psednos  (Teleostei:  Liparidae) 
from  the  western  North  Atlantic  Ocean 

251-263  Chiang,  Wei-Chuan,  Chi-Lu  Sun,  Su-Zan  Yeh, 

and  Wei-Cheng  Su 

Age  and  growth  of  sailfish  Ustiophorus  platypterus) 
in  waters  off  eastern  Taiwan 


264-277  Clark,  Randall  D.,  John  D.  Christensen,  and  Mark  E.  Monaco, 

Philip  A.  Caldwell,  Geoffrey  A.  Matthews, 
and  Thomas  J.  Minello 

A  habitat-use  model  to  determine  essential  fish  habitat 
for  juvenile  brown  shrimp  (Farfantepenaeus  aztecus) 
in  Galveston  Bay,  Texas 

278-288  Delgado,  Gabriel  A.,  Claudine  T.  Bartels,  Robert  A.  Glazer, 

Nancy  J.  Brown-Peterson,  and  Kevin  J.  McCarthy 

Translocation  as  a  strategy  to  rehabilitate  the  queen  conch 
(Strombus  gigas)  population  in  the  Florida  Keys 

289-297  Lage,  Christopher,  Kristen  Kuhn,  and  Irv  Kornfield 

Genetic  differentiation  among  Atlantic  cod  (Gadus  morhua) 
from  Browns  Bank,  Georges  Bank, 
and  Nantucket  Shoals 


298-305  Lenihan,  Hunter  S.,  and  Charles  H.  Peterson 

Conserving  oyster  reef  habitat  by  switching  from  dredging  and 
tonging  to  diver-harvesting 


Fishery  Bulletin  102(2) 


306-327  Macewicz,  Beverly  J.,  John  R.  Hunter,  Nancy  C.  H.  Lo,  and  Erin  L.  LaCasella 

Fecundity,  egg  deposition,  and  mortality  of  market  squid  (Loli/go  opalescens) 

328-348  Orr,  James  W.,  and  James  E.  Blackburn 

The  dusky  rockfishes  (Teleostei:  Socrpaeniformes)  of  the  North  Pacific  Ocean 

resurrection  of  Sebastes  variabilis  (Pallas,  1814)  and  a  redescnption  of  Sebastes  ci/iatus  (Tilesius,  1813) 

349-365  Powers,  Joseph  E. 

Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 

366-375  Szedlmayer,  Stephen  T.,  and  Jason  D.  Lee 

Diet  shifts  of  juvenile  red  snapper  (Lut/anus  campechanus)  with  changes  in  habitat  and  fish  size 

376-388  Webb,  Stacey,  and  Ronald  T.  Kneib 

Individual  growth  rates  and  movement  of  juvenile  white  shrimp  (Litopenaeus  setiferus)  in  a  tidal  marsh  nursery 


Notes 

389-392  Forsythe,  John,  Nuutti  Kangas,  and  Roger  T.  Hanlon 

Does  the  California  market  squid  (Loligo  opalescens)  spawn  naturally  during  the  day  or  at  night? 
A  note  on  the  successful  use  of  ROVs  to  obtain  basic  fisheries  biology  data 

393-399  Kotas,  Jorge  E.,  Silvio  dos  Santos,  Venancio  G.  de  Azevedo,  Berenice  M.  G.  Gallo, 

and  Paulo  C.  R.  Barata 

Incidental  capture  of  loggerhead  (Caretta  caretta)  and  leatherback  (Dermochelys  conacea)  sea  turtles 
by  the  pelagic  longline  fishery  off  southern  Brazil 

400-405  Yang,  Mei-Sun 

Diet  changes  of  Pacific  cod  (Gadus  macrocephalus)  in  Pavlof  Bay  associated  with  climate  changes  in  the 
Gulf  of  Alaska  between  1980  and  1995 

406  Subscription  form 


233 


Abstract— We  estimated  the  total 
number  of  pantropical  spotted  dolphin 
(Stenella  attenuata)  mothers  killed 
without  their  calves  ("calf  deficit")  in 
all  tuna  purse-seine  sets  from  1973-90 
and  1996-2000  in  the  eastern  tropical 
Pacific.  Estimates  were  based  on  a 
tally  of  the  mothers  killed  as  reported 
by  color  pattern  and  gender,  several 
color-pattern-based  frequency  tables, 
and  a  weaning  model.  Over  the  time 
series,  there  was  a  decrease  in  the  calf 
deficit  from  approximately  2800  for 
the  western-southern  stock  and  5000 
in  the  northeastern  stock  to  about  60 
missing  calves  per  year.  The  mean 
deficit  per  set  decreased  from  approxi- 
mately 1.5  missing  calves  per  set  in 
the  mid-1970s  to  0.01  per  set  in  the 
late-1990s.  Over  the  time  series  exam- 
ined, from  75%  to  95%  of  the  lactating 
females  killed  were  killed  without  a 
calf.  Under  the  assumption  that  these 
orphaned  calves  did  not  survive  with- 
out their  mothers,  this  calf  deficit  rep- 
resents an  approximately  14%  increase 
in  the  reported  kill  of  calves,  which  is 
relatively  constant  across  the  years 
examined.  Because  the  calf  deficit  as 
we  have  defined  it  is  based  on  the  kill 
of  mothers,  the  total  number  of  mis- 
sing calves  that  we  estimate  is  poten- 
tially an  underestimate  of  the  actual 
number  killed.  Further  research  on 
the  mechanism  by  which  separation 
of  mother  and  calf  occurs  is  required 
to  obtain  better  estimates  of  the  unob- 
served kill  of  dolphin  calves  in  this 
fishery. 


Annual  estimates  of  the  unobserved 
incidental  kill  of  pantropical  spotted  dolphin 
{Stenella  attenuata  attenuato)  calves 
in  the  tuna  purse-seine  fishery 
of  the  eastern  tropical  Pacific 

Frederick  Archer 

Tim  Gerrodette 

Susan  Chivers 

Alan  Jackson 

Southwest  Fisheries  Science  Center 

National  Marine  Fisheries  Service 

8604  La  Jolla  Shores  Dr. 

La  Jolla,  California  92037 

E-mail  address  (for  F  Archer):  enc.archeriainoaa.gov 


Manuscript  approved  for  publication 
7  January  2004  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Sceintific  Publications  Office. 

Fish.  Bull.  102:233-244  (2004). 


In  the  eastern  tropical  Pacific  (ETP), 
yellowfin  tuna  (Thunnus  albacares) 
are  frequently  found  swimming  under 
schools  of  pantropical  spotted  ( Stenella 
attenuata)  and  spinner  (S.  longirostris) 
dolphins.  For  the  past  four  decades, 
the  ETP  yellowfin  tuna  fishery  has 
made  use  of  this  association  by  chasing 
the  more  visible  dolphins  at  the  sur- 
face and  using  purse-seines  to  encircle 
the  schools  "carrying"  the  tuna  (NRC, 
1992).  The  large  bycatch  of  dolphins  in 
this  fishery  has  become  widely  known 
as  the  "tuna-dolphin  issue"  (Gerro- 
dette, 2002).  During  the  1960s,  the 
number  of  dolphins  killed  by  the  fishery 
was  estimated  to  be  200,000-500,000 
per  year  (Wade,  1995),  and  two  stocks 
of  spotted  and  spinner  dolphins  were 
reduced  to  fractions  of  their  previous 
sizes  (Smith,  1983;  Wade  et  al.1).  Along 
history  of  technological  innovations  by 
fishermen,  laws  and  fishing  regula- 
tions, dolphin  quotas,  eco-labeling  of 
"dolphin-safe"  tuna,  and  a  comprehen- 
sive international  observer  program 
(Gosliner,  1999;  Hall  et  al,  2000;  Ger- 
rodette, 2002)  has  reduced  the  dolphin 
bycatch  to  less  than  1%  of  its  former 
level.  The  reported  bycatch  in  recent 
years  is  less  than  2000  dolphins  per 
year  for  all  species  combined  (IATTC, 
2002). 

Although  the  reported  kill  has  dra- 
matically decreased,  recent  studies 


suggest  that  there  is  little  evidence 
that  the  stocks  are  growing  close  to 
expected  rates  (Wade  et  al.1).  One  hy- 
pothesis for  this  lack  of  recovery  has 
been  that  there  are  unobserved  kills  of 
dolphins  during  tuna  purse-seine  sets. 
Archer  et  al.  (2001)  presented  evidence 
of  an  under-representation  of  suckling 
spotted  and  spinner  dolphin  calves  in 
a  sample  of  tuna  purse-seine  sets  in 
the  eastern  tropical  Pacific.  Given  that 
some  of  these  missing  calves  are  still 
dependent  on  their  mothers  for  nutri- 
tion, it  is  likely  that  once  separated 
they  would  die  and  this  under-repre- 
sentation represents  some  degree  of 
unobserved  kill. 

In  Archer  et  al.  (2001),  the  sample 
of  sets  examined  was  limited  to  those 
sets  in  which  all  of  the  animals  killed 
had  biological  data  collected  by  techni- 
cians aboard  the  tuna  vessel.  Calves 
still  dependent  on  their  mothers  in  the 
kill  were  identified  by  five  intervals  of 
body  length,  chosen  to  cover  a  range  of 


1  Wade,  P.  R..  S.  B.  Reilly.  and  T.  Gerro- 
dette. 2002.  Assessment  of  the  popula- 
tion dynamics  of  the  northeastern  offshore 
spotted  and  the  eastern  spinner  dolphin 
populations  through  2002.  National 
Oceanographic  and  Atmospheric  Admin- 
istration Administrative  Report  LJ-02- 
13.  58  p.  Southwest  Fisheries  Science 
Center.  8604  La  Jolla  Shores  Dr.,  La  Jolla, 
CA  92037. 


234 


Fishery  Bulletin  102(2) 


calf  sizes.  Because  of  this  approach,  it  was  not  possible  to 
derive  a  single  estimate  of  the  number  of  missing  calves 
or  to  extrapolate  their  estimate  to  sets  not  used  in  this 
analysis. 

In  the  current  study,  we  present  a  different  method  of 
estimating  the  number  of  missing  calves  in  each  set  where 
offshore  spotted  dolphins  (S.  attenuate!  attenuata)  were 
killed.  For  brevity,  we  call  the  shortage  of  calves  in  the  kill 
in  relation  to  the  number  of  lactating  females  in  the  kill 
the  "calf  deficit."  We  examined  the  western-southern  and 
northeastern  offshore  stocks  separately  according  to  the 
geographic  boundaries  described  by  Dizon  et  al.  (1994). 
As  they  age,  spotted  dolphins  change  color  through  five 
color  phases  (Perrin,  1970).  We  used  the  color-phase 
frequency  distribution  of  the  kill  in  conjunction  with  age- 
and  color-based  frequency  distributions  from  a  sample  of 
the  kill  to  estimate  the  total  number  of  missing  calves  in 
each  stock,  along  with  confidence  intervals  derived  from 
bootstrap  replications.  This  method  also  allowed  us  to 
examine  the  calf  deficit  from  sets  in  recent  years  from 
which  we  did  not  have  biological  samples  and  to  examine 
the  time  series  of  available  years  for  evidence  of  a  trend 
in  the  calf  deficit. 


Methods 

Since  1973,  observers  have  been  randomly  placed  on  tuna 
purse-seine  vessels.  For  each  spotted  dolphin  killed  during 
an  observed  set,  observers  attempted  to  record  the  sex  and 
the  color  phase  of  the  dolphin  ( neonate,  two-tone,  speckled, 
mottled,  and  fused,  see  Perrin,  1970).  From  the  National 
Marine  Fisheries  Service  (NMFS)  set  log  database,  we 
obtained  the  number  of  northeastern  and  western-south- 
ern offshore  spotted  dolphins  (by  gender  and  color  phase) 
killed  in  every  observed  set  from  1973  to  1990.  The  Inter- 
American  Tropical  Tuna  Commission  (IATTC)  provided 
the  same  data  from  1996  to  2000. 

Proration 

In  each  set,  color  phase  or  gender  (or  both)  may  not  have 
been  recorded  for  some  dolphins.  Assuming  that  the  distri- 
bution of  the  demographic  composition  of  this  missing  data 
is  equivalent  to  the  overall  demographic  composition  of  the 
kill,  we  allocated  the  number  of  dolphins  cf  unknown  color 
phase  (nu)  to  unknown  gender  in  each  color  phase  (jigu) 
according  to  the  following  formula, 


ngu:  =  ngu,  + 


N. 


I", 


(1) 


where  c  =  one  of  the  five  color  phases  (neonate  to  fused  I; 
Nc  =  the  total  number  of  dolphins  in  each  color 
phase  in  the  entire  data  set;  and 
ngu\.  =  the  new  number  of  dolphins  in  each  color  phase 
where  gender  is  unknown,  including  the  indi- 
viduals of  prorated  unknown  color  phase 


The  number  of  male  (nm'c)  or  female  (nf'c)  dolphins  in  a 
color  phase  was  calculated  as 


nm,.  =  nm   + 


ngu,  ■ 


Nm. 


Nni  +  Nf\  j 


nfc'=nft  + 


ngu 


w 


Nmc+Nfr 


(2) 


(3) 


where  Nmc  and  Nfc  are  the  total  number  of  males  and 
females,  respectively,  observed  in  that  color  phase  in 
the  entire  data.  Table  1  gives  the  sample  size  of  sets  for 
both  stocks  by  year,  as  well  as  the  fraction  of  the  kill  of 
unknown  gender  and  color  phase  that  were  prorated  as 
described  above. 

Number  of  suckling  calves 

As  time  permitted,  NMFS  observers  would  also  collect 
biological  data  from  a  subset  of  the  kill.  For  this  study, 
we  used  ages  estimated  from  teeth  collected  for  a  study  of 
spotted  dolphin  growth  and  reproduction  (Myrick  et  al., 
1986 ).  The  specimens  used  were  a  random  sample  of  all 
male  and  female  spotted  dolphins  collected  between  1973 
and  1978  for  which  total  body  length  was  recorded  and 
teeth  were  collected.  However,  additional  specimens  with 
lengths  less  than  150  cm  were  selected  in  order  to  match 
as  closely  as  possible  the  length  distribution  of  the  aged 
sample  to  the  underlying  length  distribution  of  the  spotted 
dolphins  in  the  kill.  This  was  necessary  because  observ- 
ers did  not  generally  collect  teeth  from  smaller,  younger 
animals.  Later,  another  sample  of  female  spotted  dolphins 
was  selected  from  specimens  collected  in  1981.  Specimens 
were  aged  as  described  in  Myrick  et  al.  ( 1986 ). 

The  final  data  set  used  in  our  analyses  included  age 
estimates  for  1094  female  spotted  dolphin  specimens  and 
798  male  specimens.  Of  these,  649  females  and  457  males 
belonged  to  the  northeastern  stock  and  had  color  phase  re- 
corded. These  1 106  dolphins  were  used  to  generate  the  age 
frequency  distribution  for  each  color  phase  (F  ,  Table  2). 


(4) 


'""Is,, 


where  Sac  =  the  number  of  samples  of  age  a  in  color  phase  c. 

The  oldest  age  recorded  was  36  years. 

To  derive  an  age  distribution  for  the  dolphins  killed  in 
each  tuna  set,  we  estimated  the  number  of  dolphins  in  each 
age  class  (na)  as 


»,,=x^„ 


i  5) 


where  n' 


the  sum  of  nm'c  and  «/'  (the  number  of  males 
and  females  in  each  color  phase  after  prora- 
tion from  Equations  2  and  3). 


Archer  et  al.:  Estimates  of  the  incidental  kill  of  Stenella  attenuata  attenuata  calves  in  the  tuna  purse-seine  fishery 


235 


Table  1 

Sample 

sizes  of  NMFS  (1973-1990) 

and  IATTC  (1996-2000)  observed  sets  with  spotted  dolphin  kill 

made  on  two  stocks  of  pan- 

tropical 

spotted  dolphins  iStenella  a 

ttenuata)  by  yea 

r. 

Northeastern  stock 

Western-southern  stock 

Fraction  of 

Fraction  of 

Fraction  of 

Fraction  of 

kill  of 

kill  of 

kill  of 

kill  of 

Number  of 

Observed 

unknown 

unknown 

Number  of 

Observed 

unknown 

unknown 

Year 

sets  with  kill 

kill 

color  phase 

gender 

sets  with  kill 

kill 

color  phase 

gender 

1973 

332 

5242 

0.09 

0.31 

75 

1199 

0.17 

0.34 

1974 

515 

5864 

0.16 

0.23 

92 

1715 

0.10 

0.31 

1975 

554 

8073 

0.31 

0.19 

75 

1702 

0.30 

0.20 

1976 

239 

2376 

0.24 

0.25 

356 

6293 

0.27 

0.23 

1977 

467 

2146 

0.23 

0.26 

528 

3358 

0.18 

0.32 

1978 

224 

1016 

0.18 

0.41 

329 

3998 

0.37 

0.34 

1979 

218 

1045 

0.38 

0.27 

168 

1262 

0.40 

0.14 

1980 

165 

1132 

0.45 

0.28 

106 

1206 

0.73 

0.13 

1981 

121 

815 

0.46 

0.13 

112 

1346 

0.48 

0.12 

1982 

171 

1696 

0.51 

0.22 

159 

1966 

0.37 

0.38 

1983 

12 

177 

0.80 

0.08 

35 

148 

0.32 

0.35 

1984 

43 

294 

0.37 

0.25 

71 

961 

0.48 

0.15 

1985 

186 

2625 

0.39 

0.40 

54 

381 

0.49 

0.13 

1986 

150 

1816 

0.48 

0.28 

132 

1818 

0.60 

0.22 

1987 

630 

3327 

0.25 

0.31 

175 

1768 

0.62 

0.14 

1988 

207 

1142 

0.18 

0.27 

107 

479 

0.36 

0.34 

1989 

293 

1096 

0.29 

0.25 

323 

2793 

0.48 

0.14 

1990 

157 

515 

0.16 

0.31 

121 

829 

0.35 

0.13 

1996 

273 

724 

0.27 

0.44 

161 

374 

0.18 

0.54 

1997 

163 

393 

0.15 

0.42 

274 

738 

0.24 

0.48 

1998 

161 

260 

0.21 

0.51 

125 

236 

0.19 

0.46 

1999 

189 

317 

0.18 

0.58 

88 

159 

0.11 

0.56 

2000 

146 

291 

0.23 

0.47 

115 

250 

0.20 

0.61 

In  Equation  4,  an  age  distribution  was  generated  for  each 
color  phase,  and  then  the  number  of  dolphins  in  each  age 
class  was  summed  across  all  color  phases. 

To  estimate  the  number  of  calves  in  each  set,  we  used 
this  age  distribution  in  conjunction  with  a  weaning  model 
developed  from  a  study  of  the  stomach  contents  and  ages 
of  calves  (Archer  and  Robertson,  in  press).  The  model 
predicts  the  probability  that  an  animal  of  a  given  age  (a) 
will  be  suckling: 


Pi  milk) 


(6) 


1  +  e1 


The  estimated  number  of  calves  (JV„„;f)  in  a  set  is  then 


N. 


calf 


IK 


calf 


Pmilk). 


(7) 


In  our  estimate  of  Ncal?  we  chose  to  use  only  the  first 
four  age  classes  (0  to  3)  because  P(milk)4  was  extremely 
small  (2xl0~4).  These  age  classes  allowed  us  to  decrease 
computational  time  without  significantly  affecting  the 
estimates. 


Number  of  lactating  females 

Observers  visually  examined  the  mamillaries  of  the  649 
females  used  in  the  age  distribution  above  (Eq.  4)  for  the 
presence  of  milk  as  part  of  the  suite  of  biological  data 
collected.  Using  these  data  in  conjunction  with  the  color 
phase  of  these  females,  we  calculated  the  fraction  of  lactat- 
ing females  in  each  color  phase  (Flacv), 


Flac 


■S/ac 
Sfem, 


(8) 


where  Slacv  and  Sfemc 


the  number  of  females  that  were 
lactating  and  the  total  number 
of  females  in  color  phase  c  of  the 
samples  examined. 


Flacc  was  0.00,  0.01,  0.04,  0.22,  and  0.50  for  neonate,  two- 
tone,  speckled,  mottled,  and  fused  specimens,  respectively. 
The  estimated  number  of  lactating  females  (Nlac)  in  a  set 
was  then 


Nlai.  =  ^(nf;  Flacv 


(9) 


236 


Fishery  Bulletin  102(2) 


Calf  deficit 

As  described  in  Archer  et  al.  (2001),  the  calf  deficit  (D) 
in  each  set  was  calculated  by  subtracting  the  number  of 
calves  (iVca;J  from  the  number  of  lactating  females  (N[ac). 
If  this  value  was  zero  or  less,  then  D  was  set  to  zero  to 
indicate  that  there  were  enough  calves  to  account  for  all 
lactating  females  killed  ( Fig.  1 1, 


D- 


0  if7V„„<iV„i;/- 


(10) 


We  calculated  three  deficit-based  fractions:  1)  the  mean 
deficit  per  set  (Ds);  2)  the  mean  deficit  per  dolphin  killed 
(Dk);  and  3),  the  mean  deficit  per  lactating  female  killed 
CD,): 


D 


D,  = 


D,= 


1° 

ObsSets 

ObsKill 

!*> 

EstLacKill 


(11) 


(12) 


(13) 


where   ZD  =  the  total  observed  calf  deficit  in  each  year; 
ObsSets  =  the  number  of  observed  sets  used  in  the 
analysis,  including  those  sets  without  a 
dolphin  kill; 
ObsKill  =  the  number  of  dolphins  killed  in  the  observed 
sets;  and 
EstLacKill  =  the  total  estimated  number  of  lactating 
females  killed. 

The  above  analysis  was  conducted  each  year.  Estima- 
tion error  was  evaluated  with  20,000  bootstrap  replicates 
for  each  year.  For  each  replicate,  the  sets  within  that  year 
were  randomly  resampled.  The  frequency  tables  Fni.  and 
Flact  were  also  recalculated  by  resampling  the  list  of  bio- 
logical specimens.  The  parameters  for  the  weaning  model, 
P(milk)a,  were  estimated  again  by  resampling  the  29 
calves  and  by  fitting  the  logistic  model  to  the  new  data  set 
as  described  in  Archer  and  Robertson  (in  press).  All  resa- 
mpling was  done  with  replacement.  Nralp  Nlac,  and  D  were 
estimated  as  described  above  for  each  set,  and  Ds,  Dh,  and 
Dj  were  calculated  for  the  replicate.  The  95°;  confidence 
intervals  for  Af  n//,  Nlac,  D,  Ds,  Dk,  and  D/  were  estimated 
from  the  2.5';  and  97.5%  quantiles  of  the  distributions  of 
the  bootstrap  replicate  values. 

The  total  calf  deficit  (Dtotal)  was  estimated  as  the  deficit 
per  dolphin  killed  (Dk)  multiplied  by  the  total  number  of 
dolphins  killed  [NkiUed)  by  stock  each  year, 


Table  2 

Age- 

class  frequency  distribution  for  e 

ich  color  phase  CFac). 

Age 

Two- 

(yr) 

Neonate 

tone 

Speckled 

Mottled 

Fused 

0 

0.80 

0.12 

0 

0 

0 

1 

0.20 

0.32 

0 

0 

0 

2 

0 

0.31 

0.04 

0 

0 

3 

0 

0.16 

0.18 

0.01 

0 

4 

0 

0.05 

0.14 

0.02 

0 

5 

0 

0.02 

0.13 

0.03 

0 

6 

0 

0 

0.13 

0.04 

0.01 

7 

0 

0 

0.06 

0.05 

0 

8 

0 

0 

0.10 

0.06 

0 

9 

0 

0 

0.06 

0.07 

0.01 

10 

0 

0 

0.01 

0.10 

0.01 

11 

0 

0 

0.01 

0.14 

0.03 

12 

0 

0 

0.01 

0.08 

0.02 

13 

0 

0 

0.04 

0.07 

0.03 

14 

0 

0 

0.03 

0.07 

0.03 

15 

0 

0 

0 

0.06 

0.06 

16 

0 

0.01 

0.01 

0.06 

0.07 

17 

0 

0 

0.01 

0.03 

0.07 

18 

0 

0 

0 

0.01 

0.07 

19 

0 

0 

0 

0.03 

0.09 

20 

0 

0 

0 

0.03 

0.07 

21 

0 

0 

0 

0.01 

0.08 

22 

0 

0 

0 

0 

0.06 

23 

0 

0 

0.01 

0 

0.07 

24 

0 

0 

0 

0.01 

0.04 

25 

0 

0 

0 

0.01 

0.04 

26 

0 

0 

0 

0 

0.04 

27 

0 

0 

0 

0.01 

0.03 

28 

0 

0 

0 

0.01 

0.02 

29 

0 

0 

0 

0 

0.01 

30 

0 

0 

0 

0.01 

0.02 

31 

0 

0 

0 

0 

0.01 

32 

0 

0 

0 

0 

0 

33 

0 

0 

0 

0 

0.01 

34 

0 

0 

0 

0 

0 

35 

0 

0 

0 

0 

0 

36 

0 

0 

0 

0 

0.01 

For  the  period  1973-84,  annual  values  of  Nhlllcil  for  each 
stock  were  provided  by  the  IATTC  (Joseph2).  For  1984-90 
and  1996-2000.  values  were  published  by  IATTC  (2002). 
In  the  bootstrap  estimation  of  the  959?  CI  around  Dlntal,  for 
the  1973-90  period,  each  replicate  was  randomly  sampled 
from  a  normal  distribution  by  using  the  estimated  total 
kill  standard  error.  For  1996-2000,  the  total  kill  was 
reported  to  be  exact;  therefore  the  total  kill  was  used 
without  variance  in  all  replicates. 


D 


total 


Dl;  *  Nkmd 


(1  li 


-  Joseph.  J.  1994.  Letter  of  September  6  to  Michael  Tillman. 
2  p.  Southwest  Fisheries  Science  Center.  8604  La  Jolla  Shores 
Dr.,  LaJolla.  CA  92037. 


Archer  et  al.:  Estimates  of  the  incidental  kill  of  Stene/la  attenuata  attenuate/  calves  in  the  tuna  purse-seine  fishery 


237 


Fraction  of  females  lactating, 
by  color  (1973-78,  1981):  Flac 


Number  of  lactating 

females  killed:  Nh„ 


Tally  of  females, 
by  color:  nj ' 


Tally  of  dolphins  killed,  by  color  and 
sex,  from  set  log  ( 1973-90.  1996-2000) 


Tally  of  dolphins, 
by  color:  n\ 


Fraction  of  dolphins  in  age  class, 
by  color  (1973-78,  1981):  Fat 


Number  of  dolphins 
killed,  by  age:  Na 


Probability  of  suckling, 
by  age  (1989-91):  P(milk)a 


Calf  deficit:  D 


Number  of  suckling 
calves  killed:  Ncay 


Figure  1 

Diagram  of  the  analytical  method  used  to  estimate  the  spotted  dolphin  (Stenella  attenuata  attenuata) 
calf  deficit  in  each  set  as  described  in  the  text.  Boxes  identify  original  data  that  were  bootstrapped  to 
produce  confidence  intervals.  Values  in  parentheses  are  years  for  which  data  were  available. 


In  a  subset  of  the  sets  that  we  examined,  every  indi- 
vidual killed  had  been  examined  and  biological  samples 
had  been  collected  from  it;  therefore,  we  knew  the  actual 
number  of  lactating  females  killed.  There  were  1108  of 
these  "100%  sampled"  sets  on  the  northeastern  stock,  and 
697  on  the  western-southern  stock  from  1973  to  1990.  We 
evaluated  the  accuracy  of  our  frequency-based  method 
by  conducting  a  paired  /-test  between  our  estimate  of  the 
number  of  lactating  females  and  the  number  observed  in 
each  of  these  sets. 

Stomach-content  data  were  not  available  for  every 
animal  in  these  100%  -sampled  sets;  therefore,  we  did  not 
know  the  actual  number  of  suckling  calves.  However,  we 
also  used  paired  /-tests  to  compare  our  estimate  of  the 
number  of  suckling  calves  in  each  set  with  the  number  of 
animals  smaller  than  122  cm,  which  was  the  estimated 
length  at  which  the  probability  of  milk  in  the  stomach 
was  0.5,  given  the  weaning  model  of  Archer  and  Robertson 
(in  press).  Likewise,  our  estimate  of  the  calf  deficit  was 
compared  with  the  deficit  as  estimated  by  using  a  cutoff 
length  of  122  cm.  These  tests  were  done  to  determine  if  the 
method  in  the  present  study  would  produce  significantly 
different  results  from  the  method  used  in  the  previous 
study  Paired  /-tests  were  conducted  for  each  year  sepa- 
rately, as  well  as  for  all  years  combined.  A  power  analysis 
was  also  performed  for  these  paired  /-tests  to  determine 
the  minimum  detectable  difference  at  which  we  could  re- 
ject the  null  hypothesis  of  no  difference  between  methods 
given  observed  sample  sizes  and  variability. 


Results 

The  calf  deficit  as  a  fraction  of  the  number  of  dolphins 
killed  (Dk)  increased  slightly  during  the  mid-1970s  but 
remained  relatively  constant  throughout  the  rest  of  the 
time  series  at  approximately  0.14  missing  calves  per  dol- 
phin killed  for  both  stocks  (Fig.  2).  The  total  calf  deficit 
(Dtotal)  as  estimated  from  the  annual  kill  decreased  from 
highs  of  approximately  5000  in  the  mid-1970s  down  to 
2000-3000  by  the  early  1980s  (Fig.  3).  In  the  late  1980s, 
this  value  increased  to  approximately  5000  in  northeast- 
ern spotted  dolphins  (Table  3A)  and  approximately  2800 
in  the  western-southern  stock  (Table  3B),  reflecting  an 
increase  in  the  reported  kills.  In  the  last  five  years  of  the 
time  series  (1996-2000),  the  estimated  total  deficit  was 
approximately  60  missing  calves. 

The  mean  deficit  per  set  (Z),)  for  northeastern  spotters 
over  all  years  was  1.03  missing  calves  per  set,  and  the  me- 
dian was  0.30  (Fig.  4).  For  western-southern  spotted  dol- 
phins, the  mean  was  1.28  missing  calves  per  set,  and  the 
median  was  0.33.  The  estimated  mean  deficit  per  set  was 
approximately  1.5  in  the  mid-1970s  and  decreased  over 
time  to  0.01-0.02  at  the  end  of  the  time  series  (Fig.  4).  For 
both  stocks,  75-  95%  of  lactating  females  killed  were  not 
killed  with  their  calf  (Fig.  5). 

In  the  sets  that  were  100% -sampled,  for  all  years  com- 
bined, there  was  no  significant  difference  between  the 
observed  and  the  estimated  number  of  lactating  females 
killed  in  either  stock  (Table  4).  The  results  of  paired  /-tests 


238 


Fishery  Bulletin  102(2) 


0.3- 

Northeastern 

0.2- 

•:.[•• 

"          "»"..| 

1" 

0.1  - 

\\-  |t 

"  ""         f          ^T 

ll' 

o.o- 

0.3- 

Western-southern 

0.2- 

1 1 

I     " 

}ll|ttftf     ll 

.       i 

0.1  - 

o.o- 

— 1 1 1 1— 

1970 


1980 


1990 


2000 


Year 


Figure  2 

Calf  deficit  per  spotted  dolphin  (Stenella  attenuata 
attenuata)  killed  (D/; )  by  year.  Vertical  lines  indicate 
95%  confidence  intervals. 


12000 


8000  -- 


4000 


0 


=    12000  ■■ 


8000 


4000  "- 


+■ 


Northeastern 


••At 


Western-southern 


»t  Air  * 


-+- 


-i- 


1970 


1980  1990 

Year 


2000 


Figure  3 

Total  estimated  calf  deficit  ^Dlotal)  by  year.  Vertical 
indicate  95%  confidence  intervals. 


by  year  indicated  that  the  observed  number  of  lactating 
females  in  each  set  was  significantly  greater  (P<0.05) 
than  the  estimated  number  in  1977  for  the  northeastern 
and  the  western-southern  stocks  and  in  1979  for  the  west- 
ern-southern stock.  The  difference  was  significantly  less 
in  1984  for  the  western-southern  stock.  Using  0.1  as  our 
type-2  error  level,  we  determined  through  power  analysis 
that  the  minimum  detectable  difference  («=0.05)  between 
the  mean  observed  and  estimated  number  of  lactating 
females  per  set  across  all  years  was  approximately  0.08 
and  0.09  in  the  northeastern  and  western-southern  stocks 
respectively. 

The  observed  number  of  calves  per  set,  defined  as  the 
number  of  dolphins  less  than  122  cm,  was  significantly 
greater  for  both  stocks,  for  all  years  combined,  than  the 
values  estimated  in  this  paper  ( Table  5 ).  The  overall  mean 
difference  was  0.17  calves  per  set  for  the  northeastern 
stock  and  0.12  for  the  western-southern  stock.  About 
half  of  the  years  showed  a  significant  difference  for  each 
stock.  In  the  comparison  of  the  calf  deficit  by  year,  only  a 
few  years  showed  significant  differences  in  either  stock 
(Table  5).  However,  the  estimated  deficit  tended  to  be 
larger  than  the  observed  deficit.  The  paired  t-test  for  all 


years  combined  was  significant  for  the  northeastern  stock, 
although  the  mean  difference  was  only  -0.06  missing 
calves  per  set.  The  minimum  detectable  difference  from 
the  power  analysis  for  the  mean  number  of  calves  per  set 
and  mean  calf  deficit  per  set  across  all  years  was  0.06  and 
0.08  respectively  for  both  stocks. 


Discussion 

In  the  present  study,  we  present  an  estimate  of  the  number 
of  missing  dependent  northeastern  and  western-southern 
offshore  spotted  dolphin  calves  in  the  tuna  purse-seine 
kill  from  1973  to  1990  and  from  1996  to  2000.  The  total 
number  of  missing  calves  decreased  through  the  time 
series,  which,  because  we  estimated  the  calf  deficit  as  a 
function  of  the  size  of  the  kill,  was  a  direct  result  of  the 
large  reduction  in  the  annual  dolphin  kill  by  the  fishery. 
Between  1973  and  2000,  the  shortage  of  calves  in  the 
kill  remained  at  a  relatively  constant  fraction  of  the  kill, 
about  14ri  ,  for  both  stocks  of  pantropical  spotted  dolphins 
(Fig.  2).  On  the  assumption  that  suckling  calves  do  not 
survive  separation  from  their  mother  ( Archer  et  al.,  2001; 


Archer  et  al.:  Estimates  of  the  incidental  kill  of  Stenella  attenuata  attenuate  calves  in  the  tuna  purse-seine  fishery 


239 


3.0 

2.5 
2.0 
1.5 
1.0 
2°    0.5 

CD 
if) 

I    0.0 

I     3.0 

a 

1     2.5 

c 
a 

|     2.0  f 
1.5 
1.0 
0.5" 
0.0" 


-+- 


Northeastern 


•(.tt 


It 

•  * 


Western-southern 


1 1 

■+- 


-H 


-t- 


1970 


1980  1990 

Year 


2000 


Figure  4 

Mean  calf  deficit  per  set  (Ds)  by  year.  Vertical 
lines  indicate  959?  confidence  intervals. 


1.0" 

1 

0.9" 

■ 

,, 

.1 

",,      " 

ii                                                         ii 

ti                                             '  ■ 

0.8" 

'< 

" 

0.7- 

Q 

§     0.6" 

5 

Northeastern 

CD 

;?    0.5- 

CD 

"5.    10- 

c 

J?     0.9  - 

i  ■ 

1     '  ,, 

Deficit  per 

o 
co 

,      ". 

ti 

,, 

0.7  1 

0.6" 

Western-southern 

0.5- 

970            1980            1990            2000 

Year 

Figure  5 

Calf  deficit  per  lactating  female  killed  (D,)  by  year. 

Vertical  lines  indicate  95^  confidence  intervals. 

Edwards3!,  the  estimated  calf  deficit  represents  an  approx- 
imately 147c  underestimate  of  the  reported  kill. 

The  calf  deficit  in  the  present  study  was  estimated  from 
the  number  of  dependent  calves  and  lactating  females 
killed  by  using  age-color  frequency  tables  and  data  on  the 
stomach  contents  of  weaning  calves.  Specimens  used  to 
derive  the  age  and  color  table  were  collected  from  1973  to 
1978  and  1981,  and  specimens  used  for  the  weaning  model 
were  collected  between  1989  and  1991.  If  the  distributions 
of  these  samples  were  not  representative  of  all  years  that 
we  examined,  then  our  results  may  be  biased.  However,  the 
results  of  a  study  to  construct  the  annual  age  distribution  of 
the  kill  (Archer  and  Chivers4 )  indicated  that  there  is  no  sig- 
nificant difference  in  the  age-color  frequency  table  across 
years.  The  sample  size  for  the  stomach  data  ( 29  calves)  was 
too  small  to  examine  differences  between  years. 


Our  finding  of  no  significant  difference  between  our  esti- 
mates of  the  number  of  lactating  females  and  the  observed 
tally  of  lactating  females  in  sets  where  the  entire  kill  was 
sampled  validates  this  portion  of  our  estimation  proce- 
dure. However,  because  the  number  of  suckling  calves 
present  in  these  100%  -sampled  sets  was  not  recorded,  we 
were  unable  to  validate  the  method  used  to  generate  these 
estimates  in  a  similar  manner. 

The  results  of  our  paired  Ntests  indicated  that  the  ob- 
served number  of  animals  smaller  than  122  cm  tended  to 
be  greater  than  the  number  we  estimated.  This  is  most 
likely  a  result  of  the  difference  between  how  calves  were 
counted  in  each  method.  Archer  et  al.  ( 2001 )  considered  all 
animals  under  a  series  of  cutoff  values  to  be  calves  that 
were  dependent  on  suckling  for  survival.  In  the  present 
study,  the  weaning  model  that  we  used  (Archer  and  Rob- 


3  Edwards,  E.  F.  2002.  Behavioral  contributions  to  separa- 
tion and  subsequent  mortality  of  dolphin  calves  chased  by  tuna 
purse-seiners  in  the  eastern  tropical  Pacific  Ocean.  National 
Oceanographic  and  Atmospheric  Administration  Administra- 
tive Report  LJ-02-28,  34  p.  Southwest  Fisheries  Science 
Center,  8604  La  Jolla  Shores  Dr.,  La  Jolla,  CA  92037. 


4  Archer.  F..  and  S.  J.  Chivers.  2002.  Age  structure  of  the 
northeastern  spotted  dolphin  incidental  kill  by  year  for  1971  to 
1990  and  1996  to  2000.  National  Oceanographic  and  Atmo- 
spheric Administration  Administrative  Report  LJ-02-12,  18  p. 
Southwest  Fisheries  Science  Center,  8604  La  Jolla  Shores  Dr., 
La  Jolla.  CA  92037. 


240 


Fishery  Bulletin  102(2) 


Table  3 

Estimated  calf  deficit  per  kill  (Dt) 

and  total  calf  deficit 

Total  number  of 

spotted  dolphins  killed  reported  by  the  I ATTC  ( 2002 )  and 

Joseph  (footnote  2  in  the  general  text).  Values  in  parentheses  are  95%  lower  and  upper  confidence  intervals. 

Mean  calf 

Total  number 

Estimated  calf 

deficit 

of  NE  spotted 

Estimated 

Stock  and 

deficit  in 

Observed 

per  kill 

dolphins  killed 

total  calf 

year 

observed  sets 

dolphin  kill 

(Dk) 

(±SE) 

deficit 

A  Northeastern  (NE)  stock 

1973 

599 

5242 

0.11 

49928  ±8899 

5709 

(464,964) 

(3947,6820) 

(0.10,0.16) 

(3972,9532) 

1974 

634 

5864 

0.11 

37410  ±4222 

4046 

(583,10271 

(4943,6916) 

(0.10,0.16) 

(3573,6708) 

1975 

1014 

8073 

0.13 

49399  ±8809 

6206 

1618,12691 

(6578,9965) 

(0.08,0.14) 

(3297.8254) 

1976 

300 

2376 

0.13 

20443  ±4721 

2583 

(196,408) 

(1786.3079) 

(0.09,0.15) 

(1284.3903) 

1977 

341 

2146 

0.16 

5937  ±690 

943 

(249,416) 

(1743.26221 

(0.13,0.18) 

(656,1167) 

1978 

148 

1016 

0.15 

4226  ±827 

616 

(83,209) 

(684,1431) 

(0.11,0.16) 

(336,8361 

1979 

138 

1045 

0.13 

4828  ±817 

640 

(96.226) 

(680,1629) 

(0.11,0.17) 

(428,963) 

1980 

178 

1132 

0.16 

6468  ±962 

1016 

(107,239) 

(724.1637) 

(0.12,0.18) 

(622,13001 

1981 

137 

815 

0.17 

8096  ±1508 

1366 

(84,173) 

(560,1122) 

(0.12,0.18) 

(753,1774) 

1982 

212 

1696 

0.12 

9254  ±1529 

1155 

(155,347) 

(1126,23951 

(0.11,0.17) 

(833,1840i 

1983 

27 

177 

0.15 

2460  ±659 

377 

(7,59) 

(35,410) 

(0.11,0.23) 

(169.678) 

1984 

38 

294 

0.13 

7836  ±1493 

1017 

(26,57) 

(191,417) 

(0.10,0.17) 

(608,1602) 

1985 

337 

2625 

0.13 

25975  ±3210 

3338 

(235,508) 

(1839.3529) 

(0.11,0.16) 

(2447,4748) 

1986 

290 

1816 

0.16 

52035  ±8134 

8297 

H19.478) 

(859.3440) 

(0.10,0.17) 

14496,9935) 

1987 

497 

3327 

0.15 

35366  ±4272 

5280 

(397,667) 

(2777,4002) 

(0.13.0.18) 

(3949.71061 

1988 

182 

1142 

0.16 

26625  ±2744 

4234 

(122,215) 

(880,1462) 

(0.12.0.17) 

(2825.4907) 

1989 

165 

1096 

0.15 

28898  ±3108 

4357 

(120.217) 

(871,1371) 

(0.12.0.17) 

(3186,54921 

1990 

65 

515 

0.13 

22616  ±2575 

2875 

(53.90) 

(421,632) 

(0.11,0.17) 

(2176,4085) 

1996 

88 

724 

0.12 

818 

99 

(76.142) 

(568,926) 

(0.12.0.17) 

(96.139) 

1997 

49 

393 

0.13 

721 

91 

(42,69) 

(331,461) 

(0.11,0.17) 

(81,121) 

1998 

33 

260 

0.13 

298 

38 

(26,41) 

(230,296) 

(0.10.0.16) 

(30,46) 

1999 

36 

317 

0.11 

358 

40 

(30,48) 

(282,357) 

(0.10.0.15) 

(35,53) 

2000 

43 

291 

0.15 

295 

44 

(32.58) 

(247.342) 

(0.12,0.18) 

(35.541 
continued 

Archer  et  al.:  Estimates  of  the  incidental  kill  of  Stenella  attenuata  attenuate  calves  in  the  tuna  purse-seine  fishery 


241 


Table  3  (continued) 


Stock  and 
year 


Mean  calf 

Total  number 

Estimated  calf 

deficit 

of  NE  spotted 

Estimated 

deficit  in 

Observed 

per  kill 

dolphins  killed 

total  calf 

observed  sets 

dolphin  kill 

CD*) 

(±SE) 

deficit 

141 

1199 

0.12 

51,712  ±10.721 

6076 

(110,229) 

(836,1638) 

(0.10,0.17) 

(3993-10,633) 

254 

1715 

0.15 

35,499  ±10.309 

5254 

(100,318) 

(939,2733) 

(0.07.0.15) 

(1554,6890) 

197 

1702 

0.12 

48,837  ±10,055 

5664 

(123.322) 

11104,2434) 

(0.09,0.15) 

(3285,9121) 

795 

6293 

0.13 

52,206  ±8883 

6595 

(524,1036) 

(4925,7860) 

(0.09,0.15) 

(3833,9223) 

491 

3358 

0.15 

11.260  ±1186 

1647 

(345,563) 

(2860,3906) 

(0.11.0.16) 

(1098,1959) 

660 

3998 

0.17 

11.610  ±2553 

1917 

(342,949) 

(2508,5922) 

(0.12.0.18) 

(932.2614) 

157 

1262 

0.12 

6.254  ±1229 

776 

1104.216) 

(939.1643) 

(0.09,0.15) 

(438.1138) 

144 

1206 

0.12 

11.200  ±2430 

1339 

(59.3441 

(411.2542) 

(0.10,0.17) 

(831,2320) 

191 

1346 

0.14 

12.512  ±2629 

1775 

(90,340) 

(577.2416) 

(0.11.0.17) 

(1010,2682) 

306 

1966 

0.16 

9869  ±1146 

1536 

(198,474) 

(1337,2734) 

(0.13,0.19) 

(1156,2088) 

23 

148 

0.16 

4587  ±928 

724 

(15.33) 

(99.206) 

(0.12.0.20) 

(418,1087) 

114 

961 

0.12 

10.018  ±2614 

1183 

(80.224) 

(526,1513) 

(0.12,0.18) 

(712,2352) 

52 

381 

0.14 

8089  ±951 

1105 

(32,791 

(225.5701 

(0.11.0.17) 

(781,1524i 

275 

1818 

0.15 

20,074  ±2187 

3037 

(143,373) 

(1065.2784) 

(0.10.0.17) 

(1776,3617) 

271 

1768 

0.15 

19,298  ±2899 

2959 

(147,374) 

(1068,2661) 

(0.11,0.16) 

(1754,3695) 

75 

479 

0.16 

13,916  ±1741 

2166 

(51,96) 

(368,605) 

(0.12,0.18) 

(1453,2785) 

392 

2793 

0.14 

28,560  ±2675 

4011 

(242,589) 

(1819,4277) 

(0.11,0.16) 

(2861.4977) 

123 

829 

0.15 

12,578  ±1015 

1864 

(78,160) 

(582.1128) 

(0.11,0.17) 

(1283,2236) 

53 

374 

0.14 

545 

77 

(42,711 

(308.448) 

(0.12,0.18) 

(64,97) 

89 

738 

0.12 

1044 

126 

(72,132) 

(598.931) 

(0.11.0.16) 

(112,165) 

31 

236 

341 

44 

0.13 

(25,42) 

(192.288) 

(0.11,0.17) 

(38,58) 

22 

159 

0.14 

253 

35 

(16.32) 

(123,209) 

(0.11,0.18) 

(28,44) 

28 

250 

0.11 

435 

48 

(22.44) 

(189,330) 

(0.10.0.15) 

(42.67) 

B  Western- 
1973 

1974 

1975 

1976 

1977 

1978 

1979 

1980 

1981 

1982 

1983 

1984 

1985 

1986 

1987 

1988 

1989 

1990 

1996 

1997 

1998 

1999 

2000 


southern  (WS)  stock 


242 


Fishery  Bulletin  102(2) 


Table  4 

Annual  mean  observed  and  mean  estimated 
are  95%  lower  and  upper  confidence  intervals 
ference  from  zero  (P<0.05)  in  the  paired  t-test 

number  of  lactating  females  per 
assuming  a  normal  distribution 

s. 

set  in  100%  sampled  sets.  Values  in  parentheses 
of  differences.  Bold  type  indicates  significant  dif- 

Year 

Northeastern  stock 

Western-southern  stock 

No.  of 
sets 

Observed 

Estimated 

Difference 
(959S  CI) 

No.  of 
sets 

Observed 

Estimated 

Difference 
(95%  CD 

1973 

116 

0.55 

0.61 

-0.06  1-0.17.0.051 

21 

1.19 

1.30 

-0.11  (-0.63,0.421 

1974 

98 

0.51 

0.54 

-0.03  1-0.13.0.07) 

16 

0.75 

0.81 

-0.061-0.36.0.24) 

1975 

99 

0.57 

0.48 

0.09  (-0.05,0.22) 

14 

1.07 

0.92 

0.15  (-0.46.0.77) 

1976 

51 

0.28 

0.35 

-0.08  1-0.18.0.02) 

90 

0.500 

0.502 

-0.002  (-0.119.0.115) 

1977 

167 

0.55 

0.46 

0.09  (0.01.0.15) 

163 

0.49 

0.37 

0.12  (0.03,0.21) 

1978 

82 

0.37 

0.40 

-0.03  1-0.14,0.08) 

93 

0.50 

0.52 

-0.02  (-0.19,0.13) 

1979 

75 

0.47 

0.46 

0.01  (-0.13,0.14) 

61 

0.64 

0.47 

0.17  (0.01,0.33) 

1980 

54 

0.39 

0.38 

0.01  (-0.11.0.13) 

34 

0.50 

0.44 

0.06  1-0.09,0.20) 

1981 

41 

0.53 

0.74 

-0.21  (-0.81,0.38) 

38 

0.66 

0.64 

0.02  1-0.16.0.19) 

1982 

36 

0.62 

1.18 

-0.56  (-1.40,0.27) 

33 

0.30 

0.44 

-0.14  1-0.37,0.10) 

1983 

33 

1.33 

2.14 

-0.8K-7.89.6.28) 

6 

0.17 

0.57 

-0.40  (-1.57,0.77) 

1984 

4 

0.25 

0.49 

-0.24  1-0.67,0.18) 

29 

0.48 

1.08 

-0.60  (-0.96,-0.23) 

1985 

70 

0.34 

0.50 

-0.16  1-0.36,0.061 

17 

0.35 

0.50 

-0.15  1-0.49,0.20) 

1986 

45 

0.71 

0.47 

0.24  1-0.04,0.51) 

28 

0.61 

0.42 

0.19  1-0.01.0.38) 

1987 

121 

0.43 

0.46 

-0.03  (-0.18,0.11) 

30 

0.27 

0.46 

-0.19  1-0.44.0.06) 

1988 

6 

0.44 

0.57 

-0.13  (-0.59,0.35) 

— 

— 

— 

— 

1989 

24 

0.96 

1.03 

-0.07  1-0.59,0.44) 

15 

0.93 

0.96 

-0.03  (-0.68,0.64) 

1990 

16 

0.56 

0.47 

0.09  1-0.25,0.44) 

9 

0.67 

0.93 

-0.26  1-0.94,0.42) 

All 

1108 

0.50 

0.53 

-0.03  1-0.08,0.02) 

697 

0.545 

0.546 

-0.001  (-0.053,0.051) 

ertson,  in  press)  estimated  the  probability  that  a  calf  of  a 
given  age  class  was  still  suckling.  Given  that  body  length 
has  a  near  linear  relationship  with  age  in  these  young 
age  classes  (Perrin,  1976),  this  meant  that  for  any  chosen 
length  of  independence,  each  individual  smaller  than  that 
cutoff  value  would  only  be  counted  fractionally,  in  effect 
correcting  for  the  probability  that  an  animal  of  a  given 
age  is  not  suckling.  This  procedure  caused  the  method  in 
this  paper  to  tally  fewer  "calves"  in  each  set  than  in  the 
previous  study.  A  secondary  result  of  this  effect  was  that 
the  mean  deficit  per  set  estimated  in  the  present  study 
tended  to  be  slightly  higher  than  that  presented  by  Archer 
etal.  in  2001. 

We  estimated  the  total  number  of  missing  calves  as  a 
function  of  the  number  of  dolphins  killed  in  each  stock 
(Table  3).  Prior  to  1995,  only  a  fraction  of  the  purse-seine 
trips  carried  scientific  observers.  To  estimate  the  number 
killed  in  each  stock,  kill  rates  from  the  observed  trips  were 
applied  to  unobserved  trips,  stratified  by  area  and  stock 
(IATTC,  2002;  Joseph,  19942).  Since  1995  it  has  been  re- 
ported that  all  dolphin  sets  have  been  observed,  and  that 
the  number  of  dolphins  killed  is  therefore  known  without 
error  (IATTC,  2002). 

The  total  calf  deficit  could  also  be  estimated  as  a  function 
of  the  number  of  sets  by  multiplying  the  total  number  of 
sets  made  on  each  stock  by  Ds  (Fig.  4).  In  the  only  study  to 
estimate  the  number  of  sets  made  on  each  stock  annually. 


Archer  et  al.5  used  a  relatively  simple  proration  scheme  of 
unobserved  sets  derived  from  ratios  of  the  number  of  sets 
made  on  each  stock  in  observed  sets.  However,  because 
Archer  et  al.5  did  not  stratify  unobserved  sets  by  area,  bas- 
ing the  total  calf  deficit  on  these  estimates  would  produce 
a  different  result  from  that  presented  in  Table  3.  Because 
the  estimates  of  the  kill  by  stock  included  stratification 
by  area,  estimates  of  the  total  calf  deficit  calculated  by 
multiplying  the  kill  estimates  by  D,.  are  likely  to  be  more 
accurate.  It  is  important  to  realize  that  the  deficit  that  we 
present  is  directly  related  to  the  kill  observed  in  the  sets 
that  we  used.  In  other  words,  if  proration  schemes  for  un- 
observed sets  were  the  same  for  the  number  of  sets  made 
and  the  number  of  dolphins  killed,  estimates  of  the  total 
calf  deficit  with  either  D^  or  Dk  would  be  equivalent. 

Wade  et  al.1  explored  the  effects  of  50%  and  100'  <  ad- 
ditional fisheries-related  mortality  on  the  assessment  of 
the  northeastern  spotted  dolphin  stock.  The  assumption  of 
additional  mortality  led  to  higher  estimates  of  maximum 


Archer.  F..  T.  Gerrodette.  and  A.  Jackson.  2002.  Prelim- 
inary estimates  of  the  annual  number  of  sets,  number  of 
dolphins  chased,  and  number  of  dolphins  captured  by  stock 
in  the  tuna  purse-seine  fishery  in  the  eastern  tropical  Pacific. 
1971-2000.  National  Oceanographic  and  Atmospheric  Admin- 
istration Administrative  Report  LJ-02-10.  26  p.  Southwest 
Fisheries  Science  Center,  8604  La  Jolla  Shores  Dr.,  La  Jolla, 
CA  92037. 


Archer  et  al.:  Estimates  of  the  incidental  kill  of  Stenella  attenuata  attenuata  calves  in  the  tuna  purse-seine  fishery 


243 


Table  5 

Annua 

1  mean  number  of  dolphins  killed  sl22  cm  (calves  killed  based 

on  length)  and  estimated  number  of 

suckling  calves  (calves 

based 

3n  weaning  model  1  per  set  in  100rt  sampled  sets  (first  line  for  each  year!.  Mean  deficit  per  set  using 

.22  cm  as  cutoff  length 

(calf  deficit  based  on  length 

)  and  calf  deficit 

as  estimated  in  this  article  (calf  deficit  based  on  weaning  model)  on  second  line  for 

each  year.  Values  in  parentheses  are  959!  lower  and  upper  confidence  intervals  assuming  a  normal  dist 

•ibution  of  differences. 

Differences  in  bold  indicate 

significant  difference  from  zero  (PsO.05) 

in  the  paired  Mest. 

Northeastern  stock 

Western-southern 

stock 

Calves  killed 

Calves  killed 

Calves  killed 

Calves  killed 

based  on 

based  on 

based  on 

based  on 

length 

weaning  model 

length 

weaning  model 

No. 

Calf  deficit 

Calf  deficit 

No. 

Calf  deficit 

Calf  deficit 

of 

based  on 

based  on 

Difference 

of 

based  on 

based  on 

Difference 

Year 

sets 

length 

weaning  model 

(95%  CD 

sets 

length 

weaning  model 

(95%  CD 

1973 

116 

0.54 

0.21 

0.33  (0.18,0.50) 

21 

0.33 

0.06 

0.27  (0.01,0.55) 

0.35 

0.48 

-0.13  (-0.26,-0.03) 

1.00 

1.25 

-0.25  1-0.79.0.29) 

1974 

98 

0.39 

0.05 

0.34  (0.20,0.47) 

16 

0.56 

0.09 

0.47  1-0.53.1.47) 

0.36 

0.50 

-0.14  (-0.26,-0.03) 

0.56 

0.74 

-0.181-0.61,0.25) 

1975 

99 

0.57 

0.15 

0.42  (0.20,0.64) 

14 

0.29 

0.11 

0.18(-0.03.0.39) 

0.46 

0.40 

0.04  1-0.05.0.161 

0.93 

0.83 

0.10(-0.45,0.66i 

1976 

51 

0.18 

0.11 

0.07  1-0.01,0.15) 

90 

0.13 

0.07 

0.06  (0.001,0.13) 

0.28 

0.31 

-0.031-0.14.0.06) 

0.49 

0.47 

0.02  1-0.10.0.15) 

1977 

167 

0.10 

0.03 

0.07  (0.02,0.12) 

163 

0.17 

0.06 

0.11  (0.06,0.16) 

0.51 

0.45 

0.06  (-0.01,0.14) 

0.46 

0.35 

0.11(0.03,0.20) 

1978 

82 

0.17 

0.03 

0.14(0.05,0.23) 

93 

0.18 

0.05 

0.13  (0.04,0.23) 

0.35 

0.39 

-0.04  (-0.14.0.07) 

0.43 

0.50 

-0.07  1-0.22.0.09) 

1979 

75 

0.09 

0.04 

0.05  (-0.02,0.13) 

61 

0.31 

0.13 

0.18(0.04,0.32) 

0.44 

0.43 

0.01  (-0.11,0.13) 

0.51 

0.37 

0.14  1-0.03,0.31) 

1980 

54 

0.16 

0.03 

0.13  (0.02,0.25) 

34 

0.00 

0.01 

-0.01  (-0.02,-0.003) 

0.373 

0.371 

0.002-0.115,0.119) 

0.50 

0.44 

0.061-0.08.0.21) 

1981 

41 

0.105 

0.110 

-0.005  1-0.194,0.185) 

38 

0.05 

0.04 

0.01  (-0.04.0.07) 

0.53 

0.65 

-0.121-0.57,0.31) 

0.63 

0.62 

0.01  (-0.17,0.20) 

1982 

36 

0.44 

0.21 

0.23  (-0.10.0.55) 

33 

0.06 

0.02 

0.04  1-0.04,0.12) 

0.44 

1.00 

-0.56(-1.27,0.14i 

0.27 

0.42 

-0.15  (-0.37,0.08) 

1983 

33 

0.00 

0.14 

0.14  1-0.64.0.36) 

6 

0.17 

0.04 

0.13  1-0.31.0.57) 

1.33 

2.00 

-0.67  (-7.25.5.91) 

0.17 

0.56 

-0.39  1-1.56,0.76) 

1984 

4 

0.00 

0.02 

-0.02  1-0.08,0.04) 

29 

0.14 

0.04 

0.10  1-0.01,0.21) 

0.25 

0.49 

-0.24  1-0.67,0.18) 

0.35 

1.04 

-0.69  (-1.13,-0.26) 

1985 

70 

0.13 

0.04 

0.09(0.02,0.15) 

17 

0.06 

0.04 

0.02  1-0.06.0.10) 

0.29 

0.47 

-0.181-0.39,0.03) 

0.35 

0.49 

-0.14  (-0.48,0.21) 

1986 

45 

0.13 

0.04 

0.09  (0.01,0.17) 

28 

0.04 

0.03 

0.01  (-0.04,0.06) 

0.64 

0.44 

0.20  1-0.04.0.44) 

0.57 

0.39 

0.181-0.02,0.38) 

1987 

121 

0.14 

0.02 

0.12  (0.05,0.20) 

30 

0.23 

0.08 

0.15(0.02,0.30) 

0.38 

0.45 

-0.07  1-0.22,0.07) 

0.27 

0.43 

-0.16  (-0.41,0.09) 

1988 

6 

0.11 
0.33 

0.12 
0.50 

-0.01  (-0.23.0.22) 
-0.17  1-0.62,0.28) 

— 

— 

— 

— 

1989 

24 

0.22 

0.13 

0.09  (-0.11,0.29) 

15 

0.47 

0.20 

0.27  (0.05,0.49) 

0.87 

0.95 

-0.08  (-0.60,0.43) 

0.73 

0.82 

-0.09  1-0.66,0.48) 

1990 

16 

0.31 

0.21 

0.10  (-0.18,0.38) 

9 

0.89 

0.17 

0.72  1-0.18,1.62) 

0.56 

0.41 

0.15  (-0.20.0.51) 

0.33 

0.77 

-0.44  (-1.15,0.27) 

All 

1108 

0.25 

0.08 

0.17  (0.13,0.20) 

697 

0.18 

0.06 

0.12  (0.09,0.16) 

0.42 

0.48 

-0.06  (-0.10,-0.01) 

0.49 

0.51 

-0.02  (-0.08,0.03) 

244 


Fishery  Bulletin  102(2) 


growth  rates  and  lower  estimates  of  the  current  size  of 
the  population  in  relation  to  carrying  capacity.  Wade  et 
al.1  did  not  model  the  calf  deficit  estimated  in  our  present 
study,  but  the  effect  of  14/r  additional  mortality  would 
probably  be  less  than  the  50f>  additional  mortality  that 
was  modeled.  The  50^  mortality  was  spread  over  all  age 
classes,  and  additional  mortality  due  to  missing  calves 
should  be  assigned  to  the  first  two  year  classes  only.  The 
important  question  is  whether  the  calf  deficit  in  the  kill 
represents  the  main  effect  of  mother-calf  separation  by 
the  fishing  process.  As  outlined  in  Archer  et  al.  (2001t, 
the  mechanism  by  which  suckling  calves  are  separated 
from  their  mothers  is  unknown.  If  separation  is  simply  a 
function  of  the  number  of  lactating  females  killed,  then  the 
deficit  presented  here  is  an  accurate  representation  of  the 
number  of  "missing"  calves. 

However,  there  is  some  evidence  that  separation  can 
occur  without  the  mother  being  killed.  In  the  early  days 
of  the  backdown  procedure,  purse-seine  skippers  reported 
that  "Babies  swim  around  the  outside  of  the  net  pushing  to 
get  back  in  probably  because  their  mothers  are  still  inside" 
i  Gehresp  (.  It  is  unclear  whether  these  calves  were  sepa- 
rated prior  to  encirclement  or  were  released  early  during 
backdown,  prior  to  their  mothers.  Regardless,  given  that 
dolphins  exhibit  some  of  their  fastest  swimming  during 
a  set  immediately  upon  release  from  the  net  tChivers  and 
Scott' ),  separated  calves  waiting  immediately  outside  the 
net  may  risk  separation  if  their  mothers  join  the  rest  of  the 
school  rapidly  swimming  away  from  the  net.  If  this,  or  any  of 
the  other  scenarios  regarding  the  manner  in  which  perma- 
nent separation  can  occur  without  the  mother  being  killed 
i  Archer  et  al..  2001 1.  then  the  calf  deficit  underestimates  the 
actual  number  of  orphaned  calves.  Future  research  should 
focus  on  the  mechanism  of  calf  separation  because  a  better 
understanding  of  this  process  is  the  only  way  we  will  be  able 
to  estimate  the  magnitude  of  the  unobserved  calf  mortality 
and  its  subsequent  effects  on  the  population. 


Acknowledgments 

The  authors  wish  to  thank  Michael  Scott  and  Xick  Vogel 
of  the  IATTC  for  providing  data  as  well  as  Jay  Bar- 


B  Gehres.  L.  E.  1971.  Letter  of  July  2  to  Alan  R.  Longhurst. 
2  p.  Southwest  Fisheries  Science  Center,  8604  La  Jolla  Shores 
Dr..  La  Jolla.  CA  92037. 

7  Olivers.  S.  J.,  and  M.  D.  Scott.  2002.  Tagging  and  tracking 
of  Stenella  spp.  during  the  2001  Chase  Encirclement  S 
Studies  cruise.  National  Oceanographic  and  Atmospheric 
Administration  Administrative  Report  LJ-02-33.  21  p.  South- 
west Fisheries  Science  Center,  8604  La  Jolla  Shores  Dr..  La 
Jolla.  CA  92037. 


low  and  Bill  Perrin  for  helpful  reviews  and  analytical 
suggestions. 


Literature  cited 

Archer.  F..  T.  Gerrodette,  A.  Dizon.  K.  Abella  and  S.  Southern. 

2001.  Unobserved  kill  of  nursing  dolphin  calves  in  a  tuna 
purse-seine  fishery      Mar  Mamm.  Sci.  17:540-554. 

Archer.  F..  and  K.  M.  Robertson. 

In  press.  Age  and  length  at  weaning  and  development  of 
prey  preferences  of  pantropical  spotted  dolphins.  Stenella 
attenuata,  from  the  eastern  tropical  Pacific  Mar  Mamm. 
Sci. 

Dizon.  A.  E..  W.  F.  Perrin.  and  P.  A.  Akin. 

1994.  Stocks  of  dolphins  iStenella  spp.  and  Delphmux 
delphis)  in  the  eastern  tropical  Pacific:  a  phylogeographic 
classification.     NOAA  Tech.  Rep.  NMFS-119.  20  p. 

Gerrodette.  T. 

2002.  Tuna-dolphin  issue.  In  Encyclopedia  of  marine 
mammals  (W.  F.  Perrin.  B.  Wursig.  and  J.  G.  M.  Thewis- 
sen,  eds.l.  p.  1269-1273.  Academic  Press.  San  Diego. 
CA. 

Hall.  M.  A.,  L.  A.  Dayton,  and  K.  I.  Metuzals. 

2000.     Bycatch:  problems  and  solutions.     Mar.  Poll.  Bull. 
41:204-19. 
IATTC  i  Inter-American  Tropical  Tuna  Commission). 

2002.     Annual  report  of  the  Inter-American  Tropical  Tuna 
Commission  2000,  171  p.     IATTC.  La  Jolla.  CA. 
Gosliner,  M.  L. 

1999.     The  tuna-dolphin  controversy.     In  Conservation  and 
management  of  marine  mammals  (J.  R.  Twiss  Jr..  and  R. 
R.  Reeves,  eds.l,  p.  120-155.     Smithsonian  Institution 
Press.  Washington  and  London. 
Myrick.  Jr..  A.  C,  A.  A.  Hohn.  J..  Barlow,  and  P.  A.  Sloan. 

1986.     Reproductive  biology  of  female  spotted  dolphins. 
Stenella  attenuata,  from  the  eastern  tropical  Pacific.     Fish. 
Bull.  84:247-259. 
NRC  t  National  Research  Council  I. 

1992.     Dolphins  and  the  tuna  industry.  176  p.     National 
Academy  Press.  Washington.  D.C. 
Perrin.  W.  F. 

1970.     Color  patterns  of  the  eastern  Pacific  spotted  por- 
poise Stenella  graffmani  LbnnbeglCetacea,  DelphinidaeL 
Zoologica   NY)  54:135-149. 
Perrin.  W,  F..  J.  M.  Coe,  and  J.  R.  Zweifel. 

1976.     Growth  and  reproduction  of  the  spotted  porpoise. 
Stenella  attenuata,  in  the  offshore  eastern  tropical  Pa- 
cific.    Fish.  Bull.  74:229-269. 
Smith,  T. 

1983.     Changes  in  size  of  three  dolphin  [Stenella  spp.1 
populations  in  the  eastern  tropical  Pacific.     Fish.  Bull. 
81:1-13. 
Wade,  P.  R. 

1995.  Revised  estimates  of  incidental  kill  of  dolphins  I  Del- 
phinidae1  by  the  purse-seine  tuna  fishery  in  the  eastern 
tropical  Pacific.  1959-1972.     Fish.  Bull.  93:345-354. 


245 


Abstract— Psednos  rossi  new  species 
(Teleostei:  Liparidaei  is  described  from 
two  specimens  collected  in  the  North 
Atlantic  Ocean  off  Cape  Hatteras, 
North  Carolina,  at  depths  of  500- 
674  m.  Psednos  rossi  belongs  to  the 
P.  christinae  group,  which  includes 
six  other  species  and  is  characterized 
by  46-47  vertebrae  and  the  absence 
of  a  coronal  pore.  Psednos  rossi  dif- 
fers from  those  six  species  by  having 
characters  unique  within  the  genus: 
straight  spine,  body  not  humpbacked 
at  the  occiput,  and  a  very  large  mouth 
with  a  vertical  oral  cleft.  Other  distin- 
guishing characters  include  a  notched 
pectoral  fin  with  15-16  rays,  eye 
17-19%  SL,  and  color  in  life  orange- 
rose.  With  P.  rossi,  the  genus  Psednos 
as  currently  known  includes  26  de- 
scribed and  five  undescribed  species  of 
small  meso-  or  bathypelagic  liparids 
from  the  Atlantic.  Pacific,  and  Indian 
Oceans. 


A  remarkable  new  species  of  Psednos 
(Teleostei:  Liparidae)  from  the 
western  North  Atlantic  Ocean 


Natalia  V.  Chernova 

Zoological  Institute 

Russian  Academy  of  Sciences 

Unlversitetskaya  nab-  1 

St.  Petersburg  199034.  Russia 


David  L.  Stein 

NOAA/NMFS  Systematlcs  Laboratory 

Smithsonian  Institution 

P.O.  Box  37012 

National  Museum  of  Natural  History,  MRC-0153 

Washington,  DC.  20013-7012 

E-mail  address  (for  D  L.  Stem,  contact  author):  david.stenvanoaa  gov 


Manuscript  approved  for  publication 
7  January  2004  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:245-250  (2004). 


The  liparid  genus  Psednos  Barnard 
1927  is  a  group  of  meso-  and  bathype- 
lagic snailfishes  distinguished  from  the 
genus  Paraliparis  by  having  the  infra- 
orbital canal  of  the  cephalosensory 
system  interrupted  behind  the  eye  and 
usually  having  a  pronounced  dorsal 
curvature  of  the  spine,  producing  a 
"humpbacked"  body.  Psednos  are  small, 
easily  damaged,  and  often  misidenti- 
fied  as  juvenile  Paraliparis.  Until  1978, 
the  genus  was  known  only  from  two 
specimens  of  a  single  species  (Psednos 
micrurus  Barnard  1927)  collected  off 
Cape  Point,  South  Africa.  Two  addi- 
tional specimens  were  collected  in  the 
southern  Indian  Ocean  and  reported 
by  Stein  (1978).  No  further  specimens 
or  species  were  described  until  Andria- 
shev  (1992)  described  another  new 
species.  Since  then,  active  searches  for 
material  from  collections  around  the 
world  have  yielded  many  specimens 
from  the  Atlantic.  Pacific,  and  Indian 
Oceans.  To  date,  25  species  have  been 
described  (Andriashev,  1992,  199.3; 
Chernova.  2001;  Stein  et  al..  2001; 
Chernova  and  Stein,  2002)  and  an 
additional  five  are  undescribed  (one  in 
Stein  et  al.,  2001,  three  in  Chernova 
and  Stein,  2002,  all  in  poor  condition; 
and  another  that  is  currently  being 
described  by  Stein).  In  this  article,  we 
describe  an  especially  noteworthy  spe- 


cies of  the  genus  from  two  specimens 
collected  from  the  North  Atlantic  off 
Cape  Hatteras,  North  Carolina. 


Materials  and  methods 

All  characters  available  for  both  speci- 
mens were  studied.  Characters  and 
terms  used  were  described  by  Andria- 
shev (1992),  Chernova  (2001),  Stein 
et  al.  (2001),  and  Chernova  and  Stein 
(2002 1.  Counts  were  made  from  a  radio- 
graph of  the  holotype  and  from  each 
specimen  where  possible;  vertebral 
counts  include  the  urostyle.  The  first 
caudal  vertebra  is  that  with  the  haemal 
spine  supporting  the  first  anal-fin  ray. 
The  posterior  tip  of  the  lower  jaw  in 
Psednos  forms  a  distinct  and  promi- 
nent ventrally  directed  angle,  the  ret- 
roarticular  process  (Chernova,  2001). 
Counts  and  proportions  are  given  as  a 
percentage  of  standard  length  ( SL)  and 
head  length  (HL).  Nonstandard  mea- 
surements are  the  following:  distance 
from  mandible  to  anus  (from  ante- 
rior tip  of  mandible  to  center  of  anus); 
distance  from  anus  to  anal-fin  origin 
(from  center  of  anus  to  anal-fin  origin); 
interorbital  width  (measured  between 
upper  margins  of  eyes);  postocular 
head  length  (distance  from  posterior 
margin  of  eye  to  tip  of  opercular  flap). 


246 


Fishery  Bulletin  102(2) 


Figure  1 

Psednos  rossi  n.sp.,  paratype,  USNM  372727.  Adult,  51.8  mm  SL,  57.2  mm  TL.  Sta.  CH-01-047, 
off  Cape  Hatteras.  Scale  5  mm.  Infraorbital  pore  6  not  shown  owing  to  damage. 


We  selected  the  smaller  specimen  to  serve  as  the  holo- 
type  owing  to  its  better  condition  (skin,  pores,  shape  of 
head)  and  the  availability  of  more  characters.  Unfortu- 
nately, it  is  distorted  and  does  not  look  natural;  therefore 
the  undistorted  larger  (adult)  specimen,  the  paratype,  is 
illustrated.  It  is  also  more  useful  to  have  a  drawing  of  an 
adult  for  comparison  with  other  Psednos  specimens. 

In  these  small  fishes,  precise  counts  of  number  of  tooth 
rows  are  possible  only  in  disarticulated  cleared  and 
stained  specimens;  thus,  we  provide  approximate  counts. 
Similarly,  the  drawing  of  the  gill  arch  of  the  paratype  was 
made  without  dissection  by  viewing  through  an  opening  in 
the  branchiostegal  membrane. 

Although  Andriashev  (1986)  and  Andriashev  and 
Stein  (1998)  demonstrated  the  importance  of  the  pectoral 
girdle  in  distinguishing  among  species  and  in  explain- 
ing liparid  relationships,  we  did  not  dissect,  clear,  and 
stain  a  pectoral  girdle  from  these  specimens  owing  to  the 
high  probability  of  damaging  them  and  destroying  other 
characters  (Chernova,  2001;  Chernova  and  Stein,  2002). 
The  new  species  is  so  easily  distinguished  from  congeners 
that  it  is  not  necessary  for  a  diagnosis  of  the  species  to 
look  at  additional  characters  that  the  pectoral  girdle  can 
provide.  Future  specimens  should  be  used  to  study  these 
characters. 

The  holotype  and  paratype  are  permanently  deposited 
in  the  Division  of  Fishes,  Smithsonian  Institution,  Na- 
tional Museum  of  Natural  History  (USNM  collection). 


Results 

Psednos  rossi,  n.sp. 

Holotype 

USNM  372726,  juvenile,  37.2  mm  SL.  TL?,  Sta.  EL-00-033, 
off  Cape  Hatteras  (The  Point),  35°30.036'N,  74°46.497'W, 
500-674  m  over  about  900  m  depth,  23  July  2000,  Tucker 
trawl.  Good  condition  but  distorted. 


Paratype 

USNM  372727,  adult  (sex  not  identified),  51.8  mm  SL, 
57.2  mm  TL,  Sta.  CH-01-047,  off  Cape  Hatteras  (The  Point), 
35°28.93'N,  74°45.93'W,  628-658  m  over  1090-704  m  depth, 
24  Aug.  2001,  Tucker  trawl.  Throat  slightly  damaged,  head 
slightly  compressed,  skin  on  head  partly  missing. 

Diagnosis 

Vertebrae  47,  dorsal-fin  rays  42-44,  coronal  pore  absent. 
Mouth  vertical,  symphysis  of  upper  jaw  above  level  of 
eye.  Body  not  humpbacked,  vertebral  column  not  curved 
behind  cranium.  Gill  cavity  enlarged.  Anus  on  vertical 
behind  head.  Pectoral  fin  notched,  rays  8+2+5-6.  Eye 
17-19%  HL. 

Description 

Counts  and  proportions  are  given  in  Table  1.  Head  large, 
about  one-third  SL,  its  depth  less  than,  and  its  width 
equal  to  or  a  little  greater  than,  its  length  (Fig.  1).  Head 
depth  slightly  greater  than  its  width.  Mouth  very  large, 
distinctly  superior.  Jaws  almost  vertical,  at  angle  of  about 
90°  to  horizontal.  Symphysis  of  upper  jaw  above  level  of 
eye.  Ascending  process  of  premaxilla  horizontal,  its  distal 
end  almost  above  center  of  eye.  Posterior  tip  of  lower  jaw- 
exactly  below  symphysis  of  upper  jaw.  Posterior  (lower) 
end  of  mouth  cleft  well  below  level  of  lower  margin  of  eye. 
When  mouth  closed,  ventral  surface  of  lower  jaw  forms 
entire  frontal  surface  of  head.  Lower  jaw  included.  Sym- 
physeal  process  present  at  lower  jaw  symphysis,  projecting 
forward  prominently;  retroarticular  processes  of  lower 
jaw  large,  acute,  directed  anteroventrally  (Fig.  2  A  I.  Teeth 
large,  sharp,  spear-shaped,  strongly  curved  inward  (Fig. 
2B),  in  (smaller)  holotype  in  approximately  22  and  24  (32 
and  35)  rows  on  upper  and  lower  jaw;  5  (8-9)  teeth  in  first 
full  row  near  symphyses  of  both  jaws.  Snout  short,  1.5  ( 1.0) 
times  eye  diameter.  Olfactory  rosette  (7  lobes)  and  nostril 
above  anterior  third  of  eye.  Eyes  not  large,  close  to  upper 


Chernova  and  Stein:  A  new  species  of  Psednos  from  the  western  North  Atlantic  Ocean 


247 


Table  1 

Counts  and  proportions  for  the  holotype  and  paratype  of  Psednos  rossi  new  species.  Proportions  are  in  %  of  standard  length  (SL) 
followed  by  %  head  length  (HL,  in  parentheses). 


Vertebrae 

Dorsal-fin  rays 

Anal-fin  rays 

Pectoral-fin  rays 

Caudal-fin  rays 

Gill  rakers 

Head  length 

Head  width 

Head  depth 

Body  depth 

Body  depth  at  anal-fin  origin 

Predorsal-fin  length 

Preanal-fin  length 

Mandible  to  anus 

Anus  to  anal  fin  origin 

Upper  pectoral-fin  lobe  length 

Pectoral-fin  notch  ray  length 

Lower  pectoral-fin  lobe  length 

Eye  diameter 

Snout  length 

Interorbital  width 

Postocular  head  length 

Upper  jaw  length 

Lower  jaw  length 

Gill  opening  length 

Opercle  length 


USNM  372726 

Holotype  37.2  mm  SL 

47 

44 

35 

16  [L]  15  [R] 

6 

32.3 

22.0(68.1) 

23.7(73.4) 

21.5(66.6) 

13.4(41.5) 

29.6(91.6) 

47.8(148.0) 

34.9(108.0) 

23.7(73.4) 

13.4(41.5) 

8.1(25.1) 

9.4(29.1) 

5.4(16.7) 

8.1(25.0) 

13.4(42.0) 

18.8(58.0) 

16.1(49.8) 

16.1(49.8) 

5.4(16.7) 

13.4(41.5) 

USNM  372727 
Paratype  51.8  mm  SL 


42 
33 

15  [L,  R] 
6 
10 

29.9 

13.5(45.2) 
17.4(58.2) 
25.1(83.9) 
17.0(56.8) 
26.6(89.0) 
48.3(161.5) 
36.7(122.7) 
21.2(70.9) 
13.5(45.2) 


5.8(19.4) 
7.7(25.8) 
11.2(37.4) 
19.3(64.5) 
12.5(41.8) 
13.5(45.2) 
5.4(18.1) 
12.5(41.8) 


contour  of  head.  Interorbital  space  flat,  2.5  (1.9)  times  eye 
diameter.  Gill  opening  short,  1.0  (0.9)  times  eye  diameter, 
at  45°  angle,  entirely  above  pectoral-fin  base  and  slightly 
anterior  to  it  (distance  between  ventral  end  of  gill  opening 
and  base  of  upper  pectoral  ray  about  equal  to  length  of  gill 
opening).  Opercular  flap  small,  acute.  Opercle  very  long, 
directed  ventrally  and  posteriorly,  its  tip  below  level  of  pos- 
terior end  of  lower  jaw.  Interopercle  of  similar  length,  vis- 
ible externally,  its  anterior  tip  projecting  anteriorly  from 
ventral  contour  of  head  (Fig.  1).  Long  opercle,  interopercle 
and  elongated  branchiostegal  rays  support  membranes 
of  enlarged  branchial  cavity  that  appears  externally  as 
a  black  posterior  part  of  head.  Branchial  cavity  length 
slightly  more  than  half  head  length.  Branchiostegal  rays 
(4+2)  long  and  distinctly  visible  externally.  Gill  rakers 
modified,  closely  but  irregularly  set,  mostly  alternating 
(especially  on  gill  arch  one),  often  paired  on  the  outer 
and  inner  sides  of  each  gill  arch  (central  part  of  arches 
two  and  three);  plates  flattened,  triangular,  similar  in 
shape  to  those  in  P.  pallidus  or  Psednos  sp.l  of  Chernova 


and  Stein  (2002,  Figs.  9  and  13).  spinule-bearing  surface 
directed  internally,  flat  and  longitudinally  oval.  Spinules 
closely  set,  usually  in  two  longitudinal  rows,  each  of  five 
to  eight  spinules,  often  with  a  few  additional  spinules  in 
between  (Fig.  2C). 

Sensory  pores  difficult  to  see  because  of  thin  transpar- 
ent skin  (damaged  in  paratype).  Nasal  pores  2,  the  poste- 
rior on  a  vertical  through  center  of  eye.  Paired  nasal  bones 
(through  which  the  nasal  canals  run)  long,  tubular,  and 
visible  externally.  Coronal  pore  absent.  Lacrimal  bones 
(bearing  anterior  portion  of  infraorbital  canal)  large,  vis- 
ible externally,  slightly  prominent  anteriorly.  Infraorbital 
canal  (better  preserved  in  holotype)  interrupted  behind 
eye,  infraorbital  pores  6  (5+1),  posteriormost  above  poste- 
rior margin  of  eye  (Fig.  2A).  In  paratype,  skin  behind  eye 
missing.  Preoperculomandibular  pores  6  (3  on  lower  jaw 
+  3  on  preopercular  area).  Two  temporal  pores  present:  tx 
a  short  distance  behind  posterior  margin  of  eye,  and  tsb, 
the  suprabranchial  pore,  above  and  in  front  of  gill  opening 
(Fig.  2A). 


248 


Fishery  Bulletin  102(2) 


Pectoral  fin  notched,  of  16  (15)  rays.  Upper  lobe  of  8  (8) 
rays,  the  2  (2)  notch  rays  more  widely  spaced  and  placed 
exactly  at  middle  of  fin  base.  In  holotype,  left  lower  pec- 
toral lobe  with  6,  on  right  5,  rays.  In  paratype,  5  rays  on 
each  side.  Bases  of  lower-lobe  rays  stronger  and  thicker 
than  those  of  upper-lobe  rays.  Level  of  uppermost  pectoral 
ray  below  horizontal  through  lower  end  of  upper  jaw.  Base 
of  pectoral  fin  close  to  vertical,  lowest  ray  almost  directly 
below  uppermost.  Upper-lobe  rays  not  reaching  anal  fin 
origin,  lower-lobe  rays  not  reaching  vertical  through  ends 
of  upper  lobe  rays.  In  holotype,  length  of  notch  rays  1.7 
times  in  upper  pectoral-fin  lobe  length,  lower  pectoral-fin 
lobe  1.4  times  in  it. 

Body  not  humpbacked,  dorsal  contour  of  back  almost 
straight;  spine  horizontal,  its  anterior  end  not  dorsally 


"'Vt 


B 


Figure  2 

Details  of  anatomy  of  Psednos  rossi.  (A)  Cephalic  pores 
and  prominent  features  of  head.  Portions  of  sensory  canals 
passing  through  bones  are  stippled.  N  =  nostril  and  olfac- 
tory rosette;  io  =  infraorbital  pores,  n  =  nasal  pores,  t  = 
temporal  pores;  S  =  symphyseal  knob;  R  =  retroarticular 
process.  (B)  Teeth  of  paratype:  (leftl  frontal  view;  (right) 
lateral  view.  Tooth  length  about  0.25  mm.  (C)  First  gill 
arch  of  paratype,  USNM  372727.  right  side;  view  from 
inside  of  gill  cavity.  Raker  height  about  0.3  mm. 


curved  (Fig.  3).  Neural  spines  of  vertebrae  1-4  neither 
longer  nor  broader  than  those  posterior,  unlike  other  spe- 
cies (see  Fig.  5  in  Chernova,  2001).  Maximum  body  depth 
4.2  (4.0)  times  in  standard  length  and  1.6(1.5)  times  depth 
at  anal-fin  origin.  In  holotype,  occiput  slightly  swollen 
(Fig.  3);  in  paratype,  dorsal  outline  of  head  and  back  in 
front  of  dorsal  fin  origin  almost  perfectly  flat  (Fig.  1),  pos- 
sibly an  age-related  difference.  Abdominal  part  of  body 
long,  preanal  length  almost  half  of  standard  length.  Inter- 
neural  of  first  dorsal-fin  ray  between  neural  spines  3  and 
4.  Dorsal  and  anal  fins  moderately  deep,  maximum  depth 
of  erect  dorsal  fin  in  paratype  8.9  times  in  SL,  2.7  times  in 
head  length  (damaged  in  holotype).  Dorsal  and  anal  fins 
overlapping  about  one-third  of  caudal-fin  length.  Anus  on 
vertical  behind  head,  slightly  behind  base  of  uppermost 
pectoral  ray.  Skin  transparent.  Gelatinous  subcutaneous 
tissue  weakly  developed.  In  holotype  (smaller  specimen) 
body  not  as  deep  and  jaws  longer  than  in  the  paratype 
(larger  specimen).  Differences  in  head  width  and  interor- 
bital  width  are  great  because  head  of  paratype  was  slightly 
compressed  during  fixation.  Other  proportions  similar  to 
those  of  holotype. 

Body  color  in  alcohol  pale;  under  magnification,  slightly 
dusky  blotches  of  dots  present  caudally  in  paratype  and 
absent  in  holotype.  Head  musculature  pale.  Black  perito- 
neum visible  through  body  wall.  Mouth  and  gill  cavities, 
gill  arches,  tongue,  and  both  jaws  black;  gill  rakers  pale. 
Musculature  of  pectoral  girdle  appears  pale  on  lateral 
surface  of  belly.  Color  in  life  orange-rose. 

Distribution 

Western  North  Atlantic  off  Cape  Hatteras,  mesopelagic 
at  depths  of  500-674  m. 

Etymology 

The  patronym  "rossi"  after  Steve  W.  Ross,  who  initially 
notified  us  of  the  captures  and  furnished  the  specimens 
to  us  for  examination. 

Comparative  notes 

Psednos  rossi  seems  to  belong  to  the  P.  christinae  group 
(see  Chernova,  2001;  Chernova  and  Stein,  2002),  includ- 
ing P.  andriashevi,  P.  barnardi,  P.  christinae.  P.  dentatus, 
P.  groenlandicus,  and  P.  harteli.  Species  of  this  group  are 
characterized  by  vertebrae  46-47,  dorsal-fin  rays  38-42, 
anal-fin  rays  33-35,  and  coronal  pore  absent  (versus  the 
P.  micrurus  group  having  vertebrae  40-44.  dorsal-fin  rays 
34-38,  anal-fin  rays  28-33,  and  coronal  pore  present) 
(Chernova,  2001).  Psednos  rossi  distinctly  differs  from  the 
other  species  of  the  christinae  group  in  at  least  having 
occiput  not  swollen  (vs.  greatly  swollen),  not  humpbacked 
because  the  vertebral  column  is  straight  (vs.  humpbacked 
owing  to  the  greatly  curved  anterior  part  of  the  spine), 
mouth  vertical  with  jaws  at  90°  to  horizontal,  symphysis 
of  upper  jaw  above  level  of  eye  (vs.  a  more  or  less  oblique 
mouth  at  an  angle  of  30-45°  and  the  upper  jaw.  symphysis 
on  a  horizontal  with  the  lower  half  of  the  eve);  and  anus 


Chernova  and  Stem:  A  new  species  of  Psednos  from  the  western  North  Atlantic  Ocean 


249 


Figure  3 

Radiograph  of  Psednos  rossi  n.sp.,  holotype.  USNM  372726.  Juvenile,  37.2  mm  SL.  Sta.  EL-00-033,  off  Cape  Hatteras. 


behind  the  head  (vs.  anus  below  the  posterior  third  of  the 
head).  The  very  oblique,  almost  vertical  mouth  occurs 
often  in  species  of  the  P.  micrurus  group,  five  of  which 
have  the  mouth  at  75-85°  to  the  horizontal  (P.  anoderkes, 
P.  cathetostomus,  P.  microps,  P.  mirabilis,  P.  sargassicus). 
However,  they  all  differ  as  described  above. 


Discussion 

The  physical  features  of  Psednos  rossi  are  unique  in  the 
genus.  The  straight  vertebral  column  and  body  are  outside 
the  previous  diagnosis  of  the  genus,  because  all  previously 
known  species  are  humpbacked  owing  to  the  curved  spinal 
column.  Nevertheless,  P.  rossi  clearly  belongs  in  Psednos 
rather  than  Paraliparis  because  it  has  the  other  generic 
characters  of  Psednos  (Chernova,  2001);  particularly, 
its  sensory  canal  system  and  pores  are  of  Psednos  type, 
having  an  interrupted  infraorbital  canal  behind  the  eye. 
We  suggest  that  its  remarkable  body  shape  is  an  extreme 
transformation  of  the  usual  Psednos  body  shape  and  is 
associated  with  the  change  of  the  mouth  from  oblique  and 
of  normal  size  to  vertical  and  very  large.  In  this  process 
the  anterior  movement  of  the  bony  elements  of  the  jaws 
greatly  enlarges  the  branchial  cavity. 

The  morphology  of  Psednos  rossi  invites  speculation 
about  its  ecology.  The  very  large  superior  mouth  with  verti- 
cal jaws,  eyes  located  close  to  the  dorsal  contour  of  the  head 
and  oriented  to  look  forward  and  up,  and  straight  body  sug- 
gest adaptation  to  feeding  on  detritus  and  animals  (such  as 
copepods)  above  it  in  the  water  column.  These  adaptations, 
similar  to  those  of  hatchetfishes  (family  Sternoptychidae), 
are  highly  advantageous  for  a  mesopelagic  mode  of  life. 
Sudden  opening  of  the  very  large  vertical  lower  jaw  could 
produce  a  strong  orobranchial  suction,  simultaneously 
bringing  food  into  the  mouth  and  thus  saving  energy  for 
this  fish,  which  is  probably  a  poor  swimmer. 


Work  over  the  last  several  years  has  made  it  clear  that 
Psednos  species  exist  at  mesopelagic  depths  in  the  North 
Atlantic,  Indian,  North  Pacific,  and  South  Pacific  Oceans. 
We  confidently  expect  discovery  of  additional  species  from 
meso-  and  bathypelagic  waters. 


Acknowledgments 

We  wish  to  thank  S.  W.  Ross,  K.  J.  Sulak,  and  J.  V. 
Gartner  Jr.  for  collecting  the  specimens,  bringing  them 
to  our  attention,  and  loaning  them  to  us  for  description. 
Collections  were  supported  by  the  U.S.  Geological  Survey, 
State  of  North  Carolina,  North  Carolina  Coastal  Reserve 
Program,  and  the  Duke/UNCW  Oceanographic  Consor- 
tium. The  figures  are  drawn  by  the  senior  author,  who 
was  supported  by  the  Russian  Science  Foundation  Grants 
02-04-48669  and  00-15-07794. 


Literature  cited 

Andriashev,  A.  P. 

1986.  Review  of  the  snailfish  genus  Paraliparis  (Scorpaeni- 
formes:  Liparididae)  of  the  Southern  Ocean,  204  p.  The- 
ses Zoologicae  7,  Koeltz  Scientific  Books,  Koenigstein 

1992.  Morphological  evidence  for  the  validity  of  the  anti- 
tropical  genus  Psednos  Barnard  ( Scorpaeniformes,  Lipari- 
didae) with  a  description  of  a  new  species  from  the  eastern 
North  Atlantic.     UO,  Tokyo  41:1-18. 

1993.  The  validity  of  the  genus  Psednos  Barnard  (Scor- 
paeniformes, Liparidae)  and  its  antitropical  distribution 
area.  Vopr.  Ikhtiol.  33(11:5-15  [in  Russian]  J.  Ichthyol. 
33  (5):81-98.  [English  translation.] 

Andriashev,  A.  P.,  and  D.  L.  Stein. 

1998.  Review  of  the  snailfish  genus  Careproctus  (Lipari- 
dae, Scorpaeniformes)  in  Antarctic  and  adjacent  waters. 
Contr.  Sci.  Nat.  Hist  Mus.  Los  Angeles  Cty.  470:1-63. 


250 


Fishery  Bulletin  102(2) 


Chernova.  N.  V. 

2001.  A  review  of  the  genus  Psednos  (Pisces,  Liparidae) 
with  description  often  new  species  from  the  North  Atlantic 
and  southwestern  Indian  Ocean.  Bull.  Mus.  Comp.  Zool. 
155:477-507. 

Chernova,  N.  V.,  and  D.  L.  Stein. 

2002.  Ten  new  species  of  Psednos  (Pisces,  Scorpaeni- 
formes:  Liparidae)  from  the  Pacific  and  North  Atlantic 
Oceans.     Copeia  2002  (3):755-778. 


Stein,  D.  L. 

1978.     The  genus  Psednos  a  junior  synonym  of  Paraliparis, 
with  a  redescription  of  Paraliparis  mierurus  (Barnardi 
(Scorpaeniformes:  Liparidae).     Matsya  4:5-10. 
Stein,  D.  L.,  N.  V.  Chernova,  and  A.  P.  Andriashev. 

2001.  Snailfishes  (Pisces:  Liparidae)  of  Australia,  includ- 
ing descriptions  of  30  new  species.  Rec.  Austr.  Mus. 
53:341-406. 


251 


Abstract— Age  and  growth  of  sailfish 
(Jstiophorus  platypterus)  in  waters  off 
eastern  Taiwan  were  examined  from 
counts  of  growth  rings  on  cross  sections 
of  the  fourth  spine  of  the  first  dorsal  fin. 
Length  and  weight  data  and  the  dorsal 
fin  spines  were  collected  monthly  at  the 
fishing  port  of  Shinkang  (southeast 
of  Taiwanl  from  July  1998  to  August 
1999.  In  total.  1166  dorsal  fins  were 
collected,  of  which  1135  (97r£>  (699 
males  and  436  females)  were  aged  suc- 
cessfully. Trends  in  the  monthly  mean 
marginal  increment  ratio  indicated 
that  growth  rings  are  formed  once  a 
year.  Two  methods  were  used  to  back- 
calculate  the  length  of  presumed  ages, 
and  growth  was  described  by  using 
the  standard  von  Bertalanffy  growth 
function  and  the  Richards  function. 
The  most  reasonable  and  conserva- 
tive description  of  growth  assumes 
that  length-at-age  follows  the  Rich- 
ards function  and  that  the  relationship 
between  spine  radius  and  lower  jaw  fork 
length  ( LJFL I  follows  a  power  function. 
Growth  differed  significantly  between 
the  sexes;  females  grew  faster  and 
reached  larger  sizes  than  did  males. 
The  maximum  sizes  in  our  sample  were 
232  cm  LJFL  for  female  and  221  cm 
LJFL  for  male. 


Age  and  growth  of  sailfish  Ustiophorus  platypterus) 
in  waters  off  eastern  Taiwan 


Wei-Chuan  Chiang 

Chi-Lu  Sun 

Su-Zan  Yeh 

Institute  of  Oceanography 

National  Taiwan  University 

No  1,  Sec.  4,  Roosevelt  Road 

Taipei,  Taiwan  106 

E-mail  address  (for  C  L.  Sun,  contact  author):  chilufiintu  edu.tw 

Wei-Cheng  Su 

Taiwan  Fisheries  Research  Institute 
No.  199,  Ho-lh  Road 
Keelung,  Taiwan  202 


Manuscript  approved  for  publication 
22  December  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102(2):  251-263  (2004). 


The  sailfish  (Istiophorus  platypterus) 
is  distributed  widely  in  the  tropical 
and  temperate  waters  of  the  world's 
oceans.  According  to  data  from  longline 
catches,  sailfish  are  usually  distributed 
between  30°S  and  50°N  in  the  Pacific 
Ocean,  and  highest  densities  are  found 
in  the  warm  Kuroshio  Current  and 
its  subsidiary  currents.  This  species 
has  a  tendency  to  be  found  close  to  the 
coast  and  near  islands  (Nakamura, 
1985).  During  the  1990s  the  annual 
landings  of  sailfish  off  Taiwan  ranged 
between  600  and  2000  metric  tons,  of 
which  approximately  54%  came  from 
waters  off  Taitung  (eastern  Taiwan). 
Sailfish  are  seasonally  abundant  from 
April  to  October  (peak  abundance  from 
May  to  July)  and  contribute  substan- 
tially to  the  economic  importance  of 
the  eastern  coast  of  Taiwan  where  this 
species  is  taken  primarily  by  drift  gill 
nets,  although  they  are  also  caught  by 
set  nets,  harpoons,  and  as  incidental 
bycatch  in  inshore  longline  fisheries. 

Age  and  growth  of  sailfish  caught 
in  recreational  fisheries  in  the  Atlan- 
tic Ocean  have  been  studied  by  using 
various  methods,  including  length- 
frequency  analysis  (de  Sylva,  1957). 
analysis  of  release-recapture  data  (Far- 
ber1),  and  inferences  from  observed 
marks  on  hard  parts,  such  as  spines 
(Jolley,  1974,  1977;  Hedgepeth  and 
Jolley,  1983)  and  otoliths  (Radtke  and 
Dean,  1981;  Radtke,  1983;  Prince  et  al., 
1986).  In  contrast,  very  few  attempts 


have  been  made  to  age  sailfish  in  the 
Pacific  Ocean.  Koto  and  Kodama  ( 1962 ) 
estimated  the  growth  of  sailfish  caught 
with  longlines  from  1952  to  1955  in  the 
East  China  Sea  using  length-frequency 
analysis,  and  Alvarado-Castillo  and  Fe- 
lix-Uraga  (1996,  1998)  used  the  fourth 
spine  of  the  first  dorsal  fin  to  estimate 
age  and  growth  of  sailfish  caught  from 
1989  to  1991  in  the  recreational  fishery 
off  Mexico.  However,  western  Pacific 
sailfish  have  not  been  aged  with  calci- 
fied structures  in  any  previous  study. 

The  aging  of  fishes,  and  consequently 
the  determination  of  their  growth  and 
mortality  rates,  is  an  integral  compo- 
nent of  modern  fisheries  science  (Paul. 
1992).  Mortality  and  growth  rates  pro- 
vide quantitative  information  on  fish 
stocks  and  are  needed  for  stock  assess- 
ment methods  such  as  yield-per-recruit 
and  cohort  analysis  (Powers.  1983). 

The  objectives  of  this  study  were  to 
estimate  age  and  growth  of  sailfish  by 
counting  growth  rings  on  cross  sections 
of  the  fourth  spine  of  the  first  dorsal  fin 
and  to  determine  which  of  the  Richards 
function  and  the  standard  von  Berta- 
lanffy growth  function  best  represents 
growth  of  sailfish  in  waters  off  eastern 


1  Farber,  M.I.  1981.  Analysis  of  Atlantic 
billfish  tagging  data:  1954-1980  Unpubl. 
manuscr.  ICCAT  workshop  on  billfish, 
June  1981.  Southeast  Fisheries  Center 
Miami  Laboratory.  National  Marine  Fish- 
eries Service,  NOAA,  75  Virginia  Beach 
Drive,  Miami,  FL  33149. 


252 


Fishery  Bulletin  102(2) 


120 


125 


130 


135 


140  E 


Figure  1 

Fishing  grounds  of  the  gillnet  (cross  lines)  and  longline  (oblique  lines)  fish- 
ing boats  based  at  Shinkang  fishing  port. 


Taiwan.  This  information  could  be  used  to  determine  the 
age  composition  of  the  catch  and  to  assess  the  status  of 
sailfish  in  these  waters  by  using  yield-per-recruit  or  se- 
quential population  analysis  techniques. 


Materials  and  methods 

Materials 

Data  on  total  length  (TL),  eye  fork  length  (EFL),  lower 
jaw  fork  length  (LJFL)  (in  cm),  round  weight  (RW)  (in  kg) 
and  the  first  dorsal  fins  of  male  and  female  sailfish  were 
collected  monthly  at  the  fishing  port  of  Shinkang  (Fig.  1) 
from  July  1998  to  August  1999.  In  total,  304  TLs,  1166 
LJFLs,  1166  RWs,  and  1166  dorsal  fins  were  collected. 
The  dorsal  fins  were  kept  in  cold  storage  before  being 
boiled  to  remove  surrounding  tissue  and  to  separate  the 
fourth  spines.  Three  cross  sections  (thickness  0.75  mm) 
were  taken  successively  along  the  length  of  each  spine 
with  a  low-speed  "ISOMET"  saw  (model  no.  11-1280)  and 
diamond  wafering  blades,  at  a  location  equivalent  to  1/2  of 
the  maximum  width  of  the  condyle  base  measured  above 
the  line  of  maximum  condyle  width  (Fig.  2A)  (Ehrhardt 
et  al.,  1996;  Sun  et  al.,  2001,  2002).  The  sections  were 


immersed  in  95'/J  ethanol  for  several  minutes  for  cleaning, 
placed  on  disposable  paper  to  air  dry,  and  then  stored  in  a 
labeled  plastic  case  for  later  reading.  Spine  sections  were 
examined  with  a  binocular  dissecting  microscope  (model: 
Leica-MZ6)  under  transmitted  light  at  zoom  magnifica- 
tions of  10-20x  depending  on  the  sizes  of  the  sections.  The 
most  visible  one  of  these  three  sections  was  read  twice, 
approximately  one  month  apart.  If  the  two  ring  counts 
differed,  the  section  was  read  again,  and  if  the  third  ring 
count  differed  from  the  previous  two  ring  counts,  the  spine 
was  considered  unreadable  and  discarded.  The  precision 
of  reading  was  evaluated  by  using  average  percent  error 
(APE)  (Beamish  and  Fournier,  1981;  Campana,  2001)  and 
coefficient  of  variation  (CV)  (Campana,  2001)  statistics. 

Images  of  the  cross  sections  were  captured  by  using  the 
Image-Pro  Image  analysis  software  package  (Media  Cy- 
bernetics, Silver  Spring  MD,  1997)  in  combination  with  a 
dissecting  microscope  equipped  with  a  charged  coupled  de- 
vice (CCD)  camera  (model:  Toshiba  IK-630)  and  a  Pentium 
II  computer  equipped  with  a  640x480  pixel  frame  grab 
card  and  a  high-resolution  (800x600  pixel)  monitor. 

The  distance  from  the  center  of  the  spine  section  to  the 
outer  edge  of  each  growth  ring  was  measured  in  microns 
with  the  Image-Pro  software  package  after  calibration 
against  an  optical  micrometer.  The  center  of  the  spine 


Chiang  et  al.:  Age  and  growth  of  Istiophorus  platypterus  in  waters  off  eastern  Taiwan 


253 


SPINE 
SHAFT 


VASCULARIZED 
CORE 


SECTION 
AREA 

0.75  MM 

CONDYLE 
BASE 


CROSS  SECTION 


Figure  2 

Schematic  diagram  of  the  fourth  dorsal  spine  of  sailfish  [I.  platypterus)  and  the 
location  of  the  cross  section  (A),  and  a  cross  section  showing  the  measurements 
taken  for  age  determination  of  sailfish  (B).  W=  maximum  width  of  condyle  base,  R 
=  radius  of  spine,  r,  =  radius  of  ring  i,  d  =  diameter  of  spine,  dl  =  diameter  of  ring  i. 
The  vascularized  core  and  growth  rings  (1,  2,  3,  4,  5)  are  also  shown. 


section  was  estimated  according  to  the  methods  of  Cayre 
and  Diouf  (1983)  (Fig.  2B).  The  distances  (d,l  were  then 
converted  into  radii  (r,0  by  using  the  equation  (Megalofo- 
nou,  2000;  Sun  et  al.,  2001): 


r;  =  d:  -  W/2), 


where  ri 
d 


radius  of  the  ring  i; 

distance  from  the  outside  edge  of  ring  i  to  the 
opposite  edge  of  the  cross  section;  and 
d    =  diameter  of  the  spine. 


False  growth  rings  were  defined  according  to  criteria 
of  Berkeley  and  Houde  (1983),  Tserpes  and  Tsimenides 
( 1995 ),  and  Ehrhardt  et  al.  ( 1996 ). 


core  of  the  spine.  The  number  of  early  but  missing  growth 
rings  was  therefore  estimated  by  the  replacement  method 
applied  to  Pacific  blue  marlin  (Makaira  nigricatis)by  Hill  et 
al.  (1989).  This  method  involved  first  compiling  ring  radii 
statistics  from  younger  specimens  that  had  at  least  the  first 
or  second  ring  visible.  Radii  of  the  first  four  visible  rings 
from  samples  that  had  missing  early  rings  were  then  com- 
pared with  the  radii  for  these  younger  specimens.  When 
the  radii  of  at  least  two  successive  rings  of  the  first  four 
visible  rings  each  fitted  well  within  one  standard  deviation 
from  the  mean  radii  of  each  of  two  or  more  rings  from  the 
data  compiled  from  the  younger  specimens,  the  number  of 
missing  rings  was  computed  as  the  difference  between  the 
ring  counts  for  the  matched  radii  compiled  from  younger 
specimens  and  those  for  the  specimen  of  interest. 


Accounting  for  missing  early  rings 

The  first  several  growth  rings  of  the  larger  specimens  may 
be  obscured  because  of  the  large  size  of  the  vascularized 


Validation 

The  marginal  increment  ratio  (MIR),  which  was  used 
to  validate  the  rings  as  annuli,  was  estimated  for  each 


254 


Fishery  Bulletin  102(2) 


specimen  by  using  the  following  equation  (Hayashi,  1976. 
Prince  et  al.,  1988;  Sun  et  al.,  2002): 

MIR  =  {R-rn)l{rn-rn_1), 

where     R  =  spine  radius;  and 

rn  and  rnl  =  radius  of  rings  n  and  re— 1. 

The  mean  MIR  and  its  standard  error  were  computed 
for  each  month  by  sex  for  all  ages  combined,  and  also  for 
the  ages  1-5  and  6-11  for  males  and  1-5  and  6-12  for 
females. 

Growth  estimation 

Growth  for  males  and  females  was  estimated  by  back-cal- 
culation of  lengths  at  presumed  ages.  Two  methods  were 
used.  Method  1  was  based  on  the  assumption  that  the  rela- 
tionship between  spine  radius  (R)  and  LJFL  (L)  is  linear, 
i.e., L=a1+61i?  (Berkeley  and  Houde,  1983;  Sun  etal,  2002), 
whereas  method  2  was  based  on  the  assumption  that  this 
relationship  is  a  power  function,  i.e.,  L=a0Rh-  (Ehrhardt, 
1992;  Sun  et  al.,  2002).  The  parameters  of  the  relationships 
were  estimated  by  maximum  likelihood,  assuming  log-nor- 
mally distributed  errors.  Akaike's  information  criterion 
(AIC,  Akaike,  1969)  was  used  to  select  which  of  the  linear 
and  power  functions  best  represented  the  data: 

AIC  =  -21nL  +  2p, 

where  InL  =  logarithm  of  likelihood  function  evaluated 
at  the  maximum  likelihood  estimates  for  the 
model  parameters,  and 
p  =  number  of  model  parameters. 

The  equations  used  to  back-calculate  the  lengths  at 
presumed  ages  were 


where  L,  =  the  mean  LJFL  at  age  t; 
Lx  =  the  asymptotic  length; 


o 


the  hypothetical  age  at  length  zero; 


M/?  <L_a,) 

h, 

5-]   L 

R) 


linear  relationship 
power  relationship 


where  Ln  =  LJFL  when  ring  n  was  formed; 
L    =  LJFL  at  time  of  capture;  and 
rn  =  radius  of  ring  n. 

The  standard  von  Bertalanffy  growth  function  (stan- 
dard VB)  (von  Bertalanffy.  1938)  and  the  Richards  func- 
tion (Richards,  1959)  were  then  fitted  to  the  mean  back- 
calculated  male  and  female  lengths-at-age  from  methods 
1  and  2,  assuming  additive  error. 

Standard  VB: 

L,=L  (l-c'""  ■»), 
Richards  function: 

L(=L.(l-e-K"  '"•)''"'  , 


k  and  A'  =  the  growth  coefficients;  and 

m   =  the  fourth  growth-equation  parameter. 

An  analysis  of  residual  sum  of  squares  lARSS)  was  used  to 
test  whether  the  growth  curves  for  the  two  sexes  were  dif- 
ferent (Chen  et  al.,  1992;  Tserpes  and  Tsimenides.  1995; 
Sun  et  al.,  2001 ),  and  the  log-likelihood  ratio  test  was  used 
to  determine  whether  the  Richards  function  provided  a 
statistically  superior  fit  to  the  data  than  the  length-at-age 
standard  VB  growth  function. 


Results 

Of  the  1166  dorsal  spines  sampled,  1135  (97%)  (699  males 
and  436  females)  were  read  successfully.  The  average  per- 
cent error  (APE)  was  6.31%  (5.91%  for  males  and  6.93%  for 
females)  and  the  coefficient  of  variation  (CV)  was  8.93% 
(8.36%  for  males  and  9.81%  for  females).  Of  the  31  spines 
that  could  not  be  read,  22  were  considered  unreadable 
because  the  existence  of  multiple  rings  made  the  identifi- 
cation of  annuli  difficult  or  resulted  in  aging  discrepancies 
between  readings,  and  the  remaining  nine  spines  were 
unreadable  because  of  abnormal  growth. 

The  length-frequency  and  weight -frequency  distribu- 
tions for  the  1166  individuals  are  shown  in  Figure  3. 
These  individuals  ranged  from  78  to  221  cm  LJFL 
(mean=177.62,  SD  =  16.13,  «=720l  or  1  to  49  kg  RW 
(mean=22.13,  SD  =  5.68)  for  the  males  and  from  80  to  232 
cm  LJFL  (mean=179.96,  SD=17.90,  n  =  446)  or  2  to  58  kg 
RW  (mean=23.65,  SD=7.34)  for  the  females.  The  females 
were  significantly  larger  than  the  males  (r-test,  P<0.05). 
Table  1  summarizes  the  relationships  between  EFL  and 
LJFL  and  TL,  and  that  between  LJFL  and  weight.  The 
latter  relationship  differed  significantly  between  males 
and  females  (analysis  of  covariance;  P<0.05). 

At  least  the  first  or  second  ring  in  417  (60%)  of  male 
spines  and  300  1 69%)  of  female  spines  was  visible.  The 
ring  radii  statistics  by  sex  is  summarized  in  Figure  4.  All 
other  specimens  were  assigned  inner  rings  and  final  age 
estimates  based  upon  these  data.  The  mean  ring  radii  by- 
age  group,  for  males  and  females,  after  correction  for  miss- 
ing early  rings,  are  listed  in  Table  2.  The  maximum  age 
of  the  sampled  sailfish,  after  correction  for  missing  early 
rings,  was  11  years  for  males  and  12  years  for  females. 
The  maximum  ages  before  correction  were  8  years  for 
both  sexes. 

The  monthly  means  of  the  marginal  increment  ratio 
(MIR)  for  males  of  all  ages  during  May-August  were  high 
(-0.72)  but  declined  markedly  thereafter  and  reached  a 
minimum  of  0.46  in  November  (Fig.  5).  Similarly,  the  MIR 
for  females  dropped  from  0.71  in  September  to  a  minimum 
of  0.47  in  November  (Fig.  6).  The  monthly  means  of  MIR 
did  not  differ  significantly  from  each  other  over  the  period 
December-March  (ANOVA,  P^O.86,  P9=0.96).  However, 
the  monthly  means  of  MIR  from  April  through  August  for 
males  and  from  April  through  September  for  females  were 


Chiang  et  al.:  Age  and  growth  of  Istiophorus  p/atypterus  in  waters  off  eastern  Taiwan 


255 


100  -i 


80 


60 


40 


20  - 


□  Male         (n=720) 
■   Female     (n=446) 


p  pH    ,  M  ^ 


L*^ 


75   85   95   105  115  125  135  145  155  165  175  185  195  205  215  225  235 

Lower  jaw  fork  length  (cm) 


120 


90 


60 


30 


r?   f   P  ^ 


2   16   20   24   28   32   36   40   44   48   52   56   60 

Round  weight  (kg) 

Figure  3 

The  size-frequency  distributions  by  5-cm  intervals  (upper  figure)  and  by 
2-kg  intervals  (lower  figure)  for  male  and  female  sailfish  (/.  platypterus) 
collected  from  the  waters  off  eastern  Taiwan. 


significantly  higher  than  those  from  September  through 
November  for  males  U-test,  P<0.001)  and  from  October 
through  November  for  females  (f-test,  P<0.001).  Also,  the 
mean  MIR  in  November  was  significantly  lower  than  that 
in  December  (f-tests,  PCT<0.05,  P9<0.05).  The  trends  in  the 
monthly  means  of  MIR  when  the  data  were  split  into  ages 
1-5  and  6+  were  similar  to  those  for  all  ages  combined. 
The  results  in  Figures  5  and  6  indicate  that  one  growth 
ring  is  formed  each  year,  most  likely  from  September  to 
November  for  males  and  from  October  to  November  for 
females. 

Figure  7  shows  the  sex-specific  relationships  between 
LJFL  and  spine  radius  based  on  method  1  (linear  regres- 
sion) and  method  2  (power  function).  The  relationships  for 
males  and  females  are  significantly  different  (method  1: 


=  56.07,  P<0.01;  method  2:  F, 


-  =  59.93,  P<0.01). 


According  to  AIC,  the  power  function  provides  a  better  fit 
to  the  data  (4AIC  =  38.57  and  30.96  for  males  and  females, 


respectively).  Therefore,  the  most  parsimonious  repre- 
sentation of  the  data  is  the  power  function  with  separate 
parameters  for  males  and  females. 

The  mean  back-calculated  lengths-at-age  obtained  from 
methods  1  and  2  are  listed  in  Table  3.  After  the  first  year 
of  life,  the  growth  rates  of  both  sexes  slow  appreciably. 
However,  females  still  grow  faster  and  consequently  reach 
larger  sizes  than  males.  The  standard  VB  and  the  Rich- 
ards function  for  males  and  females  are  shown  in  Figure  8 
and  the  corresponding  parameter  estimates  are  listed  in 
Table  4.  The  growth  curves  for  males  differ  significantly 
from  those  for  females  (F=99.86  P<0.05  and  P=107.38 
P<0.05  for  the  standard  VB  curve  [methods  1  and  2],  and 
P=144.01  P<0.05  and  F=48.43  P<0.05  for  the  Richards 
function  [methods  1  and  2]).  The  Richards  function  pro- 
vides a  statistically  superior  fit  to  the  data  (log-likelihood 
ratio  test;  P<0.001)  when  method  2  is  used  to  back-calcu- 
late length-at-age  but  not  when  method  1  is  used. 


256 


Fishery  Bulletin  102(2) 


Table  1 

Linear  relationships  (Y=a+bX)  among  total  length  (TL,  cm),  lower  jaw  fork  length 
and  the  log-linear  length-weight  (round  weight,  RW,  kg)  relationships  for  sailfish 
parentheses  are  standard  errors. 

(LJFL,  cml  and  eye  fork  length  (EFL,  cm), 
in  the  waters  off  eastern  Taiwan.  Values  in 

Y 

X 

a 

b 

n 

LJFL  range  (cm)           RW  range  ikg)              r- 

Male 

TL 

LJFL 

19.660 
(6.334) 

1.205 
(0.037) 

184 

78-211                                                       0.854 

TL 

EFL 

24.782 
(6.176) 

1.364 
(0.042) 

184 

78-211                                                    0.854 

EFL 

LJFL 

-5.196 

(0.772) 

0.893 
10.004) 

720 

78-221                                                       0.983 

log10RW 

log10LJFL 

-5.381 
(0.080) 

2.985 
(0.036) 

720 

78-221                            1-46                    0.906 

Female 

TL 

LJFL 

6.728 
(9.351) 

1.286 
(0.055) 

120 

109-210                                                    0.824 

TL 

EFL 

6.754 
(9.505) 

1.489 

(0.064) 

120 

109-210                                                       0.820 

EFL 

LJFL 

-2.209 

(0.802) 

0.876 
(0.004 

446 

80-232                                                    0.989 

log]0RW 

log10LJFL 

-5.338 
(0.103) 

2.970 
(0.0461 

446 

80-232                          2-58                  0.905 

Discussion 

Age  estimate  determined  from  dorsal-fin  spines 

Dorsal-fin  spines  appear  to  be  useful  for  aging  sailfish. 
They  are  easily  sampled  without  reducing  the  economic 
value  of  the  fish  and  can  also  be  read  easily  (the  growth 
rings  stand  out  clearly).  In  contrast,  scales  cannot  be 
used  to  age  sailfish  because  scale  deposition  patterns 
change  as  sailfish  age  (Nakumura,  1985),  and  otoliths  are 
extremely  small  and  fragile  and  are  often  difficult  to  locate 
(Radtke,  1983).  Reading  otoliths  is  more  time  consuming 
and  expensive  than  reading  spines  and  spines  can  also 
be  easily  stored  for  future  re-examination  (Compean- 
Jimenez  and  Bard,  1983;  Sun  et  al„  2001,  2002). 

The  problems  associated  with  the  fin-spine  aging  meth- 
od used  in  this  study  were  the  possible  existence  of  false 
rings  and  the  presence  of  the  vascularized  core  which  can 
obscure  early  growth  rings  in  larger  fish.  These  problems 
were  also  noted  by  Berkeley  and  Houde  (1983),  Hedge- 
pet  h  and  Jolley  (1983),  Tserpes  and  Tsimenides  (1995), 
Megalofonou  (2000),  and  Sun  et  al.  (2001,  2002).  However, 
Tserpes  and  Tsimenides  (1995)  and  Megalofonou  (2000) 
noted  that  experienced  readers  can  overcome  the  problem 
of  multiple  rings  by  determining  whether  the  rings  are 
continuous  around  the  circumference  of  the  entire  spine 
section  and  by  judging  their  distance  from  the  preceding 
and  following  rings.  We  observed  false  rings  in  spines  for 
all  age  classes  larger  than  age  two,  which  we  read  with- 
out problem  by  using  these  guidelines.  The  missing  early 


growth  rings  in  larger  specimens  were  accounted  for  by 
compiling  ring  radii  statistics  for  younger  specimens  for 
which  at  least  the  first  or  second  ring  was  visible  and  by 
comparing  the  radii  of  the  first  several  visible  rings  of  the 
specimens  that  had  missing  early  rings  to  the  mean  radii 
and  standard  deviations  of  the  compiled  data.  Similar  ap- 
proaches for  solving  the  problem  of  missing  rings  have  also 
been  used  for  Pacific  blue  marlin  (Hill  et  al.,  1989). 

Marginal  increment  ratio  (MIR)  analysis  is  the  most 
commonly  applied  method  for  age  validation  (Campana, 
2001).  The  MIR  analysis  conducted  for  sailfish  suggested 
that  one  growth  ring  is  formed  each  year  from  September 
to  November  for  males  and  from  October  to  November  for 
females.  Spawning  for  sailfish  in  the  waters  east  of  Taiwan 
lasts  from  April  through  September  (Chiang  and  Sun-). 
This  is  exactly  the  period  when  growth  is  low,  as  indicated 
by  the  narrow  and  translucent  rings.  Similar  findings 
have  been  reported  for  skipjack  tunatAntoine  et  al.,  1983), 
swordfish  (Ehrhardt,  1992;  Tserpes  and  Tsimenides, 
1995),  and  bigeye  tuna  (Sun  et  al.,  2001).  Although  the 
timing  of  annulus  formation  coincides  with  spawning  sea- 
son for  sailfish  in  the  eastern  Taiwan,  annulus  deposition 


-  Chiang,  W.  C,  and  C.  L.  Sun.  2000.  Sexual  maturity  and  sex 
ratio  of  sailfish  {Istiophorus platypterus)  in  the  eastern  Taiwan 
waters.  Abstracts  of  contributions  presented  at  the  2000 
annual  meeting  of  the  Fisheries  Society  of  Taiwan,  Keelung. 
Taiwan,  16-17  December  2000,  15  p.  The  Fisheries  Society  of 
Taiwan,  199  Hou-Ih  Road,  Keelung,  202  Taiwan. 


Chiang  et  al.:  Age  and  growth  of  Istiophorus  platypterus  in  waters  off  eastern  Taiwan 


257 


123456789 
Ring  number 

Figure  4 

Mean  (±1  SD)  ring  radius  for  male  and  female  sailfish  (/.  platypterus)  collected 
from  the  waters  off  eastern  Taiwan  that  had  at  least  the  first  or  second  ring 
present.  The  numbers  above  the  vertical  bars  are  the  sample  sizes. 


may  also  be  related  to  sailfish  migration  and  environmen- 
tal factors,  as  suggested  by  Sun  et  al.  (2002)  for  swordfish. 
The  MIR  analysis  provides  only  a  partial  age  validation; 
complete  validation  requires  either  mark-recapture  data 
or  the  study  of  known-age  fish  (Beamish  and  McFarlane. 
1983;  Prince  et  al.,  1995;  Tserpes  and  Tsimenides,  1995; 
Sun  et  al.,  2001,  2002). 

Selection  of  a  growth  curve 

Female  sailfish  are  typically  larger  for  similar  ages  in 
males  and  grow  faster  than  males,  and  the  length-weight 
relationship  differs  significantly  between  the  sexes. 
Similar  results  have  been  reported  for  east  Pacific  Ocean 


sailfish  (Hernandez-Herrera  and  Ramirez-Rodriguez, 
1998 ).  Indian  Ocean  sailfish  (Williams,  1970 )  and  Atlantic 
Ocean  sailfish  (Beardsley  et  al.,  1975;  Jolley.  1974,  1977; 
Hedgepeth  and  Jolley,  1983). 

The  Richards  function  appears  to  fit  the  data  better 
than  the  standard  VB  curve  (Fig.  8)  and  provides  a  more 
realistic  description  of  growth  for  animals  of  age  0.  The 
standard  VB  curve  is  commonly  used  to  describe  asymp- 
totic growth  in  fish  but  did  not  fit  the  back-calculated 
lengths  for  fish  younger  than  three  (Table  4,  Fig.  8). 

Further  discussion  of  growth  curves  will  likely  focus 
on  method  2  (i.e.,  a  power  function  relationship  between 
spine  radius  and  LJFL)  because  it  provides  a  better  fit  to 
the  data  than  method  1.  Ehrhardt  (1992),  Ehrhardt  et  al. 


258 


Fishery  Bulletin  102(2) 


3 

z 


u 

^ 

■n 

c 

fl 

" 

- 

cfl 

5   ? 


c 

3 

o 

« 

h(l 

IN 

C 

0/ 

a 
£ 

o 

C 

CS 

,« 

CO 

co 

"S 

I 


& 

.. 

_ 

— 

c 

CD 

[-] 

en 

m 

— 

■S 

c3 
co 

c 

OJ 

— 

1  s. 

™      CO 


c  -2 


X 


X 


X 


X 


«  £ 

5?  t- 

Qj 

Q. 

a> 

o    aj 

s 

N 

cd 

UJ 

CO 

1     / 
to  co 


N    n    n 


o    _    — 

CO      •*      CN      -—      — 
rH      c~      cN      00      CN 


-  rt  to  ra  - 

CO  [^  31  N  ^]  _ 

C£  ^h  rH  rj  CD  ^_ 

*-<  ao  co  o  ai  as 

Q  [^  X  CO  N  CO 

CN  CN  CN  CN  CM  CN 


I   I 


00 

CO 

t> 

00 

00 
I— 1 

ai 

£ 
CO 

iH 

m 

CO 

o 

Tp 

CN 

<N 

CN 

CN 

CN 

CN 

CN 

-     ^     CJ      CO      - 

co    co    in    co    co    « 

h    to    h    h    n    ^ 


CO 

o 

CN 

o 

CN 

o 

o 
o 

CN 

O 

CO 

O0 

CO 

as 

CN 

CN 

CN 

CN 

CM 

rH 

r"1 

-q 

-p 

CN 

CN 
CO 

■* 

o 

1— 1 

p 

TP 

in 

CO 

in 

ao 

00 
TP 

CO 
•* 

CN 

■sf1 

■5C 

.-I     "tf     X      00 
.-.     CO 


00     o 

05      CO 


I   I   I 


I   I   I 


I   I   I 


CN      O)      CO 


ai    ^_ 


oo    ai    o    i-t 


^H        O 


00 

CO 
CO 

in 

in 

- 

CM 

■* 

■0" 

■* 

o 

o 

£ 

CN 

£ 

£ 

05 
CN 

CO 
CN 

o 

CN 

T- 1 

CM 

o 

CM 

^ 

•* 

^p 

■* 

o 

o 

01 

o 

CN 

O 

00 

CO 

Tp 

m 

o 

CO 
CN 

CO 
CN 

■*p 

"* 

CO 

CO 

■»* 

o 

o 

CM 

00 

00 

o 

in 

in 

CO 
CM 

OO 

CO 
CM 

CO 

CO 

CO 

CO 

CO 

CO 

co 

o 

o 

£ 

- 

o 

CO 

5> 
c~ 

CM 
CN 

£ 

CM 

00 

o 
in 

CO 

CO 

01 
O 

o 

00 

CN 
CN 

CM 

CO 

CO 

CO 

CO 

CO 

CO 

CM 

CO 

o 

o 

1  - 

£ 
ao 

CO 
CM 

CM 

0- 

£ 

■* 

t> 

00 

"CP 

CM 

O 

1 

1 

CO 

o 

CN 

CO 

CO 

CO 

CO 

CO 

CO 

CO 

o 

o 

CM      O      O 


CO 

Tp 

CO 

1~i 

CO 

CM 

o 

o 

o 


Tp        O 

r-i    d 


Q     C 


CO 

in 


m 

oo 

T 

C 

CO 

■^ 
-* 

■* 

m 

m 

o 

o 

i> 

3 

5 

in 

m 

00 

CD 
CO 

■* 

m 

CN 

■* 

■<* 

•<* 

Tf 

o 

o 

[  - 

r*- 

»— i 

00 
CO 

CO 
CN 

^p 

ao 

rp 

LO 
CO 

CM 

CO 

in 

CN 

m 

CO 

t- 

3 

3 

CC 

o 

00 

IO 
CM 

CM 

o 

ao 

CN 

CO 
CN 

Tp 

Tp 

CO 

TP 

■* 

TP 

o 

o 

C£  CO  ^H  [>  ^H  y-4 

a  cd  ©  ^j  oi  co 

co  co  t^  ^  co  co 

CO  CO  CO  CO  CO  CO 


o  ^  "^  t—  --^  ^^ 

i-t  CO  CO  i-t  i^-  ^h 

X  O  CO  O  O  "-' 

rt  CD  UC  rr  ^-t  CO 

CO  CO  CO  CO  CO  CO 


00 
CO 

■* 

o 

^p 

CO 

a> 

CN 

o 

o 

CO 

o 

CO 

^p 

00 

in 
ao 

CO 

ao 

CN      CN      CN      CN      CN      CN 


^  -  CD  W  ' 

h~  <X)  ^t  O  CTi  O  -— - 

t-4  CO  .-I  i-i  iO  i-l  Tf 

t-  Oi  CD  UO  iO  O  CD 

tJ-  CO  CO  ^  U0  rj«  CO 

CN  CN  CN  CN  CN  CN  CN 


iH 

^~ 

CO 

— 

t> 

CO 

■>* 

iH 

CO 
CM 

o 
o 

o 
o 

CO 

ao 

m 
o 

CO 

o 

CM 
O 

01 

3 

3 

in 

t— 

CO 

^H 

CN 

3 

in 

00 

in 

tp 

00 

ao 

CN 

in 

CM 

in 

CO 

in 

H        H        1^        CO        CO 


I  I  I 


I  I  I 


I  I  I 


I  I  I  I 


s     _i 


^HCNco^incot^ooaio 


CO  o  o 


CO  o  o 


c    o 

■* 

CO 

CD 

^ 

H       ^ 

CO 

TP 

^    I> 

ao 

r^ 

^     1 

1          ^ 

^* 

CO 

CM 

o 

CO       1 

CM 

CO 

0      CO 

CO 

CO 

CO 

CM 

CO 

o 

o 

co  co  ao 

CO   CM   CO 

cm  d  d 


^p 

in 

CN 

CO 

CN 

O 

o 

CM  o  o 


TP  o 
i-i  d 


3  S 


Chiang  et  al.:  Age  and  growth  of  Ist/ophorus  p/atypterus  in  waters  off  eastern  Taiwan 


259 


09   - 

Male                                                      All  ages  combined 

0.8   - 
0.7   - 
06   - 

f 

43       117                158 
3                   18        Ti^A 

T  /"I       ^^-^r      \  48                      8 
12  S\                               T^  32 

05    - 

I\  17/ 

0.4   - 

o 

|      09   - 

I      08   - 

E 

§      07- 

c 

75      0,6   - 

c 

CD 

£      05   - 
5 

04    - 

i 

T       50      81               Ages  1-5 

/1^\I^^I^    73 

'            A                           1/       ^             T^^^^k         10 

r                              ^v  T                       3 

/                                                    \J           6                     T 

"\            7  /                                        P\  T             A 

i — f                                   r\  7  / 

09   -i 

08    ■ 
07    ■ 
06    ■ 

i 

1              , 

\ 

Ages  6-11 

67                  f 
;       '7      28       T                 I 

l_ — f^^^^T           \    38        26                   T 

05   ■ 

i — T\  10  y\ 

0.4   - 

JFMAMJJASOND 

Month 

Figure  5 

Monthly  means  of  marginal  increment  ratio  for  male  sailfish  (/.  platy- 
pterus)  in  the  waters  off  eastern  Taiwan  for  all  ages  combined  and  for 
age  classes  1-5  and  6-11,  respectively.  Vertical  bars  are  ±1  SE;  numbers 
above  the  vertical  bars  are  sample  sizes. 

Table  3 

Mean  back-calculated  lower  jaw  fork  lengths  at  age  for  sailfish  in  the  waters  off  eastern  Taiwan. 

Back-calculated  length  (cm) 


Method  1 

Method  2 

Age  (yr) 

Method  1 

Method  2 

Age  lyr) 

Male 

Female 

Male 

Female 

Male 

Female 

Male 

Female 

1 

108.53 

113.41 

99.90 

103.51 

7 

181.11 

185.36 

181.86 

186.09 

2 

125.70 

130.79 

121.79 

126.32 

8 

188.99 

192.82 

189.84 

193.67 

3 

138.82 

143.90 

137.27 

141.96 

9 

194.98 

200.60 

196.59 

201.47 

4 

150.80 

156.02 

150.56 

155.54 

10 

200.78 

207.85 

201.74 

208.81 

5 

161.78 

166.22 

162.12 

166.38 

11 

208.05 

213.29 

209.14 

214.66 

6 

171.63 

176.60 

172.18 

177.12 

12 

217.15 

219.05 

260 


Fishery  Bulletin  102(2) 


0.9   I 

Female 

All  ages  combined 

0.8  ■ 
0.7  - 
0.6  • 
0.5   • 

2 

J.          2 

2 

20 

52 

93 

116 

88       25 

1\                             m 

\   19 

NL    10  /i 

04  - 
03   - 

1 

— i — i — i — i — i 

0.9 

08 
0.7   - 
0.6 
05    - 
0.4    - 


0.3 

09 
08 
07 
06 
05 
0.4 
03 
02 


40  Ages  1-5 


- 1 r~ 


—l 1 1 1 


Ages  6-12 


— i 1 1 1 1- 


JFMAMJJASOND 
Month 

Figure  6 

Monthly  means  of  marginal  increment  ratio  for  the  female  sailfish  (/. 
platypterus)  in  the  waters  off  eastern  Taiwan  for  all  ages  combined  and 
for  age  classes  1-5  and  6-12,  respectively.  Vertical  bars  are  ±1  SE;  num- 
bers above  the  vertical  bars  are  sample  sizes. 


(1996),  and  Sun  et  al.  (2002)  favored  method  2  because 
they  believed  it  to  be  more  biologically  realistic.  When 
the  back-calculated  lengths-at-age  are  generated  with 
this  method  the  Richards  function  provides  a  statistically 
superior  fit  to  the  length-at-age  data.  Therefore,  the  pa- 
rameter estimates  for  the  Richards  function  with  method  2 
listed  in  Table  4  are  recommended  as  the  most  appropriate 
for  calculating  the  age  composition  of  sailfish  in  the  waters 
to  the  east  of  Taiwan.  It  is  perhaps  worth  noting  that  the 
tu  values  estimated  for  the  Richards  function  with  method 
2  are  much  closer  to  zero  than  those  estimated  for  the 
Richards  function  with  method  1. 

Comparison  with  previous  studies 

Figure  9  compares  the  age-length  relationships  of  this 
paper  with  those  for  Atlantic  (de  Sylva,  1957;  Hedgepeth 


and  Jolley,  1983;  Farber1)  and  Pacific  sailfish  (Koto  and 
Kodama,  1962;  Alvarado-Castillo  and  Felix-Uraga.  1998). 
De  Sylva  ( 1957 )  and  Koto  and  Kodama  ( 1962 1  used  length- 
frequency  analysis  and  concluded  that  sailfish  are  a  very 
fast  growing  and  short-lived  species.  However,  they  likely 
underestimated  age  and  overestimated  growth  rate  when 
their  results  are  compared  with  those  of  other  more  recent 
studies. 

The  maximum  ages  found  in  this  study  (11  years  for 
males  and  12  years  for  females)  are  close  to  the  maximum 
longevity  of  at  least  13  years  proposed  by  Prince  et  al. 
(1986)  based  on  tagging  data.  Farber1  analyzed  Atlantic 
billfish  tagging  data  and  suggested  that  the  asymptotic 
size  was  essentially  reached  by  age  3  (Hedgepeth  and  Jol- 
ley, 1983).  whereas  the  present  study  found  a  more  gradual 
increase  in  length  with  age.  in  common  with  the  results  of 
Hedgepeth  and  Jolley  (1983). 


Chiang  et  al.:  Age  and  growth  of  Istiophorus  platypterus  in  waters  off  eastern  Taiwan 


261 


250  -I 

Male 

200  - 

p 

aij*||»|Pe§>o' 

ylliliP^0    * 

150  ■ 

0o     >^P^ 

_^l°J  - 

100  ■ 

°                               n  =  699 

o                                 --  LJFL  =64.825  +  30.471  R 

r2  =  0.704 

-J         50  ■ 
ll 

-J 

—  LJFL  =  79.833  R06'2 
r2  =  0.720 

length 
o 

i                     i                     i                     i                     i                     i 

1       25°- 

Female 

S 

TO, 

S       200  - 
o 

iJ0^^ 

150  - 

<0^ 

100  - 

o 

0 

n  =436 

o                           --  LJFL  =70.31 2  +  30.093  R 

r  =  0.731 

50  - 

—  LJFL  =83.461  R0596 

r2  =  0.750 

0                    12                    3                    4                    5                    6 

Spine  radius  (R,  mm) 

Figure  7 

Relationship  between  lower  jaw  fork  length  and  spine  radius  for 

male  and  female  sailfish  (/.  platypterus)  in  the  waters  off  eastern 

Taiwan. 

Table  4 

Parameter  estimates  and  standard  errors  (in 

parenthesis)  for  the  standai 

•d  von  Berta 

anffy  growth  function  and  the  Richards 

function  for  sailfish  in  the  waters  off  eastern  Taiwan. 

Standard  von  Bertalan 

ffy  growth 

function 

Richards 

function 

Method  1 

Method  2 

Method  1 

Method  2 

Parameter                 Male 

Female 

Male 

Female 

Male 

Female 

Male 

Female 

L,                             252.6 

261.4 

240.4 

250.3 

271.8 

280.4 

294.0 

343.8 

(3.652) 

(3.397) 

(3.794) 

(4.278) 

(22.713) 

(19.882) 

(29.607) 

(47.921) 

k                                 0.115 

0.110 

0.145 

0.138 

(0.005) 

(0.004) 

(0.008) 

(0.008) 

t0                              -3.916 

-4.207 

-2.781 

-2.990 

-2.473 

-2.608 

-0.704 

-0.468 

(0.143) 

(0.147) 

(0.154) 

(0.186) 

(0.931) 

(0.896) 

(0.279) 

(0.186) 

A' 

0.051 

0.049 

0.023 

0.011 

(0.034) 

(0.030) 

(0.013) 

(0.007) 

m 

-0.551 

-0.578 

-1.288 

-1.639 

(0.472) 

(0.436) 

(0.308) 

(0.243) 

262 


Fishery  Bulletin  102(2) 


250 


H    200 


150- 


100- 


50- 


Male 


Standard  VB  -  method  1 
Standard  VB  -  method  2 
Richards  function  -  method  1 
Richards  function  -  method  2 


zsu- 

Female 

o 

200- 

i  L-4^**^^ 

150- 

100- 

4/ 

Standard  VB- method  1 

-  -  -  -  Standard  VB  -  method  2 

50- 

■'    **/          / 

Richards  function  •  method  1 

f              | 

Richards  function  -  method  2 

n  - 

1 

-5    -4    -3    -2    -1     0 


1      2     3     4     5     6     7     8     9    10   11    12  -5-4-3-2-1 

Age  (year) 


9    10   11    12 


Figure  8 

Observed  and  back-calculated  length-at-age  and  standard  von  Bertalanffy  and  Richards  function  model-predicted  growth  curves 
for  male  and  female  sailfish  <■!.  platypterus)  in  the  waters  off  eastern  Taiwan. 


300  -i 


250 


200 


150 


100  - 


50  - 


-de  Sylva  (1957)  -sexes  combined' 

-  Koto  and  Kodama  (1 962)  -  sexes  combined' 

-  Farber  (1981)  -  sexes  combined* 

-  Hedgepeth  and  Jolley  (1 983)  -  male" 

-  Hedgepeth  and  Jolley  (1 983)  -  female' 

-  Alvarado-C  and  Felix-U.  (1998)  -  sexes  combined 

-  Present  study  -  male 

-  Present  study  -  female 


6         7         8 
Age  (year) 


10 


12       13 


Figure  9 

A  comparison  of  the  growth  curves  for  sailfish  (/.  platypterus)  esti 
by  different  authors.  I  ■  Data  from  Table  1  of  Hedgepeth  and  Jolley 


mated 
1983.1 


use  in  stock  assessments  of  the  sailfish  popu- 
lation in  the  western  Pacific  Ocean. 


Acknowledgments 

The  authors  express  sincere  gratitude  to 
Andre  Punt,  School  of  Aquatic  and  Fishery 
Sciences.  University  of  Washington,  for  his 
valuable  comments  and  comprehensive  edit- 
ing of  the  manuscript.  This  study  was  in 
part  supported  financially  by  the  "Fisheries 
Agency,  Council  of  Agriculture,  Taiwan," 
through  grant  91AS-2.5.1-FK7)  to  Chi-Lu 
Sun. 


Literature  cited 


Even  though  the  aging  method  used  in  the  present  study 
is  the  same  as  that  of  Hedgepeth  and  Jolley  (1983)  and 
Alvarado-Castillo  and  Felix-Uraga  (1998),  there  are  nev- 
ertheless differences  in  the  estimated  length-at-age.  This 
difference  could  be  due  to  spatial  differences  in  growth, 
the  range  of  ages  and  sizes  used  in  the  analysis,  or  the 
form  of  the  growth  model  applied.  The  size  range  in  the 
present  study  is  broader  than  those  in  previous  studies  and 
the  growth  curve  is  based  on  the  Richards  function  rather 
than  the  standard  VB  function.  Therefore,  we  believe  that 
our  growth  parameter  estimates  are  more  appropriate  for 


Akaike.  H. 

1969.     Fitting  autoregressive  models  for  pre- 
diction.    Ann.  Inst.  Stat.  Math.  21:243- 
247. 
Alvarado-Castillo,  R.  M..  and  R.  Felix-Uraga. 

1996.     Age  determination  in  Istiophorus 
platypterus  (Pisces:  Istiophoridael  in  the 
south  of  the  Gulf  of  California,  Mexico.     Rev.  Biol.  Trop. 
44:233-239. 
1998.     Growth  of Istiophorus  platypterus  (Pisces:  Istiophori- 
dael from  the  south  of  the  Gulf  of  California.     Rev.  Biol. 
Trop.  46:115-118. 
Antoine,  L.  M..  J.  J.  Mendoza,  and  P.  M.  Cayre. 

1983.     Progress  of  age  and  growth  assessment  of  Atlantic 
skipjack  tuna.  Euthynnus  pelamis.  from  dorsal  fin  spines. 
NOAA  Tech.  Rep.  NMFS  8:91-97. 
Beamish,  R.  J.,  and  D.  A.  Fournier. 

1981.     A  method  for  comparing  the  precision  of  a  set  of  age 
determinations.     Can.  J.  Fish.  Aquat.  Sci.  38:982-983. 


Chiang  et  al.:  Age  and  growth  of  Istiophorus  platypterus  in  waters  off  eastern  Taiwan 


263 


Beamish,  R.  J.,  and  G.  A.  McFarlane. 

1983.     Validation  of  age  determination  estimates,  the  forgot- 
ten requirement.     NOAA  Tech.  Rep.  NMFS  8:29-33. 
Beardsley,  G.  L„  N.  R.  Merrett,  and  W.  J.  Richards. 

1975.  Synopsis  of  the  biology  of  sailfish,  Istiophorus  pla- 
typterus (Shaw  and  Nodder,  1791).  NOAA  Tech.  Rep. 
NMFS  SSRF  675:95-120. 

Berkeley,  S.  A.,  and  E.  D.  Houde. 

1983.     Age  determination  of  broadbill  swordfish,  Xiphias 
gladius.  from  the  Straits  of  Florida,  using  anal  fin  spine 
sections.     NOAA  Tech.  Rep.  NMFS  8:137-146. 
Campana,  S.  E. 

2001.     Accuracy,  precision  and  quality  control  in  age  deter- 
mination, including  a  review  of  the  use  and  abuse  of  age 
validation  methods.     J.  Fish.  Biol.  59:197-242. 
Cayre,  P.  M„  and  T.  Diouf. 

1983.     Estimated  age  and  growth  of  little  tunny,  Euthyn- 
nus  alletteratus,  off  the  coast  of  Senegal,  using  dorsal  find 
spine  sections.     NOAA  Tech.  Rep.  NMFS  8:105-110. 
Chen,  Y.,  D.  A.  Jackson,  and  H.  H.  Harvey. 

1992.     A  comparison  of  von  Bertalanffy  and  polynomial 
functions  in  modelling  fish  growth  data.     Can.  J.  Fish. 
Aquat.  Sci.  49:1228-1235. 
Compean-Jimenez,  G.,  and  F  X.  Bard. 

1983.     Growth  of  increments  on  dorsal  spines  of  eastern 
Atlantic  bluefin  tuna,  Thunnus  thynnus,  and  their  pos- 
sible relation  to  migratory  patterns.     NOAA  Tech.  Rep. 
NMFS  8:77-86. 
de  Sylva,  D.  P. 

1957.     Studies  on  the  age  and  growth  of  the  Atlantic  sail- 
fish,  Istiophorus  americanus  (Cuvier),  using  length-fre- 
quency curves.     Bull.  Mar.  Sci.  Gulf  Caribb.  7:1-20. 
Ehrhardt,  N.  M. 

1992.     Age  and  growth  of  swordfish,  Xiphias  gladius,  in 
the  northwestern  Atlantic.     Bull.  Mar.  Sci.  50:292-301. 
Ehrhardt,  N.  M.,  R.  J.  Robbins,  and  F.  Arocha. 

1996.     Age  validation  and  estimation  of  growth  of  sword- 
fish,  Xiphias  gladius,  in  the  northwest  Atlantic.     ICCAT 
(International  Commission  for  the  Conservation  of 
Tunas)  Col.    Vol.  Sci.  Pap.  45(21:358-367. 
Hayashi,  Y. 

1976.  Studies  on  the  red  tilefish  in  the  east  China  sea — I. 
A  fundamental  consideration  for  age  determination  from 
otoliths.     Bull.  Jap.  Soc.  Sci.  Fish.  42:1237-1242. 

Hedgepeth,  M.  Y,  and  J.  W.  Jolley  Jr. 

1983.     Age  and  growth  of  sailfish,  Istiophorus  platypterus, 
using  cross  section  from  the  fourth  dorsal  spine.     NOAA 
Tech.  Rep.  NMFS  8:131-135. 
Hernandez-Herrera,  A.,  and  M.  Ramirez-Rodriguez. 

1998.     Spawning  seasonality  and  length  at  maturity  of 
sailfish  (Istiophorus  platypterus)  off  the  Pacific  coast  of 
Mexico.     Bull.  Mar.  Sci.  63:459-467. 
Hill,  K.  T.,  G.  M.  Cailliet,  and  R.  L.  Radtke. 

1989.     A  comparative  analysis  of  growth  zones  in  four 
calcified  structures  of  Pacific  blue  marlin,  Makaira 
nigricans.     Fish.  Bull.  87:829-843. 
Jolley,  J.  W„  Jr. 

1974.  On  the  biology  of  Florida  east  coast  Atlantic  sailfish, 
(Istiophorus  platypterus).  NOAA  Tech.  Rep.  NMFS  SSRF 
675:81-88. 

1977.  The  biology  and  fishery  of  Atlantic  sailfish,  Istiopho- 
rus platypterus,  from  southeast  Florida.  Fla.  Mar.  Res. 
Publ.  28,  31  p. 

Koto,  T.,  and  K.  Kodama. 

1962.     Some  considerations  on  the  growth  of  martins,  using 


size-frequencies  in  commercial  catches.  I.  Attempts  to 
estimate  the  growth  of  sailfish.     Rep.  Nankai.  Reg.  Fish. 
Res.  Lab.  15:97-108. 
Megalofonou,  P. 

2000.  Age  and  growth  of  Mediterranean  albacore.  J. 
Fish.  Biol.  57:700-715. 

Nakamura,  I. 

1985.  FAO  species  catalogue.  Billfish  of  the  world.  An 
annotated  and  illustrated  catalogue  of  marlins,  sailfishes, 
spearfishes,  and  swordfishes  known  to  date.  FAO  Fish. 
Synop.  125(5),  65  p 

Paul.  L.  J. 

1992.     Age  and  growth  studies  of  New  Zealand  marine 
fisheries,  1921-90:  a  review  and  bibliography.     Aust.  J. 
Mar.  Freshw.  Res.  43:879-912. 
Powers,  J.  E. 

1983.     Some  statistical  characteristics  of  ageing  data  and 
their  ramification  in  population  analysis  of  oceanic  pelagic 
fishes.     NOAA  Tech.  Rep.  NMFS  8:19-24. 
Prince,  E.  D.,  D.  W.  Lee,  C.  A.  Wilson,  and  S.  A.  Berkeley. 

1988.     Use  of  marginal  increment  analysis  to  validate  the 
anal  spine  method  for  ageing  Atlantic  swordfish  and  other 
alternatives  for  age  determination.     ICCAT  Col.  Vol.  Sci. 
Pap.  27:194-201. 
Prince,  E.  D..  D.  W.  Lee,  C.  A.  Wilson,  and  J.  M.  Dean. 

1986.  Longevity  and  age  validation  of  a  tag-recaptured 
Atlantic  sailfish,  Istiophorus  platypterus,  using  dorsal 
spines  and  otoliths.     Fish.  Bull.  84:493-502. 

Prince,  E.  D.,  D.  W.  Lee,  C.  J.  Cort,  G.  A.  McFarlane, 
and  A.  Wild. 

1995.  Age  validation  evidence  for  two  tag-recaptured  Atlan- 
tic albacore,  Thunnus  alalunga,  based  on  dorsal,  anal, 
and  pectoral  finrays,  vertebrae,  and  otoliths.  In  Recent 
developments  in  fish  otolith  research  (D.  H.  Sector,  J.  M. 
Dean,  and  S.  E.  Campana,  eds.),  p.  375-396.  Univ.  South 
Carolina  Press,  Columbia,  SC. 
Radtke,  R.  L. 

1983.     Istiophorid  otoliths:  extraction,  morphology,  and  pos- 
sible use  as  ageing  structures.     NOAA  Tech.  Rep.  NMFS 
8:123-129. 
Radtke,  R.  L.,  and  J.  M.  Dean. 

1981.     Morphological  features  of  the  otoliths  of  the  sailfish, 
Istiophorus  platypterus,  useful  in  age  determination.     Fish. 
Bull.  79:360-367. 
Richards,  F.  J. 

1959.     A  flexible  growth  function  for  empirical  use.     J.  Exp. 
Bot.  10:290-300. 
Sun,  C.  L.,  C.  L.  Huang,  and  S.  Z.  Yeh. 

2001.  Age  and  growth  of  the  bigeye  tuna,  Thunnus  obesus, 
in  the  western  Pacific  Ocean.     Fish.  Bull.  99:502-509. 

Sun,  C.  L.,  S.  P.  Wang,  and  S.  Z.  Yeh. 

2002.  Age  and  growth  of  swordfish  (Xiphias  gladius  L.,) 
in  the  waters  around  Taiwan  determined  from  anal-fin 
rays.     Fish.  Bull.  100:822-835. 

Tserpes,  G.,  and  N.  Tsimenides. 

1995.     Determination  of  age  and  growth  of  swordfish, 
Xiphias  gladius  L.,  1758,  in  the  eastern  Mediterranean 
using  anal-fin  spines.     Fish.  Bull.  93:549-602. 
von  Bertalanffy,  L. 

1938.     A  quantitative  theory  of  organic  growth  (Inquiries 
on  growth  laws.  II).     Hum.  Biol.  10:181-213. 
Williams,  F. 

1970.  The  sport  fishery  for  sailfish  at  Malindi,  Kenya, 
1958-1968,  with  some  biological  notes.  Bull.  Mar.  Sci. 
20(4):830-852. 


264 


Abstract— A  density  prediction  model 
for  juvenile  brown  shrimp  (Farfan- 
tepenaeus  aztecus)  was  developed  by 
using  three  bottom  types,  five  salinity 
zones,  and  four  seasons  to  quantify  pat- 
terns of  habitat  use  in  Galveston  Bay, 
Texas.  Sixteen  years  of  quantitative 
density  data  were  used.  Bottom  types 
were  vegetated  marsh  edge,  submerged 
aquatic  vegetation,  and  shallow  non- 
vegetated  bottom.  Multiple  regression 
was  used  to  develop  density  estimates, 
and  the  resultant  formula  was  then 
coupled  with  a  geographical  informa- 
tion system  (GIS)  to  provide  a  spatial 
mosaic  (map)  of  predicted  habitat  use. 
Results  indicated  that  juvenile  brown 
shrimp  (<100  mm)  selected  vegetated 
habitats  in  salinities  of  15-25  ppt  and 
that  seagrasses  were  selected  over 
marsh  edge  where  they  co-occurred. 
Our  results  provide  a  spatially  resolved 
estimate  of  high-density  areas  that  will 
help  designate  essential  fish  habitat 
(EFH)  in  Galveston  Bay.  In  addition, 
using  this  modeling  technique,  we  were 
able  to  provide  an  estimate  of  the  over- 
all population  of  juvenile  brown  shrimp 
(<100  mm)  in  shallow  water  habitats 
within  the  bay  of  approximately  1.3 
billion.  Furthermore,  the  geographic 
range  of  the  model  was  assessed  by 
plotting  observed  (actual)  versus 
expected  (model)  brown  shrimp  densi- 
ties in  three  other  Texas  bays.  Similar 
habitat-use  patterns  were  observed 
in  all  three  bays — each  having  a  coef- 
ficient of  determination  >0.50.  These 
results  indicate  that  this  model  may 
have  a  broader  geographic  application 
and  is  a  plausible  approach  in  refining 
current  EFH  designations  for  all  Gulf 
of  Mexico  estuaries  with  similar  geo- 
morphological  and  hydrological  char- 
acteristics. 


A  habitat-use  model  to  determine  essential 
fish  habitat  for  juvenile  brown  shrimp 
(Farfantepenaeus  aztecus)  in  Galveston  Bay,  Texas 


Randall  D.  Clark 

John  D.  Christensen 

Mark  E.  Monaco 

Biogeography  Program 

Center  for  Coastal  Monitoring  and  Assessment 

National  Center  for  Coastal  Ocean  Science 

National  Ocean  Service.  NOAA 

Silver  Spring,  Maryland  20910 

E-mail  address  (For  R.  D  Clark)  Randy  Clarkfflnoaa  gov 

Philip  A.  Caldwell 

Geoffrey  A.  Matthews 

Thomas  J.  Minello 

Fishery  Ecology  Branch 

Galveston  Southeast  Fisheries  Science  Center  Laboratory 

National  Marine  Fisheries  Service,  NOAA 

Galveston,  Texas  77550 


Manuscript  approved  for  publication 
22  December  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:264-277  (200  1 1 


Shallow  estuarine  habitats,  whose  com- 
plexity promotes  survival  and  growth, 
are  used  by  many  young  fish  and  macro- 
invertebrate  species  (Boesch  and 
Turner,  1984).  A  complete  understand- 
ing of  how  these  habitats  sustain  spe- 
cies productivity  is  unknown  and  has 
become  a  focal  point  of  federal  fishery 
management  programs.  The  National 
Marine  Fisheries  Service  (NMFS) 
has  developed  guidelines  to  identify 
essential  fish  habitat  (EFH)  for  all 
federally  managed  species  based  on 
four  levels  of  available  information 
that  encompass  the  ecological  linkages 
between  habitats  and  fishery  produc- 
tion. Examination  of  habitat-use  pat- 
terns (habitat-related  densities)  are 
needed  to  determine  which  habitats 
are  likely  to  be  most  essential.  These 
patterns  are  measurable  and  can  be 
reasonable  indicators  of  habitat  value. 
Relative  habitat  values  have  been  esti- 
mated by  comparing  animal  densities 
under  the  assumption  that  high  densi- 
ties reflect  greater  habitat  quality  and 
preferred  habitat  (Pearcy  and  Myers, 
1974;  USFWS,  1981;  Zimmerman  and 
Minello,  1984;  Sogard  and  Able,  1991; 
Baltzetal.,  1993). 


Considerable  bottom-type  variation 
exists  in  northern  Gulf  of  Mexico  estu- 
aries, including  intertidal  marsh,  sub- 
merged aquatic  vegetation,  oyster  reef, 
mangroves,  tidal  mudflats,  and  sub- 
tidal  bay  bottom.  Within  each  of  these 
habitats,  environmental  and  structural 
gradients  may  affect  the  functional 
role  or  importance  of  these  habitats 
for  particular  species.  To  understand 
these  relationships,  fisheries  indepen- 
dent monitoring  (FIM)  data  are  needed 
to  determine  species-habitat  affini- 
ties that  provide  evidence  that  not  all 
habitats  are  of  equal  importance  for  the 
maintenance  of  a  population  (Monaco 
et  al.,  1998;  Minello  1999;  Beck  et  al., 
2001).  Habitat  affinities  may  change 
with  spatial  and  temporal  fluctuations 
of  environmental  variables,  such  as  sa- 
linity and  temperature  (Copeland  and 
Bechtel,  1974;  Baltz  et  al.,  1998). 

In  this  study  we  developed  predictive 
models  that  estimate  brown  shrimp 
{Farfantepenaeus  aztecus,  formerly 
Penaeus  aztecus  [see  Perez-Farfante 
and  Kensley,  1997])  habitat-use  pat- 
terns and  interactions  as  a  function 
of  density-independent  processes  in 
Galveston  Bay,  Texas.  Previous  com- 


Clark  et  al.:  A  habitat-use  model  for  juvenile  Farfantepenaeus  aztecus  In  Galveston  Bay 


265 


29  5°  N 


29°  N 


North 

A 

■i- 

%&*% 

y 

Trinity 
Bay 

C^l            Galveston             ^ 
Bay      "^*^- 
East 

<_k  ■:■  • 

i. 

j? 

v. ■  ' 

.-  '■ 

...  r'    v     . 

\. 

)  7^ 

'  i     j& 

m 

vf- 

*-\Christmas 

Kilometers 
i 1 

Bay 

0                  10 

95    W 

Figure  1 

Map  of  Galveston  Bay,  Texas. 


94.5° W 


parisons  of  brown  shrimp  densities  among  different  bot- 
tom types  in  Louisiana  and  Texas  estuaries  have  been 
conducted  within  limited  temporal  and  spatial  scales 
(Peterson  and  Turner,  1994;  Zimmerman  et  al.,  1984;  Zim- 
merman et  al.,  1990b;  Rozas  and  Minello,  1998;  Minello, 
1999). 

Our  work  expands  upon  these  studies  by  developing  a 
multivariate  bottom-type  use  and  environmental  model 
incorporated  into  a  geographic  information  system  (GISi 
that  provides  a  spatial  assessment  of  habitat  use.  In  ad- 
dition, the  model  is  designed  to  be  transferable  to  other 
northern  Gulf  of  Mexico  estuaries  and  thus  would  allow 
fishery  managers  to  identify  the  relative  importance  of 
habitat  types  for  population  maintenance  and  recruitment 
into  the  fishery. 


Materials  and  methods 

Geographic  setting 

The  Galveston  Bay  complex  (Fig.  1)  encompasses  approxi- 
mately 2020  km2  and  is  one  of  the  largest  estuaries  in 
the  northern  Gulf  of  Mexico  (NOAA,  1989).  Comprising 
several  major  embayments,  including  Trinity,  Galveston, 
East,  and  West  bays,  the  complex  contains  many  smaller 
interconnecting  subbays,  rivers,  streams,  tidal  creeks, 
wetlands,  reefs,  and  tidal  flats  around  its  periphery. 


The  bay  bottom  is  mostly  flat  and  shallow  (mean  depth 
is  approximately  2  m)  and  has  slightly  elevated  oyster 
reefs,  elevated  dredge  material  areas,  river  channels,  and 
deeper  dredged  navigation  channels. 

Data  collection 

Sixteen  years  (1982-97)  of  brown  shrimp  density  data 
were  analyzed  to  quantify  areas  of  potential  EFH.  A  total 
of  46,080  brown  shrimp  were  captured  during  this  time 
period  with  a  mean  total  length  of  27.5  mm  (Fig.  2).  Data 
from  published  studies  by  Czapla  (1991),  Minello  et  al. 
( 1991),  Minello  and  Zimmerman  ( 1992),  Minello  and  Webb 
( 1997 ),  Rozas  and  Minello  ( 1998 ),  Zimmerman  et  al.  (1984, 
1989,  1990a,  1990b),  Zimmerman  and  Minello  (1984), 
and  various  unpublished  sources  from  the  Galveston 
Laboratory  of  the  National  Marine  Fisheries  Service  were 
combined  to  comprise  a  comprehensive  density  database  of 
associated  bottom-type  and  environmental  data  that  would 
support  model  development  and  GIS  analyses.  All  samples 
were  collected  by  using  a  drop  trap  sampler,  described  in 
Zimmerman  et  al.  (1984),  which  employs  large  cylinders 
( 1.0  or  2.6  m2  area)  released  from  a  boom  affixed  to  a  boat 
to  entrap  organisms.  This  quantitative  technique  samples 
fishes  and  macro-invertebrates  in  highly  structured  shal- 
low-water habitats  such  as  salt  marshes,  seagrass  beds, 
and  oyster  reefs  where  the  efficiency  of  conventional  trawl 
and  bag-seine  gear  is  diminished. 


266 


Fishery  Bulletin  102(2) 


Habitat  mapping 

The  underlying  spatial  framework  for  incorporating 
model  predictions  into  the  GIS  consisted  of  six  maps: 
four  salinity  periods,  one  bathymetric  map,  and  one 
map  defining  bottom-type  distribution.  All  GIS  maps 
were  developed  in  Universal  Transverse  Mercator 
projection,  UTM,  datum-1983,  zone-15,  using  ArcView 
3.1  (Redlands,  CA)  software.  Each  map  consisted  of 
10  x  10  m  grid  cells  where  each  cell  contained  pertinent 
salinity,  depth,  or  bottom-type  information. 

Salinity  maps  were  developed  from  depth-aver- 
aged salinity  models  by  using  historical  Galveston 
Bay  data  collected  during  1979-90  (Orlando  et  al., 
1993).  Four  salinity  periods  were  identified  to  rep- 
resent typical  salinity  conditions  under  average  sea- 
sonal freshwater  inflow:  low  (March-June),  increas- 
ing (July),  high  (August-October),  and  decreasing 
(November-February).  Five  isohalines  were  developed  to 
display  spatial  salinity  distribution  (Christensen  et  al.1): 
0-0.5,  0.51-5,  5.1-15,  15.1-25,  and  >25  parts  per  thou- 
sand (ppt)  (Fig.  3). 

Bottom  types  from  the  drop  sample  database  were  di- 
vided into  three  categories: 


Marsh  edge  (ME) 


Submerged 
aquatic 
vegetation  (SAV) 


Shallow  non- 
vegetated 

bottom  (SNB) 


intertidal  marsh  within  5  meters  of 
open  water  habitat.  This  category 
consisted  primarily  of  saltmarsh  cord 
grass  (Spartina  alterniflora),  and 
smaller  proportions  of  salt  meadow- 
grass  {Spartina  patens),  black  needle- 
rush  {Juncus  roemerianus),  salt  grass 
(Distichlis  spicata),  bullrushes  iScir- 
pus  spp.),  and  cattails  (Typha  spp.); 

consisted  primarily  of  shoalgrass 
{Halodule  wrightii),  wigeongrass 
iRuppia  maritima),  and  a  sporadic 
distribution  of  wild  celery  (Vallisneria 
americana); 

generally  restricted  to  waters  less 
than  1  meter  deep,  including  creeks, 
ponds,  shoreline,  and  open  bay  habitat. 


Density  data  for  other  bottom  types  were  limited  and  were 
not  used  in  the  analysis. 

Wetland  maps,  used  in  the  creation  of  the  bottom  type 
map  in  the  GIS,  were  obtained  from  the  U.S.  Fish  and  Wild- 
life Service's  national  wetland  inventory  (NWI).  The  NWI 
maps  were  obtained  as  vector  files,  created  by  digitizing 
boundaries  between  wetland  types  from  1989  aerial  photo- 
graphs and  classified  by  using  the  classification  scheme  of 
Cowardin  et  al.  ( 1979).  Regularly  flooded  emergent  vegeta- 
tion and  submerged  aquatic  vegetation  distributions  from 


Christensen,  J.  D.,  T.  A.  Battista,  M.  E.  Monaco,  and  C.  J. 
Klein.  1997.  Habitat  suitability  modeling  and  GIS  technol- 
ogy to  support  habitat  management:  Pensacola  Bay,  Florida 
Case  Study,  58  p.  NOAA/NOS  Strategic  Environmental 
Assessments  Division,  Silver  Spring,  MD. 


40      50      60       70 
Total  length  (mm) 

Figure  2 

Total-length  frequency  distribution  for  juvenile  brown  shrimp 
captured  in  drop  traps  within  Galveston  Bay  (1982-971. 


the  NWI  maps  of  Galveston  Bay  were  chosen  to  represent 
ME  and  SAV,  respectively,  from  the  drop  sample  database. 
Nonvegetated  open  water  areas  with  depths  greater  than 
1  m  were  eliminated  throughout  the  bay  to  reflect  depth 
range  from  the  drop  sample  database.  This  elimination 
was  done  by  plotting  approximately  400,000  depth  sound- 
ings obtained  from  the  National  Geophysical  Data  Center 
(NGDC ),  and  a  bathymetric  grid  map  was  developed  in  1-m 
contours  with  ArcView  3.1  (6  nearest  neighbors,  power=2). 
The  nonvegetated  open  water  map  from  NWI  was  overlaid 
with  the  bathymetric  map  and  only  those  areas  within  the 
1-m  contour  were  extracted  and  added  to  the  bottom-type 
map  (Fig.  4). 

Two  maps  were  used  to  plot  (map)  seasonal  model 
predictions,  bottom  type,  and  the  respective  salinity 
period.  The  salinity  maps  did  not  completely  correspond 
temporally  with  seasons  defined  by  cluster  analysis  of  in 
situ  temperature  recordings  from  the  density  database. 
Salinity  periods  were  chosen  to  correlate  with  temporal 
seasons  based  on  maximum  monthly  overlap  to  develop  the 
seasonal  prediction  maps:  low  salinity  (spring);  increas- 
ing salinity  (summer);  high  salinity  (fall);  and  decreasing 
salinity  (winter). 

The  total  area  of  Galveston  Bay  (2020  km2)  was  deter- 
mined by  combining  the  total  areas  for  regularly  flooded 
emergent  vegetation,  irregularly  flooded  emergent  vegeta- 
tion, SAV,  and  open  water  classifications  from  NWI  data. 
The  bottom-type  map  reflects  the  study  area  and  totaled 
565.6  km2  after  excluding  all  areas  >1  m  in  depth  and  with 
irregularly  flooded  emergent  vegetation:  SNB  =  476.2  km2. 
ME  =  84.9  km2,  and  SAV  =  4.5  km2.  Initially,  NWI's  SAV 
classification  totaled  5.7  km2,  but  the  final  SAV  coverage 
was  reduced  to  4.5  km2  based  on  SAV  mapping  by  White 
etal.  (1993). 

Regression  modeling 

ANOVA  and  Tukey-Kramer  multiple  means  comparisons 
were  used  to  determine  if  mean  density  varied  significantly 
by  bottom  type,  salinity  zone,  and  season.  Multiple  regres- 
sion with  significant  predictors  was  used  to  predict  mean 
log  density.  The  model  was  then  applied  to  the  GIS  maps 


Clark  et  al.:  A  habitat-use  model  for  juvenile  Farfantepenaeus  aztecus  in  Galveston  Bay 


267 


Low  salinity  (April-June) 


Increasing  salinity  (July) 


&m 


High  salinity  (August-October) 


Decreasing  salinity  (November-March) 


Salinity  Zone  (ppt) 


□  0-0.5      CZZI  0  51-5       CD  51"15       CZ]  15  1-25       H  >  25 


Figure  3 

Galveston  Bay  seasonal  salinity  distribution  maps. 


to  spatially  display  model  predictions  in  each  10  x  10  m 
cell.  The  resulting  values  for  each  cell  (predicted  mean  log 
density)  were  converted  to  numbers/m2  and  reclassified 
into  5  percentiles  based  on  their  resultant  distribution: 
0-20%,  21-40%,  41-60%,  61-80%,  and  81-100%.  All 
statistical  analyses  were  conducted  with  JMP  statistical 
software  (SAS  Institute,  Cary,  NO. 

Due  to  difficulties  in  creating  continuous  salinity  and 
temperature  contour  maps  in  GIS,  these  variables  were 
classified  as  follows:  salinity  was  classified  by  one  of  the 
five  isohaline  zones  described  previously  and  analyzed  as 
such  to  determine  its  influence  on  brown  shrimp  distri- 
bution; and  water  temperature  was  classified  by  season 
determined  by  cluster  analysis  and  analyzed  to  examine 
possible  temporal  effects  of  brown  shrimp  distribution. 

Spatial  patterns  were  evaluated  by  comparing  the  pre- 
dicted mean  log  density  values  with  the  observed  mean 
log  density  values  from  Galveston  Bay  drop  samples.  Addi- 
tionally, the  model's  predictive  performance  was  assessed 
by  comparing  the  predicted  mean  log  density  values  with 
observed  mean  log  density  values  from  samples  collected 
in  Matagorda,  Aransas,  and  San  Antonio  bays  using  the 


same  collection  method.  With  this  approach,  the  assump- 
tion was  made  that  brown  shrimp  modeled  in  Galveston 
Bay  respond  similarly  to  the  range  of  biotic  and  abiotic 
factors  in  the  other  bay  systems. 

Drop  sample  data  collected  during  July-September  1984 
(/i=128),  and  April- June  1985  (re=144)  from  West  Bay  (ME, 
SNB)  and  Christmas  Bay  (ME,  SAV,  and  SNB)  were  used 
to  examine  bottom-type  preference  or  selectivity.  Tukey- 
Kramer  multiple  comparisons  test  was  used  to  compare 
log  density  patterns  in  areas  where  ME  and  SAV  occurred 
together  and  in  areas  where  SAV  was  not  present. 


Results 

Brown  shrimp  model 

ANOVA  and  Tukey-Kramer  pair-wise  comparisons 
showed  significant  differences  in  brown  shrimp  log  density 
between  the  three  bottom  types,  five  salinity  zones,  and 
four  seasons  (Fig.  5).  Multiple  regression  models  were  run 
with  these  discreet  variables  (Mahon  and  Smith,  1989; 


268 


Fishery  Bulletin  102(2) 


"T    .^ 


Nonvegetated  bottom  (SNB) 

Marsh  edge  (ME) 

Submerged  aquatic  vegetation  (SAV) 


Figure  4 

Spatial  distribution  of  Galveston  Bay  bottom  types  used  in  the  multivariate 
regression  model. 


Results  of  the  least  squares  multiple 
*  =  significant  at  P  <  0.05. 

regression  model  for 

Table  1 

predicting 

seasonal  brown 

shrimp  density  in  Galveston  Bay,  Texas. 

Model  fit 

r2 

Mean 

Observations  in) 

Mean  square  error 

0.73 

0.47 

47 

0.20 

Source 

ANOVA 

df 

Sum  of  squares 

Mean  square 

F  ratio 

Prob  >  F 

Model 
Error 
Total 

17 
29 
46 

5.74 
1.61 
6.90 

0.33 
0.04 

8.43 

<0.0001* 

Source 

Effects 

df 

Sum  of  squares 

F  ratio 

Prob  >  F 

Season 
Bottom  type 
Salinity  zone 
Bottom  typex 
Salinity  zone 

3 
2 

4 

8 

1.85 
0.61 
3.15 

0.86 

15.43 

7.57 
19.68 

2.69 

<0.0001* 

0.0023* 

<0.0001* 

0.0242* 

Krumgalz  et  al.,  1992;  Garrison,  1999)  and  we  tested 
for  possible  interactions  between  the  variables.  Only  the 
interaction  between  bottom  type  and  salinity  zone  yielded 


statistically  significant  results.  ANOVA  results  for  the 
model  including  the  bottom-type  and  salinity-zone  interac- 
tion term  (Table  1)  and  variable  coefficients  (Table  2  (fitted 


Clark  et  al.:  A  habitat-use  model  for  juvenile  Farfantepenaeus  aztecus  in  Galveston  Bay 


269 


Table  2 

Variable  coefficients  (log  +1)  derived  from  brown  shrimp  multivariate  regression  model.    ME 
aquatic  vegetation;  SNB  =  shallow  nonvegetated  bottom. 


marsh  edge;  SAV  =  submerged 


y-intercept 


Bottom  type 


Season 


0.335 


0.113      (ME) 

0.043      (SAV) 

-0.156      (SNB) 


0.239  (spring) 

0.165  (summer) 

-0.045  (fall) 

-0.359  (winter) 


Salinity  zone 


Bottom  type  x  salinity  zone 


-0.525 

-0.147 

0.079 

0.286 

0.307 


(0-0.5) 

(0.5-5) 

(5-15) 

(15-25) 

(>25) 


-0.104 
-0.055 

0.159 

0.273 
-0.396 

0.123 
-0.030 

0.049 
-0.018 
-0.119 

0.288 
-0.168 
-0.018 

0.114 
-0.096 


(ME/0-0.5) 

(SAV/0-0.5) 

(SNB/0-0.5 

(ME/0.5-5) 

(SAV/0.5-5) 

(SNB/0.5-5) 

(ME/5-15) 

(SAV/5-15) 

(SNB/5-15) 

(ME/15-25) 

(SAV/15-25) 

(SNB/15-25) 

(ME/>25) 

(SAV/>25) 

(SNB/>25) 


the  data  well  (r2=0.73,  n=47).  Overall,  density  predictions 
were  highest  in  the  spring,  declined  through  summer  and 
fall,  and  reached  the  lowest  values  during  winter  (Fig.  6). 
SNB  density  predictions  were  highest  in  the  >25  ppt  salin- 
ity zone  and  declined  as  salinity  declined  in  the  estuary. 
ME  density  predictions  exhibited  similar  density  predic- 
tion trends;  however,  a  smaller  peak  was  observed  in  the 
0.5-5  ppt  salinity  zone.  This  result  may  be  an  artifact  of 
two  fall  samples  that  exhibited  high  density  within  this 
salinity  zone.  Density  predictions  within  SAV  were  near 
zero  in  the  lower  two  salinity  zones,  peaked  in  the  15-25 
ppt  salinity  zone,  and  slightly  decreased  in  the  >25  ppt 
salinity  zone. 

Model  prediction  maps 

For  all  seasons,  highest  density  predictions  corresponded 
with  ME  and  SAV  bottom  types  within  the  region  of  the  bay 
with  highest  salinity  —  Christmas  and  West  bays  (Fig.  7). 
Density  predictions  decreased  within  all  bottom  types  as 
salinity  declined  in  the  middle  and  upper  regions  of  the 
bay.  Spring  density  predictions  were  the  highest;  maxi- 
mum values  were  predicted  within  ME  (6.14/m2)  and  SAV 
(14.49/m2)  located  in  Christmas  and  West  bays  (Fig.  7). 
Density  predictions  steadily  declined  through  the  middle 
bay  and  declined  to  1/m2  or  less  within  SAV  and  SNB  in 
the  upper  region  of  the  bay  (Trinity  Bay)  where  salinities 
were  less  than  5  ppt.  Density  predictions  during  summer, 
fall,  and  winter  were  lower  than  those  observed  during  the 
spring  but  exhibited  similar  spatial  trends — higher  pre- 
dictions within  the  high  salinity  vegetated  bottom  types, 
and  decreasing  with  decreasing  salinity. 

Model  performance 

Spatial  patterns  were  assessed  by  plotting  predicted  mean 
density  values  from  the  model  and  observed  mean  density 


A  1" 

K 

08 

* 

0  6 

0.4  ■ 

* 

0.2  ■ 

/•2=0.16 
P<0.0001 

SAV                              ME                               SNB 

Bottom  type 

B  08 

* 

>.     06 
I      0.4- 

* 

3      0.2 

r2=0.06           $ 
P<0.0001 

0-0.5             0-5.5              5-15             15-25              >25 

Salinity  zone 

c  '■ 

0  8  - 

0.6  - 

* 

* 

04  - 

* 

02  ■ 

r2=0.13 

P<0.0001 

spring              summer                  fall                    winter 

Season 

Figure  5 

Analysis  of  variance  and  Tukey-Kramer  pair-wise 

comparisons  of  brown  shrimp  density  between  (A) 

bottom  type,  (B)  salinity  zone  and,  (C)  season.  Mean 

densities  are  represented  by  solid  diamonds  and  lines 

determine  standard  error.  SAV  =  submerged  aquatic 

vegetation;  ME  =  marsh  edge;  SNB  =  shallow  non- 

vegetated  bottom. 

270 


Fishery  Bulletin  102(2) 


SNB 


1.2 

1 

0.8 
0.6 
0.4 
0.2 
0 


li    ifll 


spring  summer  fall  winter 

Salinity  zone 

|a  0-0.5  D  0.5-5  □  5-15  G15-25B>  25  | 

Figure  6 

Seasonal  density  predictions  for  brown  shrimp  (F.  aztecus)  by  bottom  type  and 
salinity  zone.  ME  =  marsh  edge;  SAV  =  submerged  aquatic  vegetation;  SNB  = 
shallow  nonvegetated  bottom. 


values  from  drop  sample  data  collected  in  Galveston  Bay. 
Regression  analysis  from  this  plot  exhibited  a  strong  posi- 
tive relationship  (r2=0.83,  P<0.0001)  between  predicted 
and  observed  density  data  (Fig.  8).  This  analysis  was  per- 
formed to  verify  how  the  model  represented  the  observed 
density  data. 

Model  performance  and  transferability  were  assessed 
by  regressing  predicted  mean  density  values  from  the 
Galveston  Bay  model  on  observed  mean  density  values 
from  drop  sample  data  collected  in  Matagorda,  San 
Antonio,  and  Aransas  bays  (Fig.  9).  Regression  analy- 
sis produced  a  positive  relationship  for  the  entire  drop 
sample  data  from  these  bays  combined  (r2=0.56)  and  in- 
dividually: Matagorda — r2=0.54;  San  Antonio — r2=0.57; 
and  Aransas  —  r2  =  0.56.  In  Aransas  and  San  Antonio 
bays,  brown  shrimp  densities  were  greatest  during  the 
spring  within  the  SAV  bottom  type  and  within  salinities 
>15  ppt.  In  Matagorda  Bay,  brown  shrimp  densities  were 
greatest  in  the  spring  within  ME  bottom  types  in  waters 
>15  ppt.  No  SAV  samples  were  taken  in  this  estuarine 
system. 


Use  of  bottom  types 

Results  from  spring  (1985)  and  fall  (1984)  drop  samples 
within  Christmas  and  West  Bay  (in  lower  Galveston  Bay) 
bottom  types  revealed  significantly  greater  brown  shrimp 
densities  in  Christmas  Bay  SAV  than  adjacent  ME  and 
SNB  (P<0.0001).  Brown  shrimp  densities  in  West  Bay 
ME  were  not  significantly  different  from  Christmas  Bay 
SAV  but  were  significantly  greater  than  densities  within 
adjacent  SNB  and  Christmas  Bay  ME  and  SNB  bottom 
types  (Fig.  10). 

The  model  results  were  also  used  to  roughly  estimate 
an  overall  population  of  approximately  1.3  billion  juve- 
nile brown  shrimp  in  Galveston  Bay  during  the  spring 
season,  by  multiplying  predicted  densities  by  bottom-type 
area  (Table  3).  Total  area  of  bottom  types  in  Galveston 
Bay  were  as  follows:  4.5  km2  (SAV);  84.9  km2  of  marsh 
edge  (ME);  and  1627.2  km2  of  nonvegetated  bottom  (29% 
[476.2  km2]  of  the  latter  area  was  considered  SNB).  On 
the  basis  of  predicted  densities  in  different  salinity  regimes, 
we  estimated  that  there  would  be  51.0  million  shrimp 


Clark  et  al.:  A  habitat-use  model  for  juvenile  Farfantepenaeus  aztecus  in  Galveston  Bay 


271 


spring 


summer 


winter 


Predicted  density  (#/m2 ) 

m  0-0.12 

0.13-1.51 
1.52-2.72 
2.73-5.46 
■I   5.47-14.85 


Figure  7 

Seasonal  spatial  distribution  maps  of  predicted  densities  for  brown  shrimp  (F.  aztecus). 


in  SAV  and  858.7  million  shrimp  in  SNB.  We  used  marsh 
edge  densities  to  estimate  473.5  million  shrimp  in  regu- 
larly flooded  vegetation  or  about  55,700  shrimp  per  hectare 
of  this  habitat  type. 


Discussion 

Various  factors  are  considered  important  in  defining 
nursery  areas  for  juvenile  estuarine-dependent  organ- 
isms; however,  the  specific  contributions  of  these  factors 
are  poorly  understood  (Beck  et  al.,  2001).  Specific  combi- 
nations of  physiochemical  conditions  and  cyclic  primary 
production  that  are  related  to  food  availability,  growth, 
and  sanctuary  from  predation  often  define  optimal  envi- 
ronments (Miller  and  Dunn,  1980).  Barry  et  al.  (1999) 
considered  prey  availability  to  be  a  necessary  component 
defining  the  nursery  function  of  estuarine  habitats. 


Shrimp  and  blue  crab  production  has  been  correlated  with 
the  availability  of  wetland  habitat  in  estuaries  (Turner, 
1977;  Zimmerman  et  al.,  2000).  In  the  present  study, 
brown  shrimp  were  most  abundant  in  the  lower  bay  where 
vegetated  habitats  were  most  abundant.  Zimmerman  et  al. 
( 1990b)  reported  that  benthic  infauna  are  most  abundant 
in  vegetated  habitats  within  lower  Galveston  Bay  and 
are  nutritionally  important  for  penaeids  (Zein-Eldin  and 
Renaud,  1986;  McTigue  and  Zimmerman,  1991,  1998). 
In  addition,  field  and  laboratory  experiments  have  shown 
that  brown  shrimp  growth  is  positively  correlated  with 
the  abundance  of  marsh  epiphytes  and  phytoplankton 
(Gleason  and  Zimmerman,  1984). 

Most  estuarine  nekton  are  adaptable  to  the  highly 
dynamic  environmental  conditions  exhibited  within  es- 
tuaries (Gifford,  1962;  Tagatz,  1971;  Zimmerman  et  al., 
1990b).  These  organisms  are  commonly  found  in  a  wide 
range  of  salinities  and  temperatures  and  are  most  affected 
by  sudden  changes  in  these  environmental  conditions 


272 


Fishery  Bulletin  102(2) 


Table  3 

Estimated  area  (km2)  of  each  bottom  type 

and  salinity  zone  combination  sampled  dur 

ing  spring  (March- 

-May), 

and  estimated 

brown  shrimp  population  based 

on  spring 

density  predictions  from  the  model.  ME  = 

marsh  edge;  SAV 

=  submerged  aquatic 

vetetation;  SNB  = 

shallow  nonve 

getated  bottom. 

Salinity  zone 

Bottom  type 

area 

Density  estimate 

Population  estimate 

Shrimp/ha. 

Bottom  type 

(ppt) 

(km2) 

(number/m2) 

(millions) 

(thousands) 

ME 

0-0.5 

1.4 

0.14 

0.2 

1428 

0.5-5 

1.6 

5.50 

8.8 

55,000 

5-15 

22.4 

4.44 

99.4 

44,375 

15-25 

59.5 

6.14 

365.3 

61,394 

>25 

0 

8.46 

0 

0 

Total 

84.9 

473.5 

55,771 

SAV 

0-0.5 

1.0 

0.09 

0.09 

9000 

0.5-5 

0.03 

0.18 

0.005 

1667 

5-15 

0.02 

4.56 

0.09 

45,000 

15-25 

3.5 

14.52 

50.8 

145,142 

>25 

0 

9.91 

0 

0 

Total 

4.5 

51.0 

114,680 

SNB 

0-0.5 

29.6 

0 

0 

0 

0.5-5 

54.2 

1.01 

54.7 

10,092 

5-15 

183.6 

1.61 

295.6 

16,100 

15-25 

203.3 

2.41 

489.9 

24,097 

>25 

5.5 

3.37 

18.5 

33,636 

Total 

476.2 

858.7 

18,032 

Total 

565.6 

1383.2 

24,455 

(Christensen  et  al.,  1997).  In  laboratory  experiments, 
Zein-Eldin  and  Aldrich  (1965)  concluded  that  higher  sa- 
linities are  more  favorable  for  brown  shrimp.  Salinities  of 


ME        +   SAV        x  SNB 


0.2  0.4  0.6  0.( 

Observed  log  density 

Figure  8 

Relationship  between  predicted  and  observed  densities  of 
brown  shrimp  (F.  aztecus)  in  Aransas,  Matagorda  and  San 
Antonio  bays  and  predicted  densities  from  the  Galveston  Bay 
model.  ME  =  marsh  edge;  SAV  =  submerged  aquatic  vegetation; 
SNB  =  shallow  nonvegetated  bottom. 


20  ppt  or  greater  were  considered  optimum  in  data  from 

Louisiana  (Barrett  and  Gillespie,  1973). 

In  the  present  study,  brown  shrimp  were  captured 
throughout  Galveston  Bay,  but  highest  densities 
were  observed  in  the  lower  bay  where  salinities  were 
greater  than  15  ppt.  This  spatial  trend  was  further 
strengthened  by  greater  abundance  of  vegetated 
bottom  types  in  the  lower  portions  of  the  bay,  where 
nearly  half  of  the  total  marsh  edge  and  90%  of  sea- 
grass  beds  are  located  (Fig.  4).  These  bottom  types 
are  regularly  inundated  and  provide  stable  substrate 
for  brown  shrimp  prey  (epiphytic  algae  and  infauna), 
whereas  seasonal  oligohaline  marsh  and  SAV  habitats 
in  the  upper  bay  may  not  promote  favorable  condi- 
tions for  prey  organisms  (Zimmerman  et  al.,  1990b). 
Therefore,  salinity  effects  and  the  greater  availability 
of  vegetated  habitats  in  the  lower  bay  may  work  in  a 
complementary  manner  to  provide  nursery  areas  for 
brown  shrimp  in  Galveston  Bay. 

Previous  attempts  to  examine  spatial  patterns 
of  abundance  and  to  determine  linkages  between 
organisms  and  habitat  included  the  development  of 
habitat  suitability  index  (HSI)  models.  Early  methods 
were  derived  by  the  U.S.  Fish  and  Wildlife  Service 
(USFWS)  for  freshwater  species,  where  the  HSI  was 
defined  as  a  numerical  index  that  represented  the 
capacity  of  a  given  habitat  to  support  a  selected  spe- 
cies. The  scale  of  HSI  values  (0-1.0)  reflects  a  linear 
relationship  between  suitability  and  carrying  capacity 


Clark  et  al.:  A  habitat-use  model  for  juvenile  Farfantepenaeus  aztecus  in  Galveston  Bay 


273 


1.2 
1.0 

0.8 
0.6 
0.4 
02 
0.0 
-0.2 

1.0 
0.8 
0.6 
0.4 
0.2 
0.0 
-0.2 


+            V 

+                 jr 

■              >^              ■ 
+    ■   +                    S 
■             y*       * 
-                                  ■      ■           ■   ./ 

•  X      * 

x  .   •/*&     s                + 

x    is          « 

_       ■>♦ 

All  Bays 

^      ♦                                 n  =  63 

r2  =  0.56 

+ 

+             yT 
X            /^         ' 

x S        ♦ 

X 

Aransas  Bay 

n  =  9 

r2  =  0.56 

x       ■          _, X 

x   •\^    x 

x^^^'                X 

-     X 

> 

Matagorda  Bay 
n  =  25 

r2  =  0.54 

I         I         I         i         i         i 

+ 
+                   s 

*s*            * 

x   ^r 
X       v^ 

San  Antonio  Bay 

^        +                                       n  =  29 

r2  =  0.57 
i         i         I         i         i         i 

0      0.2     0.4    0.6     0.8  1.0     1.2 


0.2     0.4    0.6     0.8  1.0     1.2     1.4 


Observed  log  density 


ME        +  SAV       x  SNB 


Figure  9 

Relationships  between  observed  densities  of  brown  shrimp  iF.  aztecus)  in 
Aransas,  Matagorda,  and  San  Antonio  Bays  and  predicted  densities  from 
the  Galveston  Bay  model.  Relationships  for  all  bays  combined  are  shown 
in  the  upper  left  graph.  For  each  relationship,  the  r2  is  shown  for  the  least 
squares  regression,  and  the  number  of  observations  (n )  and  the  total  number 
of  samples  in  parentheses.  ME  =  marsh  edge;  SAV  =  submerged  aquatic 
vegetation;  SNB  =  shallow  nonvegetated  bottom. 


(USFWS,  1981).  Recently,  Christensen  et  al.1  and  Brown 
et  al.,  2000,  developed  suitability  indices,  based  on  lit- 
erature reviews  and  expert  opinion,  and  raster-based  GIS 
models  that  produce  a  spatial  view  of  relative  suitability. 
The  Florida  Fish  and  Wildlife  Conservation  Commis- 
sion-Marine Research  Institute  (FMRI)  and  the  National 
Ocean  Service's  Center  for  Coastal  Monitoring  and  As- 
sessment (NOS/CCMA)  collaborated  to  develop  a  suite  of 
quantitative  HSI  modeling  approaches,  using  fisheries- 
independent  monitoring  catch-per-unit-of-effort  (CPUE) 
data  (Rubec  et  al.,  1998,  1999,  2001).  These  studies  used 
an  unweighted  geometric  mean  formula  as  part  of  the  HSI 
models  to  assess  overall  suitability.  This  approach  assigns 
equal  weight  to  all  factors  by  using  scaled  suitability  indi- 
ces as  inputs  to  the  model.  The  regression  approach  used 
in  this  study  more  appropriately  weights  density  according 
to  the  factors  in  the  model  and  allows  a  more  robust  tech- 
nique to  elucidate  spatial  patterns  of  habitat  use  by  using 
actual  CPUE  data.  In  addition,  the  method  described 
in  our  study  can  support  more  complex  analyses,  such 
as  interaction  effects  or  trophic  relationships  (or  both). 


Our  ANOVA  (Table  1)  revealed  that  season,  bottom  type, 
salinity,  and  the  interaction  between  salinity  and  bottom 
type  are  significant  factors  that  influence  the  distribution 
of  juvenile  brown  shrimp  in  Galveston  Bay.  The  addition  of 
the  interaction  effect  to  the  model  increases  the  coefficient 
of  determination  from  0.63  to  0.73.  Without  this  term  in 
the  model,  predicted  values  for  brown  shrimp  density  are 
overestimated  compared  to  the  observed  density  data. 
Seagrass  beds  in  salinities  greater  than  15  ppt  supported 
significantly  greater  densities  of  brown  shrimp  than  did 
marsh  edge.  However,  in  locations  with  salinities  less 
than  15  ppt,  brown  shrimp  densities  were  not  significantly 
different  between  the  two  bottom  types.  These  results  in- 
dicate significantly  lower  use  among  all  the  bottom  types 
analyzed  in  the  fresher  portion  of  the  estuary.  It  is  likely 
that  salinity  and  a  combination  of  other  environmental 
factors  directly  or  indirectly  (or  directly  and  indirectly) 
affect  abundance  on  bottom  types  and  habitat  quality  in 
this  region.  The  results  indicate  that  SAV  supports  greater 
brown  shrimp  density  than  do  ME  and  SNB;  however,  SAV 
accounts  for  less  than  1%  of  the  total  bottom  type  within 


274 


Fishery  Bulletin  102(2) 


April-June  1985 

1.5  - 

* 

1.0  - 

i 

0.5  - 

i 

r2=0.46 

* 

* 

P<0.0001 

July-September  1984 

1.5   - 

i 

i 

1.0   - 

$ 

0.5   - 

i 

r2=o.59 

i 

0  - 

P<0.0001 

i                      I 

i 

i 

West  Bay 
ME 


West  Bay 
SNB 


Xmas  Bay 
ME 


Xmas  Bay 
SAV 


Xmas  Bay 
SNB 


Figure  10 

Brown  shrimp  (F.  aztecus)  observed  log  density  and  standard  deviation  for  bottom 
types  in  Christmas  Bay  and  West  Bay.  ME  =  marsh  edge:  SAV  =  submerged  aquatic 
vegetation;  SNB  =  shallow  nonvegetated  bottom. 


Galveston  Bay.  Our  data  suggest  that  brown  shrimp  select 
SAV  over  ME  when  these  habitats  co-occur  (Christmas 
Bay)  and  select  ME  when  grassbeds  are  absent  (West  Bay) 
( Fig.  10 ).  Habitat  submergence  time  may  explain  high  SAV 
use  in  Christmas  Bay  (Rozas  and  Minello,  1998).  Subtidal 
grassbeds  may  provide  more  continuous  refuge  and  food 
supply  at  both  low  and  high  tides  than  the  marsh  surface, 
which  can  be  accessed  only  during  high  tides.  Additionally, 
brown  shrimp  were  significantly  smaller  in  SAV  (,v  =  17 
mm)  than  in  ME  (5=25  mm)  (t-test,  P<0.001>.  which  may 
imply  ontogenetic  changes  in  habitat  or  trophic  require- 
ments (Conrow  et  al.,  1990;  Thomas  et  al.,  1990;  Rozas  and 
Minello,  1999).  Differences  in  the  use  of  bottom  types  may 
correspond  with  the  population's  size  distribution  at  the 
time  of  sampling.  Additional  research  is  needed  to  reveal 


ontogenetic  habitat  shifts  and  relationships  among  shal- 
low estuarine  bottom  types  (Mclvor  and  Rozas,  1996). 

Assessment  of  the  model  performance  was  based  on 
FWS  HSI  theory  where  there  is  a  positive  relationship 
between  HSI  value  and  the  carrying  capacity  of  the  avail- 
able habitat.  In  the  present  study,  the  relationship  equates 
high  brown  shrimp  densities  with  optimal  habitat  condi- 
tions that  promote  high  carrying  capacity.  Therefore,  low 
densities  would  reflect  a  low  suitability  or  a  low  capacity 
to  support  the  population.  Comparisons  of  predicted  den- 
sity with  that  of  observed  values  from  Galveston  Bay,  and 
other  Texas  bays  (Figs.  7  and  8)  agree  with  FWS  theory 
by  exhibiting  a  strong  relationship  between  density  and 
suitable  habitat  as  determined  from  the  model.  Model  per- 
formance and  transferability  were  examined  by  applying 


Clark  et  al.:  A  habitat-use  model  for  juvenile  Farfantepenaeus  aztecus  in  Galveston  Bay 


275 


the  Galveston  Bay  model  (with  interaction  term)  to  brown 
shrimp  density  data  from  Aransas,  Matagorda,  and  San 
Antonio  bays.  The  results  indicated  similar  habitat-use 
patterns  in  Aransas  and  San  Antonio  bays;  there  were 
higher  densities  in  high-salinity  seagrass  beds  and  a  de- 
clining density  as  salinity  decreased  in  these  bay  systems. 
No  SAV  samples  were  taken  in  Matagorda  Bay;  however, 
the  model  performed  well  in  predicting  greater  brown 
shrimp  density  in  higher-salinity  marsh-edge  habitats. 
Our  analysis  suggests  that  although  the  empirical  model 
is  complex,  it  is  general  enough  to  be  applicable  across  a 
broader  range  of  habitat  types.  The  model  results  may, 
however,  have  some  geographic  limitations.  For  instance, 
the  model  may  not  perform  well  within  the  Laguna  Madre 
in  south  Texas,  where  freshwater  inflow  is  diminished  and 
hypersaline  conditions  exist.  This  conclusion  is  consistent 
with  Rubec  et  al.  (1999),  who  used  similar  methods  to 
demonstrate  that  HSI  models  are  applicable  across  estuar- 
ies in  central  Florida.  Our  results  are  promising  in  view 
of  previous  efforts  where  predictions  of  nekton  abundance 
with  empirical  models  have  proven  difficult. 

Currently,  estuarine  EFH  for  most  federally  managed 
species  in  the  Gulf  of  Mexico  exists  as  mapped  estimates 
of  relative  abundance  from  NOS's  estuarine  living  marine 
resources  (ELMR)  database  (GMFMC,  1998;  Nelson  and 
Monaco,  2000).  The  entire  Galveston  Bay  complex  was 
considered  EFH  for  brown  shrimp  based  on  ELMR  relative 
abundance  data.  Our  model,  generated  by  using  brown 
shrimp  density  data,  provides  a  more  spatially  resolved 
delineation  of  EFH  (in  waters  <1  m  depth)  for  brown 
shrimp  <100  mm. 

The  analyses  described  in  the  present  study  focused 
on  bottom  types  in  waters  less  than  1  m  which  comprise 
about  25%  of  the  available  habitat  in  Galveston  Bay. 
Trawl  CPUE  data  from  Texas  Parks  and  Wildlife  De- 
partment (TPWD)  were  analyzed  to  compare  abundance 
and  distribution  patterns  in  waters  >1  m.  These  trawls 
(3.8-cm  stretched  mesh)  do  not  capture  small  size  classes 
(<50  mm  TL)  of  brown  shrimp  efficiently;  thus  the  trawl 
analysis  provides  information  only  on  larger  size  classes 
(mean=89  mm).  However,  few  individuals  in  smaller  size 
classes  of  shrimp  (<50  mm  TL)  are  likely  to  inhabit  deeper 
bay  waters;  density  estimates  of  small  nekton,  including 
shrimp,  decline  rapidly  with  depth  (Mock,  1966;  Baltz  et 
al.,  1993;  Rozas,  1993;  Rozas  and  Zimmerman,  2000).  In 
addition,  these  CPUE  values  are  likely  underestimates  of 
brown  shrimp  density;  catch  efficiency  for  shrimp  in  trawls 
can  be  roughly  estimated  at  20f/f  (Zimmerman  et  al.,  1984; 
Rozas  and  Minello,  1997).  Despite  these  problems,  shrimp 
abundance  estimates  in  water  >1  m  appear  low;  abun- 
dance estimates  from  TPWD  trawl  data  in  deep  open-bay 
waters  were  almost  two  orders  of  magnitude  lower  than 
densities  in  shallow  water  habitats. 

Brown  shrimp  population  estimates  from  the  present 
study  (Table  3)  were  highest  in  the  lower  bay  (224,568  per 
ha.).  Seagrass  beds  accounted  for  more  than  607i  of  the  es- 
timate ( 145,142  per  ha.)  and  marsh  edge  and  nonvegetated 
bottom  types  combined  were  estimated  at  approximately 
79,000  per  ha.  As  noted  earlier,  the  NWI  regularly  flooded 
emergent  vegetation  classification  is  not  all  marsh  edge  but 


is  a  complex  of  SNB,  marsh  edge,  and  inner  marsh  with 
different  shrimp  densities  associated  with  each  of  these 
microhabitat  types.  Minello  and  Rozas  (in  press)  modeled 
small-scale  density  patterns  on  the  marsh  surface  in  a 
437-ha.  salt  marsh  of  lower  Galveston  Bay  and  applied 
these  data  to  a  GIS  analysis  of  marsh  landscape  patterns. 
In  this  highly  fragmented  marsh  complex  that  was  37% 
SNB  and  63%  marsh  vegetation,  they  estimated  brown 
shrimp  populations  at  37,000  per  ha.  We  could  not  estimate 
brown  shrimp  populations  in  irregularly  flooded  emergent 
vegetation,  although  the  areal  coverage  of  this  habitat  type 
was  large.  Compared  with  the  regularly  flooded  wetlands, 
overall  densities  of  brown  shrimp  in  these  irregularly 
flooded  systems  should  be  relatively  low  because  of  higher 
marsh  surface  elevations  (Rozas  and  Reed,  1993;  Minello 
et  al.,  1994;  Minello  and  Webb,  1997)  and  restricted  tidal 
access  (Rozas  and  Minello,  1999).  We  also  were  unable  to 
assess  the  contribution  of  oyster  reef  as  habitat  for  brown 
shrimp.  Coen  et  al.  ( 1999),  however,  reported  brown  shrimp 
on  oyster  reefs,  and  Powell  ( 1993 )  estimated  that  there  was 
108  km2  of  this  habitat  in  Galveston  Bay. 

Our  modeling  results  provide  evidence  that  estuarine 
habitat  types  are  discriminately  used  by  brown  shrimp. 
The  success  of  transferring  our  empirical  model  from 
Galveston  Bay  to  adjacent  bay  systems  in  Texas  suggests 
that  the  model  has  a  broad  application  and  can  possibly 
be  used  to  simulate  patterns  of  habitat  use  in  systems 
that  lack  sufficient  density  data.  Continuing  collections 
of  density  data  in  Gulf  estuaries  are  necessary  to  make 
additional  interestuary  comparisons  and  to  determine 
whether  these  habitat-use  patterns  differ  throughout  the 
distributional  range  of  brown  shrimp.  The  use  of  other 
habitat  types  also  needs  to  be  examined.  For  example, 
other  available  habitat  types  from  Galveston  Bay,  such  as 
oyster  reef  and  inner  marsh,  and  from  other  Gulf  estuar- 
ies, such  as  mangrove,  calcium  carbonate  rock  formations, 
and  sponge  communities,  may  be  important  habitats  for 
this  federally  managed  species. 


Acknowledgments 

Funding  and  support  for  this  work  was  provided  by  the 
Southeast  Region  of  NOAA's  National  Marine  Fisheries 
Service,  The  Southeast  Fisheries  Science  Center,  and  the 
Biogeography  Program  of  the  National  Ocean  Service.  We 
would  like  to  thank  Pete  Sheridan,  Lawrence  Rozas,  Ken 
Heck,  and  Roger  Zimmerman  for  providing  access  to  pub- 
lished and  unpublished  data  sets.  John  Boyd  helped  with 
construction  of  the  nekton  density  database. 


Literature  cited 

Baltz,  D.  M.,  J.  W.  Fleeger,  C.F.  Rakocinski,  and  J.  N.  McCall. 
1998.     Food,  density,  and  microhabitat:  factors  affecting 
growth  and  recruitment  potential  of  juvenile  saltmarsh 
fishes.     Environ.  Biol.  Fish.  53:89-103. 
Baltz,  D.  M.,  C.  Rakocinski,  and  J.  W.  Fleeger. 

1993.     Microhabitat  use  by  marsh-edge  fishes  in  a  Louisiana 
estuary.     Environ.  Biolog.  Fish.  36:109-126. 


276 


Fishery  Bulletin  102(2) 


Barret,  B.  B.,  and  M.  C.  Gillespie. 

1973.  Primary  factors  which  influence  commercial  shrimp 
production  in  coastal  Louisiana.  La.  Wild  Life  Fish. 
Comm,  Tech.  Bull.  9,  28  p. 

Barry,  J.  P.,  M.  M.  Yoklavich,  G.  M.  Cailliet,  D.  A.  Ambrose, 
and  B.  S.  Antrim. 

1999.  Trophic  ecology  of  the  dominant  fishes  in  Elkhorn 
Slough,  California,  1974-1980.     Estuaries  19:115-138. 

Beck,  M.  W.,  K.  L.  Heck  Jr.,  K.  Able,  D.  Childers,  D.  Eggleston, 
B.  M.  Gillanders,  B.  Halpern,  C.  Hays,  K.  Hoshino, 
T.  Minello,  R.  Orth.  P.  Sheridan,  and  M.  Weinstein. 

2001.     The  identification,  conservation,  and  management 
of  estuarine  and  marine  nurseries  for  fish  and  inverte- 
brates.    Bioscience  51:633-641. 
Boesch,  D.  F.,  and  R.  E.  Turner. 

1984.     Dependence  of  fishery  species  on  salt  marshes:  The 
role  of  food  and  refuge.     Estuaries  7:460-468. 
Brown,  S.  K.,  K.  R.  Buja,  S.  H.  Jury,  M.  E.  Monaco,  and 
A.  Banner. 

2000.  Habitat  suitability  index  models  for  eight  fish  and 
invertebrate  species  in  Casco  and  Sheepscot  Bays,  Maine. 
N.  Am.  J.  Fish.  Manag.  20:408-435. 

Christensen,  J.  D.,  M.  E.  Monaco,  and  T.  A.  Lowery. 

1997.  An  index  to  assess  the  sensitivity  of  Gulf  of  Mexico 
species  to  changes  in  estuarine  salinity  regimes.  Gulf 
Res.  Rep.  9(4):219-229. 

Coen,  L.  D.,  M.  W.  Luckenbach,  and  D.  L.  Breitburg. 

1999.     The  role  of  oyster  reefs  as  essential  fish  habitat:  a 

review  of  current  knowledge  and  some  new  perspectives.     In 

Fish  habitat:  essential  fish  habitat  and  rehabilitation  (L. 

R.  Benaka,  ed.),  p.  438-454.     Am.  Fish.  Soc,  Symposium 

22,  Bethesda,  MD. 
Conrow,  R.  A.,  V.  Zale,  and  R.  W.  Gregory. 

1990.  Distributions  and  abundances  of  early  life  stages 
of  fishes  in  a  Florida  lake  dominated  by  aquatic  macro- 
phytes.     Trans.  Am.  Fish.  Soc.  119:521-528. 

Copeland,  B.  J.,  and  T.  J.  Bechtel. 

1974.  Some  environmental  limits  of  six  gulf  coast  estuarine 
organisms.     Contrib.  Mar.  Sci.  18:169-204. 

Cowardin,  L.  J.,  V.  Carter,  F.  C.  Golet,  and  E.  T.  Laroe. 

1979.     Classification  of  wetlands  and  deepwater  habitats 
of  the  United  States.     U.S.  Fish  Wildl.  Serv.  Biol.  Serv. 
Program  FWS/OBS-79/31,  131  p. 
Czapla,  T.  E. 

1991.  Diets  and  prey  selection  of  pinfish  and  southern  floun- 
der in  a  Halodule  wnghtii  seagrass  meadow.  Ph.D.  diss., 
119  p.    Texas  A&M  Univ.,  College  Station,  TX 

Garrison,  L.  P. 

1999.     Vertical  migration  behavior  and  larval  transport  in 
brachyuran  crabs.     Mar.  Ecol.  Prog.  Ser.  176:103-113. 
Gifford,  C.  A. 

1962.     Some  aspects  of  osmotic  and  ionic  regulation  in  the 
blue  crab.  Callinectes  sapidus,  and  the  ghost  crab,  Ocypode 
albicans.     Publ.  Inst.  Mar.  Sci.  8:97-125. 
Gleason,  D.  F.,  and  R.  J.  Zimmerman. 

1984.     Herbivory  potential  of  postlarval  brown  shrimp  as- 
sociated with  salt  marshes.     J.  Exp.  Mar.  Biol.  Ecol.  84: 
235-246. 
GMFMC  (Gulf  of  Mexico  Fishery  Management  Council). 

1998.  Generic  amendment  for  addressing  essential  fish 
habitat  requirements.  Prepared  by  the  GMFMC,  October 
1998,  34  p. 

Krumgalz,  B.  S.,  G.  Fainshtein,  and  A.  Cohen. 

1992.  Grain  size  effect  on  anthropogenic  trace  metal  and 
organic  matter  distribution  in  marine  sediments.  Sci. 
Total  Environ.  116(l-2):15-30. 


Mahon.  R.  and  R.  W.  Smith. 

1989.     Comparison  of  species  composition  in  a  bottom 

trawl  calibration  experiment.     J.  Northw.  Atl.  Fish.  Sci. 

9:73-79. 
Mclvor,  C.  C,  and  L.  P.  Rozas. 

1996.  Direct  nekton  use  of  intertidal  saltmarsh  habitat  and 
linkage  with  adjacent  habitats:  a  review-  from  the  south- 
eastern United  States.  In  Estuarine  shores:  evolution, 
environments  and  human  alterations  (K.  F.  Nordstrom 
and  C.  T.  Roman,  eds.  I,  p.  31 1-334.  John  Wiley  and  Sons. 
Ltd.,  Chichester,  England. 

McTigue,  T  A.,  and  R.  J.  Zimmerman. 

1991.  Carnivory  versus  herbivory  in  juvenile  Penaeus  setfi- 
erus  (Linnaeus)  and  Penaeus  aztecus  (Ives).  J.  Exp.  Mar. 
Biol.  Ecol.  15:1-16. 

1998.  The  use  of  infauna  by  juvenile  Penaeus  aztecus  ( Ives) 
and  P.  setiferus  (Linnaeus).     Estuaries  21:160-175. 

Miller.  J.  M.,  and  M.  L.  Dunn. 

1980.  Feeding  strategies  and  patterns  of  movement  in  juve- 
nile estuarine  fishes,  hi  Estuarine  perspectives  (V.  S.  Ken- 
nedy, ed.),  p.  437-448.     Academic  Press.  New  York.  NY. 

Minello,  T.  J. 

1999.  Nekton  densities  in  shallow  estuarine  habitats  of 
Texas  and  Louisiana  and  the  identification  of  essential 
fish  habitat.  In  Fish  habitat:  essential  fish  habitat  and 
rehabilitation  (L.  Beneka,  ed.).  p.  43-75.  Am.  Fish.  Soc. 
Bethesda,  MD. 

Minello,  T.  J.,  and  L.  P.  Rozas. 

In  press.     Nekton  populations  in  Gulf  Coast  wetlands: 

fine-scale  spatial  distributions,  landscape  patterns,  and 

restoration  implications.     Ecol.  Appl. 
Minello,  T.  J.,  and  J.  W.  Webb  Jr. 

1997.  Use  of  natural  and  created  Spartinu  alterniflora 
salt  marshes  by  fishery  species  and  other  aquatic  fauna 
in  Galveston  Bay,  Texas,  USA.  Mar.  Ecol.  Prog.  Ser. 
151:165-179. 

Minello.  T.  J..  J.  W.  Webb  Jr.,  R.  J.  Zimmerman,  R.  B.  Wooten, 
J.  L.  Martinez,  T.  J.  Baumer,  and  M.  C.  Patillo. 

1991.  Habitat  availability  and  utilization  by  benthos  and 
nekton  in  Hall's  Lake  and  West  Galveston  Bay.  NOAA 
Tech.  Memo.  NMFS-SEFC-275,  37  p. 

Minello.  T.  J.,  and  R.  J.  Zimmerman. 

1992.  Utilization  of  natural  and  transplanted  Texas  salt 
marshes  by  fish  and  decapod  crustaceans.  Mar.  Ecol. 
Prog.  Ser.  90:273-285. 

Minello,  T.  J.,  R.  J.  Zimmerman,  and  R.  Medina. 

1994.     The  importance  of  edge  for  natant  macrofauna  in  a 
created  salt  marsh.     Wetlands  14:184-198. 
Mock,  C.  R. 

1966.     Natural  and  altered  estuarine  habitats  of  penaeid 
shrimp.     Proceedings  of  the  Gulf  and  Caribbean  Fisheries 
Institute,  19lh  annual  session,  p.  86-98.     Gulf  and  Carib- 
bean Fish.  Inst.,  Fort  Pierce,  FL. 
Monaco,  M.  E.,  S.  B.  Weisberg,  and  T.  A.  Lowery. 

1998.  Summer  habitat  affinities  of  estuarine  fish  in  US 
mid-Atlantic  coastal  systems.  Fish.  Manag.  Ecol.  5:161- 
171. 

NOAA  (National  Oceanic  and  Atmospheric  Administration). 

1989.  Estuaries  of  the  United  States:  vital  statistics  of  a 
natural  resource  base,  79  p.  Strategic  Environmental 
Assessments  Division,  National  Ocean  Service  (NOS), 
NOAA,  Rockville,  MD. 

Nelson,  D.  M.,  and  M.  E.  Monaco. 

2000.  National  overview  and  evolution  of  NOAA's  estua- 
rine living  marine  resources  (ELMR)  Program.  NOAA 
Tech.  Memo.  NOS  NCCOS  CCMA  144,  60  p.     Center  for 


Clark  et  al.:  A  habitat-use  model  for  juvenile  Farfantepenaeus  aztecus  in  Galveston  Bay 


277 


Coastal  Monitoring  and  Assessment,  NOS,  NOAA,  Silver 
Spring,  MD. 
Orlando,  S.  P.  Jr.,  L.  P.  Rozas,  G.  H.  Ward,  and  C.  J.  Klein. 

1993.  Salinity  characteristics  of  Gulf  of  Mexico  estuaries, 
209  p.  Office  of  Ocean  Resources  and  Conservation  and 
Assessment,  NOAA,  Silver  Spring,  MD. 

Pearcy,  W.  G.,  and  S.  S.  Myers. 

1974.     Larval  fishes  of  Yaquina  Bay,  Oregon:  a  nursery 
ground  for  marine  fishes?     Fish.  Bull.  72:201-213. 
Perez-Farfante,  I.,  and  B.  Kensley. 

1997.     Penaeoid  and  sergestoid  shrimps  and  prawns  of 
the  world;  keys  and  diagnoses  for  the  families  and  gen- 
era.    Memoires  du  Museum  National  d'Histoire  Naturelle, 
tome  175,  233  p. 
Peterson,  G.  W.,  and  R.  E.  Turner. 

1994.  The  value  of  salt  marsh  edge  vs.  interior  as  a  habi- 
tat for  fish  and  decapod  crustaceans  in  a  Louisiana  tidal 
marsh.     Estuaries  17:235-262. 

Powell,  E.  N. 

1993.  Status  and  trends  analysis  of  oyster  reef  habitat  in 
Galveston  Bay.  In  Proceeding,  second  state  of  the  bay 
symposium  (R.  W.  Jensen  et  al.,  eds.),  p.  207-209.  Galves- 
ton Bay  National  Estuary  Program,  Houston,  TX. 

Rozas,  L.  P. 

1993.  Nekton  use  of  salt  marshes  of  the  southeast  region  of 
the  United  States.  /;;  Proc.  8th  symp.  coastal  and  ocean 
management  (O.  T  Magoon,  W.  S.  Wilson,  H.  Converse, 
and  L.  T.  Tobin,  eds. ),  p.  528-537.  Coastal  Zone  '93  Con- 
ference, Am.  Soc.  Civil  Eng.,  New  Orleans,  LA. 

Rozas,  L.  P.,  and  T  J.  Minello. 

1997.  Estimating  densities  of  small  fishes  and  decapod 
crustaceans  in  shallow  estuarine  habitats:  a  review  of 
sampling  design  with  focus  on  gear  selection.  Estuaries 
20:  199-213. 

1998.  Nekton  use  of  salt  marsh,  seagrass,  and  nonveg- 
etated  habitats  in  a  south  Texas  (USA)  estuary.  Bull. 
Mar.  Sci.  63(31:481-501. 

1999.  Effects  of  structural  marsh  management  on  fish- 
ery species  and  other  nekton  before  and  during  a  spring 
drawdown.     Wetl.  Ecol.  Manag.  7:  121-139. 

Rozas,  L.  P.,  and  D.  J.  Reed. 

1993.     Nekton  use  of  marsh-surface  habitats  in  Louisiana 

( LISA)  deltaic  salt  marshes  undergoing  submergence.     Mar. 

Ecol.  Prog.  Ser.  96:147-157. 
Rozas,  L.  P.,  and  R.  J.  Zimmerman. 

2000.  Small-scale  patterns  of  nekton  use  among  marsh 
and  adjacent  shallow  nonvegetated  areas  of  the  Galves- 
ton Bay  estuary,  Texas  (USA).  Mar.  Ecol.  Prog.  Ser. 
193:217-239. 

Rubec,  P.  J.,  M.  S.  Coyne,  R.  H.  McMichael  Jr.,  and 
M.  E.  Monaco. 

1998.  Spatial  methods  being  developed  in  Florida  to  deter- 
mine essential  fish  habitat.     Fisheries  23(71:21-25. 

Rubec.  P.  J.,  J.  C.  W.  Bexley.  H.  Norris,  M.  S.  Coyne, 
M.  E.  Monaco,  S.  G.  Smith,  and  J.  S.  Ault. 

1999.  Suitability  modeling  to  delineate  habitat  essential  to 
sustainable  fisheries.     Am.  Fish.  Soc.  Symp.  22:108-133. 

Rubec,  P.  J.,  S.  G.  Smith,  M.  S.  Coyne,  M.  White,  A.  Sullivan, 
T  MacDonald,  R.  H.  McMichael  Jr.,  M.  E.  Monaco, 
and  J.  S.  Ault. 

2001.  Spatial  modeling  of  fish  habitat  suitability  in  Florida 
estuaries.     In  Spatial  processes  and  management  of 


marine  populations  (G.  H.  Kruse,  N.  Bez,  A.  Booth,  M.  W. 
Dorn,  S.  Hills,  R.  N.  Lipcus.  D.  Pelltier,  C.  Roy,  S.  J.  Smith, 
andD.  Witherell,  eds.),  p.  1-18.     Sea  Grant  report  AK-SG- 
01-02.     Univ.  Alaska.  Fairbanks,  AK. 
Sogard,  S.  M.,  and  K.  W.  Able. 

1991.     A  comparison  of  eelgrass,  sea  lettuce  macroalgae, 
and  marsh  creeks  as  habitats  for  epibenthic  fishes  and 
decapods.     Estuar.  Coast.  Shelf  Sci.  33:501-519. 
Tagatz,  M.  E. 

1971.     Osmoregulatory  ability  of  blue  crabs  in  different  tem- 
perature-salinity combinations.     Ches.  Sci.  12:14-17. 
Thomas,  J.  L.,  R.  J.  Zimmerman,  and  T.  J.  Minello. 

1990.     Abundance  patterns  of  juvenile  blue  crabs  ( Callinectes 
sapidus)  in  nursery  habitats  of  two  Texas  bays.     Bull.  Mar. 
Sci.  46:115-125. 
Turner.  R.  E. 

1977.     Intertidal  vegetation  and  commercial  yields  of 
penaeid  shrimp.     Trans.  Am.  Fish.  Soc.  106:411-416. 
USFWS  (U.S.  Fish  and  Wildlife  Servicei. 

1981.     Standards  for  the  development  of  habitat  suit- 
ability index  models  for  use  with  the  habitat  evaluation 
procedures.     Report  103,  ESM  release  1-81,  66  p.     Divi- 
sion of  Ecological  Services,  USFWS,  Washington,  DC. 
White,  W.  A.,  T.  A.  Tremblay,  E.  G.  Wermund  Jr..  and 
L.  R.  Handley. 

1993.     Trends  and  status  of  wetland  and  aquatic  habitats  in 
the  Galveston  Bay  System,  Texas,  225  p.     Galveston  Bay 
National  Estuary  Program,  Galveston.  TX. 
Zein-Eldin,  Z.  P.,  and  D.  V.  Aldrich. 

1965.     Growth  and  survival  of  postlarval  Penaeus  azte- 
cus under  controlled  conditions  of  temperature  and 
salinity.     Biol.  Bull.  (Woods  Hole)  129:199-216. 
Zein-Eldin,  Z.  P.,  and  M.  L.  Renaud. 

1986.     Inshore  environmental  effects  on  brown  shrimp, 
Penaeus  aztecus,  and  white  shrimp.  P.  setiferus,  popula- 
tions in  coastal  waters,  particularly  Texas.     Mar.  Fish. 
Rev.  48:9-19. 
Zimmerman,  R.  J.,  and  T  J.  Minello. 

1984.     Densities  of  Penaeus  aztecus,  P.  setiferus  and  other 
natant  macrofauna  in  a  Texas  salt  marsh.     Estuaries  7: 
421-433. 
Zimmerman,  R.  J.,  T  J.  Minello.  T.  J.  Baumer,  and 
M.  C.  Castiglione. 

1989.     Oyster  reef  as  habitat  for  estuarine  macrofauna. 
NOAA  Tech.  Memo.,  NMFS-SEFC-249,  16  p. 
Zimmerman,  R.  J.,  T.  J.  Minello,  M.  C.  Castiglione,  and 
D.  L.  Smith. 

1990a.  The  use  ofJuncus  and  Spartina  marshes  by  fisher- 
ies species  in  Lavaca  Bay,  Texas,  with  reference  to  effects  of 
floods.  NOAA  Tech.  Memo.,  NMFS-SEFC-251,  40  p. 
1990b.  Utilization  of  marsh  and  associated  habitats  along  a 
salinity  gradient  in  Galveston  Bay.  NOAA  Tech.  Memo., 
NMFS-SEFC-250,  68  p. 
Zimmerman,  R.  J.,  T.  J.  Minello,  and  L.  P.  Rozas. 

2000.  Salt  marsh  linkages  to  productivity  of  penaeid 
shrimps  and  blue  crabs  in  the  northern  Gulf  of  Mexico.  In 
Concepts  and  controversies  in  tidal  marsh  ecology  (M.  P. 
Weinstein  and  D.A.  Kreeger,  eds.),  p.  293-314.  Kluwer 
Academic  Publ.,  Dordrecht,  The  Netherlands. 
Zimmerman,  R.  J.,  T.  J.  Minello,  and  G.  Zamora. 

1984.     Selection  of  vegetated  habitat  by  Penaeus  aztecus  in 
a  Galveston  Bay  salt  marsh.     Fish.  Bull.  82:325-336. 


278 


Abstract— Queen  conch  (Strombus 
gigas)  stocks  in  the  Florida  Keys  once 
supported  commercial  and  recreational 
fisheries,  but  overharvesting  has 
decimated  this  once  abundant  snail. 
Despite  a  ban  on  harvesting  this  spe- 
cies since  1985,  the  local  conch  popu- 
lation has  not  recovered.  In  addition, 
previous  work  has  reported  that  conch 
located  in  nearshore  Keys  waters  are 
incapable  of  spawning  because  of  poor 
gonadal  condition,  although  reproduc- 
tion does  occur  offshore.  Queen  conch 
in  other  areas  undergo  ontogenetic 
migrations  from  shallow,  nearshore 
sites  to  offshore  habitats,  but  conch  in 
the  Florida  Keys  are  prevented  from 
doing  so  by  Hawk  Channel.  The  pres- 
ent study  was  initiated  to  determine 
the  potential  of  translocating  non- 
spawning  nearshore  conch  to  offshore 
sites  in  order  to  augment  the  spawning 
stock.  We  translocated  adult  conch 
from  two  nearshore  sites  to  two  off- 
shore sites.  Histological  examinations 
at  the  initiation  of  this  study  confirmed 
that  nearshore  conch  were  incapable  of 
reproduction,  whereas  offshore  conch 
had  normal  gonads  and  thus  were  able 
to  reproduce.  The  gonads  of  nearshore 
females  were  in  worse  condition  than 
those  of  nearshore  males.  However,  the 
gonadal  condition  of  the  translocated 
nearshore  conch  improved,  and  these 
animals  began  spawning  after  three 
months  offshore.  This  finding  suggests 
that  some  component  of  the  nearshore 
environment  (e.g..  pollutants,  tem- 
perature extremes,  poor  food  or  habitat 
quality)  disrupts  reproduction  in  conch, 
but  that  removal  of  nearshore  ani- 
mals to  suitable  offshore  habitat  can 
restore  reproductive  viability.  These 
results  indicate  that  translocations 
are  preferable  to  releasing  hatchery- 
reared  juveniles  because  they  are  more 
cost-effective,  result  in  a  more  rapid 
increase  in  reproductive  output,  and 
maintain  the  genetic  integrity  of  the 
wild  stock.  Therefore,  translocating 
nearshore  conch  to  offshore  spawn- 
ing aggregations  may  be  the  key  to 
expediting  the  recovery  of  queen  conch 
stocks  in  the  Florida  Keys. 


Translocation  as  a  strategy  to  rehabilitate 
the  queen  conch  (Strombus  gigas)  population 
in  the  Florida  Keys 

Gabriel  A.  Delgado 

Claudine  T.  Bartels 

Robert  A.  Glazer 

Florida  Fish  and  Wildlife  Conservation  Commission 

Florida  Marine  Research  Institute 

2796  Overseas  Highway,  Suite  119 

Marathon,  Florida  33050 

E-mail  address  (for  G.  A  Delgado)  gabneLdelgado;g>fwc. state  fl  us 

Nancy  J.  Brown-Peterson 

Department  of  Coastal  Sciences 
College  of  Science  and  Technology 
The  University  of  Southern  Mississippi 
P.O.  Box  7000 
Ocean  Springs,  Mississippi  39566 

Kevin  J.  McCarthy 

National  Marine  Fisheries  Service,  NOAA 
75  Virginia  Beach  Drive 
Miami,  Florida  33149 


Manuscript  approved  for  publication 
24  November  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:278-288(20041. 


The  queen  conch  (Strombus  gigas)  is 
a  large  marine  gastropod  harvested 
intensively  throughout  the  Caribbean 
for  its  meat  and  shell.  In  the  Florida 
Keys,  conch  once  supported  commercial 
and  recreational  fisheries,  but  overhar- 
vesting severely  depleted  the  popula- 
tion. The  harvesting  of  conch  has  been 
banned  in  Florida  since  1985,  but  the 
population  has  not  recovered  to  levels 
that  can  support  exploitation  (Glazer 
and  Berg.  1994;  Berg  and  Glazer,  1995; 
Glazer  and  Delgado,  2003).  Intensive 
fishing  may  invoke  depensatory  mecha- 
nisms as  densities  are  reduced,  limit- 
ing the  ability  of  conch  to  locate  mates 
and  increasing  the  chance  of  recruit- 
ment failure  (Appeldoorn,  1995).  This 
seems  to  be  the  case  in  Florida  because 
the  lack  of  recovery  has  been  attrib- 
uted to  diminished  recruitment  due  in 
part  to  small  spawning  aggregations 
(Stoner  et  al.,  1997;  Stoner  and  Ray- 
Culp,  2000). 

Queen  conch  occur  in  the  various 
oceanside  habitats  of  the  Florida  Keys 
archipelago  with  the  exception  of  Hawk 


Channel  (Glazer  and  Berg,  1994).  This 
naturally  occurring  deep-water  channel 
runs  parallel  to  the  Florida  Keys,  be- 
tween the  island  chain  and  the  offshore 
reef  tract.  The  substrate  on  the  bottom 
of  Hawk  Channel  is  predominantly  soft 
sediment,  which  is  poor  conch  habitat; 
consequently,  Hawk  Channel  serves 
as  a  barrier  to  migration  and  isolates 
nearshore  from  offshore  conch  aggre- 
gations (Glazer  and  Berg,  1994).  We 
have  been  monitoring  queen  conch 
stocks  throughout  the  Florida  Keys 
since  1987,  and  despite  extensive  sur- 
veys, we  have  never  observed  reproduc- 
tive activity  among  conch  in  nearshore 
aggregations  (Glazer  and  Berg,  1994). 
Conversely,  reproductive  behavior  has 
been  commonly  observed  among  conch 
in  offshore  aggregations  (Glazer  and 
Berg,  1994).  Moreover,  a  preliminary 
histological  examination  of  conch  from 
these  two  regions  indicated  that  the 
gonads  of  offshore  conch  were  capable 
of  undergoing  gametogenesis,  whereas 
the  gonads  of  nearshore  conch  were 
nonfunctional  (Glazer  and  Quintero, 


Delgado  et  al.:  Translocation  of  Strombus  gigas  as  a  strategy  to  rehabilitate  the  Florida  Keys  conch  population 


279 


DK 

Key  West   4jSt-^*' 


PS 


Atlantic  Ocean 


I    I  Hawk  Channel 
Reef  Tract 


25   00'  N 


24   30' N 


8130'W 


81    00' W 


80  30'  W 


Figure  1 

Queen  conch  lStroi7ibus  gigas)  translocation  sites  in  the  Florida  Keys  (adapted  from 
McCarthy  et  al.  2002).  The  nearshore  region  is  the  stretch  of  water  on  the  landward 
side  of  Hawk  Channel;  the  offshore  region  is  the  stretch  of  water  on  the  other  side 
of  the  channel,  contiguous  with  the  Atlantic  Ocean.  Nearshore  conch  were  translo- 
cated from  Tinglers  Island  (TI)  to  Alligator  Reef  (AR)  and  from  Duck  Key  (DK)  to 
Pelican  Shoal  (PS). 


1998;  McCarthy  et  al.,  2002).  In  a  metapopulation  context, 
the  nearshore  region  in  the  Florida  Keys  can  be  considered 
a  "blackhole  sink"  for  larval  recruitment  because  conch 
that  settle  there  do  not  spawn  and  thus  do  not  contribute 
to  the  reproductive  output  of  the  stock  (se?isu  Morgan  and 
Botsford,  2001). 

In  1990,  the  Florida  Fish  and  Wildlife  Conservation 
Commission's  (FWC)  Florida  Marine  Research  Institute 
constructed  an  experimental  hatchery  to  test  the  feasi- 
bility of  rehabilitating  queen  conch  stocks  in  the  Florida 
Keys  by  releasing  hatchery-reared  juveniles.  A  series  of 
experiments  to  determine  the  best  size  of  juveniles,  time 
of  release,  and  area  to  release  hatchery-reared  juvenile 
conch  were  conducted,  and  a  cost-benefit  analysis  was 
performed.  Unfortunately,  the  high  mortality  of  conch 
after  release,  coupled  with  high  production  costs,  caused 
us  to  examine  alternate  strategies  (Glazer  and  Delgado, 
2003). 

Translocation  is  defined  as  the  intentional  introduction 
or  reintroduction  of  animals  in  an  attempt  to  establish, 
reestablish,  or  augment  a  population  in  order  to  aid  in  the 
recovery  of  a  native  species  whose  numbers  have  been  re- 
duced by  overharvesting  or  habitat  loss  (or  both)  (Griffith 
et  al.,  1989).  This  method  of  population  recovery  has  been 
used  to  facilitate  the  recovery  of  numerous  species  of  birds 
and  mammals  (Griffith  et  al.,  1989)  and  several  aquatic 
species,  including  cutthroat  trout  (Harig  et  al.,  2000)  and 
corals  (Edwards  and  Clark,  1999;  Rinkevich,  1995;  van 
Treeck  and  Schuhmacher,  1997).  Nest  translocations 
have  also  proven  effective  in  efforts  to  recover  sea  turtles 
(Garcia  et  al.,  1996). 


The  present  study  was  initiated  to  determine  the  po- 
tential of  translocating  nonspawning  nearshore  conch  to 
offshore  sites  as  a  method  to  augment  spawning  aggrega- 
tions and  as  an  aid  in  the  recovery  of  the  queen  conch 
population  in  the  Florida  Keys.  However,  this  strategy  will 
be  beneficial  only  if  the  translocated  conch  regain  their 
reproductive  capacity.  To  test  this  approach,  we  translo- 
cated adult  conch  from  the  nearshore  region  into  existing 
offshore  breeding  aggregations  and  examined  changes 
in  reproductive  behavior  (i.e.,  mating  and  spawning)  and 
gonadal  development. 


Materials  and  methods 

Translocations  and  reproductive  behavior 

During  March  1999,  we  translocated  adult  conch  from 
nearshore  aggregations  to  aggregations  offshore.  Near- 
shore  aggregations  were  located  at  Tinglers  Island 
(24°41'N,  8r05'W;  water  depth  <l-2  m)  and  Duck  Key 
(24°45'N,  80°55'W;  water  depth  <l-2  m)  (Fig.  1).  The 
habitat  at  the  two  nearshore  sites  was  characterized  as  a 
matrix  of  hard-bottom  and  Thalassia  testudinum  patches. 
Offshore  aggregations  were  located  at  Alligator  Reef 
(24°51'N,  80°37'W;  water  depth  9-11  m)  and  Pelican  Shoal 
(24°30'N,  81°37'W;  water  depth  5-7  m)  (Fig.  1).  The  habitat 
at  the  offshore  sites  consisted  of  back-reef  rubble,  sandy 
plains,  and  patches  of  Thalassia  testudinum. 

We  tagged  44  adult  conch  at  Tinglers  Island;  23  were 
translocated  to  Alligator  Reef,  and  21  were  rereleased  at 


280 


Fishery  Bulletin  102(2) 


Table  1 

The  number  of  gonadal  tissue  samples  taken  from  resident  nears 
conch,  by  sex  and  season. 

hore,  resident  offshore,  and  translocated  nearshore  queen 

Spring 

Summer 

Fall 

Females                 Males 

Females                  Males 

Females                  Males 

Resident  nearshore                                 13                          12 
Resident  offshore                                  22                        20 
Translocated  nearshore 

14                           12 
19                           20 
12                         12 

10                            6 
25                           15 
13                            4 

Table  2 

Index  and  definitions 

used  to  quantify  gonadal  maturity  in  queen  conch.  This  index  is  patterned  after  the  maturity  scale  devel- 

oped  by  Egan  (1985). 

The  dashed  line 

separates  the  scores  1-5  from  6-8  that  were  combined  for  statistical  analyses. 

Gonadal  condition 

Score 

Definition 

Early  development 

1 

primary  and  cortical  alveolar  oocytes  in  females;  spermatogonia  and  spermato- 
cytes in  males 

Mid  development 

2 

vitellogenesis  beginning  in  females;  spermatozoa  present  in  males 

Late  development 

3 

fully  developed  oocytes  in  females,  none  in  oviduct;  all  stages  of  spermatogenesis, 
no  spermatozoa  in  vas  deferens 

Ripe 

4 

oocytes  in  oviduct  for  females;  spermatozoa  in  vas  deferens  for  males 

Spent 

5 

reabsorption  of  vitellogenic  oocytes  in  females;  empty  lobules,  residual  spermato- 
zoa in  males 

Atresia 

6 

reabsorption  of  oocytes  and  no  vitellogenesis  in  females;  reabsorption  of  spermato- 
zoa in  males 

Regressed 

7 

only  primary  oocytes  in  females;  only  primary  spermatogonia  in  males 

No  tissue 

8 

no  gonadal  tissue  development  and  no  germ  cells  present;  this  is  an  abnormal 
condition  in  adult  females  and  males 

Tinglers  Island.  We  also  tagged  132  adult  conch  at  Duck 
Key;  73  were  translocated  to  Pelican  Shoal,  and  59  were 
re-released  at  Duck  Key.  In  addition.  100  resident  offshore 
conch  were  tagged  in  situ  at  both  Alligator  Reef  and  Peli- 
can Shoal.  Conch  were  tagged  with  individually  numbered 
tags  that  were  secured  to  the  shell  spires  by  Monel  wire; 
in  addition,  colored  flagging  tape  was  similarly  attached 
to  facilitate  recapture. 

Reproductive  behavior  of  tagged  queen  conch  was  moni- 
tored at  each  of  the  four  sites  on  a  weekly  basis,  weather 
permitting,  from  March  1999  through  November  1999.  Off- 
shore sites  were  surveyed  by  using  SCUBA;  nearshore  sites 
were  surveyed  by  snorkeling.  Mating  activity  was  quanti- 
fied by  counting  the  number  of  tagged  individuals  (both 
males  and  females)  copulating;  spawning  activity  was 
quantified  by  counting  the  number  of  tagged  females  laying 
egg  masses.  Data  from  the  two  nearshore  sites  were  pooled 
and  data  from  the  two  offshore  sites  were  pooled.  Data  were 
also  pooled  by  season:  spring  consisted  of  March,  April,  and 
May;  summer  consisted  of  June,  July,  and  August;  and  fall 
consisted  of  September,  October,  and  November. 


Histological  examinations 

Gonadal  tissue  samples  from  adult  conch  were  collected 
for  histological  examination  at  the  initiation  of  the 
study  (spring;  the  start  of  the  breeding  season),  during 
July-August  (summer;  breeding  season),  and  during 
October  (fall;  the  end  of  the  breeding  season)  in  order  to 
assess  gonadal  development  in  relation  to  time  after  trans- 
location. We  collected  approximately  40  resident  offshore 
conch  during  each  season  (Table  1).  However,  because 
of  the  small  size  of  the  nearshore  aggregations  and  the 
small  number  of  nearshore  conch  translocated  offshore, 
we  collected  about  20  individuals  from  these  two  groups 
each  season  (Table  1).  We  did  not  determine  the  sex  of  the 
animals  before  sample  collection;  therefore  the  breakdown 
by  sex  is  not  exactly  even  (Table  1). 

A  one-cm:i  piece  of  tissue  from  the  middle  of  the  gonad  of 
each  animal  was  placed  in  a  labeled  plastic  cassette  and 
preserved  in  10f;i  neutral  buffered  formalin.  After  7  to  14 
days  in  fixative,  the  tissue  samples  were  rinsed  overnight 
in  freshwater.  The  samples  were  then  dehydrated  in  a  se- 


Delgado  et  al.:  Translocation  of  Strombus  gigas  as  a  strategy  to  rehabilitate  the  Florida  Keys  conch  population  281 


ries  of  graded  ethanols  (one  change  of  60%  ethanol  and  two 
changes  of  70%  ethanol  for  two  hours  each)  and  loaded  into 
an  automatic  tissue  processor  (Shandon  Hypercenter  XP, 
Shandon  Scientific  Ltd.,  Pittsburgh,  PA)  for  dehydration, 
clearing,  and  paraffin  infiltration.  Tissues  were  embedded 
in  Paraplast  Plus  (Fisher  Scientific,  Pittsburgh,  PA)  and 
sectioned  at  4  jim  with  a  rotary  microtome.  Two  serial 
sections  from  each  tissue  sample  were  mounted  on  glass 
slides,  allowed  to  dry  overnight,  and  stained  with  hema- 
toxylin 1  and  eosin  Y  (Richard  Allen  Inc.,  Richland,  MI). 
All  laboratory  procedures  followed  approved  standard  op- 
erating procedures  developed  under  the  Good  Laboratory 
Practices  guidelines  (EPA  and  FDA  guidelines). 

A  detailed  histological  inspection  of  each  sample  was 
made  to  assess  the  stage  of  gonadal  maturity  and  the 
percentage  of  gametogenic  tissue.  Each  animal  was  given 
a  score  from  1  to  8  to  quantify  gonadal  maturity  (Table  2). 
This  index  was  derived  from  a  maturity  scale  developed 
by  Egan  (1985).  Because  of  the  small  number  of  animals 
collected,  gonadal  maturity  scores  from  1  to  5  were  com- 
bined to  group  animals  that  would  be  capable  of  spawning 
or  had  recently  spawned  (Table  2).  Scores  from  6  to  8  were 
combined  to  group  animals  that  would  not  spawn  again  in 
a  season  or  were  not  capable  of  spawning  (Table  2).  In  ad- 
dition, the  percentage  of  gametogenic  tissue  present  (i.e., 
the  percentage  of  ovarian  or  testicular  tissue  occupying 
the  available  space  of  the  section)  was  visually  estimated 
by  using  the  following  index:  <25%,  25-50%,  51-75%,  and 
>75%.  For  statistical  analyses,  this  index  was  reduced  to 
two  categories:  <50%  and  >50%. 

Statistical  analyses 

We  evaluated  differences  in  reproductive  behavior  (mating 
and  spawning)  between  resident  nearshore  and  translo- 
cated nearshore  conch  for  each  season  by  using  Fisher's 
exact  test  because  it  is  not  sensitive  to  small  sample  sizes 
(Zar,  1996).  We  also  examined  differences  in  gonadal 
condition  (i.e.,  gonadal  maturity  and  the  percentage  of 
gametogenic  tissue)  between  resident  nearshore  and 
resident  offshore  conch  for  each  season  by  using  Fisher's 
exact  test.  Males  and  females  were  analyzed  separately. 
In  order  to  assess  the  effectiveness  of  the  translocations  to 
the  offshore  region,  we  used  Fisher's  exact  test  to  compare 
the  gonadal  condition  of  translocated  nearshore  conch 
with  the  gonadal  condition  of  resident  nearshore  conch 
in  summer  and  in  fall.  Again,  the  sexes  were  analyzed 
separately.  All  tests  were  run  on  SPSS  9.0  (SPSS  Inc., 
Chicago,  ID  for  Windows.  Results  were  considered  sig- 
nificant if  P<0.05. 


Results 

Reproductive  behavior:  mating 

Approximately  84%  of  the  tagged  resident  nearshore  conch. 
69%  of  the  tagged  translocated  nearshore  conch,  and  88% 
of  the  tagged  resident  offshore  conch  were  observed  at 
least  once  during  monitoring.  Resident  nearshore  conch 


Table  3 

Percentage  of  mating  (the  number  of  males  and  females 
mating  divided  by  the  total  number  of  conch  observed 
during  that  season)  and  spawning  (the  number  of  females 
spawning  divided  by  the  total  number  of  females  observed 
during  that  season  I  in  nearshore  conch  and  offshore  conch 
by  season  I  adapted  from  McCarthy  et  al..  2002 ).  Numbers 
in  parentheses  represent  the  number  of  observations; 
P  represents  the  probabilities  from  Fisher's  exact  test 
of  differences  in  reproductive  behavior  between  resi- 
dent nearshore  and  translocated  nearshore  conch.  The 
asterisk  (*)  indicates  that  the  test  was  statistically  sig- 
nificant. N/A  indicates  that  statistical  analyses  were  not 
conducted  because  no  mating  or  spawning  was  observed 
among  either  resident  nearshore  or  translocated  near- 
shore  animals. 


Offshore 
conch 


Nearshore  conch 


Resident        Resident    Translocated    P 


Mating 
Spring 
Summer 
Fall 

Spawning 
Spring 
Summer 
Fall 


5.3(95)  0.0  1 37  i 

2.4(4671  0.0(1061 

0.9(2321  0.0(20) 

46.2(39)  0.0(6) 

16.8(191)  0.0(34) 

5.2(97)  0.0(9) 


0.0(19)  N/A 

0.0(81)  N/A 

0.0(51)  N/A 

0.0(10)  N/A 

12.2(41)  0.041* 

18.5(27)  0.214 


and  translocated  nearshore  conch  were  not  observed 
mating  during  any  of  the  field  surveys;  conversely,  resi- 
dent offshore  conch  were  observed  mating  throughout  the 
study  (Table  3).  The  mating  frequency  of  resident  offshore 
conch  was  highest  during  the  spring  ( 5.3% )  and  decreased 
during  subsequent  seasons  to  0.9%  in  the  fall  (Table  3). 
All  observed  mating  occurred  between  resident  offshore 
animals. 

Reproductive  behavior:  spawning 

Neither  resident  nearshore  females  nor  translocated  near- 
shore  females  were  observed  spawning  during  the  spring 
(Table  3).  However,  by  summer,  translocated  nearshore 
females  had  attained  the  capacity  to  spawn  and  had  a 
significantly  higher  spawning  frequency  than  resident 
nearshore  females  (12.2%  vs.  0.0% ,  respectively)  (Table  3). 
During  the  fall,  spawning  frequency  of  translocated 
nearshore  females  peaked  at  18.5%,  whereas  resident 
nearshore  females  had  still  not  exhibited  any  spawn- 
ing behavior  (Table  3).  However,  this  difference  was  not 
statistically  significant  because  of  the  small  number  of 
resident  nearshore  conch  observed  (Table  3).  Looking  at 
individual  performance  instead  of  spawning  frequency, 
seven  (or  about  14%)  of  the  approximately  50  nearshore 
females  translocated  offshore  were  observed  spawning  at 
least  once  during  the  study  period. 


282 


Fishery  Bulletin  102(2) 


" "  TV.-  w  . 


'■  "0 


o 


k   O-' 


VJ, 


■■ 


B 


i      •        -  &*  <  •    >:  ^ 


it  '    ■ 


/ .  •■  S§i  s    , ; .  u 


E 


i:; 


4  ..<-_.■>  v* 


Figure  2 

Photomicrographs  of  the  gonadal  condition  of  resident  nearshore,  resident  offshore,  and  translocated  nearshore 
queen  conch  (Strombus  gigas).  (A)  Resident  nearshore  female  during  spring,  no  tissue  and  <25%  gametogenic 
tissue.  (B)  Resident  offshore  female  during  spring,  ripe  and  >75%  gametogenic  tissue.  (C)  Translocated  nearshore 
female  during  summer,  late  development  and  25-50%  gametogenic  tissue.  (D)  Resident  nearshore  male  during 
spring,  early  development  and  <25%  gametogenic  tissue.  (E)  Resident  offshore  male  during  spring,  ripe  and  >75% 
gametogenic  tissue.  (F)  Translocated  nearshore  male  during  summer,  ripe  and  25-5095  gametogenic  tissue. 


Resident  offshore  females  were  observed  spawning 
throughout  the  study  (Table  3 1.  Their  spawning  frequency 
peaked  during  the  spring  at  46.2%  and  decreased  during 
subsequent  seasons  to  5.2%  in  the  fall  (Table  3). 


Histology:  females 

Histological  examinations  revealed  that  the  gonadal  con- 
dition of  resident  nearshore  and  resident  offshore  female 


Delgado  et  al.:  Translocation  of  Strombus  gigas  as  a  strategy  to  rehabilitate  the  Florida  Keys  conch  population  283 


Females 


100 

80  • 

60 

40 

20 
100 

80 

60 

40 

20  • 
100 

80 

60 

40 

20 
0 


Spring 

n 

i 

Summer 

i  i  n 

.    n  .1  i 

Fall 

n  n Jllii 

A« 


6^  J 


Males 


Spring 

ill 

■  1 

Summer 

Ml  1 

Li 

Fall 

n 

1 

<8 


6B    s"    i 


c5-      9      d?     ■& 


resident  nearshore 
resident  offshore 


Figure  3 

Gonadal  maturity  of  resident  nearshore  and  resident  offshore  queen  conch  (Strom- 
bus  gigas)  by  sex  and  season.  The  dotted  line  separates  the  categories  that  were 
combined  for  statistical  analyses. 


conch  were  markedly  different  at  the  beginning  of  the 
study  (Fig.  2.  A  and  B).  There  were  significant  differences 
in  gonadal  maturity  between  resident  offshore  and  resi- 
dent nearshore  female  conch  during  the  spring,  summer, 
and  fall  (Table  4).  During  the  spring,  the  gonads  of  most 
resident  offshore  females  were  categorized  as  being  in 
late  development;  by  summer  most  were  ripe  and  by  fall 
most  were  either  spent,  in  atresia,  or  regressed  (Fig.  3).  In 
contrast,  the  gonads  of  most  resident  nearshore  females 
contained  no  germ  cells  during  the  spring  (Fig.  3).  By 
summer,  the  gonads  of  some  resident  nearshore  females 
were  found  to  be  in  the  early  stages  of  development,  but 
most  females  were  still  incapable  of  spawning,  and  by  fall, 
all  the  resident  nearshore  females  sampled  were  incapable 
of  spawning  (Fig.  3).  There  were  also  significant  differ- 
ences in  the  percentage  of  gametogenic  tissue  between 
resident  offshore  and  resident  nearshore  females  during 
the  spring,  summer,  and  fall  (Table  4).  The  gonads  of  most 
resident  offshore  females  contained  >75%  gametogenic 
tissue  throughout  the  study  period,  whereas  those  of  most 
resident  nearshore  females  had  <25%  (Fig.  4). 

The  gonadal  condition  of  translocated  nearshore  females 
(Fig.  2C)  improved  when  compared  with  the  gonadal 
condition  of  resident  nearshore  females  (Fig.  2A).  There 
were  significant  differences  in  gonadal  maturity  between 


Table  4 

Probabilities  from  Fisher's  exact  test 

of  differences  in 

gonadal   maturity   and 

the   percentage 

of  gametogenic 

tissue  between  resident  nearshore  and 

resident  offshore 

conch  by  sex  and  season.  n  r 

?presents 

the  total  number 

of  observations.  Asterisks  (    I 

indicate  that  the  test  was 

statistically  significant. 

Females 

Males 

n 

P 

H 

P 

Gonadal  maturity 

Spring 

35 

<0.001* 

32 

0.004* 

Summer 

33 

<0.001* 

32 

0.002* 

Fall 

35 

0.006* 

21 

<0.001* 

%  gametogenic  tissue 

Spring 

35 

<0.001* 

32 

<0.001* 

Summer 

33 

<0.001* 

32 

<0.001* 

Fall 

35 

0.002* 

21 

<0.001* 

translocated  nearshore  and  resident  nearshore  females 
during  both  the  summer  and  fall  (Table  5).  There  was 


284 


Fishery  Bulletin  102(2) 


Females 


Males 


a 


60 

40  ■ 

20  ■ 

0 

80 

i-      60 
CD 
~       40 

|       20 

D-         0 

80 

60 

40 

20 

0 


Spnng  I 


J_i Q_m 

In  rfi  J1 


&'      ,5?        &        & 


Spnng 


mmer 

It 

rail 

IL 


n 


resident  nearshore 
resident  offshore 


Figure  4 

The  percentage  of  gametogenic  tissue  of  resident  nearshore  and  resident  offshore 
queen  conch  iStrombus  gigas)  by  sex  and  season.  The  dotted  line  separates  the  cat- 
egories that  were  combined  for  statistical  analyses. 


Nearshore  females 


Nearshore  males 


100 

80 

60 

40 

20 
100 

80  • 

60  ■ 

40 

20 
0 


An 


L 


n  n 


L 


Summer 

J 

nl     I 

Fall 

[ 

]    n 

■ 

?*/ 


&      8 


<?  s* 


6* 


resident 
translocated 


Figure  5 

Gonadal  maturity  of  resident  nearshore  and  translocated  nearshore  queen  conch 
iSlrombus  gigas)  by  sex  and  season.  The  dotted  line  separates  the  categories  that 
were  combined  for  statistical  analyses. 


a  higher  percentage  of  translo- 
cated nearshore  females  in  some 
stage  of  gonadal  development  than 
resident  nearshore  females  during 
the  summer;  in  fact,  about  30%  of 
the  translocated  females  were  ripe 
(Fig.  5).  By  fall,  the  differences 
were  even  more  extreme;  over  60% 
of  the  translocated  nearshore  fe- 
males were  ripe,  whereas  all  of  the 
resident  nearshore  females  were 
incapable  of  reproducing  (Fig.  5). 
Although  there  was  a  significant 
difference  in  gonadal  maturity  be- 
tween translocated  nearshore  and 
resident  nearshore  females  during 
the  summer,  there  was  no  signifi- 
cant difference  in  the  percentage 
of  gametogenic  tissue  (Table  5  and 
Fig.  6).  However,  by  fall,  there 
were  significant  differences  in  the 
percentage  of  gametogenic  tissue 
between  translocated  nearshore 
and  resident  nearshore  females 
(Table  5 1.  Most  translocated  near- 
shore  females  had  developed  >75'  i 
of  the  gonad,  whereas  most  resi- 
dent nearshore  females  still  had 
<259c  gametogenic  tissue  (Fig.  6). 

Histology:  males 

There  were  marked  differences  in 
gonadal  condition  of  resident  near- 
shore  and  resident  offshore  male 
conch  (Fig.  2,  D  and  E).  There  were 
significant  differences  in  gonadal 
maturity  between  resident  offshore 
and  resident  nearshore  male  conch 
during  the  spring,  summer,  and  fall 
(Table  4).  During  the  spring  and 
summer,  the  gonads  of  most  resi- 
dent offshore  males  were  catego- 
rized as  ripe;  by  fall  most  were  spent 
(Fig.  3).  In  contrast,  at  least  half  of 
the  resident  nearshore  males  were 
not  capable  of  spawning  during 
the  spring  and  summer,  although 
some  were  in  the  early  stages  of  tes- 
ticular development  and  some  were 
even  ripe  (Fig.  3).  However,  all  the 
sampled  resident  nearshore  males 
were  incapable  of  spawning  by  fall 
and  none  were  identified  as  spent 
(Fig.  3).  Histological  examinations 
also  revealed  significant  differ- 
ences in  the  percentage  of  game- 
togenic tissue  between  resident 
offshore  and  resident  nearshore 
males  during  the  spring,  summer. 


Delgado  et  al.:  Translocation  of  Strombus  gigas  as  a  strategy  to  rehabilitate  the  Florida  Keys  conch  population 


285 


Nearshore  females 


Nearshore  males 


Figure  6 

The  percentage  of  gametogenic  tissue  of  resident  nearshore  and  translocated  near- 
shore  queen  conch  {Strombus gigas)  by  sex  and  season.  The  dotted  line  separates  the 
categories  that  were  combined  for  statistical  analyses. 


and  fall  (Table  4).  Most  resident  offshore  males  had  >75% 
gametogenic  tissue  throughout  the  study  period,  whereas 
most  resident  nearshore  males  had  <25%  (Fig.  4). 

The  gonadal  condition  of  translocated  nearshore  males 
(Fig.  2F)  improved  in  relation  to  the  gonadal  condition  of 
resident  nearshore  males  (Fig.  2D).  There  were  significant 
differences  in  gonadal  maturity  between  translocated 
nearshore  and  resident  nearshore  males  during  both  the 
summer  and  fall  (Table  5).  Almost  80%  of  the  translocated 
nearshore  males  were  ripe  during  the  summer,  whereas 
about  half  of  the  resident  nearshore  males  were  incapable 
of  reproducing  (Fig.  5).  By  fall,  most  translocated  near- 
shore  males  were  still  capable  of  reproduction,  whereas 
none  of  the  resident  nearshore  males  were  (Fig.  5).  There 
were  also  significant  differences  in  the  percentage  of 
gametogenic  tissue  between  resident  nearshore  and 
translocated  nearshore  males  during  the  summer  and 
fall  (Table  5).  During  the  summer,  the  gonads  of  most 
of  the  resident  nearshore  males  contained  <25%  game- 
togenic tissue,  whereas  translocated  nearshore  males 
were  divided  equally  among  the  four  gametogenic  tissue 
categories  (Fig.  6).  During  the  fall,  the  gonads  of  most  of 
the  resident  nearshore  males  still  had  <25%  gametogenic 
tissue;  however,  most  translocated  nearshore  males  had 
developed  >50'/r  of  the  gonad  (Fig.  6). 


Discussion 

In  the  nearshore  region  of  the  Florida  Keys,  adult  queen 
conch  had  severe  deficiencies  in  reproductive  behavior  and 
gonadal  development.  Histological  examinations  of  resi- 


Table  5 

Probabilities  from  Fisher's  exact  test  of  differences  in 
gonadal  maturity  and  the  percentage  of  gametogenic 
tissue  between  resident  nearshore  and  translocated  near- 
shore  conch  by  sex  and  season,  n  represents  the  total 
number  of  observations.  Asterisks  (*)  indicate  that  the 
test  was  statistically  significant. 


Females 


Males 


Gonadal  maturity 
Summer 
Fall 

7o  gametogenic  tissue 
Summer 
Fall 


26  0.019* 

23  <0.001* 

26  0.130 

23  0.038* 


24 
10 


0.045* 
0.033* 


24        0.014* 
10        0.033* 


dent  nearshore  conch  revealed  that  most  were  incapable 
of  reproducing,  whereas  resident  offshore  conch  exhibited 
a  normal  reproductive  cycle  (as  described  by  Egan,  1985, 
and  Stoner  et  al.,  1992).  Furthermore,  our  results  suggest 
that  female  conch  may  be  more  sensitive  to  the  nega- 
tive effects  of  nearshore  conditions  than  male  conch.  For 
example,  during  the  spring  and  summer,  some  resident 
nearshore  males  were  ripe  (although  their  reproductive 
output  would  have  been  severely  reduced  because  of  a  low 
percentage  of  gametogenic  tissue),  whereas  the  gonads  of 


286 


Fishery  Bulletin  102(2) 


most  resident  nearshore  females  contained  no  germ  cells. 
The  latter  condition  may  have  been  due  to  the  fact  that 
egg  production  is  more  costly  bioenergetically  than  sperm 
production  (Ricklefs,  1990). 

Mating  and  spawning  do  not  occur  among  resident  near- 
shore  conch  presumably  because  of  their  retarded  gonadal 
development;  however,  the  translocation  of  nearshore 
conch  to  the  offshore  region  mitigated  the  deleterious  ef- 
fects that  the  nearshore  environment  had  on  their  gonadal 
development.  The  reproductive  tissues  of  translocated 
nearshore  conch  began  to  develop  during  the  summer 
after  the  conch  had  spent  about  three  months  offshore. 
Most  translocated  female  conch  were  in  the  early  stages 
of  gonadal  development,  whereas  most  translocated  male 
conch  were  ripe.  We  believe  this  difference  in  gonadal  de- 
velopment is  due  to  the  fact  that  the  starting  gonadal  con- 
dition of  nearshore  females  was  worse  than  the  starting 
condition  of  male  conch.  By  fall,  after  six  months  offshore, 
most  translocated  females  had  become  ripe.  In  addition, 
the  percentage  of  gametogenic  tissue  in  the  gonads  of  both 
sexes  increased  through  the  summer  and  fall. 

In  conjunction  with  the  improvement  in  gonadal  condi- 
tion, nearshore  females  translocated  to  the  offshore  region 
were  observed  spawning  during  the  summer  and  fall; 
however,  no  mating  was  observed  among  nearshore  conch 
translocated  offshore.  Resident  offshore  conch  also  had  low 
mating  frequencies  (<6"7r ).  Similarly  low  mating  frequen- 
cies have  been  reported  in  the  Virgin  Islands  (Randall, 
1964)  and  the  Bahamas  (Stoner  et  al.,  1992).  We  suspect 
that  the  lack  of  observations  of  nearshore  conch  mating  in 
the  offshore  region  may  have  been  an  artifact  of  the  low 
probability  of  encountering  that  activity  due  to  the  small 
number  of  nearshore  conch  translocated  offshore.  Never- 
theless, we  believe  mating  must  have  occurred  because 
translocated  nearshore  conch  were  observed  spawning. 
However,  it  is  unknown  if  queen  conch  are  capable  of  lay- 
ing unfertilized  egg  masses. 

The  beginning  of  reproductive  activity  in  queen  conch 
is  linked  to  the  start  of  spring,  when  water  temperatures 
begin  rising  (Randall,  1964;  Stoner  et  al,  1992;  Weil  and 
Laughlin,  1984).  This  same  seasonal  pattern  was  observed 
in  our  study  with  resident  offshore  conch.  They  exhibited 
the  highest  mating  and  spawning  frequencies  during  the 
spring  and  reproductive  behavior  decreased  during  the 
ensuing  seasons.  However,  compared  with  the  spawn- 
ing pattern  of  resident  offshore  conch,  peak  spawning  in 
translocated  nearshore  conch  was  delayed;  peak  spawning 
occurred  during  the  fall.  Nevertheless,  there  was  evidence 
to  suggest  that  the  timing  of  reproductive  behavior  of 
both  resident  offshore  and  translocated  nearshore  conch 
might  eventually  become  similar.  Our  results  indicated 
that  it  takes  at  least  three  months  after  translocation  for 
the  negative  effects  of  the  nearshore  environment  to  be 
mitigated  and  for  gonadal  maturation  to  occur.  The  out- 
of-phase  spawning  may  have  been  prevented  if  the  trans- 
locations had  occurred  earlier  in  the  year  (e.g.,  January, 
instead  of  March). 

Identifying  the  causative  factor  or  factors  that  inhibit 
the  reproductive  viability  of  nearshore  queen  conch  re- 
quires further  study.  However,  the  juxtaposition  of  the 


nearshore  conch  aggregations  with  human  population  cen- 
ters suggests  that  anthropogenic  changes  to  the  nearshore 
region  may  be  partially  responsible.  Decreased  reproduc- 
tive output  caused  by  anthropogenic  contaminants  has 
been  observed  in  several  marine  invertebrates,  including 
dogwinkles  iNucella  lapillus)  (Bryan  et  al.,  1987;  Gibbs 
and  Bryan,  1986),  scallops  (Gould  etal.,  1988),  sea  urchins 
(Krause,  1994;  Thompson  et  al.,  1989),  and  shrimps  and 
crabs  (Wilson-Ormond  et  al.,  1994).  For  example,  chronic 
exposure  to  tributyltin  has  been  shown  to  sterilize  females 
of  several  species  of  mollusks  (Matthiessen  and  Gibbs, 
19981,  and  sublethal  levels  of  copper  greatly  inhibited 
gamete  production  and  maturation  in  scallops  (Gould  et 
al.,  1988).  There  have  also  been  numerous  reports  impli- 
cating eutrophication  in  nearshore  habitat  degradation 
in  the  Florida  Keys  (Lapointe  et  al.,  1990;  Lapointe  and 
Clark,  1992;  Szmant  and  Forrester,  1996);  however,  very 
little  is  known  about  the  effects  of  increased  nutrient  levels 
at  the  organismal  level. 

The  retarded  gonadal  condition  in  nearshore  queen 
conch  may  also  be  due  to  environmental  factors  such  as 
suboptimal  habitat,  poor  food  quality,  or  temperature 
extremes  associated  with  shallow  water.  Research  on 
bivalves  has  shown  that  habitat,  diet,  and  food  quality 
directly  affect  gamete  production  (Le  Pennec  et  al..  1998: 
Madrones-Ladja  et  al.,  2002).  As  they  increase  in  age  and 
size,  queen  conch  undergo  ontogenetic  migrations  from 
shallow,  nearshore  sites  to  deeper-water  habitats  (Ran- 
dall, 1964;  Sandt  and  Stoner,  1993;  Stoner,  1989;  Weil  and 
Laughlin,  1984).  It  has  been  hypothesized  that  as  conch 
grow  larger  and  require  more  food,  they  migrate  to  take 
advantage  of  the  augmented  food  supply  in  more  produc- 
tive offshore  habitats  (Sandt  and  Stoner.  1993;  Stoner. 
1989).  However,  nearshore  queen  conch  in  the  Florida 
Keys  are  prevented  from  migrating  offshore  by  Hawk 
Channel  (Glazer  and  Berg,  1994).  Therefore,  translocat- 
ing nearshore  conch  offshore  would,  in  effect,  link  these 
isolated  environments. 

The  implications  of  this  study  are  of  particular  impor- 
tance to  the  FWC-Florida  Marine  Research  Institute's 
ongoing  queen  conch  stock  restoration  program.  Trans- 
locating naturally  recruiting  nearshore  conch  to  offshore 
areas  would  be  more  cost  effective  than  hatchery  produc- 
tion of  juvenile  conch,  especially  because  production  costs 
are  eliminated  and  survival  of  translocated  conch  is  likely 
to  be  much  greater  than  that  of  hatchery  outplants  (see 
Stoner,  1997.  for  a  review  of  juvenile  mortality  in  stock 
enhancement  efforts).  Translocations  would  also  have  a 
more  immediate  effect  on  reproductive  output  than  would 
the  release  of  hatchery-reared  conch.  A  translocation  pro- 
gram would  focus  on  moving  large  juveniles  and  adults 
offshore,  whereas  a  hatchery  program  must  release  small 
juveniles  (to  minimize  production  costs)  that  would  then 
have  to  survive  to  maturity.  Consequently,  translocations 
would  quickly  alleviate  the  depensatory  mechanisms  de- 
scribed by  Appeldoorn  (1995)  that  can  affect  the  recovery 
of  queen  conch  stocks.  Finally,  translocations  provide  the 
added  benefit  of  maintaining  the  genetic  diversity  of  the 
population.  Hatchery-reared  conch  are  typically  derived 
from  a  few  egg  masses  and  there  is  a  concurrent  loss  in 


Delgado  et  al.:  Translocation  of  Strombus  gigas  as  a  strategy  to  rehabilitate  the  Florida  Keys  conch  population 


287 


rare  alleles  (Allendorf  and  Ryman,  1987).  However,  the 
use  of  wild  conch  to  enhance  the  spawning  aggregations 
eliminates  this  problem. 

Queen  conch  appear  to  be  a  prime  candidate  for  reha- 
bilitation by  translocation  because  they  meet  the  criteria 
associated  with  successful  translocations  reported  by 
Griffith  et  al.  (1989).  These  factors  include  release  within 
the  historical  range  of  the  species  or  into  areas  of  in- 
creased habitat  quality  (or  both).  Additionally,  herbivorous 
animals  stand  a  greater  chance  of  translocation  success 
than  do  carnivores  or  omnivores.  Lastly,  wild  animals 
translocate  more  successfully  than  captive-bred  animals. 
According  to  these  parameters,  queen  conch  are  ideally 
suited  for  translocations. 

However,  before  a  full-scale  translocation  program  can 
be  implemented,  there  are  some  theoretical  considerations 
that  must  be  addressed.  For  example,  Stoner  and  Ray- 
Culp  (2000)  reported  that  conch  reproductive  behavior 
reached  an  asymptotic  level  near  200  conch/ha.;  therefore, 
it  would  seem  advantageous  to  enhance  reproductive  ag- 
gregations to  that  density.  However,  without  high  habitat 
quality,  translocations  have  low  success  rates  regardless 
of  how  many  animals  are  released  (Griffith  et  al.,  1989). 
First,  we  must  ascertain  if  offshore  habitats  can  support 
the  added  number  of  conch  or  if  the  translocated  or  na- 
tive animals  (or  both)  will  simply  disperse  after  release 
because  of  density-dependent  factors  (e.g.,  intraspecific 
competition  for  limited  resources).  Conch  grazing  has  been 
shown  to  significantly  reduce  the  biomass  of  seagrass  mac- 
rodetritus  and  epiphytes  (Stoner,  1989).  In  addition,  the 
effects  of  removing  nearshore  conch  from  the  nearshore 
environment  need  to  be  investigated. 

Additionally,  if  increased  recruitment  is  the  ultimate 
goal  of  the  translocation  program,  larvae  must  survive 
and  be  retained  within  the  Florida  Keys.  At  this  point,  it 
is  unknown  whether  larvae  produced  from  translocated 
nearshore  conch  are  viable  or  as  viable  as  the  larvae  pro- 
duced by  native  offshore  conch.  Furthermore,  the  relative 
contribution  of  local  and  upstream  sources  to  recruitment 
is  unknown.  Stoner  et  al.  (1996,  1997)  suggested  that  most 
of  the  queen  conch  larvae  entering  the  Florida  Keys  come 
from  upstream  sources.  If  this  is  indeed  the  case,  then  local 
translocations  will  not  be  as  effective  as  an  international  or 
regional  management  strategy.  However,  mechanisms  for 
larval  retention  in  the  Florida  Keys  have  been  described 
by  Lee  and  Williams  (1999),  who  suggested  that  the  pe- 
riodic formation  of  gyres  in  the  lower  Keys  may  facilitate 
the  retention  and  recruitment  of  locally  produced  larvae. 
If  larvae  are  retained  within  the  Florida  Keys  system,  any 
increase  in  local  larval  production  will  increase  larval  sup- 
ply and  may  increase  recruitment.  Therefore,  translocation 
sites  should  be  located  in  the  lower  Keys  in  order  to  ensure 
maximum  larval  retention  and  recruitment. 

The  present  study  has  shown  that  translocation  may  be 
a  viable  method  for  rehabilitating  queen  conch  populations 
in  the  Florida  Keys.  We  have  demonstrated  that  nearshore 
conch  that  were  translocated  offshore  regained  some  of 
their  reproductive  capacity  and  abilities.  Therefore,  mov- 
ing conch  from  nearshore  larval  sinks  to  offshore  larval 
sources  may  be  the  key  to  expediting  the  recovery  of  queen 


conch  stocks.  Further  research  (e.g.,  larval  retention 
studies,  studies  on  the  effect  of  water  quality  on  larval 
survival,  carrying  capacity  studies)  and  monitoring  will 
determine  the  efficacy  of  this  restoration  strategy. 


Acknowledgments 

John  Hunt,  William  Sharp,  James  Colvocoresses,  Allan 
Stoner,  and  one  anonymous  reviewer  provided  insightful 
comments  on  the  manuscript.  Judy  Leiby  and  Jim  Quinn 
provided  editorial  comments.  We  thank  Mary  Enstrom 
and  Sherry  Dawson  of  The  Nature  Conservancy  (TNC)  as 
well  as  the  numerous  TNC  volunteers  who  participated  in 
the  field  surveys.  Meaghan  Darcy  and  other  staff  members 
at  the  Florida  Marine  Research  Institute  assisted  in  the 
field  and  in  sample  processing.  This  project  was  funded  by 
Partnerships  for  Wildlife  Grant  no.  P-3  from  the  U.S.  Fish 
and  Wildlife  Service  and  by  the  Florida  Fish  and  Wildlife 
Conservation  Commission. 


Literature  cited 

Allendorf,  F.  W.,  and  N.  Ryman. 

1987.  Genetic  management  of  hatchery  stocks.  In  Popula- 
tion genetics  and  fishery  management  (N.  Ryman  and  F. 
Utter,  eds.),  p.  141-59.  Washington  Sea  Grant,  Wash- 
ington D.C. 

Appeldoorn,  R.  S. 

1995.  Potential  depensatory  mechanisms  operating  on 
reproductive  output  in  gonochoristic  molluscs,  with  par- 
ticular reference  to  strombid  gastropods.  ICES  Mar.  Sci. 
Symp.  199:13-18. 

Berg,  Jr.,  C.  J.,  and  R.  A.  Glazer. 

1995.  Stock  assessment  of  a  large  marine  gastropod 
{Strombus  gigas)  using  randomized  and  stratified  towed- 
diver  censusing.     ICES  Mar.  Sci.  Symp.  199:247-258. 

Bryan,  G.  W.,  P.  E.  Gibbs,  G.  R.  Burt,  and  L.  G  Hummerstone. 
1987.     The  effects  of  tributyltin  (TBT I  accumulation  on  adult 
dog-whelks,  Nucella  lapillus:  Long-term  field  and  labora- 
tory experiments.     J.  Mar.  Biol.  Assoc.  U.K.  67:525-544. 
Edwards,  A.  J.  and  S.  Clark. 

1999.     Coral  transplantation:  a  useful  management  tool  or 
misguided  meddling?     Mar.  Pollut.  Bull.  37:474-487. 
Egan,  B.  D. 

1985.  Aspects  of  the  reproductive  biology  of  Strombus 
gigas.  M.S.  thesis,  147  p.  Univ.  British  Colombia,  Van- 
couver, Canada. 

Garcia,  A..  R.  Adaya,  and  G.  Ceballos. 

1996.  Sea  turtle  conservation  in  Cuixmala,  Jalisco:  the 
role  of  nest  translocation  in  small  beaches.  Bull.  Ecol. 
Soc.  Am.  77:155. 

Gibbs,  P.  E„  and  G.  W.  Bryan. 

1986.  Reproductive  failure  in  populations  of  the  dog-whelk, 
Nucella  lapillus.  caused  by  imposex  induced  by  tributyl- 
tin from  antifouling  paints.  J.  Mar.  Biol.  Assoc.  U.K. 
66:767-777. 

Glazer,  R.  A.,  and  C.  J.  Berg  Jr. 

1994.  Queen  conch  research  in  Florida:  an  overview.  //; 
Queen  conch  biology,  fisheries,  and  mariculture  (R.  S. 
Appeldoorn  and  B.  Rodriguez,  eds.),  p.  79-95.  Fundacion 
Cientifica  Los  Roques,  Caracas.  Venezuela. 


288 


Fishery  Bulletin  102(2) 


Glazer,  R.  A.,  and  G.  A.  Delgado. 

2003.  Towards  a  holistic  strategy  to  managing  Florida's 
queen  conch  {Strombus  gigas)  population.  In  El  Caracol 
Strombus  gigas:  conocimiento  integral  para  su  manejo 
sustentable  en  el  Caribe  ( D.  Aldana  Aranda,  ed. ),  p.  73-80. 
CYTED  (Programa  Iberoamericano  de  Ciencia  y  Tech- 
nologia  para  el  Desarrollo),  Yucatan,  Mexico. 
Glazer,  R.  A.,  and  I.  Quintero. 

1998.     Observations  on  the  sensitivity  of  queen  conch  to 
water  quality:  implications  for  coastal  development.     Proc. 
Gulf  Carib.  Fish.  Inst.  50:78-93. 
Gould,  E.,  R.  J.  Thompson,  L.  J.  Buckley,  D.  Rusanowsky,  and 
G.  R.  Sennefelder. 

1988.  Uptake  and  effects  of  copper  and  cadmium  in  the 
gonad  of  the  scallop  Placopecten  magellanicus:  concurrent 
metal  exposure.     Mar.  Biol.  97:217-223. 

Griffith,  B.,  J.  M.  Scott,  J.  W.  Carpenter,  and  C.  Reed. 

1989.  Translocation  as  a  species  conservation  tool:  status 
and  strategy.     Science  245:477-480. 

Harig,  A.  L.,  K.  D.  Fausch,  and  M.  K.  Young. 

2000.     Factors  influencing  success  of  greenback  cut- 
throat trout  translocations.     North  Am.  J.  Fish.  Manag. 
20:994-1004. 
Krause,  P.  R. 

1994.     Effects  of  an  oil  production  effluent  on  gametogenesis 
and  gamete  performance  in  the  purple  sea  urchin  {Stron- 
gylocentrotus  purpuratus  Stimpson).     Environ.  Toxicol. 
Chem.  13:1153-1161. 
Lapointe,  B.  E.  and  M.  W.  Clark. 

1992.     Nutrient  inputs  from  the  watershed  and  coastal  eutro- 
phication  of  the  Florida  Keys.     Estuaries  15:465-476. 
Lapointe,  B.  E.,  J.  D.  O'Connell,  and  G.  Garrett. 

1990.  Nutrient  couplings  between  on-site  sewage  disposal 
systems,  groundwaters,  and  nearshore  surface  waters  of 
the  Florida  Keys.     Biogeochem.  10:289-307. 

Le  Pennec,  M.,  R.  Robert,  and  M.  Avendano. 

1998.  The  importance  of  gonadal  development  on  larval 
production  in  pectinids.     J.  Shellfish  Res.  17:97-101. 

Lee,  T  N.,  and  E.  Williams. 

1999.  Mean  distribution  and  seasonal  variability  of  coastal 
currents  and  temperature  in  the  Florida  Keys  with  impli- 
cations for  larval  recruitment.     Bull.  Mar.  Sci.  64:35-56. 

Madrones-Ladja,  J.  A.,  M.  R.  dela  Pena,  and  N.  P.  Parami. 

2002.     The  effect  of  micro  algal  diet  and  rearing  condition  on 
gonad  maturity,  fecundity,  and  embryonic  development  of 
the  window-pane  shell,  Placuna  placenta  Linnaeus.     Aqua- 
culture  206:313-321. 
Matthiessen,  P.,  and  P.  E.  Gibbs. 

1998.     Critical  appraisal  of  the  evidence  for  tributyltin- 
mediated  endocrine  disruption  in  mollusks.     Environ. 
Toxicol.  Chem.  17:37-43. 
McCarthy,  K.  J.,  C.  T.  Bartels,  M.  C.  Darcy,  G.  A.  Delgado,  and 
R.  A.  Glazer. 

2002.     Preliminary  observation  of  reproductive  failure  in 
nearshore  queen  conch  {Strombus  gigas)  in  the  Florida 
Keys.     Proc.  Gulf  Carib.  Fish.  Inst.  53:674-680. 
Morgan,  L.  E.  and  L.  W.  Botsford. 

2001.  Managing  with  reserves:  Modeling  uncertainty 
in  larval  dispersal  for  a  sea  urchin  fishery.  In  Spatial 
processes  and  management  of  marine  populations  (G. 
H.  Kruse,  N.  Bez,  A.  Booth,  M.  W.  Dorn,  S.  Hills,  R.  N. 
Lipcius,  D.  Pelletier,  C.  Roy,  S.  J.  Smith,  and  D.  Wither- 


ed eds.),  p.  667-684.    Alaska  Sea  Grant,  AK-SG-01-02. 
Univ.  Alaska,  Fairbanks,  Alaska. 
Randall,  J.  E. 

1964.     Contributions  to  the  biology  of  the  queen  conch. 
Strombus  gigas.     Bull.  Mar.  Sci.  14:246-295. 
Ricklefs,  R.  E. 

1990.     Ecology.  3rd  ed.,  896  p.    W.H.  Freeman  and  Co.,  New 
York.  NY. 
Rinkevich.  B. 

1995.  Restoration  strategies  for  coral  reefs  damaged  by 
recreational  activities:  the  use  of  sexual  and  asexual 
recruits.     Restoration  Ecology  3:241-251. 

Sandt,  V.  J.  and  A.  W.  Stoner. 

1993.     Ontogenetic  shift  in  habitat  by  early  juvenile  queen 
conch,  Strombus  gigas:  patterns  and  potential  mech- 
anisms.    Fish.  Bull.  91:516-525. 
Stoner,  A.  W. 

1989.     Winter  mass  migration  of  juvenile  queen  conch  Strom- 
bus gigas  and  their  influence  on  the  benthic  environment. 
Mar.  Ecol.  Prog.  Ser.  56:99-104. 
1997.     The  status  of  queen  conch,  Strombus  gigas,  research 
in  the  Caribbean.     Mar.  Fish.  Rev.  59:14-22. 
Stoner,  A.  W„  R.  A.  Glazer,  and  P.  J.  Barile. 

1996.  Larval  supply  to  queen  conch  nurseries:  relationships 
with  recruitment  process  and  population  size  in  Florida 
and  the  Bahamas.    J.  Shellfish  Res.  15:407-420. 

Stoner,  A.  W.,  N.  Mehta,  and  T.  N.  Lee. 

1997.  Recruitment  of  Strombus  veligers  to  the  Florida  Keys 
reef  tract:  relation  to  hydrographic  events.  J.  Shellfish 
Res.  16:19-29. 

Stoner,  A.  W.  and  M.  Ray-Culp. 

2000.     Evidence  for  allee  effects  in  an  over-harvested  ma- 
rine gastropod:  density-dependent  mating  and  egg  pro- 
duction.    Mar.  Ecol.  Prog.  Ser.  202:297-302. 
Stoner,  A.  W.,  V.  J.  Sandt.  and  I.  F.  Boidron-Metairon. 

1992.     Seasonality  in  reproductive  activity  and  larval 
abundance  of  queen  conch,  Strombus  gigas.     Fish.  Bull. 
90:161-170. 
Szmant.  A.  M„  and  A.  Forrester. 

1996.  Water  column  and  sediment  nitrogen  and  phos- 
phorous distribution  patterns  in  the  Florida  Keys,  USA. 
Coral  Reefs  15:21-41. 

Thompson,  B.  E.,  S.  M.  Bay,  J.  W.  Anderson,  J.  D.  Laughlin, 
D.  J.  Greenstein,  and  D.  T.  Tsukada. 

1989.     Chronic  effects  of  contaminated  sediments  on  the  ur- 
chin Lytechinus  pictas.     Environ.  Toxicol.  Chem.  8:629-637 
van  Treeck,  P.  and  H.  Schuhmacher. 

1997.  Initial  survival  of  coral  nubbins  transplanted  by  a 
new  coral  transplantation  technology-options  for  reef  re- 
habilitation.    Mar.  Ecol.  Prog.  Ser.  150:287-292. 

Weil,  E.,  and  R.  Laughlin. 

1984.     Biology,  population  dynamics,  and  reproduction  of 
the  queen  conch  Strombus  gigas  Linne  in  the  Archipielago 
de  los  Roques  National  Park.     J.  Shellfish  Res.  4:45-62. 
Wilson-Ormond.  E.  A.,  M.  S.  Ellis,  and  E.  N.  Powell. 

1994.    The  effect  of  proximity  to  gas  producing  platforms 
on  size,  stage  of  reproductive  development  and  health  in 
shrimp  and  crabs.     J.  Shellfish  Res.  13:306. 
Zar.J.H. 

1996.  Biostatistical  analysis,  3rd  ed.,  662  p.  Prentice  Hall, 
Upper  Saddle  River,  NJ. 


289 


Abstract— This  study  examines  gene- 
tic variation  at  five  microsatellite  loci 
and  at  the  vesicle  membrane  protein 
locus,  pantophysin,  of  Atlantic  cod 
{Gadus  morhua)  from  Browns  Bank, 
Georges  Bank,  and  Nantucket  Shoals. 
The  Nantucket  Shoals  sample  rep- 
resents the  first  time  cod  south  of 
Georges  Bank  have  been  genetically 
evaluated.  Heterogeneity  of  allelic  dis- 
tribution was  not  observed  (P>0.05) 
between  two  temporally  separated 
Georges  Bank  samples  indicating 
potential  genetic  stability  of  Georges 
Bank  cod.  When  Bonferroni  correc- 
tions («=0.05,  P<0.017)  were  applied 
to  pairwise  measures  of  population 
differentiation  and  estimates  of  FST, 
significance  was  observed  between 
Nantucket  Shoals  and  Georges  Bank 
cod  and  also  between  Nantucket  Shoals 
and  Browns  Bank  cod.  However,  nei- 
ther significant  differentiation  nor  sig- 
nificant estimates  ofFST  were  observed 
between  Georges  Bank  and  the  Browns 
Bank  cod.  Our  research  suggests  that 
the  cod  spawning  on  Nantucket  Shoals 
are  genetically  differentiated  from 
cod  spawning  on  Browns  Bank  and 
Georges  Bank.  Managers  may  wish  to 
consider  Nantucket  Shoals  cod  a  sepa- 
rate stock  for  assessment  and  manage- 
ment purposes  in  the  future. 


Genetic  differentiation  among  Atlantic  cod 
(.Gadus  morhua)  from  Browns  Bank, 
Georges  Bank,  and  Nantucket  Shoals 

Christopher  Lage 

Department  of  Biological  Sciences 
Murray  Hall 
University  of  Maine 
Orono,  Maine  04469 

Kristen  Kuhn 

Irv  Kornfield 

School  of  Marine  Sciences 

Murray  Hall 

University  of  Maine 

Orono,  Maine  04469 

E-mail  address  (for  I  Kornfield,  contact  author):  irvk@maine.edu 


Manuscript  approved  for  publication 
5  November  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:289-297  (2004). 


The  Atlantic  cod  (Gadus  morhua)  is  a 
migratory  gadid  found  on  both  sides  of 
the  North  Atlantic.  In  the  Northwest 
Atlantic,  cod  are  distributed  nearly 
continuously  along  the  continental 
shelf  from  Greenland  to  North  Caro- 
lina, spawning  in  relatively  discrete, 
temporally  stable  areas,  and  differ- 
ent regions  are  regarded  as  different 
management  units  defined  primarily 
by  latitude  and  bathymetry  ( Ruzzante 
et  al,  1998).  Atlantic  cod  historically 
supported  economically  important 
fisheries  in  the  Northwest  Atlantic 
(Halliday  and  Pinhorn,  1996).  In  U.S. 
waters,  cod  are  assessed  and  managed 
as  two  stocks:  1)  Gulf  of  Maine  and  2) 
Georges  Bank  and  southward  (includ- 
ing Nantucket  Shoals).  Growth  rates 
differ  between  the  two  stocks;  growth 
is  slower  in  the  Gulf  of  Maine  compared 
to  growth  in  Georges  Bank  (Pentilla  et 
al.,  1989 );  each  stock  is  exploited  by  the 
same  gear  type  and  may  show  similar 
biological  responses  towards  such  gear 
selection.  Although  both  stocks  sup- 
port important  commercial  and  recre- 
ational fisheries,  each  is  overexploited 
and  remains  at  a  low  biomass  level 
(Mayo  and  O'Brien,  1998;  O'Brien  and 
Munroe,  2001;  Mayo  et  al„  2002).  Over- 
exploitation  may  result  in  significant 
life-history  changes  such  as  a  decline 
in  time  to  reproductive  maturity  which 
has  been  observed  in  Georges  Bank  cod 
(O'Brien,  1998);  such  changes  maybe  a 


compensatory  response  to  overfishing 
but  may  also  be  influenced  by  shifts  in 
underlying  genetic  control  (Policansky. 
1993). 

Commercial  fisheries  are  conduct- 
ed year  round,  using  primarily  otter 
trawls  and  gill  nets.  The  Canadian 
fishery  on  Georges  Bank  is  managed 
under  an  individual  quota  system. 
United  States  cod  fisheries  are  man- 
aged under  the  New  England  Fishery 
Management  Council's  Northeast  Mul- 
tispecies  Fishery  Management  Plan 
(FMP)1  as  implemented  by  the  U.S. 
Federal  Register,  50  CFR  Part  648 
(U.S.  Federal  Register,  2003).  Under 
this  FMP,  cod  are  included  in  a  com- 
plex of  15  groundfish  species  managed 
by  time  and  area  closures,  trip  limits, 
gear  restrictions,  minimum  size  limits, 
days-at-sea  restrictions,  and  a  permit 
moratorium.  The  FMP's  goal  is  to  re- 
duce fishing  mortality  to  levels  that 
will  allow  stocks  within  the  complex 
to  initially  rebuild  above  minimum 
biomass  thresholds,  and,  ultimately,  to 
remain  at  or  near  target  levels. 

When  ecological  and  evolutionary 
processes  are  responsible  for  stock 
structuring,  it  is  necessary  to  incorpo- 


1  New  England  Fishery  Management  Coun- 
cil. 2003.  Northeast  Multispecies 
Fishery  Management  Plan.  NEFMC. 
50  Water  St.,  Mill  2,  Newburyport.  MA, 
01950 


290 


Fishery  Bulletin  102(2) 


rate  them  into  strategies  designed  to  manage  exploited 
species  (Avise,  1998).  High  dispersive  capabilities  of  many 
marine  fish  often  correlate  with  low  levels  of  population  di- 
vergence over  vast  areas  (Ward  et  al.,  1994;  Graves,  1998) 
and  may  be  particularly  true  for  species  characterized  by 
high  fecundity,  large  population  size,  and  potentially  long- 
distance egg  and  larval  dispersal.  Although  marine  fish 
predominantly  have  high  dispersal  rates  and  low  levels  of 
population  structuring,  migratory  species  with  continuous 
distributions  may  develop  and  maintain  stock  structure 
if  they  show  fidelity  to  natal  spawning  sites  or  limited 
egg  and  larval  dispersal.  Fidelity  to  natal  grounds  has 
been  shown  in  Greenland-Iceland  cod  (Frank,  1992)  and 
Georges  Bank  haddock  (Polacheck  et  al.,  1992).  Genetic 
divergence  between  areas  originates  when  populations  are 
formed  or  through  the  restriction  of  gene  flow.  Cod  in  some 
regions  are  known  to  migrate  long  distances,  whereas  in 
other  regions  they  are  nearly  stationary  (Lear  and  Green, 
1984).  Tagging  studies  in  the  Gulf  of  Maine  show  little 
exchange  between  the  region  east  of  Browns  Bank  and 
Georges  Bank,  and  the  inner  Gulf  of  Maine  (Hunt  et  al., 
1999);  however  exchange  has  been  reported  among  Bay  of 
Fundy,  southern  Nova  Scotia,  Browns  Bank,  and  Georges 
Bank  populations  (Klein-MacPhee,  2002).  Such  exchange 
among  cod  from  different  management  areas  may  be  im- 
portant for  stock  assessments  and  management  practices. 
Determining  underlying  genetic  structure  of  spawning 
stocks  is  paramount  to  the  conservation  and  management 
of  overexploited  species. 

In  the  last  30  years  the  use  of  molecular-based  stud- 
ies in  fisheries  science  has  become  common  (Shaklee 
and  Bentzen,  1998).  In  cod,  a  number  of  studies  have 
used  allozymes  (Moller,  1968;  Jamieson,  1975;  Cross  and 
Payne,  1978;  Dahle  and  Jorstad,  1993),  but  their  use  and 
sensitivity  are  limited  because  of  weak  statistical  power 
resulting  from  low  levels  of  polymorphism  and  because 
of  processes  of  balancing  selection  (Mork  et  al.,  1985; 
Pogson  et  al,  1995).  Mitochondrial  DNA  (mtDNA)  char- 
acterization among  Northwest  Atlantic  cod  indicates  that 
there  is  limited,  albeit  significant,  population  structuring 
throughout  most  the  species'  range  (Smith  et  al.,  1989; 
Carr  and  Marshall,  1991;  Pepin  and  Carr,  1993;  Carr  et 
al.,  1995;  Arnason  and  Palsson,  1996).  Genetic  divergence 
at  the  vesicle  membrane  protein  locus,  pantophysin  (Panl), 
originally  called  GM798  and  identified  as  synaptophysin 
{SypT)  (Fevolden  and  Pogson,  1997),  has  been  reported 
among  populations  of  cod  from  the  Northwest  Atlantic 
(Pogson,  2001;  Pogson  et  al.,  2001),  Norway  and  the  Arctic 
(Fevolden  and  Pogson,  1997),  and  Iceland  (Jonsdottir  et 
al.,  1999,  2002).  High  levels  of  variation  have  been  re- 
ported at  nuclear  RFLP  loci  (Pogson  et  al.,  1995;  Pogson 
et  al.,  2001),  and  especially  at  microsatellite  loci  (Bentzen 
et  al.,  1996;  Ruzzante  et  al.,  1996a,  1996b,  1997,  1998; 
Beacham  et  al.,  1999;  Miller  et  al.,  2000;  Ruzzante  et  al., 
2000,  2001).  By  using  microsatellites,  significant  genetic 
structuring  has  been  detected  among  cod  populations  on 
major  continental  shelves  and  on  neighboring  banks  that 
are  separated  by  deep  channels  and  have  gyre-like  circula- 
tion patterns  hypothesized  to  act  as  retention  mechanisms 
for  eggs  and  larvae  (Ruzzante  et  al.,  1998).  Although  both 


Browns  and  Georges  Bank  maintain  persistent  gyre-like 
circulation  patterns  that  may  act  to  retain  eggs  and  lar- 
vae, they  are  separated  by  the  Fundian  Channel  (>260  m ) 
which  may  pose  a  barrier  to  juvenile  and  adult  migration 
(Klein-MacPhee,  2002).  Evaluation  of  Northwest  Atlantic 
haddock  by  using  microsatellites  showed  similarly  signifi- 
cant stock  structuring  from  Newfoundland  to  Nantucket 
Shoals  (Lage  et  al.,  2001).  Current  assessment  and  man- 
agement of  cod  in  U.S.  waters  combine  Georges  Bank  with 
the  regions  to  its  south  including  Nantucket  Shoals.  This 
study  investigates  genetic  stock  structure  among  cod  from 
this  region  and  provides  additional  insight  for  scientists 
and  managers. 


Materials  and  methods 

Sampling 

Samples  of  adult  cod  were  collected  through  the  U.S. 
National  Marine  Fisheries  Service  and  the  Canadian 
Department  of  Fisheries  and  Oceans  groundfish  surveys 
between  1994  and  2000.  Adult  cod  were  obtained  from  each 
of  the  following  spawning  grounds  (Fig.  1):  Browns  Bank 
(July  1994,  n  =  30),  Georges  Bank  (March  1994,  n  =  48; 
March  1999.  >*=96;  In  =  144),  and  Nantucket  Shoals 
(March  2000.  n  =  97).  Blood  or  tissue  (or  both)  was  obtained 
from  individual  fish  and  preserved  in  95%  ethanol  for 
subsequent  DNA  extraction. 

DNA  extraction,  amplification,  and  visualization 

DNA  was  extracted  by  using  either  a  Qiamp  DNA  Mini 
Kit  ( Qiagen  Inc.,  Valencia,  CA)  or  by  following  a  published 
protocol  designed  for  nucleated  blood  cells  ( Ruzzante  et  al., 
1998).  Five  microsatellite  loci — Grnol,  Gmol32  (Brooker 
et  al.,  1994),  Gmo8,  Gmol9,  Gmo34  (Miller  et  al,  2000). 
and  the  pantophysin  locus.  Panl  (Fevolden  and  Pogson, 
1997;  Pogson,  2001) — were  used  to  evaluate  genetic  diver- 
sity. Polymerase  chain  reactions  (PCR)  of  all  loci  were  per- 
formed in  an  Eppendorf  Mastercycler  Gradient  thermal 
cycler.  Final  concentrations  of  reagents  in  a  25  uL  PCR 
cocktail  were  as  follows:  -10  ng  of  genomic  DNA.  lxPCR 
buffer  pH  9.5  110  mM  KC1,  20  mM  Tris-HCl  pH  8.3, 10  mM 
(NH4)2SOJ,  1.5  mM  MgCl2,  200  /iM  each  dNTP,  0.15  nM 
forward  primer.  0.15  ,«M  reverse  primer  (unlabeled  for  the 
Panl  locus  and  5-labeled  with  a  TET,  FAM,  or  HEX  ABI 
dye  for  all  microsatellite  loci),  and  0.75  units  of  Taq  DNA 
polymerase.  PCR  conditions  were  as  follows:  initial  5  min 
at  95°C.  30  cycles  of  denaturing  at  95°C  for  1  min,  anneal- 
ing at  50°C  (Gmo8,  Gmol9,  and  Gmo34>.  55°C  (Pan I),  and 
57°C  (Gmol  and  Gmol32)  for  1  min  30  s,  and  extending 
at  72°C  for  1  min  30  s  with  a  final  extension  of  72°C  for  10 
min.  Gmol9  and  G/?2o34,  as  well  as  Gmol  and  Gmol32, 
were  multiplexed  in  two  25  «L  PCR  reactions.  Flourescent 
microsatellite  PCR  products  were  visualized  on  an  ABI377 
automated  DNA  sequencer  (Perkin-Elmer  Corporation. 
Foster  City,  CA)  and  were  analyzed  by  using  GeneScan 
(vers.  2.1)  and  Genotyper  (vers.  2.1)  software  programs 
(Perkin-Elmer  Corporation,  Foster  City,  CA).  Panl  PCR 


Lage  et  al.:  Genetic  structuring  of  Gadus  morhua 


291 


44 


«    42°   - 


40     - 


West  longitude 


Figure  1 

Map  of  Northwest  Atlantic  sampling  regions  for  Atlantic  cod  [Gadus  morhua). 
Dashed  lines  indicate  the  100-m  isobath. 


products  were  digested  with  the  restriction  endonuclease 
Dral  for  at  least  2  hours  at  37°C  and  visualized  on  2% 
agarose  gels  to  determine  presence  of  PanlA  or  PanlB  (or 
both)  allelic  variants. 

Genetic  analyses 

Samples  were  tested  for  conformation  to  Hardy- Weinberg 
equilibrium  (HWE)  expectations  by  the  Markov  chain 
method  (Guo  and  Thomson,  1992)  by  resampling  2000 
iterations  per  batch  for  200  batches  with  GENEPOP 
vers.  3. Id  (CEFE/CNRS,  Montpelier,  France;  available 
at  http://www.cefe.cnrs-mop.fr/)  (Raymond  and  Rous- 
set,  1995);  the  null  hypothesis  tested  was  random  union 
of  gametes  within  a  population.  All  loci  were  tested  for 
genotypic  disequilibrium  across  the  entire  data  set,  as  well 
as  for  individual  populations  by  using  Markov  chain  resa- 
mpling with  2000  iterations  per  batch  for  200  batches  in 
GENEPOP  vers.  3. Id;  the  null  hypothesis  tested  was  that 


the  genotypes  at  one  locus  are  independent  from  genotypes 
at  the  other  locus. 

Tests  of  allelic  and  genotypic  differentiation  among 
and  between  population  samples  were  conducted  by  using 
FSTAT  2.9.1  (UNIL,  Lausanne,  Switzerland;  available  at 
http://www.unil.ch/izea/softwares/fstat.html)  (Goudet, 
1995);  the  null  hypothesis  tested  was  homogeneous  distri- 
butions across  samples.  Because  alleles  can  be  considered 
as  independent  when  samples  conform  to  HWE,  it  is  valid 
to  permute  alleles  among  samples  to  test  for  population 
differentiation.  On  the  other  hand,  when  HWE  is  rejected 
within  samples,  alleles  within  an  individual  cannot  be 
considered  independent,  and  thus  permuting  genotypes 
among  samples  is  the  only  valid  permutation  scheme.  In 
both  cases,  contingency  tables  were  generated  and  classi- 
fied by  using  the  log-likelihood  statistic  G  (Goudet  et  al, 
1996).  Estimates  of  among-  and  between-sample  FST's 
were  generated  according  to  Weir  and  Cocherham  (1984) 
with  FSTAT  vers.  2.9.1  and  GENETIX  vers.  4.04  (available 


292 


Fishery  Bulletin  102(2) 


Table  1 

Genetic  variation  in  sampled  populations  of  Atlantic  cod  (Gadus  morhua)  and  P-value  (in  parentheses  I  for  among  sample  popula- 
tion differentiation,  n  =  observed  number  of  alleles;  H0  =  observed  heterozygosity;  FST  =  the  among-sample  P-valueJ;  bp  =  base 
pairs;  *  =  significant  deviation  from  HWE  («=0.05.  P<0.0083);  t  =  P  <0.05;  ?  P=  sO.OOl 

Locus 

In 

Allelic  range 

Browns  Bank 

Georges 

Bank 

Nantucket  Shoals 

Fst 

Differentiation 

H0 

n 

Ho 

n 

Ho 

n 

Panl 

2 

PcmlA/PanlB 

0.0385 

2 

0.0397 

2 

0.0222 

2 

-0.0052(0.767) 

0.7820 

Gmol 

5 

96-110  bp 

0.1000 

4 

0.1319 

5 

0.1505 

5 

0.0019(0.307) 

0.3210 

Gmo8 

23 

118-201  bp 

0.8929 

17 

0.8370 

19 

0.8444* 

20 

0.0001(0.437) 

0.5200 

Gmol9 

26 

120-237  bp 

0.8846 

17 

0.8148* 

25 

0.7975* 

23 

-0.0021(0.857) 

0.8320 

Gmo34 

11 

82-120  bp 

0.7778 

5 

0.5683 

11 

0.6630 

7 

-0.0033(0.797) 

0.8400 

Gmol32 

21 

105-155  bp 

0.8333 

13 

0.8214 

17 

0.7727 

16 

0.0255(0.000)? 

0.0010? 

All  loci 

88 

— 

— 

— 

— 

— 

— 

— 

0.0047(0.001)? 

0.0240t 

at  http://www.univ-montp2.fr/~genetix/genetix/genetix. 
htm)  (Belkhir  et  al.2).  Significance  of  FST  estimates  was 
determined  with  2000  randomizations.  Tests  of  population 
differentiation  and  estimations  of  FST  were  calculated  at 
each  locus  individually  and  at  all  loci  combined.  To  correct 
for  simultaneous  comparisons,  standard  Bonferroni  cor- 
rections were  applied  by  using  a  global  significance  level 
of  0.05  (Rice,  1989). 


Results 

Genetic  variation 

Observed  numbers  of  alleles,  allelic  ranges,  heterozygosi- 
ties, and  deviations  from  HWE  are  presented  in  Table  1. 
All  tests  of  genotypic  linkage  disequilibrium  were  non- 
significant at  the  global  and  population  levels.  When 
Bonferroni  corrections  for  multiple  tests  were  applied  to 
tests  of  HWE  (a=0.05,  P<0.0083),  the  pooled  Georges 
Bank  sample  deviated  significantly  at  Gwol9,  and  the 
Nantucket  Shoals  sample  deviated  at  Gmo8  and  at  Gmol9. 
Interestingly,  these  two  loci  have  the  greatest  variation 
based  on  number  of  alleles  and  heterozygosity.  In  each 
case,  the  cause  of  deviation  was  due  to  an  excess  of  homo- 
zygotes.  Population  samples  that  generally  conform  to 
expectations  of  random  mating  but  show  a  lack  of  concor- 
dance to  HWE  at  one  or  more  loci  may  be  due  to  a  number 
of  processes  including  null  alleles,  genetic  drift,  admix- 
ture, selection,  and  insufficient  sampling  (e.g.,  Ruzzante, 
1998).  Possible  explanations  of  homozygote  excess  include 
sample  admixture  ( Wahlund  effect)  or  drift;  however  these 
explanations  are  unlikely  because  one  would  expect  to  see 
similar  results  at  all  loci.  More  likely  explanations  are  the 


-  Belkhir  K.,  P.  Borsa,  L.  Chikhi,  N.  Raufaste,  and  F.  Bon- 
homme.  2002.  GENETIX  4.04,  logiciel  sous  Windows  TM 
pour  la  genetique  des  populations.  Laboratoire  Genome, 
Populations,  Interactions,  CNRS  UMR  5000,  Universite  de 
Montpellier  II,  Montpellier  i  France). 


presence  of  null  alleles  or  selection.  Deviations  of  HWE 
at  Gmo8  and  Gmol9  were  not  observed  in  all  population 
samples,  indicating  that  null  alleles  were  not  present  at  a 
global-level  but  may  be  present  at  the  population-level  for 
these  two  loci.  Subsequently,  any  significant  population 
structuring  observed  at  these  loci  should  be  viewed  with 
caution  (see  below). 

Population  structure 

Tests  of  population  structure  are  shown  in  Table  2.  Al- 
though Gmo8  and  Gmol9  showed  significant  deviations 
from  HWE,  they  did  not  support  any  significant  population 
structuring  even  when  tests  of  population  differentiation 
were  performed  without  assuming  conformation  to  HWE 
(i.e.,  permuting  among  genotypes  rather  than  alleles). 
Heterogeneity  of  allelic  distribution  was  not  observed 
(P>0.05)  between  the  1994  and  1999  Georges  Bank 
samples  at  each  locus  individually  and  at  all  loci  combined, 
thus  indicating  potential  genetic  stability  of  Georges  Bank 
cod.  These  samples  were  subsequently  pooled  to  form 
a  single  Georges  Bank  population  sample  to  facilitate 
statistical  analyses  by  allowing  for  better  estimations  of 
allele  frequencies  and  by  reducing  the  number  of  pairwise 
tests.  Tests  of  population  differentiation  among  samples 
showed  significant  divergence  at  Gmol32  (P<0.01)  and 
at  all  loci  combined  (P<0.05).  When  Bonferroni  correc- 
tions were  applied  to  pairwise  measures  of  divergence 
(«=0.05,  P<0.017),  significance  was  observed  between 
Nantucket  Shoals  and  Georges  Bank  at  Gwol32  and  at 
all  loci  combined,  and  also  between  Nantucket  Shoals  and 
Browns  Bank  at  Gwol32.  No  significant  differentiation 
was  observed  between  individual  or  pooled  Georges  Bank 
samples  and  the  Browns  Bank  sample. 

Significant  among-population  FST  values  were  esti- 
mated at  Gmol32  (0.0255,  P<0.001)  and  at  all  loci  com- 
bined (0.0047,  P<0.01).  When  Bonferroni  corrections  were 
applied,  significant  pairwise-population  FST  values  were 
estimated  between  Nantucket  Shoals  and  Browns  Bank 
at  Gwol32  (0.0624,  P<0.001)  and  at  all  loci  combined 


Lage  et  al.:  Genetic  structuring  of  Gadus  morhua 


293 


Table  2 

Genetic  structuring  in 

sampled  populations  of  Atlantic  cod  ^Gadus  morhua):  Above  diagonal 

are 

P 

values  for 

pairwise  differen- 

tiation.  Below  diagona 

are  pairwise 

FST  values;  upper  va 

ues  are  for  all  loci  combined;  lower  values  are  for  Cm 

3132.  *=P<0.0167 

1 1<  =  0.05  for  three  comparisons);  **  = 

P 

sO.001. 

Browns  Bank 

Georges  Bank 

Nantucket  Shoals 

Browns  Bank 

— 

0.5440 
0.1120 

0.2970 
0.0020* 

Georges  Bank 

0.0012 
0.0124 



0.0030* 
0.0010** 

Nantucket  Shoals 

0.0114* 
0.0624** 

0.0045* 
0.0226** 

— 

(0.0114,  P<0.017),  and  between  Nantucket  Shoals  and 
Georges  Bank  at  Gmol32  (0.0226,  P<0.001)  and  at  all 
loci  combined  (0.0045,  P<0.0i7).  Estimates  of  FST  values 
between  Browns  Bank  and  Georges  Bank  samples  were 
all  nonsignificant.  No  significant  genetic  structuring  was 
observed  in  any  comparison  when  Gmol32  was  excluded 
from  the  analysis. 


Discussion 

Georges  Bank,  a  large,  shallow  offshore  bank  located  along 
the  southern  edge  of  the  Gulf  of  Maine  off  the  U.S.  and 
Canadian  coasts  (Fig.  II,  supports  a  large  fish  biomass. 
High  primary  productivity  and  tightly  bound  system 
energetics  on  the  bank  result  in  relatively  stable  levels  of 
overall  biomass  and  total  fish  production,  although  major 
shifts  in  species  composition  routinely  occur  (Fogarty  and 
Murawski,  1998).  The  largest  spawning  aggregation  of 
cod  on  Georges  Bank  is  found  on  the  Northeast  Peak,  a 
gravel  region  that  is  an  important  habitat  for  the  early 
demersal  phase  of  cod,  and  may  represent  a  limiting 
resource  for  this  stock  (Lough  and  Bolz  1989;  Langton  et 
al..  1996).  The  bank  maintains  its  own  circulation  pattern 
in  a  slow  clockwise  gyre  which  may  act  as  a  transportation 
and  retention  mechanism  for  planktonic  eggs  and  larvae 
(Smith  and  Morse,  1984;  Lough  and  Bolz,  1989).  There 
may  be  exchange  of  biota  among  regions  by  episodic  fluxes 
of  shelf  water  carrying  eggs  and  larvae  away  from  the  Sco- 
tian  Shelf  and  Browns  Bank  onto  Georges  Bank  (Cohen 
et  al.,  1991;  Townsend  and  Pettigrew,  1996;  Bisagni  and 
Smith,  1998).  Once  on  Georges  Bank,  planktonic  eggs 
and  larvae  may,  depending  on  depth,  be  entrained  and 
transported  to  gravel  settlement  sites  along  the  western 
edge  of  Georges  Bank  (Smith  and  Morse  1984;  Lough  and 
Bolz,  1989;  Werner  et  al.,  1993).  However,  wind-driven 
advection  may  cause  egg  and  larval  loss  from  the  North- 
east Peak  and  southern  flank  of  Georges  Bank  (Lough 
et  al.,  1989).  Cod  spawned  in  the  Gulf  of  Maine  usually 
drift  southeasterly  towards  Georges  Bank  because  of  the 
counterclockwise  Gulf  of  Maine  gyre,  but  the  extent  of  egg 
and  larval  exchange  between  these  regions  is  unknown 
(Serchuketal.,  1994). 


Cod  have  been  found  from  the  surface  to  depths  greater 
than  450  meters;  however  few  cod  proximate  to  the  Gulf 
of  Maine  occur  deeper  than  180  meters  (Klein-MacPhee, 
2002).  Browns  Bank  and  Georges  Bank  are  bathymetri- 
cally  separated  by  the  relatively  deep  (>260  meters)  Fun- 
dian  Channel  which  may  act  as  a  barrier  to  adult  migra- 
tion, whereas  Georges  Bank  and  Nantucket  Shoals  are 
separated  by  the  relatively  shallow  (<100  meters)  Great 
South  Channel.  Although  the  latter  channel  is  probably 
not  a  significant  barrier  to  adult  migration,  it  is  an  area  of 
strong  recirculation  towards  Georges  Bank  and  could  limit 
egg  and  larval  dispersal.  Tagging  studies  show  little  ex- 
change of  adults  between  the  inner  Gulf  of  Maine  and  the 
region  east  of  Browns  Bank  and  Georges  Bank  (Hunt  et 
al.,  1999),  but  limited  exchange  has  been  reported  among 
the  Bay  of  Fundy,  southern  Nova  Scotia,  Browns  Bank, 
and  Georges  Bank  (Klein-MacPhee,  2002). 

The  likelihood  of  determining  correct  population  struc- 
ture increases  when  population  differentiation  is  sta- 
ble over  time  (Waples,  1998).  Results  from  this  study  are 
concordant  with  observations  of  temporal  stability  of  mic- 
rosatellite  variation  observed  in  Atlantic  cod  ( Ruzzante  et 
al.,  1996a,  1997,  2001).  Tests  of  population  differentiation 
and  subdivision  cannot  reject  the  maintenance  of  genetic 
homogeneity  among  Georges  Bank  cod  from  1994  to  1999 
and  thus  may  indicate  some  degree  of  temporal  genetic 
stability  among  adult  Georges  Bank  cod. 

Our  results  indicate  that  cod  from  Nantucket  Shoals 
are  genetically  distinct  from  those  from  Browns  Bank 
and  Georges  Bank,  and  cod  from  the  two  Banks  are  more 
genetically  similar.  The  observed  lack  of  heterogeneity  be- 
tween Browns  Bank  and  Georges  Bank  is  consistent  with 
gene  flow — perhaps  due  to  episodic  larval  transport  and 
some  level  of  limited  adult  exchange.  Nantucket  Shoals 
cod  may  be  genetically  distinct  because  of  egg  and  larval 
isolation  by  entrainment  in  the  Georges  Bank  gyre  or  be- 
cause of  limited  movement  of  adults  between  regions  (or 
a  combination  of  both).  Eggs  and  larvae  spawned  on  Nan- 
tucket Shoals  most  likely  do  not  enter  the  Georges  Bank 
gyre  system;  these  early  life  history  forms  may  be  retained 
on  the  shoals  or  transported  to  the  southwest  by  prevailing 
circulation  (Fogarty  and  Murawski,  1998).  Some  North 
Atlantic  cod  stocks  have  shown  substantial  differences  in 


294 


Fishery  Bulletin  102(2) 


growth  rate,  reproductive  capacity,  and  maturity  sched- 
ules related  to  temperature  (Brander,  1994).  Cod  within 
our  study  zone  generally  avoid  water  temperatures  greater 
than  10°C,  but  Nantucket  Shoals  cod  are  abundant  in 
temperatures  as  warm  as  15°C  (Klein-MacPhee,  2002). 
This  differential  thermal  tolerance  may  support  genetic 
structuring  of  Nantucket  Shoals  cod  by  selecting  against 
individuals  from  other  areas. 

Closely  related  gadid  species  such  as  cod  and  haddock 
may  exhibit  similar  patterns  of  population  genetic  struc- 
turing associated  with  similar  life  histories,  selective 
pressures,  and  ecological  constraints.  Our  results  are 
concordant  with  a  previous  study  suggesting  that  had- 
dock from  Browns  Bank  and  Georges  Bank  are  genetically 
similar  and  that  haddock  from  Nantucket  Shoals  are  dis- 
tinct (Lage  et  al.,  2001).  However,  Ruzzante  et  al.  (1998) 
observed  significant  genetic  differentiation  between  cod 
from  Browns  Bank  and  Georges  Bank.  Our  results  do  not 
agree  with  this  previously  observed  heterogeneity  between 
Browns  Bank  and  Georges  Bank  and  may  be  due  to  the 
examination  of  different  loci,  different  sampling  compari- 
sons, or  small  sample  sizes  used  in  both  studies  (or  to  a 
combination  of  these  variables)  (Ruzzante,  1998;  Smouse 
and  Chevillon,  1998). 

Among  loci,  the  greatest  genetic  differentiation  was 
observed  at  locus  Gmol32.  Indeed,  observed  statistical 
significance  of  population  differentiation  and  FST  depends 
entirely  on  Gmol32.  Length  variation  at  Gmol32  is  a  func- 
tion of  mutations  in  the  repetitive  array  and  of  an  indel  in 
a  flanking  region  (Ruzzante  et  al.,  1998)  causing  bimodal 
allele  distributions  in  some  populations.  When  compared  to 
other  microsatellite  loci,  Gwol32  has  shown  the  greatest 
differentiation  among  other  Northwest  Atlantic  cod  popu- 
lations (Bentzen  et  al.,  1996;  Ruzzante  et  al,  1998,  2001) 
and  among  Northwest  Atlantic  haddock  populations  (Lage 
et  al.,  2001)  by  an  order  of  magnitude.  Other  loci  examined 
have  not  shown  similarly  strong  measures  of  population 
structuring.  Observed  genetic  structuring  may  be  due  to 
forces  currently  determining  regional  larval  and  adult 
distributions,  including  bathymetry  and  oceanographic 
patterns.  However,  because  similar  genetic  structuring 
is  not  observed  at  all  loci,  another  potential  explanation 
is  that  structuring  at  Gmol32  is  due  to  forces  that  acted 
during  the  formation  of  populations  rather  than  to  forces 
presently  maintaining  strong  reproductive  isolation.  Once 
genetic  structure  was  generated  during  the  formation  of 
these  populations  subsequent  to  the  last  ice  age.  biological 
and  oceanographic  forces  may  have  maintained  such  struc- 
ture; other  loci  may  show  an  absence  of  structure  simply  be- 
cause it  may  not  have  been  present  when  populations  were 
formed.  Pogson  et  al.  (2001)  reported  that  the  recent  age 
of  populations,  rather  than  extensive  gene  flow,  may  be  re- 
sponsible for  weak  population  structure  in  Atlantic  cod,  and 
that  interpreting  limited  genetic  differences  among  popula- 
tions as  reflecting  high  levels  of  ongoing  gene  flow  should 
be  made  with  caution.  This  suggests  that  the  observed  lack 
of  heterogeneity  between  Browns  Bank  and  Georges  Bank 
may  not  be  due  to  high  levels  of  ongoing  gene  flow,  but  to 
similarities  between  recently  generated  populations  main- 
tained by  small  but  adequate  levels  of  gene  flow. 


Alternatively,  significant  structuring  associated  with 
Gmol32  in  both  cod  and  haddock  may  suggest  that  selec- 
tion is  acting  at  this  or  at  a  linked  locus.  Although  micro- 
satellites  themselves  may  be  generally  considered  neutral, 
there  is,  in  theory,  potential  for  physical  linkage  or  drift- 
generated  linkage  disequilibrium  between  microsatellite 
and  functional  loci.  There  is  however,  recent  evidence  of 
selection  acting  directly  on  microsatellite  loci  in  tilapia  in 
high-salinity  environments.  Streelman  and  Kocher  ( 2002 ) 
found  a  strong  functional  genotype-environment  interac- 
tion and  suggested  that  microsatellite  repeats  of  varying 
length  might  induce  promoter  conformations  that  differ  in 
their  capacity  to  bind  transcriptional  regulators.  A  poten- 
tial selective  mechanism  to  support  the  observed  genetic 
structuring  of  Nantucket  Shoals  cod  (and  haddock)  may 
be  differential  thermal  tolerance,  although  this  hypothesis 
remains  untested. 

There  is  strong  evidence  for  an  unusual  mix  of  balanc- 
ing and  directional  selection  at  the  pantophysin  (Pa«I )  lo- 
cus in  cod  but  no  evidence  of  stable  geographically  varying 
selection  among  North  Atlantic  populations  ( Pogson,  2001; 
Pogson  et  al.,  2001).  In  the  present  study,  the  Paul  locus 
showed  little  variation  and  no  significant  genetic  structur- 
ing (Table  1).  The  observed  lack  of  geographic  structuring 
at  Panl  provides  no  evidence  for  local  adaptation.  How- 
ever, our  observations  may  be  due  to  strong  balancing 
selection  among  the  geographically  proximate  populations 
examined  or,  if  Panl  is  not  under  selection,  insufficient 
variation  to  resolve  genetic  structure.  Alternatively,  this 
observed  lack  of  genetic  divergence  at  Panl  could  be  due 
to  similarities  among  recently  generated  populations  of 
North  Atlantic  cod. 

Our  research  suggests  that  the  cod  spawning  on  Nan- 
tucket Shoals  are  genetically  differentiated  from  cod 
spawning  on  Browns  Bank  and  Georges  Bank.  Managers 
may  wish  to  consider  Nantucket  Shoals  cod  as  a  separate 
stock  for  assessment  and  management  purposes  in  light  of 
current  practices  that  combine  Georges  Bank  with  regions 
to  the  south  as  one  management  unit.  Cod  from  within  the 
Gulf  of  Maine  can  potentially  migrate  along  the  coast  to 
Nantucket  Shoals  where  there  is  little  geographic  barrier 
to  adult  movement.  If  this  is  true,  the  Nantucket  Shoals 
sample  that  we  analyzed  may  actually  be  representative  of 
a  mixed  Gulf  of  Maine  and  Nantucket  Shoals  population. 
Additional  analyses  are  needed  to  evaluate  the  hypothesis 
that  Nantucket  Shoals  cod  are  genetically  distinct  from 
cod  spawning  within  the  Gulf  of  Maine.  Further  studies 
should  address  the  issues  of  temporal  stability  and  robust 
sampling  and  should  incorporate  cod  samples  from  within 
the  Gulf  of  Maine. 


Acknowledgments 

We  thank  three  anonymous  reviewers  for  their  insightful 
comments.  We  thank  the  captains,  crews,  and  scientists 
from  the  Canadian  Department  of  Fisheries  and  Oceans 
and  the  United  States.  National  Marine  Fisheries  Service. 
In  particular,  we  would  like  to  thank  Chris  Taggart.  Nina 
Shepard,  and  Holly  McBride  for  assistance  in  obtaining 


Lage  et  al.:  Genetic  structuring  of  Gadus  morhua 


295 


samples.  IK  thanks  Maryhaven  for  support.  This  work 
was  funded  by  OCE9806712  from  the  U.S.  National  Sci- 
ence Foundation. 


Literature  cited 

Arnason,  E.,  and  S.  Palsson. 

1996.     Mitochondrial  cytochrome  b  DNA  sequence  variation 

of  Atlantic  cod  {Gadus  morhua)  from  Norway.     Mol.  Ecol. 

5:715-724. 
Avise,  J.  C. 

1998.  Conservation  genetics  in  the  marine  realm.  J.  Hered. 
89:377-382. 

Beacham,  T.  D..  J.  Brattey.  K.  M.  Miller.  L.  D.  Khai,  and 
R.  E.  Withler. 

1999.  Population  structure  of  Atlantic  cod  {Gadus  morhua) 
in  the  Newfoundland  and  Labrador  area  based  on  micro- 
satellite  variation.  DFO  Canadian  Stock  Assessment 
Secretariat  Research  Document  99/35:1-32. 

Bentzen,  P.,  C.  T.  Taggart,  D.  E.  Ruzzante.  and  D.  Cook. 

1996.  Microsatellite  polymorphism  and  the  population 
structure  of  Atlantic  cod  (Gadus  morhua )  in  the  northwest 
Atlantic.     Can.  J.  Fish.  Aquat.  Sci.  53:2706-2721. 

Bisagni.  J.  J.,  and  P.  C.  Smith. 

1998.  Eddy-induced  flow  of  Scotian  Shelf  water  across 
Northeast  Channel,  Gulf  of  Maine.  Continental  Shelf 
Res.  18:515-539. 

Brander,  K.  M. 

1994.  Patterns  of  distribution,  spawning,  and  growth  in 
North  Atlantic  cod:  the  utility  of  inter-regional  compari- 
sons. International  Council  for  the  Exploration  of  the 
Seas  (ICES)  Marine  Science  Symposia  198:406-413. 

Brooker,  A.  L.,  D.  Cook,  P.  Bentzen,  J.  M.  Wright,  and  R.  W.  Doyle. 

1994.  The  organization  of  microsatellites  differs  between 
mammals  and  cold  water  fishes.  Can.  J.  Fish.  Aquat.  Sci. 
51:1959-1966. 

Carr,  S.  M..  and  H.  D.  Marshall. 

1991.  Detection  of  intraspecific  DNA  sequence  variation 
in  the  mitochondrial  cytochrome  b  gene  of  Atlantic  cod 
(Gadus  morhua)  by  the  polymerase  chain  reaction.  Can. 
J.  Fish.  Aquat.  Sci.  48:48-52. 

Carr,  S.  M.,  A.  J.  Snellen,  K.  A.  Howse,  and  J.  S.  Wroblewski. 

1995.  Mitochondrial  DNA  sequence  variation  and  genetic 
stock  structure  of  Atlantic  cod  (Gadus  morhua)  from  bay 
and  offshore  locations  on  the  Newfoundland  continental 
shelf.     Mol.  Ecol.  4:79-88. 

Cohen,  E.  B.,  D.  G.  Mountain,  and  R.  O'Boyle. 

1991.     Local-scale  versus  large-scale  factors  affecting 
recruitment.     Can.  J.  Fish.  Aquat.  Sci.  48:1003-1006. 
Cross,  T.  F.,  and  R.  H.  Payne. 

1978.     Geographic  variation  in  Atlantic  cod,  Gadus  morhua, 
off  Eastern  North  America:  a  biochemical  systematics 
approach.    J.  Fish.  Res.  Board  Can.  35:117-123. 
Dahle,  G.,  and  K.  E.  Jorstad. 

1993.     Haemoglobin  variation  in  cod-  a  reliable  marker  for 
Arctic  cod  (Gadus  morhua  L.i.     Fish.  Res.  16:301-311. 
Fevolden,  S.  E.,  and  G.  H.  Pogson. 

1997.  Genetic  divergence  at  the  synaptophysin  ( SypI )  locus 
among  Norwegian  coastal  and  north-east  Arctic  popula- 
tions of  Atlantic  cod.     J.  Fish.  Biol.  51:895-908. 

Fogarty,  M.  J.,  and  S.  A.  Murawski. 

1998.  Large-scale  disturbance  and  the  structure  of  marine 
systems:  fishery  impacts  on  Georges  Bank.  Ecol.  Appl. 
8:S6-S22. 


Frank,  K.  T. 

1992.     Demographic  consequences  of  age-specific  dispersal 

in  marine  fish  populations.     Can.  J.  Fish.  Aquat.  Sci. 

49:2222-2231. 
Goudet,  J. 

1995.  FSTAT  (vers.  1.2):  a  computer  program  to  calculate 
F-statistics.     J.  Hered.  86:485-486. 

Goudet,  J.,  M.  Raymond,  T.  De  Meeus.  and  F.  Rousset. 

1996.  Testing  divergence  in  diploid  populations.  Genetics 
144:933-940. 

Graves,  J.  E. 

1998.  Molecular  insights  into  population  structure  of  cos- 
mopolitan marine  fishes.     J.  Hered.  89:427-437. 

Guo,  S.  W„  and  E.  A.  Thomson. 

1992.     Performing  the  exact  test  of  Hardy-Weinberg  propor- 
tions for  multiple  alleles.     Biometrics  48:361-372. 
Halliday,  R.  G.,  and  A.  T  Pinhorn. 

1996.     North  Atlantic  fishery  management  systems:  a  com- 
parison of  management  methods  and  resource  trends.     J. 
Northw.  Atl.  Fish.  Sci.  20:1-143. 
Hunt,  J.  J.,  W.  T.  Stobo,  and  F.  Almeida. 

1999.  Movement  of  Atlantic  cod,  Gadus  morhua,  tagged  in 
the  Gulf  of  Maine  area.     Fish.  Bull.  97:842-860. 

Jamieson.  A. 

1975.     Enzyme  type  of  Atlantic  cod  stocks  on  the  North 
American  banks.     In  Isozymes  IV:  Genetics  and  evolu- 
tion (C.  L.  Markert  (ed.),  p.  491-515.     Academic  Press, 
New  York,  NY. 
Jonsdottir.  O.  D.  B.,  A.  K.  Imsland,  A.K.,  Danielsdottir  and 
G.  Marteinsdottir. 

2002.     Genetic  heterogeneity  and  growth  properties  of  dif- 
ferent genotypes  of  Atlantic  cod  (Gadus  morhua  L.)  at  two 
spawning  sites  off  south  Iceland.     Fish.  Res.  55:37-47. 
Jonsdottir,  O.  D.  B.,  A.  K.  Imsland,  A.K.,  Danielsdottir, 
V.  Thorsteinsson,  and  G.  Naevdal. 

1999.     Genetic  differentiation  among  Atlantic  cod  in 
south  and  south-east  Icelandic  waters:  synaptophysin 
(SypI)  and  haemoglobin  (Hbl)  variation.     J.  Fish.  Biol. 
54:1259-1274. 
Klein-MacPhee.  G. 

2002.     Cods.  Family  Gadidae.     In  Bigelow  and  Shroeder's 
fishes  of  the  Gulf  of  Maine  III  (B.  B.  Collette  and  G.  Klein- 
MacPhee,  eds.),  p.  223-260.     Smithsonian  Institution 
Press,  Washington  D.C. 
Lage,  C.  R.,  M.  Purcell,  M.  Fogarty,  and  I.  Kornfield. 

2001.     Microsatellite  evaluation  of  haddock  (Melano- 
grammus  aeglefinus)  stocks  in  the  northwest  Atlantic 
ocean.     Can.  J.  Fish.  Aquat.  Sci.  58:982-990. 
Langton,  R.  W.,  R.  S.  Steneck,  V.  Gotceitas,  F.  Juanes,  and 
P.  Lawton. 

1996.     The  interface  between  fisheries  research  and  habitat 
management.     N.  Am.  J.  Fish.  Manag.  16:1-7. 
Lear,  W.  H.,  and  J.  M.  Green. 

1984.     Migration  of  the  "northern"  Atlantic  cod  and  the 
mechanisms  involved.     In  Mechanisms  of  migration  in 
fishes  (J.  D.  McCleave.  ed.),  p.  309-315.     Plenum,  New 
York,  NY. 
Lough,  R.  G..  and  G.  R.  Bolz. 

1989.     The  movements  of  cod  and  haddock  larvae  onto 
the  shoals  of  Georges  Bank.     J.  Fish.  Biol,  (suppl.  A) 
35:71-79 
Lough.  R.  G..  P.  C.  Valentine,  D.  C.  Potter,  P.  J.  Auditore. 
G.  R.  Bolz,  J.  D.  Jielson,  and  R.  I.  Perry. 

1989.  Ecology  and  distribution  of  juvenile  cod,  and  haddock 
in  relation  to  sediment  type  and  bottom  currents  on  east- 
ern Georges  Bank.     Mar.  Ecol.  Prog.  Ser.  56:1-12. 


296 


Fishery  Bulletin  102(2) 


Mayo,  R.  K.,  and  L.  O'Brien. 

1998.     Atlantic  cod.     In  Status  of  fishery  resources  off  the 
northeastern  United  States  for  1998  (S.  H.  Clark,  ed.),  p. 
49-52.     NOAA  Tech.  Memo.  NMFS-NE-115. 
Mayo,  R.  K.,  E.  M.  Thunberg,  S.  E.  Wigley,  and  S.  X.  Cadrin. 
2002.     The  2001  assessment  of  the  Gulf  of  Maine  Atlantic 
cod  stock.     Northeast  Fish.  Sci.  Cent.  Ref.  Doc.  02-02, 
154  p. 
Miller,  K.  M.,  K.  D.  Le,  and  T.  D.  Beacham. 

2000.  Development  of  tri-  and  tetranucleotide  repeat  micro- 
satellite  loci  in  Atlantic  cod  (Gadus  morhua).  Mol.  Ecol. 
9:238-239. 

Moller,  D. 

1968.     Genetic  diversity  in  spawning  cod  along  the  Norwe- 
gian coast.     Hereditas  60:1-32. 
Mork,  J.,  Ryman,  N.,  Stahl,  G„  Utter,  F.  and  G.  Sundness. 

1985.     Genetic  variation  in  Atlantic  cod  iGadus  morhua) 
throughout  its  range.     Can.  J.  Fish.  Aquat.  Sci.  42:1580- 
1587. 
O'Brien,  L. 

1998.     Factors  influencing  rates  of  maturation  in  the 
Georges  Bank  and  the  Gulf  of  Maine  Atlantic  cod  stocks. 
NAFO  Sci.  Counc.  Res.  Doc.  98/104,  34  p. 
O'Brien,  L.,  and  N.  J.  Munroe. 

2001.  Assessment  of  the  Georges  Bank  Atlantic  cod  stock 
for  2001.  Northeast  Fish.  Sci.  Cent.  Ref.  Doc.  01-10, 
126  p. 

Pentilla,  J.  A.,  G.  A.  Nelson,  and  J.  M.  Burnett  III. 

1989.     Guidelines  for  estimating  lengths  at  age  for  18  north- 
west Atlantic  finfish  and  shellfish  species.     NOAA  Tech. 
Memo.  NMFS-F/NEC-66,  39  p. 
Pepin,  R,  and  S.  M.  Carr. 

1993.     Morphological,  meristic,  and  genetic  analysis  of 
stock  structure  in  juvenile  Atlantic  cod  (Gadus  morhua) 
from  the  Newfoundland  shelf.     Can.  J.  Fish.  Aquat.  Sci. 
50:1924-1933. 
Pogson,  G.  H. 

2001.     Nucleotide  polymorphism  and  natural  selection  at 
the  pantophysin  (Panl)  locus  in  the  Atlantic  cod,  Gadus 
morhua  (L.).     Genetics  157:317-330. 
Pogson,  G.  H.,  K.  A.Mesa,  and  R.  G.  Boutilier. 

1995.     Genetic  population  structure  and  gene  flow  in  the 
Atlantic  cod,  Gadus  morhua:  a  comparison  of  allozyme  and 
nuclear  RFLP  loci.     Genetics  139:375-385. 
Pogson,  G.  H„  C.  T.  Taggart,  K.  A.  Mesa,  and  R.  G.  Boutilier. 
2001.     Isolation  by  distance  in  the  Atlantic  cod,  Gadus 
morhua,  at  large  and  small  geographic  scales.     Evolution 
55:131-146. 
Polacheck,  T,  D.  Mountain.  D.  McMillan,  W.  Smith,  and 
P.  Berrien. 

1992.  Recruitment  of  the  1987  year  class  of  Georges  Bank 
haddock  (Melanogrammus  aeglefinus):  the  influence  of 
unusual  larval  transport.  Can.  J.  Fish.  Aquat.  Sci.  49: 
484-496. 

Policansky,  D. 

1993.  Fishing  as  a  cause  of  evolution  in  fishes.  /;;  The 
exploitation  of  evolving  populations  IT.  K.  Stokes,  J.  M. 
McGlade,  and  R.  Law,  eds.),  p.  2-18.  Springer-Verlag, 
Berlin. 

Raymond,  M.,  and  F.  Rousset. 

1995.     GENEPOP  (vers.  1.21:  population  genetics  software 
for  exact  tests  and  eumenism.     J.  Hered.  86:248-249. 
Rice,  W.  R. 

1989.  Analyzing  tables  of  statistical  tests.  Evolution  43: 
223-225. 


Ruzzante,  D.  E. 

1998.  A  comparison  of  several  measures  of  genetic  dis- 
tance and  population  structure  with  microsatellite  data: 
bias  and  sampling  variance.  Can.  J.  Fish.  Aquat.  Sci. 
55:1-14. 

Ruzzante,  D.  E„  C.  T.  Taggart,  and  D.  Cook. 

1996a.  Spatial  and  temporal  variation  in  the  genetic 
composition  of  a  larval  cod  (Gadus  morhua)  aggregation: 
cohort  contribution  and  genetic  stability.  Can.  J.  Fish. 
Aquat.  Sci.  53:2695-2705. 

Ruzzante,  D.  E.,  C.  T.  Taggart,  and  D.  Cook. 

1998.  A  nuclear  basis  for  shelf-  and  bank-scale  population 
structure  in  northwest  Atlantic  cod  (Gadus  morhua):  Lab- 
rador to  Georges  Bank.     Mol.  Ecol.  7:1663-1668. 

Ruzzante,  D.  E.,  C.  T.  Taggart,  D.  Cook,  and  S.  Goddard. 

1996b.  Genetic  divergence  between  inshore  and  offshore 
Atlantic  cod  (Gadus  morhua)  off  Newfoundland:  microsat- 
ellite DNA  variation  and  antifreeze  level.  Can.  J.  Fish. 
Aquat.  Sci.  53:634-645. 

Ruzzante,  D.  E.,  C.  T.  Taggart,  D.  Cook,  and  S.  Goddard. 

1997.  Genetic  differentiation  between  inshore  and  offshore 
Atlantic  cod  (Gadus  morhua)  off  Newfoundland:  a  test  and 
evidence  of  temporal  stability.  Can.  J.  Fish.  Aquat.  Sci. 
54:2700-2708. 

Ruzzante,  D.  E.  C.  T.  Taggart.  R.  W.  Doyle,  and  D.  Cook. 

2001.  Stability  in  the  historical  pattern  of  genetic  struc- 
ture of  Newfoundland  cod  (Gadus  morhua)  despite  the 
catastrophic  decline  in  population  size  from  1964  to 
1994.     Cons.  Genetics  2:257-269. 

Ruzzante,  D.  E.,  J.  S.  Wroblewski,  C.  T  Taggart,  R.  K.  Smedbol. 
D.  Cook,  and  S.  V.  Goddard. 

2000.     Bay-scale  population  structure  in  coastal  Atlantic 
cod  in  Labrador  and  Newfoundland,  Canada.     J.  Fish. 
Biol.  56:431-447. 
Serchuk,  F.  M.,  M.  D.  Grosslein,  R.  G.  Lough,  D.  G.  Mountain, 
and  L.  O'Brien. 

1994.     Fishery  and  environmental  factors  affecting  trends 
and  fluctuations  in  the  Georges  Bank  and  Gulf  of  Maine 
Atlantic  cod  stocks:  an  overview.     ICES  Mar.  Sci.  Symp. 
198:77-109. 
Shaklee,  J.  B.,  and  P.  Bentzen. 

1998.  Genetic  identification  of  stocks  of  marine  fish  and 
shellfish.     Bull.  Mar.  Sci.  62:589-621. 

Smith,  P.  J.,  A.  J.  Birley,  A.  Janieson,  and  C.  A.  Bishop. 

1989.     Mitochondrial  DNA  in  the  Atlantic  cod,  Gadus  mor- 
hua: lack  of  genetic  divergence  between  eastern  and  west- 
ern populations.     J.  Fish.  Biol.  34:369-373. 
Smith,  W.,  and  W.  W.  Morse. 

1984.     Retention  of  larval  haddock  Melanogrammus  aeglefi- 
nus in  the  Georges  Bank  region,  a  gyre-influenced  spawn- 
ing area.     Mar.  Ecol.  Prog.  Ser.  24:1-13. 
Smouse,  P.  E.,  and  C.  Chevillon. 

1998.     Analytical  aspects  of  population-specific  DNA  finger- 
printing for  individuals.     J.  Hered.  89:143-150. 
Streelman,  J.  T,  and  T.  D.  Kocher. 

2002.  Microsatellite  variation  associated  with  prolactin 
expression  and  growth  of  salt-challenged  tilapia.  Physiol. 
Genomics  9:1-4. 

Townsend,  D.  W.,  and  N.  R.  Pettigrew. 

1996.     The  role  of  frontal  currents  in  larval  fish  transport 
on  Georges  Bank.     Deep-Sea  Res.II.  43:1773-1792. 
United  States  Federal  Register. 

2003.  50  CFR  Part  648.  [Available  at  http://www.gpoac- 
cess.gov;  accessed  1  May  2003.1 

Waples,  R.  S. 

1998.     Separating  the  wheat  from  the  chaff:  patterns  of 


Lage  et  al.:  Genetic  structuring  of  Gadus  morhua 


297 


genetic  differentiation  in  high  gene  flow  species.     J.  Hered. 
89:438-450. 
Ward.  R.  D.,  M.  Woddwark,  and  D.  O.  F.  Skibinski. 

1994.     A  comparison  of  genetic  diversity  levels  in  marine, 
freshwater,  and  anadromous  fishes.     J.  Fish.  Biol.  44: 
213-232. 
Weir,  B.  S.,  and  C.  C.  Cockerham. 

1984.     Estimating  F-statistics  for  the  analysis  of  population 
structure.     Evolution  38:1358-1370. 


Werner,  F.  E.,  F.  H.  Page,  D.  R.  Lynch,  J.  W.  Loder,  R.  G.  Lough, 
R.  I.  Perry,  D.  A.  Greenberg,  and  M.  Sinclair. 

1993.  Influences  of  mean  advection  and  simple  life  behavior 
on  the  distribution  of  cod  and  haddock  early  life  stages  on 
Georges  Bank.     Fish.  Oceanogr.  2:43-64. 


298 


Abstract -A  major  cause  of  the  steep 
declines  of  American  oyster  ( Crassos- 
trea  virginica)  fisheries  is  the  loss 
of  oyster  habitat  through  the  use  of 
dredges  that  have  mined  the  reef 
substrata  during  a  century  of  intense 
harvest.  Experiments  comparing  the 
efficiency  and  habitat  impacts  of  three 
alternative  gears  for  harvesting  oys- 
ters revealed  differences  among  gear 
types  that  might  be  used  to  help  im- 
prove the  sustainability  of  commercial 
oyster  fisheries.  Hand  harvesting  by 
divers  produced  25-32^  more  oysters 
per  unit  of  time  of  fishing  than  tradi- 
tional dredging  and  tonging.  although 
the  dive  operation  required  two  fish- 
ermen, rather  than  one.  Per  capita 
returns  for  dive  operations  may  none- 
theless be  competitive  with  returns  for 
other  gears  even  in  the  short  term  if 
one  person  culling  on  deck  can  serve 
two  or  three  divers.  Dredging  reduced 
the  height  of  reef  habitat  by  34rr .  sig- 
nificantly more  than  the  23fr  reduction 
caused  by  tonging,  both  of  which  were 
greater  than  the  6Q<  reduction  induced 
by  diver  hand-harvesting.  Thus,  con- 
servation of  the  essential  habitat  and 
sustainability  of  the  subtidal  oyster 
fishery  can  be  enhanced  by  switch- 
ing to  diver  hand-harvesting.  Man- 
agement schemes  must  intervene  to 
drive  the  change  in  harvest  methods 
because  fishermen  will  face  relatively 
high  costs  in  making  the  switch  and 
will  not  necessarily  realize  the  long- 
term  ecological  benefits. 


Conserving  oyster  reef  habitat  by  switching 
from  dredging  and  tonging  to  diver-harvesting 


Hunter  S.  Lenihan 

Bren  School  ol  Environmental  Science  and  Management 
University  of  California,  Santa  Barbara 
Santa  Barbara,  California  93106-5131 
E-mail  address:  tenihamfflbren  ucsb  edu 


Charles  H.  Peterson 

Institute  of  Marine  Sciences 

University  of  North  Carolina  at  Chapel  Hill 

Morehead  City,  North  Carolina  28557 


Manuscript  approved  for  publication 
25  November  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  office. 

Fish.  Bull.  102:298-305(2004). 


Commercial  fishing  for  demersal  fish- 
es and  benthic  invertebrates,  such  as 
mollusks  and  crabs,  is  commonly  under- 
taken with  bottom-disturbing  gear  that 
can  inflict  damage  to  seafloor  habitats 
(Dayton  et  al.,  1995;  Engel  and  Kvitek, 
1995;  Jennings  and  Kaiser,  1998;  Wat- 
ling  and  Norse,  1998).  Habitat  damage 
from  dredges  and  analogous  gear, 
designed  to  excavate  invertebrates 
that  are  partially  or  completely  buried 
beneath  the  surface  of  the  seafloor,  is 
generally  much  more  severe  than  the 
damage  caused  by  bottom  trawls  ( Collie 
et  al.,  2000).  Furthermore,  impacts  on 
and  recovery  from  bottom-disturbing 
fishing  gear  vary  with  habitat  type; 
generally  smaller  effects  and  more 
rapid  rates  of  recovery  are  found  for 
infauna  in  sedimentary  habitats  and 
the  most  severe  and  long-lasting 
damage  in  biogenic  habitats  that 
emerge  from  the  seafloor  (Peterson  et 
al.,  1987;  Collie  et  al.,  2000).  Such  bio- 
genic habitats  include  seagrass  beds, 
fields  of  sponges  and  bryozoans.  and 
invertebrate  reefs.  Biogenic  reefs  that 
provide  important  ecosystem  services 
such  as  habitat  for  other  organisms 
include  not  only  tropical  coral  reefs 
but  also  temperate  reefs  constructed 
by  oysters  (Bahr  and  Lanier,  1981; 
Lenihan  et  al.,  2001),  polychaetes  like 
Petaloproctus  (Wilson,  1979;  Reise, 
1982),  and  vermetid  gastropods  (Saf- 
riel,  1975).  The  recovery  of  such  emer- 
gent invertebrate  reefs  is  a  slow  process 
because  of  the  relative  longevity  of  the 


organisms  that  provide  structure  for 
the  reef  after  they  die  and  because  of 
the  nature  of  reefs  as  accumulations  of 
multiple  generations  of  reef  builders. 

One  widespread  temperate  reef 
builder,  the  American  oyster  iCrassos- 
trea  virginica,  also  known  as  the  "east- 
ern oyster,"  Am.  Fish.  Soc),  has  been 
especially  affected  by  bottom-disturb- 
ing fishing  gear  as  the  target  of  fisher- 
ies. More  than  one  hundred  years  of 
dredging  and  tonging  oysters  in  the 
Chesapeake  Bay  and  Pamlico  Sound 
have  caused  severe  degradation  of  the 
oyster  reef  matrix  (deAlteris,  1988; 
Hargis  and  Haven,  1988),  such  that 
reef  area  and  elevation  have  been  dra- 
matically reduced  (Rothschild  et  al., 
1994;  Lenihan  and  Peterson,  1998). 
Reduction  in  reef  height  has  a  serious 
consequence  for  the  oyster  population 
because  one  function  of  naturally  tall 
subtidal  oyster  reefs  is  to  elevate  the 
oysters  up  into  the  mixed  surface  layer 
of  the  estuary;  this  layer  of  mixed  sur- 
face water  allows  them  to  avoid  mass 
mortality  from  persistent  exposure  to 
seasonally  anoxic  and  hypoxic  bottom 
water  (Lenihan  and  Peterson.  1998). 
Reef  height  and  structure  also  control 
reef  hydrodynamics  (e.g.,  flow  speed, 
turbulent  mixing,  and  particle  delivery 
and  deposition),  which  influence  oyster 
population  dynamics  and  production 
(Lenihan,  1999).  Consequently,  har- 
vest-related reef  destruction  and  degra- 
dation are  considered  major  factors  that 
have  led  to  declines  of  American  oys- 


Lenihan  et  al.:  Conserving  oyster  reef  habitat 


299 


ters  in  many  estuaries  located  along  the  coasts  of  the  At- 
lantic Ocean  and  Gulf  of  Mexico  (Lukenbach  et  al.,  1999). 

Loss  of  oysters  and  the  biogenic  habitat  that  they  provide 
appears  from  archaeological  and  paleontological  evidence 
to  be  a  worldwide  phenomenon  in  temperate  estuaries 
(Jackson  et  al.,  2001).  Oyster  loss  hurts  not  only  the  oys- 
ter fishery  but,  more  importantly,  imperils  the  ecosystem 
services  provided  by  the  oysters.  These  include,  especially, 
the  provision  of  emergent  habitat  and  reef-dependent  prey 
resources  for  many  fish  and  crustacean  populations  of  com- 
mercial and  recreational  importance  (Peterson  et  al..  2000; 
Lenihan  et  al,  2001;  Peterson  et  al,  2003),  the  filtration 
of  estuarine  waters  (Newell,  1988),  and  the  promotion  of 
estuarine  biodiversity  by  provision  of  hard-bottom  habitat 
in  fields  of  mobile  sediments  (Wells,  1961). 

Because  of  the  importance  of  restoring  and  sustaining 
oysters  and  their  reefs  to  serve  both  the  oyster  fishery  and 
the  ecosystem,  we  designed  a  field  test  of  the  habitat  im- 
pacts of  three  oyster  harvesting  methods:  dredging,  tong- 
ing,  and  hand  extraction  by  divers  (diver-harvesting).  Our 
study  is  a  gear  comparison,  in  which  we  assess  not  only 
the  traditional  response  variable  of  quantitative  harvest 
per  unit  of  effort  with  each  gear  but  also  the  degree  of  reef 
habitat  damage  induced  by  the  extraction  of  the  oysters 
(analogous  to  Peterson  et  al.,  1983).  We  additionally  ex- 
amine the  quality  of  the  oysters  harvested  as  a  function 
of  gear  type.  The  results  indicate  that  diver-harvesting  is 
a  more  environmentally  sound  way  of  harvesting  oysters 
than  traditional  methods  with  dredges  and  tongs  and  may 
be  more  compatible  with  conserving  oyster  reef  habitat. 


Methods 

Study  site 

Gear  comparisons  were  conducted  on  subtidal  oyster  reefs 
in  the  Neuse  River  estuary.  North  Carolina  (35°00'20"N, 
76°33'50"W).  Environmental  conditions  of  this  estuary 
are  well  described  elsewhere  (Paerl  et  al.,  1998;  Lenihan, 
1999).  Briefly,  the  estuary  is  mesohaline,  an  optimal 
habitat  for  the  American  oyster,  and  was  once  an  impor- 
tant oyster  fishery  ground  (Lenihan  and  Peterson,  1998). 
The  estuary  contains  remnants  of  many  large,  natural 
subtidal  oyster  reefs  that  have  been  intensely  mined  by 
oyster  harvesting  gear  for  over  a  century.  Dredging  is  the 
most  common  fishing  practice.  Mining  of  the  reef  matrix 
has  combined  with  sediment  loading  and  eutrophication- 
associated  hypoxia  (Paerl  et  al.,  1998)  to  degrade  the 
oyster  reef  habitats  and  greatly  reduce  oyster  populations 
(Lenihan  and  Peterson,  1998).  In  harvested  areas,  reefs 
that  were  2-3  m  tall  in  quantitative  surveys  in  the  late 
1800s  (n  =  8  reefs)  were  all  <1  m  tall  in  our  survey  con- 
ducted in  1994  —  a  modification  of  habitat  caused  entirely 
by  the  removal  of  oysters  and  shells  during  harvesting 
with  dredges  and  tongs  (Lenihan  and  Peterson,  1998). 
To  help  maintain  oyster  harvests,  the  North  Carolina 
Division  of  Marine  Fisheries  (NCDMF)  restores  oyster 
reefs  throughout  many  locations  in  the  estuary  by  creat- 
ing piles  of  oyster  shell,  or  marl,  on  the  seafloor.  These 


restored  oyster  reefs  are  also  targeted  by  oyster  fishermen 
using  dredges  and,  less  often,  using  manual  oyster  tongs 
(Marshall1). 

Experimental  oyster  reefs 

Gear  comparisons  were  conducted  in  March  1996  on  16 
subtidal  oyster  reefs  that  had  previously  been  created  in 
July  1993  as  part  of  a  reef  restoration  experiment  (Leni- 
han and  Peterson,  1998;  Lenihan,  1999)  in  collaboration 
with  NCDMF.  The  experimentally  restored  reefs  (referred 
to  as  "experimental  reefs"  in  this  gear-comparison  study) 
were  piles  of  oyster  shells  1  m  tall,  6-7  m  in  diameter 
(28.3-38.5  m2  in  area),  and  generally  hemispherical  in 
shape.  Natural  subtidal  reefs  located  elsewhere  in  the 
estuary  are  typically  larger,  rectangular  biogenic  struc- 
tures, ranging  from  8-13  m  wide  and  20-30  m  long. 
Experimental  reefs  were  constructed  in  3-4  m  of  water  on 
a  firm  and  sandy  bottom,  and  were  separated  by  at  least 
50  m.  From  the  time  of  their  construction  until  use  in  our 
experiments,  the  restored  oyster  reefs  remained  research 
sanctuaries,  protected  from  commercial  and  recreational 
shellfishing. 

As  oysters  settle  and  undergo  metamorphosis  on  the 
shells  of  other  (live  and  dead)  oysters,  to  which  they  are 
attracted  by  chemical  cues  (Tamburri  et  al.,  1992),  they 
help  cement  together  and  add  to  the  shell  matrix  of  the 
reef  over  years.  Prior  to  our  harvest  treatments,  the  ex- 
perimentally restored  reefs  were  colonized  by  at  least  three 
generations  of  oysters,  many  of  which  grew  to  adult  size 
(range  of  oyster  sizes  on  experimental  reefs  at  the  start 
of  our  experiment:  2-11  cm  in  shell  height).  Consequently, 
the  shell  matrices  of  the  reefs  had  become  somewhat  cohe- 
sive, although  probably  less  so  than  natural  oyster  reefs.  In 
February  1996,  before  initiation  of  experimental  harvests, 
there  was  no  significant  difference  in  the  mean  density 
of  adult  (>1  cm  in  shell  height)  oysters  (mean  ±SD  179.1 
±18.4/m2)  among  the  four  sets  of  four  experimental  reefs 
randomly  selected  to  receive  the  four  harvesting  treatments 
(one-way  ANOVA;  F3  12=0.29;  mean  square  error=285.06; 
P=0.83).  Experimental  reefs  in  the  Neuse  River  usually 
had  slightly  higher  oyster  densities  nearer  their  base  and 
larger  oysters  near  the  crest  (see  Lenihan,  1999). 

Experimental  harvests 

We  compared  three  types  of  oyster-harvesting  techniques: 
dredging,  hand-tonging,  and  diver-harvesting.  In  March 
1996,  each  of  16  reefs  was  either  dredged,  tonged,  diver- 
harvested,  or  left  unharvested  as  a  control  (four  replicates 
of  each  treatment).  Experimental  dredging  and  hand-tong- 
ing were  conducted  in  the  manner  applied  by  commercial 
oyster  fishermen.  The  dredge,  25  kg  in  weight  and  1  m  in 
width,  was  pulled  behind  a  powerboat  operated  by  NCDMF 
personnel  with  commercial  oyster-dredging  experience. 
Hand-tonging  was  also  conducted  by  a  professional  oyster 


Marshall,  M.  1999.  Personal  commun.  North  Carolina 
Division  of  Marine  Fisheries,  3431  Arendell  St.,  Morehead 
City,  NC  28557. 


300 


Fishery  Bulletin  102(2) 


fisherman,  R.  A.  Cummings.  Oysters  and  shell  material  col- 
lected by  dredges  and  tongs  were  separated  aboard  the  boat 
on  a  culling  board,  using  the  common  culling  techniques 
(i.e.,  breaking  apart  oysters  and  shell  with  hammers,  mal- 
lets, and  chisels).  As  mandated  by  law,  oyster  shell  and 
undersized  oysters  (<7  cm  in  height)  were  thrown  overboard 
above  the  reef  from  which  they  had  been  collected. 

Hand  collections  of  oysters  were  conducted  by  scuba 
divers  (J.  H.  Grabowski  and  H.  S.  Lenihan).  Unlike  profes- 
sional oyster  divers  in  Chesapeake  Bay  and  other  areas, 
who  rake  large  quantities  of  shell  and  attached  oysters 
into  baskets  that  are  pulled  aboard  ship  to  be  culled,  the 
divers  in  this  trial  adopted  a  different  method  designed  to 
preserve  reef  habitat.  Instead  of  collecting  shell  and  oys- 
ters indiscriminately,  the  divers  chose  only  those  oysters 
that  appeared  alive  and  of  market-size.  Selected  oysters 
were  hand  picked  from  the  reef  and  placed  in  heavy  plastic 
mesh  baskets  that,  when  full,  were  subsequently  pulled 
aboard  the  boat  with  haul  lines. 

To  standardize  fishing  effort,  each  of  the  12  harvested 
reefs  was  harvested  for  2  hours,  regardless  of  the  num- 
ber of  oysters  collected.  A  2-h  harvest  period  for  each 
28.3-38.5  m2  reef  was  considered  to  be  a  thorough  but 
not  excessive  level  of  harvesting  by  the  professional 
fishermen.  The  numbers  of  oysters  collected  in  the  final 
three  or  four  dredge  hauls  and  oyster  tongs  were  typically 
lower  (by  -10-20%)  than  the  preceding  dredge  hauls  and 
tongs.  This  reduction  in  the  catch  per  unit  of  effort  was 
great  enough  that  a  fisherman  foraging  optimally  would 
normally  cease  harvesting  at  that  time  and  move  on  to 
another  reef.  Similarly,  after  2  hours  of  diver-harvesting, 
most  of  the  clearly  visible  market-size  oysters  had  been 
harvested. 

Quantifying  reef  structure 

Measurements  of  oyster  reef  height  and  diameter  were 
conducted  on  all  16  experimental  reefs  both  before  and 
after  application  of  the  three  fishing  methods.  In  Febru- 
ary 1996,  the  preharvest  height  and  radius  of  each  oyster 
reef  were  measured  by  scuba  divers  using  a  custom-made 
"square  angle,"  consisting  of  two  pieces  (2  m  and  5  m 
long)  of  3-cm  wide  steel  angle-iron,  each  with  an  attached 
1-m  long  carpenter's  level.  Both  pieces  of  angle  iron  were 
marked  at  1-cm  intervals.  The  5-m  long  (cross)  piece  was 
attached  to  the  2-m  long  (upright)  piece  by  a  roller-joint. 
The  roller-joint  allowed  the  cross  piece  to  move  up  and 
down  the  upright  piece,  thus  providing  a  measure  of  reef 
height,  and  to  move  horizontally  in  relation  to  the  upright 
piece,  thus  providing  a  measurement  of  reef  radius.  The 
2-m  long  piece  also  had  a  0.75-m  long  piece  of  angle  iron 
attached  perpendicularly  near  its  bottom  so  that  it  would 
not  sink  into  the  seafloor  when  placed  upright. 

One  diver  held  the  2-m  long  angle  iron  perpendicular 
to  the  seafloor  at  the  edge  of  a  reef,  while  the  other  diver 
placed  the  5-m  long  angle  iron  parallel  to  the  seafloor,  so 
that  one  end  rested  on  the  highest  point  of  a  reef  and  the 
other  end  met  the  upright  angle  iron  at  the  reefs  edge. 
The  height  and  radius  of  the  reef  were  then  measured 
by  recording  the  height  at  which  the  cross  piece  met  the 


upright  piece,  and  the  distance  at  which  the  upright  piece 
met  the  cross-piece.  For  each  reef,  a  mean  diameter  was 
calculated  by  measuring  three  separate  radii  (oriented 
at  three  compass  bearings,  all  120°  apart),  multiplying 
the  radii  by  two  to  estimate  diameters,  and  then  averag- 
ing the  three  diameters.  This  averaging  procedure  was 
undertaken  because  the  reefs  were  not  perfectly  circular. 
Measurements  of  reef  height  and  radius  were  repeated  in 
March,  two-five  days  after  experimental  harvests  were 
completed. 

Sampling  oyster  populations 

We  sampled  live  and  dead  oysters  on  each  treatment  and 
control  reef  before  (late  February  1996)  and  immediately 
after  (late  March)  experimental  harvests  to  estimate 
the  proportion  of  oysters  incidentally  killed  but  not  har- 
vested by  each  harvesting  treatment.  Specifically,  oyster 
data  was  collected  within  30  hours  of  the  application  of 
the  harvest  treatment  on  each  replicate  reef.  Densities  of 
live  and  dead  oysters  were  quantified  by  divers  who  hap- 
hazardly placed  eight  0.5-m'2  weighted  PVC  quadrats  on 
the  reef  surface  at  haphazard  locations  and  recorded  the 
number  of  live  and  dead  oysters  greater  >1  cm  in  height. 
The  density  of  dead  oysters  was  measured  by  count- 
ing the  number  of  oyster  shells  that  were  articulated 
and  appeared  relatively  fresh  (i.e.,  not  black  in  color  or 
decayed),  or  oysters  with  somatic  tissue  exposed  because 
of  cracked,  broken,  or  punctured  shells.  Oysters  with 
exposed  somatic  tissue  almost  certainly  die  because  of 
predation  by  fishes  and  crabs  in  the  Neuse  River  estuary 
(Lenihan,  1999;  and  see  Lenihan  and  Micheli,  2000). 
Mean  proportions  of  dead  oysters  were  computed  (dead 
oysters/dead+alive  oysters),  as  well  as  mean  densities  of 
live  and  dead  oysters  on  each  reef. 

Catch  per  unit  of  effort 

The  relative  efficiency  of  each  harvesting  method  was 
determined  by  comparing  the  numbers  of  bushels  (1 
bushel=36.4  L)  of  market-size  oysters  taken  per  hour  of 
fishing.  We  quantified  numbers  of  bushels  for  each  har- 
vesting method  aboard  the  boat  by  placing  oysters  of  legal 
size  in  premeasured  mesh  baskets.  After  being  counted, 
and  upon  termination  of  the  harvest  trial,  many  of  the 
oysters  were  returned  to  other  nearby  reefs  not  involved 
in  the  experiment. 

Statistics 

One-way  analysis  of  variance  (ANOVA)  was  used  to  com- 
pare the  following  across  harvest  treatments  and  controls: 
1)  changes  in  mean  reef  height  and  diameter;  2)  catch  per 
unit  of  effort;  3)  the  proportion  of  oysters  found  dead  on 
reefs  before  harvest;  4)  the  proportion  of  oysters  found 
dead  on  reefs  after  harvest;  and  5)  the  absolute  difference 
in  the  proportion  of  oysters  found  dead  before  versus  after 
harvesting  ([after  minus  before]).  Data  from  all  treat- 
ment (dredging,  tonging,  and  diver-harvesting;  n  =  4  for 
each  treatment)  and  the  control  Ui  =  4)  reefs  were  used  in 


Lenihan  et  al.:  Conserving  oyster  reef  habitat 


301 


Table  1 

Results  of  one-way  ANOVAs  comparing  differences  in  reef  height  (cm),  reef  diameter  (cm),  and  catch  per  unit  of  effort  (number 
of  oysters  collected  per  hour)  among  experimental  reefs  harvested  by  different  methods  (dredging,  tonging,  diver-harvesting, 
and  controls),  df  =  degrees  of  freedom;  ms  =  mean  square;  F  =  F-value;  P  =  P-value;  ss  =  sum  of  squares.  Partial  r2=  treatment 
ss/total  ss. 


Reef  height 


Reef  diameter 


partial 


Source 


df     ms 


partial 

7'2 


Catch  per  unit  of  effort 

partial 


Harvesting  treatment 

3     0.09     36.90     0.0001    0.90 

0.07     15.79 

Residual 

12    0.003 

0.005 

Total 

15                            Total  ss:  0.31 

0.0002      0.80 


0.27 


3.21 
0.08 


17.84    0.0001     0.11 


9.64 


the  ANOVA.  Before  ANOVA,  homogeneity  of  variances 
was  tested  by  using  Cochran's  method  (a=0.05).  All  data 
passed  this  test.  After  ANOVA,  post  hoc  differences  among 
means  were  compared  by  using  Student-Newman-Keuls 
(SNK)  tests  (a=0.05). 


Results 

Reef  height  and  diameter 

Dredge  harvesting  on  experimental  reefs  removed  the 
largest  amount  of  shell  material  from  the  reefs,  based  on 
the  reduction  of  reef  height  (Fig.  1A)  and  on  the  qualitative 
assessment  of  increases  in  numbers  of  oyster  shells  found 
on  the  seafloor  around  the  reefs.  Hand-tonging  removed 
an  intermediate  amount  of  reef  materials,  and  diver-har- 
vesting removed  far  less  shell  matrix  than  either  dredging 
or  tonging.  All  harvesting  methods  reduced  the  height  of 
restored  oyster  reefs  (Fig.  1A),  but  dredging  (34%  of  reef 
height)  and  tonging  (23%)  had  greater  impacts  than  did 
diver-harvesting  (6%).  ANOVA  demonstrated  significant 
differences  among  harvest  treatments  in  mean  change  in 
reef  height  (Table  1);  all  harvest  treatments  induced  a  loss 
in  reef  height  as  compared  with  unharvested  control  reefs 
(SNK;  P<0.05).  Dredging  reduced  reef  height  more  than 
any  other  treatment  (SNK,  P<0.05),  and  tonging  reduced 
reef  height  more  than  diver-harvesting  (SNK,  P<0.05). 
The  reduction  in  reef  height  caused  by  diver-harvesting 
was  small  (mean  ±SD:  6  ±3  cm).  However,  diver-har- 
vesting nearly  eliminated  the  veneer  of  live  market-size 
oysters  on  reefs,  which  provides  substantial  structure  on 
reef  surfaces. 

Oyster  harvesting  either  slightly  increased  or  slightly 
decreased  reef  diameter,  depending  upon  method  ( Fig. 
IB).  Reef  material  was  apparently  removed  from  edges 
of  reefs  by  tonging.  thereby  reducing  reef  diameter.  Shell 
was  spread  around  the  reefs  by  dredging,  thereby  increas- 
ing reef  diameter  after  application  of  that  harvesting 
method.  The  effects  of  oyster  harvesting  on  reef  diameter 
proved  significant  (Table  1).  Tonging  significantly  re- 
duced reef  size  compared  with  controls  and  the  other  two 
harvesting  treatments  (SNK;  P<0.05),  whereas  dredging 


DC 


I  A 

A 

T 

B 

1 

C 

T 

D 

]    B 

A 

1 

B 
i      T      i 

C 

1             1 

1 

D 

control     diver-harvested     tonged  dredged 

Figure  1 

Modification  of  reef  size  and  structure  caused  by  various 
harvesting  techniques.  (A)  Mean  (+SE)  reduction  in  the 
height  of  experimentally  restored  oyster  reefs  caused 
by  three  types  of  oyster  harvesting:  hand-harvesting  by 
divers,  hand  tonging.  and  dredging.  Dredges  are  pulled 
behind  power  boats.  Reefs  were  located  in  the  Neuse  River 
estuary,  North  Carolina.  Letters  represent  results  of  SNK 
post  hoc  tests:  dredged>tonged>diver-harvested>control 
at  P<0.05.  (B)  Mean  (  +  SE)  change  in  the  diameter  of 
experimental  oyster  reefs  caused  by  different  oyster- 
harvesting  techniques.  Letters  represent  results  of  SNK 
post  hoc  rests:  dredged>diver-harvested>control>tonged 
atP<0.05. 


302 


Fishery  Bulletin  102(2) 


Table  2 

Results  of  one-way  ANOVAs  comparing  differences  in  the  proportion  of  oysters  found  dead  ("mortality"  i  on  reefs  before  and  after 
harvesting  by  different  methods  (dredging,  tonging,  diver-harvesting,  and  controls),  and  the  absolute  difference  (\after-before]) 
in  the  proportion  of  dead  oyster  found  before  versus  after  harvesting,  df  =  degrees  of  freedom;  ms  =  mean  square;  F  =  F-value; 
P  =  P-value;  ss  =  sum  of  squares.  Partial  r2  =  treatment  ss/total  ss. 


Before  mortality 


After  mortality 


partial 


Source 


df     ms 


partial 
r2 


Difference  in  mortality 


partial 


Harvesting  treatment 

3     0.001 

0.49         0.69    0.11 

0.02 

Residual 

12    0.001 

0.002 

Total 

15 

Total  ss:  0.01 

7.90      0.004 


0.58 


II  us 


0.01 
0.08 


7.56    0.004      0.57 


0.04 


increased  reef  diameter  compared  to  the  other  treat- 
ments (SNK;  P<0.05).  The  increase  in  diameter  of 
diver-harvested  reefs  was  also  greater  than  that  for 
controls  (SNK;  P<0.05).  The  substantial  increase  in 
shell  material  (with  oysters  of  all  sizes)  spread  out  on 
the  seafloor  on  dredged  reefs  indicates  that  the  collec- 
tion efficiency  of  dredges  is  less  than  100%. 

Catch  per  unit  of  effort 

Catch  per  unit  of  effort  of  oysters  included  the  time 
required  to  collect  oysters  from  the  reef  and  the  time 
needed  to  separate  (i.e.,  cull)  them  from  undersized 
oysters  and  shell  material.  Two  of  the  harvesting 
methods,  hand-tonging  and  oyster  dredging,  are  one- 
man  operations  in  which  one  fisherman  can  operate 
the  harvesting  gear,  cull  oysters,  and  drive  the  boat. 
Therefore,  measurements  of  catch  per  unit  of  effort  for 
dredging  and  tonging  represent  the  numbers  of  bush- 
els of  oysters  one  fisherman  can  collect  per  hour.  In 
contrast,  scuba  diving  is  rarely  attempted  alone  and  it 
is  usually  necessary  for  someone  else  to  tend  the  diver 
(e.g.,  helping  him  or  her  in  and  out  of  the  water)  and 
to  haul  oysters  up  to  the  boat  when  given  a  signal  by 
the  diver  on  the  reef.  Divers  should  preferably  work  as  a 
team  using  the  "buddy"  system  for  safety  reasons.  Data  for 
diver-collections  are  given  in  bushels  per  hour  collected  by 
one  diver  but  hauled  up  to  the  boat  and  culled  by  a  second 
person. 

There  was  a  significant  difference  in  the  numbers  of 
bushels  collected  per  hour  by  the  different  harvesting 
techniques  (Table  1).  Diver-harvesting  had  a  higher  catch 
efficiency  than  all  other  treatments  ( SNK;  P<0.05;  Fig.  2). 
Diver-harvesting  was  about  25%  more  time  efficient  than 
dredge  harvesting  and  32%  more  efficient  than  tonging. 
There  was  no  statistically  significant  difference  in  effi- 
ciency between  dredging  and  tonging  (SNK;  P>0.05). 

Incidental  oyster  mortality 

The  proportion  of  oysters  found  dead  on  experimental 
reefs  in  February  1996  (-20%),  prior  to  experimental 


3.0  -I 

1  bushel  =  36.4  L 

-       2.5  - 
o 

1      2.0- 

A 

T 

B 

harvested 

B 

T 

I 

I      '■»■ 

sz 

CO 

™       0.5- 

C 

control    diver-harvested    tonged         dredged 

Figure  2 

Mean  (+SE)  number  of  bushels  collected  per  hour  on  experi- 

mental reefs  by  different  oyster-harvesting  techniques. 

Letters  represent  results  of  SNK  post  hoc  tests:  diver- 

harvested>dredged  and  tonged>control  at  P<0.05. 

harvesting,  was  similar  to  that  found  on  other  nearby  ex- 
perimental and  natural  reefs  in  the  Neuse  River  estuary 
in  preceding  years  (e.g.,  Lenihan  and  Peterson.  1998: 
Lenihan  1999).  In  February,  the  proportions  of  dead  oys- 
ters did  not  differ  among  the  four  sets  of  reefs  destined 
to  be  experimentally  harvested  (Table  2.  Fig.  3A).  In 
contrast,  there  was  a  large  and  statistically  significant 
difference  in  the  proportions  of  dead  oysters  on  the  reefs 
after  harvesting  (Table  2,  Fig.  3A).  The  proportions  of 
dead  oysters  on  reefs  that  had  been  tonged  and  dredged 
were  significantly  greater  than  on  diver-harvested  and 
control  reefs  (SNK;  P<0.05). 

Before-after-control-impact  (BACI)  comparison  of  the 
change  in  proportions  of  dead  oysters  from  before  to  after 
harvesting  ( [after— before] ),  a  direct  estimate  of  incidental 
mortality  caused  by  harvesting  gear,  showed  a  similar 
pattern  to  mortality  inferred  from  in  situ  proportions  of 
dead  oysters  in  March  after  harvesting  (Table  2,  Fig.  3B). 


Lenihan  et  al.:  Conserving  oyster  reef  habitat 


303 


A  significant  treatment  effect  in  the  after  period 
(Table  2)  indicated  that  the  change  over  time  in 
proportion  of  dead  oysters  varied  among  harvest 
treatments.  Tonging  and  dredging  increased  the 
fraction  of  dead  among  in  situ  oysters  on  reefs 
(SNK;  P<0.05;  Fig.  3B),  but  diver-harvesting 
did  not.  Immediately  after  harvesting,  divers 
found  that  many  oysters  on  tonged  and  dredged 
reefs  had  been  broken  open,  severely  cracked,  or 
punctured. 


Discussion 

Our  comparisons  of  gear  revealed  relatively 
unambiguous  differences  in  their  harvesting 
efficiency  for  oyster  dredges,  tongs,  and  hands 
of  divers.  Dredging  and  tonging  had  similar  and 
statistically  indistinguishable  catch  efficiencies, 
which  seems  reasonable  given  that  both  tech- 
niques are  commonly  employed  in  the  same  loca- 
tions and  times  in  the  oyster  fishery.  Presumably, 
fishermen  choose  between  these  two  gears  on  the 
basis  of  personal  preference,  history,  and  skill,  as 
well  as  on  the  basis  of  water  depth,  bottom  type, 
and  other  factors  that  did  not  vary  in  our  study. 
Diver-harvesting  of  oysters  resulted  in  higher 
rates  of  harvest  per  hour,  but  this  enhancement 
in  catch  efficiency  required  the  presence  of  two 
people,  one  diver  beneath  the  surface  and  another 
person  on  deck  involved  in  hauling  baskets  of 
oysters  onto  the  deck  and  culling  out  market- 
able oysters.  Because  the  increase  in  efficiency 
was  only  25-32%,  this  enhancement  falls  short  of 
the  100%  required  to  compensate  each  fisherman 
to  the  same  degree  that  dredging  and  tonging  pro- 
vide. Nevertheless,  the  immediate  economics  of 
diver-harvesting  could  prove  competitive  or  even 
superior  if  the  single  deckhand  could  serve  two  or 
more  divers,  which  seems  likely  from  our  experi- 
ence with  the  workload  on  deck,  and  if  the  oysters 
taken  are  priced  more  favorably  because  of  larger 
size  or  less  damage,  which  seems  possible.  A 
complete  short-term  economic  comparison  would 
need  to  include  higher  costs  for  fuel  in  dredging 
and  costs  of  filling  air  tanks  for  diving,  as  well  as 
depreciation  of  gear. 

This  discussion  of  the  basic  efficiencies  and  eco- 
nomics of  the  methods  of  commercial  oyster  fish- 
ing is  based  upon  short-term  considerations  only. 
That  short-term  time  perspective  is  the  cause  of 
failures  to  achieve  sustainability  in  fisheries  quite 
generally  (Ludwig  et  al.,  1993;  Botsford  et  al.,  1997).  We 
show  that  adoption  of  hand-harvesting  by  divers  would 
result  in  substantially  less  fishery-induced  reduction  in 
reef  height  by  a  factor  of  four  to  six,  implying  greater 
preservation  of  the  habitat  and  thus  a  more  sustainable 
fishing  practice.  Our  data  on  the  changes  in  area  covered 
by  reefs  as  a  function  of  harvest  treatment  revealed  only 
small  differences  among  treatments.  The  height  of  a  reef 


40 


30 


20- 


10 


control    diver-   longed   dredged 
harveted 

Before 


control    diver-   tonged   dredged 
harveted 

After 


14  - 

B 

12  - 

10  - 
8  - 
6  - 

A  - 

B 

T 

I 

T 

2  - 

control    diver-harvested    tonged 


dredged 


Figure  3 

Mortality  of  oysters  caused  by  various  harvesting  techniques.  (A) 
Mean  (  +  SE)  %  dead  within  oyster  populations  on  experimental 
reefs  before  and  after  being  harvested  by  three  different  harvest- 
ing techniques:  dredging,  tonging,  and  diver-harvesting.  Control 
reefs  were  not  harvested.  Letters  represent  results  of  SNK  post  hoc 
tests:  dredged,  after>tonged,  after>all  other  treatments  at  P<0.05. 
There  was  no  difference  among  treatments  before  harvesting.  (B) 
Mean  (  +  SE)  absolute  difference  in  the  %  dead  oysters  on  experi- 
mental reefs  before  and  after  harvesting.  Difference  calculated  by: 
\9o  after-%  before].  Letters  represent  results  of  SNK  post  hoe  tests: 
dredged  and  tonged>diver-harvested  and  control  at  P<0.05. 


is  a  critical  variable  in  sustaining  the  reef  as  an  engine  of 
oyster  production  because  short  reefs  can  be  easily  covered 
by  sediment  (Lenihan,  1999),  can  be  abraded  by  sediment 
transport  (Lenihan,  1999),  and  can  fail  to  extend  above 
hypoxic  bottom  waters  (Lenihan  and  Peterson,  1998). 
Tall  reefs  (i.e.,  reefs  not  degraded  by  harvesting)  produce 
faster  flow  speeds  and  more  turbulence  for  oyster  popula- 
tions, which  in  turn  increase  oyster  growth  rate,  increase 


304 


Fishery  Bulletin  102(2) 


physicalogical  condition,  reduce  disease  incidence  and 
intensity,  and  decrease  mortality  (Lenihan,  1999).  Con- 
sequently, assessment  of  economics  of  the  oyster  fishery 
over  longer  time  frames  would  likely  demonstrate  higher 
returns  from  practicing  diver-harvesting,  assuming  that 
this  technique  conserved  reef  structure.  Diver-harvesting 
also  killed  fewer  of  the  oysters  that  remained  on  the  bot- 
tom, thereby  sustaining  future  harvests  better  through 
reduced  wastage  and  by  retention  of  more  live  oysters  that 
would  produce  more  reef  material. 

Although  the  relative  advantage  of  diver-harvesting  for 
conserving  reef  structure  is  evident,  the  absolute  conser- 
vation of  reef  habitat  under  the  various  oyster  harvesting 
methods  is  not  clear  from  our  study.  Our  data  on  impacts 
of  diver-harvesting  revealed  slight  declines  in  reef  height, 
but  whether  these  same  declines  would  apply  to  an  older 
reef,  as  opposed  to  a  recently  restored  reef,  is  open  to  ques- 
tion. The  level  of  cementation  that  binds  the  shells  of  the 
reef  is  not  as  great  on  recently  restored  reefs,  making  them 
more  susceptible  to  degradation  with  physical  disturbance. 
Our  study  measured  only  the  immediate  drop  in  reef  eleva- 
tion after  fishing  at  a  level  that  removed  a  large  fraction 
of  legally  marketable  oysters.  In  a  well-managed  fishery, 
this  drop  in  reef  elevation  would  represent  virtually  an  en- 
tire season's  decline,  after  which  substantial  reef  growth 
would  occur  through  recruitment  and  growth  of  smaller 
oysters  before  a  new  harvesting  season.  Thus,  a  healthy 
oyster  reef  may  well  be  able  to  compensate  for  the  modest 
reduction  in  elevation  caused  by  diver-harvesting.  If  so, 
oyster  reef  sanctuaries  now  being  created  throughout  the 
Chesapeake  Bay  (Luckenbach  et  al.,  1999)  could  conceiv- 
ably be  opened  to  diver-harvesting  (without  implements) 
and  still  preserve  the  reef  services  to  the  ecosystem.  This 
possibility  deserves  to  be  evaluated  in  order  to  minimize 
conflicts  between  the  goals  of  restoring  oyster  reef  habitat 
for  conservation  purposes  and  restoring  oyster  reefs  for 
the  restoration  of  lost  fisheries. 

Application  of  the  results  of  our  gear  comparisons  to 
management  of  oyster  fisheries  will  likely  encounter 
some  impediments.  Although  various  artisanal  fisheries 
worldwide  have  employed  free  diving  as  a  fishing  tech- 
nique and  some  modern  fisheries,  including  the  American 
oyster  fishery,  involve  the  use  of  scuba,  diving  is  not  a  skill 
possessed  by  most  oyster  fishermen  and  probably  is  not  a 
method  under  consideration  for  oyster  fishing  in  general. 
In  addition,  the  peak  of  oyster  harvesting  season  on  the 
Atlantic  and  Gulf  coasts  is  usually  during  winter  months 
(e.g.,  November-March)  when  water  temperatures  in 
estuaries  are  quite  low  (0-10°C).  Such  conditions  require 
cold-water  diving  equipment  (e.g.,  dry-suits),  which  will 
further  increase  the  cost  of  this  new  harvesting  tech- 
nique. Thus  acceptance  of  diver-harvesting  by  the  indus- 
try would  require  training  in  diving  skills  and  safety, 
education  and  demonstration  of  the  advantages  of  this 
gear,  and  perhaps  even  investment  of  public  funds  to  de- 
fray costs  of  the  transition  from  traditional  dredges  and 
tongs  to  scuba  or  hookah.  Because  the  gains  of  switching 
to  diver-harvesting  accrue  to  the  industry  over  the  long 
term,  while  individual  fishermen  who  switch  may  suffer 
economically  in  the  short-term,  gear  choice  represents  a 


modified  example  of  the  tragedy  of  the  commons  (Ludwig 
et  al.,  1993).  Only  when  armed  with  some  form  of  owner- 
ship rights  and  an  attendant  long-term  perspective  would 
an  individual  oyster  fisherman  choose  to  switch  to  diver- 
harvesting.  The  precipitous  declines  of  over  99r<  in  oyster 
landings  in  mid-Atlantic  estuaries  (Rothschild  et  al.,  1994; 
Lenihan  and  Peterson,  1998)  mean  that  oyster  fishermen 
can  hardly  be  expected  to  bear  the  costs  of  switching  fish- 
ing methods.  Therefore,  government  intervention  would 
be  required  to  convert  subtidal  oyster  dredge  and  tong 
fisheries  into  diver-harvesting  operations  for  two  reasons; 
the  need  for  compensation  of  start-up  costs  and  the  need  to 
overcome  the  tragedy  of  the  commons.  Given  the  dire  state 
of  oyster  fisheries  today  ( Rothschild  et  al.,  1994 ),  the  habi- 
tat destruction  in  these  declines  (deAlteris,  1988;  Hargis 
and  Haven,  1998;  Rothschild  et  al.,  1994;  Lenihan  and 
Peterson,  1998).  the  broad  ecosystem  services  provided 
by  healthy  oyster  reefs  (Jackson  et  al.,  2001;  Lenihan  et 
al.  2001),  and  the  very  real  potential  for  restoring  oysters 
and  their  reefs  (Luckenbach  et  al.,  1999:  Lenihan,  1999). 
a  mandate  to  switch  fishing  methods  for  subtidal  oyster 
fisheries  could  pay  large  dividends. 


Acknowledgments 

We  thank  Mike  Marshall,  Jeff  French,  and  those  many 
NCDMF  people  working  on  deck  for  initially  creating 
experimental  reefs  to  our  specifications  and  for  later 
applying  the  experimental  dredge  harvesting  treatment. 
We  thank  Robert  A.  Cummings  for  applying  the  hand- 
tonging  treatment,  and  Jonathan  H.  Grabowski  for  help- 
ing with  diver-harvesting  of  reefs.  This  work  was  funded 
by  the  North  Carolina  General  Assembly  through  the 
Cooperative  Institute  of  Fisheries  Oceanography  (to  C.  H. 
Peterson),  and  NOAA-Chesapeake  Bay  Program  Oyster 
Disease  Program  (to  H.  S.  Lenihan,  C.  H.  Peterson,  and 
F.  Micheli) 


Literature  cited 

Bahr.  L.  M..  and  W.  P.  Lanier. 

1981.     The  ecology  of  intertidal  oyster  reefs  of  the  South 
Atlantic  coast:  a  community  profile.     U.S.  Fish  Wildl. 
Serv.  Office  of  Biological  Services  FWS/OBS-81/15. 
Botsford,  L.  W.,  J.  C.  Castilla,  and  C.  H.  Peterson. 

1997.     The  management  of  fisheries  and  marine  ecosys- 
tems.    Science  277:509-515. 
Collie,  J.  S.,  S.  J.  Hall,  M.  J.  Kaiser,  and  I.  R.  Poiner. 

2000.     A  quantitative  analysis  of  fishing  impacts  on  shelf- 
sea  benthos.    J.  Anim.  Ecol.  69:785-798. 
Dayton.  P.  K..  S.  F.  Thrush.  M.  T.  Agardy.  and  R.  J.  Hoffman. 
1995.     Environmental  effects  of  marine  fishing.     Aquat. 
Conserv.  Mar.  Fresh.  Ecosyst.  5:205-232. 
deAlteris,  J.  T. 

1988.     The  geomorphologic  development  of  Wreck  Shoal,  a 
subtidal  oyster  reef  of  the  James  River,  Virginia.     Estu- 
aries 11:240-249. 
Engel,  J.,  and  R.  G.  Kvitek. 

1995.     Effects  of  otter  trawling  on  a  benthic  community  in 


Lenihan  et  al.:  Conserving  oyster  reef  habitat 


305 


Monterey  Bay  National  Marine  Sanctuary.     Conserv.  Biol. 
12:1204-1214. 
Hargis,  W.  J.,  Jr.,  and  D.  S.  Haven. 

1988.     The  imperiled  oyster  industry  of  Virginia:  a  critical 
analysis  with  recommendations  for  restoration.     Special 
report  290  in  applied  marine  science  and  ocean  engin- 
eering, 145  p.     Virginia  Sea  Grant  Marine  Advisory 
Services,  Virginia  Institute  of  Marine  Science,  Gloucester 
Point,  VA. 
Jackson.  J.  B.  C,  M.  X.  Kirby,  W.  H.  Berger,  K.  A.  Bjorndal. 
L.  W.  Botsford,  B.  J.  Bourque,  R.  H.  Bradbury.  R-  Cooke, 
J.  Erlandson,  J.  A.  Estes,  T.  R  Hughes,  S.  Kidwell, 
C.B.  Lange,  H.  S.  Lenihan,  J.  M.  Pandolphi,  C.  H.  Peterson, 
R.  S.  Steneck,  M.  J.  Tegner,  and  R.  R.  Warner. 

2001.     Historical  overfishing  and  the  recent  collapse  of 
coastal  ecosystems.     Science  293:629-638. 
Jennings,  S.,  and  M.  J.  Kaiser. 

1998.  The  effects  of  fishing  on  marine  ecosystems.  Adv. 
Mar.  Biol.  34:201-352. 

Lenihan,  H.  S. 

1999.  Physical-biological  coupling  on  oyster  reefs:  how  hab- 
itat structure  influences  individual  performance.  Ecol. 
Monogr.  69:251-276. 

Lenihan.  H.  S.,  and  F.  Micheli. 

2000.  Biological  effects  of  shellfish  harvesting  on  oyster 
reefs:  resolving  a  fishery  conflict  by  ecological  experi- 
mentation.    Fish.  Bull.  98:86-95. 

Lenihan,  H.  S.,  and  C.  H.  Peterson. 

1998.  How  habitat  degradation  through  fishery  distur- 
bance enhances  impacts  of  hypoxia  on  oyster  reefs.  Ecol. 
Appl.  8:128-140. 

Lenihan,  H.  S.,  C.  H.  Peterson,  J.  E.  Byers,  J.  H.  Grabowski, 
G.  W.  Thayer,  and  D.  R.  Colby. 

2001.  Cascading  of  habitat  degradation:  oyster  reefs  in- 
vaded by  refugee  fishes  escaping  stress.  Ecol.  Appl. 
11:764-782. 

Luckenbach,  M.  A.,  R.  Mann,  and  J.  A.  Wesson  (eds.). 

1999.  Oyster  reef  restoration:  a  synopsis  and  synthesis  of 
approaches,  366  p.  Virginia  Institute  of  Marine  Sciences 
Press,  Gloucester  Point,  VA. 

Ludwig,  D.,  R.  Hilborn,  and  C.  Walters. 

1993.     Uncertainty,  resource  exploitation,  and  conservation: 
lessons  from  history.     Science  260:17-18. 
Newell,  R.  I.  E. 

1988.  Ecological  changes  in  the  Chesapeake  Bay:  Are  they 
the  result  of  overharvesting  the  American  oyster,  Cras- 
sostrea  virginica?  In  Understanding  the  estuary:  advances 
in  Chesapeake  Bay  research  (M.  P.  Lynch,  E.  C.  Krome, 


eds.),  p.  536-546.     Chesapeake  Bay  Research  Consor- 
tium, Publ  129  CBP/TRS  24/88,  Gloucester  Point,  VA. 
Paerl,  H.  W.,  J.  L.  Pinckney,  J.  M.  Fear,  and  B.  L.  Peierls. 

1998.     Ecosystem  responses  to  internal  and  watershed 
organic  matter  loading:  consequences  for  hypoxia  in  the 
eutrophying  Neuse  River  Estuary,  North  Carolina,  USA. 
Mar.  Ecol.  Prog.  Ser.  166:17-25. 
Peterson,  C.  H.,  H.  C.  Summerson,  and  S.  R.  Fegley. 

1983.     The  relative  efficiency  of  tow  clam  rakes  and  their 
contrasting  impacts  on  seagrass  biomass.     Fish.  Bull. 
81:429-434. 
1987.     Ecological  consequences  of  mechanical  harvesting  of 
clams.     Fish.  Bull.  85:281-298. 
Peterson,  C.  H.,  H.  C.  Summerson,  E.  Thomson,  H.  S.  Lenihan. 
J.  H.  Grabowski,  L.  Manning,  F.  Micheli,  and  G.  Johnson. 
2000.     Synthesis  of  linkages  between  benthic  and  fish 
communities  as  a  key  to  protecting  essential  fish  habitat. 
Bull.  Mar.  Sci.  66:759-774. 
Peterson,  C.  H.,  J.  H.  Grabowski,  and  S.  P.  Powers. 

In  press.     Quantitative  enhancement  of  fish  production  by 
oyster  reef  habitat:  restoration  valuation.     Mar.  Ecol. 
Prog.  Ser. 
Reise,  K. 

1982.     Long-term  changes  in  the  macrobenthic  invertebrate 
fauna  of  the  Wadden  Sea:  are  polychaetes  about  to  take 
over?    Neth.  J.  Sea  Res.  16:29-36. 
Rothschild,  B.  J.,  J.  S.  Ault,  P.  Goulletquer,  and  M.  Heral. 

1994.     Decline  of  the  Chesapeake  Bay  oyster  population: 
a  century  of  habitat  destruction  and  overfishing.     Mar. 
Ecol.  Prog.  Ser.  111:29-39. 
Safriel,  U.  N. 

1975.     The  role  of  vermetid  gastropods  in  the  formation  of 
Mediterranean  and  Atlantic  reefs.     Oecologia  20:85-101. 
Tamburri,  M.  N.,  R.  K.  Zimmer-Faust,  and  M.  L.  Tamplin. 

1992.     Natural  sources  and  properties  of  chemical  induc- 
ers mediating  settlement  of  oyster  larvae:  a  re-examin- 
ation.    Biol.  Bull.  183:327-337. 
Watling  L.,  and  E.  A.  Norse. 

1998.     Disturbance  of  the  seabed  by  mobile  fishing  gear:  a 
comparison  to  forest  clearcutting.     Cons.  Biol.  12:1180- 
1197. 
Wells,  H.  W. 

1961.     The  fauna  of  oyster  reefs  with  special  reference  to 
the  salinity  factor.     Ecol.  Monogr.  31:239-266. 
Wilson,  W.  H„  Jr. 

1979.  Community  structure  and  species  diversity  of  the 
sediment  reefs  constructed  by  Petaloproctus  socialis  (Poly- 
chaeta:  Maldanidae).     J.  Mar.  Res.  37:623-641. 


306 


Abstract— Loligo  opalescens  live  less 
than  a  year  and  die  after  a  short 
spawning  period  before  all  oocytes  are 
expended.  Potential  fecundity  iEP), 
the  standing  stock  of  all  oocytes  just 
before  the  onset  of  spawning,  increased 
with  dorsal  mantle  length  (L),  where 
EP  =  29. 8L.  For  the  average  female 
squid  (L  of  129  mm),  EP  was  3844 
oocytes.  During  the  spawning  period, 
no  oogonia  were  produced:  therefore 
the  standing  stock  of  oocytes  declined 
as  they  were  ovulated.  This  decline  in 
oocytes  was  correlated  with  a  decline 
in  mantle  condition  and  an  increase 
in  the  size  of  the  smallest  oocyte  in 
the  ovary.  Close  agreement  between 
the  decline  in  estimated  body  weight 
and  standing  stock  of  oocytes  during 
the  spawning  period  indicated  that 
maturation  and  spawning  of  eggs  could 
largely,  if  not  entirely,  be  supported 
by  the  conversion  of  energy  reserves 
in  tissue.  Loligo  opalescens,  newly 
recruited  to  the  spawning  population, 
ovulated  about  36^  of  their  potential 
fecundity  during  their  first  spawning 
day  and  fewer  ova  were  released  in 
subsequent  days.  Loligo  opalescens  do 
not  spawn  all  of  their  oocytes;  a  small 
percentage  of  the  spawning  population 
may  live  long  enough  to  spawn  78%  of 
their  potential  fecundity. 

Loligo  opalescens  are  taken  in  a 
spawning  grounds  fishery  off  Califor- 
nia, where  nearly  all  of  the  catch  are 
mature  spawning  adults.  Thirty-three 
percent  of  the  potential  fecundity  of 
L.  opalescens  was  deposited  before  they 
were  taken  by  the  fishery  (December 
1998-99).  This  observation  led  to  the 
development  of  a  management  strategy 
based  on  monitoring  the  escapement 
of  eggs  from  the  fishery.  The  strategy 
requires  estimation  of  the  fecundity 
realized  by  the  average  squid  in  the 
population  which  is  a  function  of  egg 
deposition  and  mortality  rates.  A  model 
indicated  that  the  daily  total  mortality 
rate  on  the  spawning  ground  may  be 
about  0.45  and  that  the  average  adult 
may  live  only  1.67  days  after  spawning 
begins.  The  rate  at  which  eggs  escape 
the  fishery  was  modeled  and  the  sen- 
sitivity of  changing  daily  rates  of  fish- 
ing mortality,  natural  mortality,  and 
egg  deposition  was  examined.  A  rapid 
method  for  monitoring  the  fecundity  of 
the  L.  opalescens  catch  was  developed. 


Manuscript  approved  for  publication 
19  December  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 
Fish.  Bull.  102:306-327  (2004). 


Fecundity,  egg  deposition,  and  mortality  of 
market  squid  (Loligo  opalescens) 

Beverly  J.  Macewicz 

1.  Roe  Hunter 

Nancy  C.  H.  Lo 

Erin  L.  LaCasella 

Southwest  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

8604  La  Jolla  Shores  Drive 

La  Jolla,  California  92037-1508 

E-mail  address  (for  B  J.  Macewicz):  Bev  Macewiczi5noaa.gov 


Many  loliginid  squid  populations 
depend  entirely  upon  the  reproduc- 
tive output  of  the  preceding  genera- 
tion because  individuals  live  less  than 
a  year  (Yang  et  al.,  1986;  Hatfield, 
1991,  2000;  Natsukari  and  Komine, 
1992;  Arkhipin,  1993;  Arkhipin  and 
Nekludova,  1993;  Jackson,  1993,  1994; 
Jackson  et  al.,  1993;  Boyle  et  al.,  1995; 
Jackson  and  Yeatman,  1996;  Jackson 
et  al.,  1997;  Moltschaniwskyj  and  Sem- 
mens,  2000;  Semmens  and  Moltschani- 
wskyj, 2000).  In  California  waters, 
Loligo  opalescens  (market  squid,  also 
known  as  the  opalescent  inshore  squid 
[FAO])  live  only  6-12  months  (Butler 
et  al.,  1999)  and  die  after  spawning 
(McGowan,  1954;  Fields,  1965).  Thus, 
fecundity  of  L.  opalescens  is  a  criti- 
cal life  history  trait  and,  in  addition, 
must  be  known  in  order  to  estimate  the 
biomass  with  either  egg  deposition  or 
larval  production  methods  (Hunter  and 
Lo,  1997).  Loligo  opalescens  is  one  of  the 
most  valuable  fishery  resources  in  Cali- 
fornia waters  and  is  monitored  under 
the  Coastal  Pelagics  Species  Fishery 
Management  Plan  of  the  Pacific  Fishery 
Management  Council  as  market  squid. 
Laptikhovsky  (2000)  pointed  out 
that  squid  fecundity  estimates  would 
be  biased  if  the  females  spawned  ova 
prior  to  capture,  if  oocytes  remained 
in  the  ovary  after  death,  or  if  some  of 
the  standing  stock  of  oocytes  were  lost 
because  of  atresia.  Previous  field  work 
on  squid  fecundity  has  been  limited  to 
the  traditional  method  of  simply  count- 
ing oocytes  or  ova  (or  both)  of  animals 


taken  on  the  spawning  grounds,  and 
none  of  the  biases  identified  by  Lap- 
tikhovsky (2000)  have  been  evaluated 
(Boyle  and  Ngoile,  1993;  Coelho  et  al., 
1994;  Guerra  and  Rocha,  1994;  Boyle  et 
al„  1995;  Collins  et  al.,  1995;  Moltscha- 
niskyj,  1995;  Lopes  et  al.,  1997;  Lap- 
tikhovsky, 2000).  On  the  other  hand, 
laboratory  studies  (Ikeda  et  al.,  1993; 
Bower  and  Sakurai,  1996;  Sauer  et  al., 
1999;  and  Maxwell  and  Hanlon,  2000) 
have  indicated  that  oocytes  remain  in 
the  ovaries  after  spawning  and  death. 
Additionally,  atresia  was  found  to  oc- 
cur in  all  stages  of  oocytes  of  Loligo 
vulgaris  reynaudii  (Melo  and  Sauer, 
1998).  Modern  approaches  to  estimat- 
ing lifetime  fecundity  in  fishes  take 
the  potential  biases  of  past  spawning 
history,  residual  fecundity,  and  atresia 
into  account  (Hay  et  al.,  1987;  Hunter 
et  al.,  1992;  Macewicz  and  Hunter. 
1994;  Kjesbu  et  al.,  1998).  The  initial 
objectives  of  the  present  study  were  to 
estimate  the  fecundity  of  L.  opalescens 
by  using  a  modern  approach  that  con- 
siders such  biases,  and  to  provide  a 
histological  description  of  those  aspects 
of  ovarian  structure  upon  which  mod- 
ern fecundity  analyses  are  based.  As 
our  work  progressed,  we  realized  that 
it  might  be  practical  to  manage  the 
market  squid  fishery  by  monitoring  egg 
escapement  based  on  fecundity  mea- 
surements. Thus,  we  added  two  new 
objectives:  to  conduct  a  preliminary 
evaluation  of  the  use  of  egg  escapement 
as  a  tool  for  management  of  the  market 
squid  fishery;  and  to  develop  a  method 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Loligo  opalscens 


307 


0   7^ 

Santa  Cruz  1. 

Female  Market  Squid, 
-v    Loligo  opalescens, 
V^^  Collection  sites 

34°N 

Santa  Rosa  1 

a\ 

Research  Cruises 

°  Jan.  1998  R/V  Jordan 
*  Dec.  1998  R/V  Mako 

Santa   ^-♦O 
Catalina  I. 

33° 

■^ 

121°W 


1 20° 


119c 


1 1 8: 


Figure  1 

Collection  locations  for  female  Loligo  opalescens  during  two  joint  research  cruises  during 
1998  by  California  Department  of  Fish  &  Game  (CDF&G)  and  National  Marine  Fisheries 
Service  (NMFS)  and  for  three  immature  females  (triangles)  collected  during  February 
2000  (CDF&G). 


to  monitor  the  fecundity  of  the  catch  that  avoids  the  costly 
process  of  counting  all  oocytes  and  ova. 

In  this  study  we  consider  four  aspects  of  the  fecundity 
of  L.  opalescens:  potential  fecundity,  minimum  residual 
fecundity,  maximum  fecundity,  and  the  fecundity  depos- 
ited by  the  average  female  in  the  population.  Potential 
fecundity,  or  potential  lifetime  fecundity,  is  the  standing 
stock  of  all  oocytes  in  the  ovary  just  before  the  onset  of  the 
first  ovulation.  Because  L.  opalescens  are  semelparous,  the 
standing  stock  of  all  oocytes  in  the  ovary  just  before  first 
ovulation  equals  their  potential  lifetime  fecundity.  Clearly, 
once  ovulation  and  spawning  (deposition  of  ova  in  egg  cap- 
sules on  the  sea  floor)  begin,  the  standing  stock  of  oocytes 
can  no  longer  be  considered  a  measure  of  the  potential 
fecundity  of  the  female.  Minimum  residual  fecundity  is 
the  minimum  number  of  oocytes  that  might  be  expected 
to  remain  in  the  ovary  at  death.  Because  ovaries  of  dying 
L.  opalescens  contain  oocytes  (Knipe  and  Beeman,  1978), 
only  a  portion  of  the  potential  fecundity  will  be  spawned 
in  their  lifetime.  We  use  ancillary  information  on  L.  opal- 
escens (an  index  of  mantle  condition  and  extent  of  ovarian 
development)  to  project  what  the  minimum  residual  may 
be.  Maximum  fecundity  (potential  fecundity  less  the  mini- 
mum residual  fecundity)  is  the  maximum  number  of  eggs 
a  female  might  be  expected  to  deposit  in  a  lifetime.  We  also 
estimate  the  fraction  of  the  potential  fecundity  deposited 
by  the  average  female,  a  key  vital  rate  we  approximate 
by  modeling  the  daily  rates  of  total  mortality  and  egg 
deposition.  Lastly,  the  term  "standing  stock  of  oocytes"  is 
used  throughout  this  article  to  indicate  the  total  number 
of  oocytes  at  all  stages  in  an  ovary.  Whether  the  standing 
stock  of  a  particular  female  is  to  be  considered  a  potential 


fecundity,  a  residual  fecundity,  or  something  in  between, 
depends  upon  ancillary  information  (i.e.,  presence  of  post- 
ovulatory  follicles  in  the  ovary,  ova  in  the  oviduct,  mantle 
condition,  or  the  level  of  ovarian  maturity). 


Materials  and  methods 

We  collected  Loligo  opalescens  during  two  southern  Cali- 
fornia research  cruises  in  1998  (7-15  January  and  3-10 
December)  (Fig.  1).  Most  specimens  were  taken  at  night 
by  using  trawls,  jigging,  or  by  removing  them  from  com- 
mercial purse-seine  catches  at  sea;  some  specimens  were 
collected  during  the  day  by  using  bottom  trawls.  We  mea- 
sured dorsal  mantle  length  (mm),  weighed  the  whole  body 
(g),  and  classified  the  ovary  and  preserved  it  with  viscera 
and  oviduct  attached  in  10%  neutral  buffered  formalin. 
To  determine  reproductive  state  we  decided  not  to  use  the 
familiar  ovary  classification  systems  but  rather  tabulated 
gross  anatomical  characters  and,  later  on,  selected  the 
most  useful  characteristics.  See  Table  1  for  characters 
selected  for  scoring. 

Preserved  ovaries  and  oviducts  were  reclassified  in  the 
laboratory  and  weighed  (to  nearest  0.001  g).  A  piece  of  the 
preserved  ovary  from  each  of  the  135  female  L.  opalescens 
from  January  and  the  117  females  from  December  was 
sectioned  and  stained  (hematoxylin  and  eosin).  Analyses 
of  the  histological  sections  included  identification  of  the 
oocytes  in  the  various  development  stages  (I-VI)  as  de- 
scribed by  Knipe  and  Beeman  (1978),  and  identification  of 
atresia  and  postovulatory  follicles  (Fig.  2 ).  We  use  the  term 
"ova"  to  indicate  an  ovulated  mature  oocyte  (stage  VI). 


308 


Fishery  Bulletin  102(2) 


Figure  2 

Slide  of  the  ovary  (stained  with  hematoxylin  and  eosin)  of  a  mature  spawning 
female  L.  opaleseens.  Bar  =  1.0  mm. 


Table  1 

Classification  system  for  the  gross  anatomical  characteristics  of  the  reproductive  system  of  female  market  squid  (Loligo 
opaleseens). 


Female  organs 


Character 


Grade 


Nidamental  gland 
Accessory  nidamental  gland 

Oviduct 

Ovary 

Ovary 


length 
color 

number  of  large  clear  eggs 
number  of  large  clear  oocytes 
number  of  opaque  or  white  oocytes 


millimeters 

0=clear,  l=whitish,  2  =  pink.  3=peach. 

4 =reddish-orange 

l=none.  2  =  1-20.  3=21-200,  4=>200 
l=none,  2  =  1-20,  3=21-200.  4=>200 
l=none,  2  =  1-20,  3=21-200,  4=>200 


Postovulatory  follicles  were  classified  as  either  new,  degen- 
erating, or  very  degenerative.  We  assigned  the  females  to 
one  of  the  following  reproductive  categories  on  the  basis  of 
the  histology  of  their  ovaries  (numerical  stages  in  Knipe 
and  Beeman,  1978): 

Immature  Ovary  contains  only  unyolked  oocytes; 
oocyte  development  ranged  from  stages  I 
(oogonia)  to  IV  (follicular  invagination  oocyte) 
and  requires  microscopic  examination. 

Mature  No  postovulatory  follicles  are  present.  Ova- 

preovulatory  ries  contain  oocytes  with  yolk  (stage  V, 
yolking  begins  about  1.1  mm  in  size);  ovary 
usually  contains  unyolked  oocytes. 


Mature  Ovary  contains  postovulatory  follicles 

spawning  ( POFs)  of  any  degree  of  degeneration  (none 
to  extensive);  more  than  one  degenerative 
POF  class  may  be  present.  Oocyte  develop- 
ment stages  III— VI  are  often  present  but 
stages  Ic-II  are  rare.  (29t  of  the  ovaries 
have  late  stage  Ic  oocytes  and  none  have 
any  of  the  earliest  stages,  la  or  lb.) 

In  some  histological  sections  of  ovaries  we  saw  1-10  yolking 
oocytes  (development  stage  V)  with  a  broken  follicle  layer, 
and  the  yolk  seemed  to  be  oozing  out  between  the  other 
oocytes.  Because  this  may  have  been  an  artifact  of  hand- 
ling, we  did  not  use  such  females  to  estimate  fecundity. 


Macewicz  et  at:  Fecundity,  egg  deposition,  and  mortality  of  Lo/igo  opalscens 


309 


We  used  the  gravimetric  method  (Hunter  et  al.,  1985, 
1992)  to  estimate  the  standing  stock  of  oocytes  in  98 
L.  opalescens  ovaries.  The  gravimetric  method  overes- 
timated the  total  number  of  oocytes  of  Loligo  pealeii, 
but  the  difference  between  a  count  of  all  oocytes  and  a 
weight-based  estimate  was  slight  (Maxwell  and  Hanlon, 
2000).  We  did  not  compare  our  estimates  with  a  count  of 
all  oocytes  in  the  ovary  because  we  used  a  portion  of  the 
ovary  for  our  histological  examinations,  and  each  value 
is  the  mean  of  the  counts  from  two  tissue  samples  (aver- 
age coefficient  of  variation  between  samples  was  0.12). 
All  oocytes  in  each  tissue  sample  were  macroscopically 
classified  (Fig.  3)  as  either  unyolked,  yolked,  mature,  or 
atretic;  they  were  then  counted  by  class  and  all  stages 
were  summed.  "Atretic"  was  defined  as  oocytes  in  the 
alpha  stage  of  atresia  (Hunter  and  Macewicz,  1985b), 
recognizing,  however,  that  poor  preservation  can  create 
oocytes  of  similar  macroscopic  appearance.  The  number  of 
ova  in  the  oviduct  was  also  counted  directly  (usually  when 
n  was  less  than  300)  or  the  mean  number  was  estimated 
from  two  tissue  samples  by  using  the  gravimetric  method. 
To  illustrate  the  form  of  the  oocyte-size  distribution  in 
the  ovary,  we  measured  (to  0.01  mm)  the  major  axis  of 
all  the  oocytes  in  one  tissue  sample  from  each  ovary  of 
six  females  by  using  a  digitizer  linked  by  a  video  camera 
to  a  dissection  microscope.  In  all  other  ovaries  used  for 
fecundity  estimation,  we  measured  only  the  smallest  and 
largest  oocyte  in  the  sample.  The  length  of  the  major 
axis  of  the  smallest  oocyte  (D)  was  used  as  an  index  of 
the  extent  of  ovarian  maturity.  D  is  a  crude  index  of  time 
elapsed  during  the  spawning  period  —  as  long  as  oocyte 
maturation  continues  throughout  the  spawning  period 
and  no  new  oocytes  are  produced  —  both  of  which  appear 
to  be  true  for  L.  opalescens. 

To  monitor  body  condition  we  cut  a  tissue  sample  disc 
from  the  mantle  using  a  number  11  cork  borer  (area  of 
251.65  mm2)  and  removed  the  outer  dermis  and  the  in- 
ner membrane.  The  mantle  sample  discs  were  frozen  and 
subsequently  dried  at  56°C  to  a  constant  weight.  An  index 
of  mantle  condition  (C)  was  calculated  as  the  weight  of  the 
dry  mantle  in  milligrams  divided  by  disc  surface  area  and 
is  expressed  as  mg/mm-. 

We  evaluated  the  extent  that  body  reserves  might  be 
used  to  support  egg  production  by  comparing  dry  weight 
of  the  eggs  and  capsules  to  prespawning  female  body  dry 
weight.  For  these  calculations  we  made  the  following  mea- 
surements: 1)  the  mean  dry  weight  of  one  squid  egg  was 
0.00177  g,  including  a  fraction  of  the  egg  capsule  because 
the  value  is  based  on  the  dry  weight  of  34  egg  capsules 
(1-2  days  old)  containing  2  to  403  eggs  each  (total  of  7341 
eggs,  capsules  collected  from  La  Jolla  Canyon  6  July  and 
11  September  2000);  2)  the  relationship  of  dorsal  mantle 
length  (L)  and  whole-body  wet  weight  (Ww)  for  immature 
and  mature  preovulatory  females  of  Wu.  =  0.000051L2  8086, 
where  Ww  is  in  grams  and  L  is  in  mm  (Fig.  4);  and  3)  the 
mean  wet  weight  to  dry  weight  conversion  factor  of  0.24 
(2SE  =  0.001),  based  on  the  wet  and  dry  weights  of  mantle 
tissue  sampled  from  214  mature  females.  The  latter  con- 
version factor  was  constant  regardless  of  mantle  condition 
index;  apparently,  in  L.  opalescens,  starvation  does  not 


unyolked 

j?m 

M 

.early 
yolking 

~  *>  KBHI 

* 

S 

-^  smallest 

^  oocyteM 

(unyolked) 

new-mieM 
postovulator^M 

Figure  3 

Whole  L.  opalescens  oocytes  as  viewed  under  a  dis- 
section microscope  used  for  counting  and  classifying 
oocytes.  Bar  =  1.0  mm. 


result  in  the  replacement  of  muscle  tissue  with  water  as  it 
does  in  fishes  (Woodhead,  1960). 

In  addition  to  the  specimens  taken  during  the  research 
surveys,  we  also  estimated  the  fecundity  of  60  L.  opal- 
escens from  the  commercial  catch  sampled  by  California 
Department  of  Fish  &  Game  (CDF&G)  during  1998  and 
1999.  Landed  specimens  were  not  analyzed  histologically 
because  their  ovarian  tissues  had  begun  to  deteriorate 
before  preservation.  The  60  females  were  selected  by  dor- 
sal mantle  length  and  mantle  condition  index  to  provide  a 
wide  and  uniform  distribution  of  length  and  mantle  condi- 
tion. The  number  of  oocytes  in  the  ovaries  was  estimated 
(as  described  above)  and  the  number  of  ova  in  the  oviducts 
were  predicted  from  oviduct  weight  (Fig.  5).  CDF&G  also 
provided  data  on  the  dry  mantle  disc  weights  of  1275  ma- 
ture females  taken  from  the  catch  from  December  1998 
through  December  1999  as  random  samples  taken  during 
the  Southern  Californian  Bight  market  squid  fishery. 
About  100,000  tons  of  market  squid  were  landed  during 
this  sampling  period  . 

Modeling  egg  deposition 

To  identify  egg  deposition  and  mortality  rates  most  consis- 
tent with  our  current  understanding  of  spawning  biology, 
we  developed  a  model  to  estimate  the  proportion  of  the 
potential  fecundity  deposited  by  a  cohort  in  its  lifetime. 
The  mean  proportion  of  the  potential  fecundity  deposited 
is  the  proportion  of  eggs  deposited  weighted  by  the  propor- 


310 


Fishery  Bulletin  102(2) 


tion  of  the  cohort  that  died.  Both  the  proportion  of  eggs 
deposited  and  squid  that  died  were  expressed  as  negative 
exponential  functions.  The  cumulative  eggs  deposited  up 
to  elapsed  time  t  (days I  for  a  mature  female  L.  opalescens 
is  the  difference  of  two  terms:  ESPl  =  EP-  EYDt  where  ESP/ 
is  the  total  eggs  deposited  by  one  female  up  to  time  t,  EP  is 
the  potential  fecundity,  and  EYDl  is  the  standing  stock  of 
oocytes  in  the  ovary  plus  the  standing  stock  of  ova  in  the 
oviduct  remaining  in  the  body  at  time  t.  If  we  assume  that 
EYDl  declines  at  an  exponential  rate  from  EP:  EYDl  =  EP 


e~ut,  where  v  is  the  daily  rate  of  eggs  deposited,  then  ESPt 
=  Ej,  ( l-e_1 0.  We  constructed  the  cumulative  egg  deposition 
curve  as  Qspt~  ESPl  IEP=  l-e~vt.  Assuming  the  mortality 
(survival)  curve  for  the  squid  is  e~zt ,  where  z  is  adult  daily 
total  mortality  rate  iz=m+f,  where  m  is  natural  and  f  is 
fishing  mortality),  we  computed  the  mean  fraction  of  the 
potential  fecundity  deposited  (QSPt): 


S 


80 
70 


60 


50 

40  - 
30 
20 


10 


W=  0.000051  Z_28086 

,2=0.964 
n  =  42 


50   60   70 


n 1 1 1 1 1 1 

100  110  120  130  140  150  160 


80      90 

Dorsal  mantle  length  (mm 


Figure  4 

Female  squid  whole  body  weight  ( W)  as  a  function  of  dorsal  mantle 
length  (L)  for  the  158  females  with  fecundity  analyses.  The  line 
expresses  the  length-weight  relation  of  females  before  weight 
losses  associated  with  spawning  and  was  fitted  to  the  combined 
data  for  immature  females  (solid  triangles),  mature  preovulatory 
females  (solid  circles),  and  mature  females  judged  by  their  mantle 
condition  to  be  new  recruits  to  the  spawning  ground  (solid  circles). 
Open  circles  indicate  females  that  have  spawned. 


2000 

o 

y  =  245x                                                   "  -x" 

-o     1500 

> 

o 

pseudo  r2=  0.98                                         ^S^  • 

c 

|     1000 

Number  of 

o 
o             o 

•    St** 

I       I        I       I        I       I        I       I        I        I        i        i        i        i 

0 

12              3              4              5               6        7 

Oviduct  weight  (g) 

Figure  5 

Number  oi 

ova  in  each  oviduct  shewn  as  a  function  of  the  oviduct 

weight;  n 

equals  91  mature  females,  pseudo  r-  =  1  -  residual 

ss/ total  ss. 

|  ze-zta-e-vt)dt 


Qsp 


dt 


(!) 


=  1- 


zil-e' 


) 


(z  +  y)(\-e 


)     z  +  v 


for  large  /n 


where  tmax  is  the  total  elapsed  time  (days). 

The  mean  fraction  of  the  potential  fecundity  that 
remains  in  the  average  female  (standing  stock  of 
oocytes  and  ova)  over  her  lifetime  is  1  -  Qsp,  and 
mean  QSP  is  always  less  than  one  because  of  mor- 
tality. The  mean  duration  of  the  spawning  period  in 
days  is  computed  as  the  elapsed  time  correspond- 
ing to  the  mean  fraction  of  eggs  deposited  (QSP:  Eq. 
1  and  by  setting  QSP=l-e-'''): 


^sf=ln(l-  QP )/(-«). 


(2) 


We  evaluated  various  rates  of  adult  daily  total 
mortality  (z)  and  egg  deposition  (r)  using  these 
models  to  determine  the  combination  of  rates  that 
would  provide  estimates  of  fecundity  nearest  to  our 
observed  field  data. 

Modeling  the  effect  of  fishing  effort  on 
egg  escapement 

In  theory  we  could  manage  the  market  squid  fish- 
ery by  monitoring  egg  escapement,  that  is,  the  frac- 
tion of  the  fecundity  realized  by  the  average  female. 
Under  such  a  management  scheme,  egg  escapement 
would  be  maintained  at  a  specified  level  by  chang- 
ing fishing  effort  whenever  escapement  of  eggs 
fell  below  it.  In  this  section  we  develop  a  model  to 
explore  the  relative  effects  of  fishing  effort  on  egg 
escapement.  We  use  this  model  to  discuss  some  of 
the  biological  issues  related  to  using  egg  escape- 
ment as  a  management  tool. 

In  the  modeling  process,  we  follow  one  cohort  of 
spawners.  The  elapsed  time  0  is  defined  as  the  time 
when  squid  start  spawning.  The  total  escapement 
of  eggs  for  a  given  elapsed  time  (tk  in  days)  is  the 
sum  of  three  sources  of  egg  escapement:  Ec,  the 
total  number  of  eggs  deposited  by  mature  females 
in  the  catch;  EM,  the  total  number  of  eggs  deposited 
by  mature  females  dying  of  natural  causes;  and 
EA,  the  total  number  of  eggs  deposited  by  females 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Lo/igo  opalscens 


311 


alive  and  not  taken  by  the  fishery  up  to  time  tk,  and  tk 
<  tmax.  The  egg  escapement  rate,  ReJk,  up  to  time  tk  is  the 
sum  of  the  three  sources  of  egg  escapement  divided  by  the 
total  number  of  eggs  that  would  have  been  spawned  if  no 
fishery  existed  (E): 


Ec  +  ESI  +  EA 


(3) 


Egg  escapement  rate  at  the  maximum  elapsed  time  Umax)  is 

Re,tn 


EC  +  EM 


(4) 


where  tk  =  tn 


Because  there  are  no  survivors  at  time  tmax,  no  eggs  can 
be  deposited  and  EA  is  zero. 

Each  term  in  Equation  3  can  be  expressed  as  functions 
of  the  mean  cumulative  number  of  eggs  deposited  up  to 
time  tk,  ESP  t.  =EP-  EYD  tk  =  EP  ( 1— e-"'* ),  and  total  mortal- 
ity (z)  of  the  cohort;  z  includes  both  natural  morality  (m) 
and  fishing  mortality  if).  For  practicality,  we  considered 
cases  when  tk  =  tmax,  where  EA  is  zero.  For  formulas  of  any 
tk,  see  appendix.  The  total  number  of  eggs  deposited  by  the 
females  in  the  catch  iEc)  is 


'max 

Ec=  j  E^tN0e- 


m+l  V 


fdt 


-EPN0  J  (l-e-l")e-""+'"fdt 
o 

E~N0f 


{m  +  f )( m+f  +  v) 


(5) 


(6) 


where  EP  -  the  mean  number  of  oocytes  in  the  ovary  per 
mature  female  prior  to  spawning;  and 
Nn  =  the  number  of  mature  females  at  time  0. 


EPN0m 


( m  +  f X m  +  f  +  v ) 


(9) 


The  total  eggs  that  would  be  deposited  for  the  cohort 
without  fishing  mortality  is 

'max  'max 

E=    \  E~^p~tN0me~mtdt  =  E~pN0  J  Tl-e"1''  )me-""dt    (10) 
o  o 


E  =  EPN0  ■ 


(11) 


m  +  v 


where  tmax  (days)  =  the  maximum  elapsed  time;  and 

time  0  =  the  time  at  the  onset  of  egg  deposition. 

Egg  escapement  based  on  Equation  4  is 


f 


■  +  m- 


R.. 


( m  +  /')(/??  +  f  +  v)         ( w  +  f)(m  +  f  +  v) 


(12) 


m  +  f  +  v 


Thus,  egg  escapement  reduces  down  to  a  simple  ratio, 

involving  three  daily  instantaneous  rates:  natural 

mortality  (m),  egg  deposition  (v),  and  fishing  mortality  (/"). 

R„,      =1  when  there  is  no  fishing  and  thus  i?„„<l  with 

(Vmax  °  '-■'I: 

fishing  mortality.  The  lower  bound  of  the  egg  escapement 
rate  for  the  cohort  is  equal  to  the  ratio  of  the  eggs  escap- 
ing the  fishery  (Er)  to  the  total  eggs  deposited  if  no  fishery 
existed  (£): 


R=ECIE. 


(13) 


From  the  fishery  data,  we  can  estimate  the  total  number 
of  eggs  deposited  by  the  females  in  the  catch  (Ec)  as 


Ec  =nc[ep-eyd)> 


(7) 


where  EP  and  EYD  =  sample  estimates  from  the  catch; 
and 
Nc  =  the  total  number  of  spawners  in  the 
catch. 

The  total  number  of  eggs  deposited  by  L.  opalescens  prior 
to  death  due  to  natural  mortality  iEM)  is 


EM=jESPlN0e-'"-'"mdt  = 


(8) 


EpN0J(l- 


•")e-"' 


mdt 


Results 

Oocyte  maturation  and  production 

Immature  ovaries  contain  many  small  unyolked  oocytes 
with  a  pronounced  peak  at  about  0.15  mm  in  size  distribu- 
tion ( Fig.  6A).  As  development  continues  and  vitellogenesis 
begins,  the  peak  diminishes  and  shifts  to  a  larger  size 
class  of  unyolked  oocytes  (Fig.  6B).  Just  before  the  onset 
of  spawning,  the  size  distribution  of  oocytes  becomes 
relatively  fiat  without  pronounced  modes  (Fig.  6C)  and 
remains  so  through  the  rest  of  the  spawning  period  ( Fig.  6, 
D-F).  The  standing  stock  of  oocytes  declines  throughout 
the  spawning  period.  The  minimum  size  of  oocytes  in 
the  ovary  gradually  increases  after  the  onset  of  yolking, 
indicating  that  new  oocytes  are  not  produced.  We  saw  no 
primary  oogonia  in  our  histological  sections  of  mature  ova- 
ries, another  indication  that  new  oocytes  are  not  produced 
in  mature  ovaries.  Knipe  and  Beeman  (1978)  reached  the 


312 


Fishery  Bulletin  102(2) 


A  No  yolk  in  oocytes 


86  mm   15  g 

5646  Total  oocytes 

0  ova 

0.719  =  C 


2.0       2.5       3.0 


300 


100 


B  Begun  yolking 


112  mm  39  g 

44 10  Total  oocytes 

0  ova 

0.711  =  C 


JUiiLl 


i  i  - 1 1 1 1 1 1 1 1 1  >  < 

2.5       3.0 


C  Before  1st  spawn  -  no  POF 


122  mm  39  g 

3724  Total  oocytes 

0  ova 

0.694  =  C 


0.5        1.0        1.5        2.0 


600 
500 
400 
300 
200 
100 
0 

600 

500 
400 
300 
200 
100 
0 


D  2  stages  of  POFs 


137  mm  50  g 

2988  Total  oocytes 

88  ova 

0.667  =  C 


..Illllllllllllll.ll.lh.lJ.. I..II. IllMl.... 


1.0       1.5       2.0       2.5       3.0 


E  3  stages  of  POFs 


133  mm  45  g 

1 642  Total  oocytes 

1 446  ova 

0.574  =  C 

■■lllllllllllH  I ■■■■•■   "iHl--!  ■ 


100 
0 


2.5        3.0  0  0.5 

Oocye  major  axis  diamter  (mm 


F  Diver  caught  -  Dying 

1 36  mm  33  g 

1 487  Total  oocytes 

0  ova 

0.544  =  C 


lillllllllilililn.iiiinl.lii.,liiiii..i 

1.0        1.5        2.0        2.5 


3.0 


Figure  6 

Oocyte-size  distribution  for  six  female  Loligo  opalescens.  Dorsal  mantle  length  (mm), 
body  weight  (g),  the  total  number  of  oocytes  in  the  ovary,  and  the  number  of  ova  are 
indicated  for  each  specimen.  (A)  Female  that  is  immature.  (B  and  C)  Females  that 
are  considered  to  be  mature  and  preovulatory  because  neither  has  postovulatory  fol- 
licles (POFs)  in  their  ovaries  nor  ova  in  their  oviducts.  Although  the  oocytes  have  just 
begun  yolking  in  the  ovary  of  female  B,  female  C  has  well-yolked  oocytes  and  is  close 
to  its  first  ovulation.  (D-F)  Females  are  mature  spawning  females  and  their  ovaries 
contained  postovulatory  follicles.  Female  F  was  caught  by  a  scuba  diver  and  appeared 
to  be  dying. 


same  conclusion  from  their  histological  analysis  of  L.  opal- 
escens ovaries.  Thus,  potential  fecundity  in  L.  opalescens 
probably  becomes  fixed  near  the  onset  of  spawning.  Not 
all  oocytes  are  deposited,  however,  because  all  spawning 
females  had  some  oocytes  and  many  oocytes  were  counted 
in  the  ovary  of  a  dying  female  (Fig.  6F). 

In  L.  opalescens  the  migration  of  the  oocyte  nucleus 
begins  early  in  the  maturation  process  shortly  before  the 
onset  of  vitellogenesis,  whereas  in  fishes,  migration  is 
near  the  end  of  vitellogenesis.  The  follicle  of  a  migratory- 
nucleus-stage  oocyte  (late  stage  IV)  of  L.  opalescens  has  a 
very  large  granulosa  cell  layer  (in  relation  to  the  size  of  the 
oocyte)  and  is  highly  folded  and  perhaps  fully  developed. 
Subsequent  maturation  of  the  oocyte  seems  to  consist 
primarily  of  the  massive  addition  of  yolk  and  fluid  and  the 
consequent  stretching  and  unfolding  of  the  follicle,  ending 


with  the  formation  of  a  chorion.  Apparently,  the  formation 
of  the  chorion  compacts  the  yolk  because  many  mature 
oocytes  (endpoint  of  stage  V)  have  a  smaller  major  axis 
than  advanced  yolked  oocytes  prior  to  chorion  formation. 
Thus,  maximum  oocyte  size  is  not  a  good  proxy  for  oocyte 
maturation  in  L.  opalescens  and  is  not  an  indicator  of  the 
time  remaining  before  spawning  of  the  next  batch.  More 
importantly,  the  ovary  of  L.  opalescens  seems  well  adapted 
for  rapid  oocyte  vitellogenesis,  maturation,  and  spawning 
because  nuclear  migration  and  follicle  cell  proliferation  is 
completed  at  an  early  stage. 

Ovulation  appears  to  occur  in  small  batches.  The  distri- 
bution of  oocyte  sizes  in  spawning  L.  opalescens  was  flat 
(e.g.,  Fig.  6,  D-F)  and  lacked  the  separate  and  distinct  mode 
of  hydrated  oocytes  that  is  typical  in  fishes.  Batch  sizes  of 
mature  oocytes  ranged  from  5  to  246  and  averaged  50  (n  =72 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Loligo  opalscens 


313 


females).  The  maximum  number  of  mature  oocytes  (246) 
was  never  close  to  the  maximum  number  of  ova  ( 1726 )  in  the 
oviduct.  In  addition,  spawning  females  with  900  or  more  ova 
in  their  oviduct  had  in  every  case  three  or  more  distinctly 
different  stages  of  postovulatory  follicles  in  their  ovaries 
(Table  2).  Thus  the  oviduct  is  probably  filled  by  a  series  of 
ovulation  bouts  separated  by  enough  time  to  produce  dis- 
tinct age  classes  of  degenerating  follicles  in  the  ovary. 

Potential  fecundity  (EP) 

Potential  fecundity  (EP)  is  the  standing  stock  of  oocytes 
of  all  stages  in  the  ovary  of  a  mature  female  just  prior 
to  the  first  ovulation.  Finding  females  at  this  point  in 
their  reproductive  cycle  was  difficult  because  nearly  all 
specimens  had  already  ovulated.  The  ovaries  of  94%  of 
the  247  mature  females,  from  our  research  cruises,  con- 
tained postovulatory  follicles,  indicating  that  they  had 
recently  ovulated  and  would  not  be  suitable  for  estimating 
EP.  As  can  be  seen  in  Figure  7A,  spawning  females  had 
fewer  oocytes  in  their  ovaries  than  did  mature  preovula- 
tory females.  The  relation  between  fecundity  and  squid 
size  is  best  expressed  in  terms  of  dorsal  mantle  length 
(L)  because  L.  opalescens  lose  weight  during  spawning 
(Figs.  4,  7C).  The  data  from  thirteen  mature  preovulatory 
females  were  used  to  establish  the  relationship  between 
potential  fecundity  and  L: 


EP  =  85.62L  -  6715,         [r*  =  34.3%] 


where  L  =  dorsal  mantle  length  in  mm. 


(14) 


Because  the  constant  was  not  significant  (P=0.146)  and 
the  coefficient  was  (P=  0.036),  we  forced  the  regression 
through  zero  which  resulted  in  the  equation 


EP  =  29.8L. 


(15) 


Thus,  the  average  female  (129  mm)  according  to  Equation 
15  had  a  potential  fecundity  of  3844  oocytes  (SE  =  317). 

Clearly  it  would  be  preferable  if  the  sample  size  for  the 
estimate  of  potential  fecundity  were  larger  because  thirteen 
females  may  not  accurately  represent  theL.  opalescens  stock. 
Although  the  landed  catch  provides  an  unlimited  supply  of 
specimens,  histological  detection  of  postovulatory  follicles 
is  not  possible  because  of  deterioration  of  the  ovaries.  An 
alternative  approach  is  to  use  mantle  condition  of  mature 
females  from  the  catch  as  a  proxy  for  the  preovulatory  state. 
As  can  be  seen  in  Figure  7C,  the  mantle  condition  index  (C) 
of  mature  females  declines  as  oocyte  maturation  continues 
and  females  deposit  eggs.  The  mature  preovulatory  females 
(n=ll,  two  discs  were  lost)  had  a  mean  C  of  0.73  mg/mm2 
(SE  =  0.02).  We  believe  that  the  twenty-two  mature  females 
from  the  landed  catch  with  C&0.7  mg/mm2  had  not  begun 
to  deposit  eggs  (Table  3).  Because  many  of  them  had  ovu- 
lated, we  combined  our  estimates  of  the  standing  stock  of 
oocytes  (EY)  with  those  of  ova  (ED)  to  calculate  total  fecun- 
dity (EY  +  ED  =  EYD),  and  then  regressed  total  fecundity 
on  length.  Although  the  regression  was  not  significant,  the 
average  total  fecundity  of  3890  oocytes  (Table  3)  was  within 


Table  2 

Percentage  of  s 
the  number  of 
of  ages  (stages 
(POFs)intheii 

pawning  female  market  squid  classed  by 
eggs  in  their  oviducts  and  by  the  number 
of  degeneration)  of  postovulatory  follicles 
ovaries. 

Number  of 

Number 

Percentage  of  females 

eggs  in 
the  oviduct 

of 
females 

1  or  2  ages 
of  POFs 

s3  ages 
of  POFs 

0 

1 

100 

0 

1-300 

36 

22 

78 

301-600 

20 

35 

65 

601-900 

10 

20 

80 

901-1200 

7 

0 

100 

1201-1500 

2 

0 

100 

1501-1800 

2 

0 

100 

5%  of  the  potential  fecundity  of  4083  oocytes  computed  by 
substituting  the  mean  length  of  the  twenty-two  females 
(137  mm)  in  Equation  15.  The  close  agreement  between 
these  two  values  increases  our  confidence  that  the  potential 
fecundity  equation  is  accurate  despite  the  low  /;.  On  the 
other  hand,  this  rough  comparison  is  not  a  substitute 
for  increasing  the  sample  size  of  specimens  analyzed 
histologically,  because  females  from  the  catch  may  have 
spawned  some  of  their  ova  before  they  were  captured. 

Maximum  fecundity  (EP  —  ER) 

Few  if  any  L.  opalescens  live  to  realize  their  full  potential 
fecundity  (EP).  The  literature  on  L.  opalescens  indicates 
that  females  that  were  described  as  "spawned  out,"  dying, 
or  dead  had  oocytes  in  all  stages  of  development  except  the 
earliest  previtellogenic  stage  (Knipe  and  Beeman  1978). 
In  addition,  all  the  spawning  females  that  we  collected 
had  some  oocytes  in  their  ovaries.  Thus,  the  maximum 
fecundity  that  L.  opalescens  might  be  expected  to  realize 
is  the  potential  fecundity  less  an  estimate  of  the  number 
of  oocytes  that  might  be  left  in  the  ovary  at  death  (residual 
fecundity  [ER]).  To  estimate  residual  fecundity  we  exam- 
ined the  relationship  of  the  standing  stock  of  oocytes  in  the 
spawning  period  with  mantle  condition  index  (C),  size  of 
the  smallest  oocyte  (D),  and  dorsal  mantle  length  (L). 

The  standing  stock  of  oocytes  in  ovaries  of  mature  fe- 
males declines  rapidly  with  decreasing  mantle  condition, 
between  a  C  of  0.8  and  0.6  mg/mm2,  and  more  gradually 
over  lower  mantle  conditions  (Fig.  7C).  A  curvilinear  rela- 
tionship also  exists  between  oocyte  standing  stock  and  the 
size  of  the  smallest  oocyte  (Fig.  7B).  Thus  the  number  of 
past  spawnings  (decline  in  oocyte  standing  stock)  appears 
to  be  inversely  correlated  with  C  and  directly  correlated 
with  the  extent  of  ovarian  maturation  as  measured  by  D. 
To  quantify  how  the  standing  stock  of  oocytes  changes 
during  the  spawning  period  we  fitted  a  nonlinear  model  to 
the  fecundity  data  of  75  mature  spawning  females  (Fig.  7) 
from  our  research  cruises: 


314 


Fishery  Bulletin  102(2) 


12000  r 
10000 

8000    - 

6000 

4000   - 

2000 
0 


Hrr- 

0        60 


\f^H^. 


-H-++ 


80 


T" 


t 


~T 


100  120 

Dorsal  mantle  length  (mm) 


— I — 
140 


160 


2000 

_  B 

_  o 

0000 

_  o 
_  o 

8000 

- 

6000 

4000 

2000 

0 

I        I        I        I        I 

+ 
-H- 

I   I 

+ 

I    I    I 

0.0 


6000 

5000 

4000 

3000    - 

2000 

1000 


0.8 


0.2  0.4  0.6 

Major  axis  diameter  (mm)  ot  smallest  oocyte 


1.0 


0.3 


0.4 


0.5 


— r~ 

0.6 


-I 1- 

0.7 


"i r 

0.8 


— I r 

0.9 


1.0 


Mantle  condition  index  (mg/mm2) 


O  Immature  *  Mature  preovulatory    +  Mature  spawning    a  Diver  caught 


Figure  7 

The  number  of  oocytes  in  ovaries  of  98  Loligo  opalescens  as  a  function  of 
dorsal  mantle  length  (A)  and  the  diameter  of  the  major  axis  of  the  smallest 
oocyte  (Bl.  In  (C),  the  number  of  oocytes  is  plotted  as  a  function  of  mantle 
condition  index  (the  dry  weight  per  surface  area  of  a  mantle  tissue  disci  for 
87  mature  females  (4  discs  were  lost  and  the  7  immature  were  not  included  I. 
Line  in  A  expresses  potential  fecundity  as  a  function  of  length  (£^=29.8/,) 
for  the  13  mature  preovulatory  females  (solid  circles);  open  circles  represent 
immature  females;  plus  signs  represent  spawning  females;  and  the  triangle 
represents  a  dying  mature  female. 


ER  =  30283e'_1  24W  ■  B  19('-  0.024i  +  0.059LC) 

where  C  =  mantle  condition  index; 

D  =  size  of  the  smallest  oocyte;  and 
L  =  dorsal  mantle  length. 


(16)  Substituting  into  the  model  (Fig.  8)  the  maximum  ob- 

served D  (0.771  mm)  and  the  minimum  observed  C  (0.323 
mg/mm2)  from  our  research  survey  data  set,  we  estimated 
that  a  female  L.  opalescens  with  L  of  129  mm  may  have 
a  minimum  residual  fecundity  of  834  oocytes  (CV=0.12). 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Loligo  opa/scens 


315 


Table  3 

Mean  fecundity,  gonad  weight,  and  dorsal  mantle  length  fo 
ports  December  1998  to  December  1999. 

■  60  mature 

female  market  squid  (Loligo 

opalescens)  sampled  at  the 

Mantle  condition 
(mg/mm2) 

ndex 

( 

Fecundity 
mean  numberl 

Mean 

gonad  weight 

Cg) 

Dorsal 

mantle  length 

( mm ) 

Number 

of 
females 

Oocytes 

in  ovary 

(EY)  ' 

Ova  in 

oviduct 

(ED) 

Total 
(EYD) 

Class 

Mean 

Mean 

Range 

0.347- 

-0.499 

0.432 

1134 

231 

1365 

2.215 

132 

106- 

-146 

22 

0.500 

-0.699 

0.613 

2072 

522 

2594 

4.959 

125 

102- 

-154 

16 

0.700- 

-0.951 

0.824 

2589 

1301 

3890 

8.988 

137 

106- 

-160 

22 

0.347- 

-0.951 

0.624 

1917 

701 

2618 

5.397 

132 

102- 

-160 

60 

A  129-mmL.  opalescens  with  a  potential  fecundity  of  3844 
oocytes  would  have  a  maximum  fecundity  of  3010  eggs 
(3844-834  eggs)  or  about  78%  of  the  potential  fecundity. 
Very  few  females  would  be  expected  to  deposit  78%  of 
their  potential  because  this  maximum  is  based  on  extreme 
values  for  both  mantle  condition  index  and  ovarian  matu- 
ration. In  a  much  larger  set  of  mantle  samples  from  the 
catch  (Table  4),  only  1.5%  of  the  females  had  values  of 
C  less  than  0.35  mg/mm-.  Clearly  very  few  squid  live  to 
deposit  78%  of  their  potential  fecundity. 

Another  approach  is  to  count  the  number  of  oocytes 
remaining  in  the  ovaries  of  females  presumed,  from  their 
behavior  and  appearance,  to  be  dying.  Although  L.  opal- 
escens has  been  observed  to  be  dying  or  dead  on  the  bot- 
tom on  video  from  a  remotely  operated  vehicle  (Cossio1), 
capturing  such  females  was  not  attempted  at  the  time.  A 
female  L.  opalescens  (136  mm)  believed  to  be  dying  was 
opportunistically  collected  by  a  diver  6  July  2000  on  the 
La  Jolla  Canyon  spawning  grounds  (McGowan,  1954). 
There  were  no  ova  and  the  ovary  contained  1487  oocytes  — 
substantially  more  oocytes  than  our  estimate  of  the  mini- 
mum residual  fecundity.  In  fact,  the  female  had  deposited 
only  about  63%  of  her  potential  fecundity. 

Role  of  body  reserves 

We  used  weight  relationships  to  evaluate  the  extent  to 
which  body  reserves  might  be  used  to  support  the  repro- 
duction of  spawning  female  L.  opalescens.  In  these  crude 
energetic  calculations  we  did  not  include  metabolism, 
conversion  efficiencies,  or  caloric  values  of  tissues.  We 
used  the  average  dry  weight  of  squid  eggs,  length  to  body 
weight  conversion,  potential  fecundity  equation,  and 
the  conversion  factor  from  wet  to  dry  mantle  weight.  We 
assumed  preovulatory  mantle  condition  index  (C)  for  an 
average  mature  female  of  130  mm  was  0.798  mg/mm2, 
the  mean  for  values  (/!=41)  of  C  >  0.700  mg/mm2  in  the 


Table  4 

Distribution  of  mantle  condition  index  for  1275  mature 
female  L.  opalescens  sampled  from  the  landed  catch  from 
December  1998  to  December  1999. 

Mantle  condition 
index  ( mg/mm2 1 

Mature  females 

Number 

Percentage 

0.263-0.299 

4 

0.3 

0.300-0.349 

15 

1.2 

0.350-0.399 

29 

2.3 

0.400-0.449 

54 

4.2 

0.450-0.499 

91 

7.1 

0.500-0.549 

128 

10.0 

0.550-0.599 

207 

16.2 

0.600-0.649 

210 

16.5 

0.650-0.699 

216 

16.9 

0.700-0.749 

137 

10.7 

0.750-0.799 

94 

7.4 

0.800-0.849 

53 

4.2 

0.850-0.899 

18 

1.4 

0.900-0.949 

10 

0.8 

0.950-0.999 

6 

0.5 

1.000-1.043 

3 

0.2 

1  Cossio,  A.  2000.  Personal  commun.  Southwest  Fisheries 
Science  Center,  National  Marine  Fisheries  Service.  8604  La 
Jolla  Shores  Dr.,  La  Jolla,  CA  92037 


our  fecundity  data  set.  We  calculated  that  the  potential 
fecundity  of  a  130-mm  L.  opalescens  (i.e.,  3874  encapsu- 
lated eggs)  has  a  dry  weight  of  6.86  g  which  is  equivalent 
to  64.8%  of  the  whole-body  dry  weight  (10.58  g)  of  that 
female  just  before  spawning.  If  mantle  condition  is  reduced 
in  proportion  to  the  dry  weight  of  all  the  eggs,  our  hypo- 
thetical female  would  have  a  C  of  about  0.281mg/mm2 
(0.798x[(10.58-6.86)/10.58]).  This  end  point  (C=0.281, 
egg=0)  and  the  beginning  point  for  the  mature  preovula- 
tory female  (C=0.798,  eggs=3874)  create  a  hypothetical 


316 


Fishery  Bulletin  102(2) 


2000 


1500 


1000 


500 


L=  129mm  Er 

-    C=  0.323  mg/mm2 


30283e<'1  24D-6.19C-0.024L+0.059f.C) 


Maximum  Observed  D 

(0.771)  b 

i 

I 


.Yolking 


Begin 
Ovulation 


t — i — i — i — i — i — rn — I — I — I — I — I     I     I     I     I     I     I     I 

0         0.2      0.4      0.6      0.8       1.0       1.2       1.4       1.6       1.8      2.0 
Maior  axis  of  smallest  oocyte  (D)  (mm) 


2000  - 


1500 


1000  - 


500  - 


L  =  129mm 
D=  0.771mm 

Minimum  Observed  C 
(0.323) 


0.2 


0.3 


0.4 


T 


— I — 
0.8 


0.5         0.6         0.7         0.8 
Mantle  condition  index  (C)  (mg/mm: 


"l 1 

1.0 


Figure  8 

Changes  in  the  standing  stock  of  oocytes  predicted  by  Equation 
16  (equation  also  given  at  top  of  panel)  when  major  axis  of  small- 
est oocyte  (D)  is  varied  and  mantle  condition  index  (Cl  held  con- 
stant (upper  panel),  and  when  C  is  varied  and  D  held  constant 
llower  panel).  The  major  axis  size  of  oocyte  when  yoking  begins 
and  when  ovulation  begins  is  also  indicated,  as  are  the  maximum 
observed  D  and  minimum  observed  C.  Substitution  of  the  latter 
two  values  into  the  equation  yields  the  standing  stock  of  oocytes  of 
females  close  to  the  end  of  their  reproductive  activity  and  is  consid- 
ered to  be  a  minimum  estimate  of  residual  fecundity. 


line  that  expresses  oocyte  standing  stock  for  the  average 
mature  female  of  130  mm  as  a  function  of  mantle  condi- 
tion. In  addition  to  the  hypothetical  line,  we  plotted  the 
total  standing  stock  of  oocyte  and  ova  (EYD)  and  mantle 
condition  index  for  all  147  mature  females  used  for  direct 
fecundity  determinations  (Fig.  9).  Our  hypothetical  line, 
based  on  direct  proportionality  between  egg  dry  weight 
and  body  dry  weight,  follows  the  general  trend  in  the 
data,  indicating  that  energy  reserves  in  mantle  tissue 
may  largely  support  the  production  and  spawning  of 
eggs.  Of  course,  actual  energy  costs  would  be  higher 
because  metabolism,  other  somatic  tissue,  and  conversion 
efficiency  of  mantle  tissue  to  eggs  are  not  considered.  The 
lowest  observed  C  in  the  fecundity  data  set  was  0.323  and 
the  lowest  C  observed  in  the  1275  mature  females  from 
the  landed  catch  was  0.263.  Using  the  above  preovulatory 
C  (0.798  mg/mm2),  we  determined  that  these  values  of  C 


are  equivalent  to  60%  and  67%  losses  in  body  dry  weight 
for  these  individuals.  Fields  (1965)  suggested  body  wet 
weight  declined  by  as  much  as  50% ,  which  is  consistent 
with  our  results. 

These  rough  calculations  support  the  long  held  belief 
that  oocyte  maturation  is  supported  primarily  by  body 
reserves.  Some  feeding  occurs  during  spawning;  L.  opal- 
escens  has  been  observed  feeding  under  lights  at  night 
on  the  spawning  grounds  (Butler2).  Maxwell  and  Hanlon 
(2000)  observed  L.  pealeii  feeding  between  egg-laying 
bouts  when  they  were  held  in  the  laboratory.  Feeding  be- 
tween spawning  bouts  by  the  more  robust  spawners  that 
may  migrate  on  and  off  the  grounds  each  day  seems  quite 


2  Butler,  J.  2000.  Personal  commun.  Southwest  Fisheries 
Science  Center.  National  Marine  Fisheries  Service,  8604  La 
Jolla  Shores  Dr.,  La  Jolla,  CA  92037. 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Lohgo  opalscens 


317 


7000 

> 

o 

■n 

6000 

t 

CO 

W 

5000 

>N 

4000 

o 

o 

3000 

<1) 

F 

2000 

-j 

r 

"Hi 

1000 

o 

- 

0 

-70 

A 

A 
A 

°         *A 

4        ~°          o 

a     ^       o  ^  °^  <A 

^kK^A    °: 

A      A         AO<^§^^                         O 

o2*A^r^^AO 

A±£*      \                     O             * 

^^                      30                            0     %  loss 

l          l          i          i          i                    i  of  body  wt. 

I             I             I             I             I             I             I             i 

0.2 


0.4  0.6  OS 

Mantle  condition  index  (mg/mm2) 


1.0 


Figure  9 

Standing  stock  of  oocytes  and  ova  as  a  function  of  mantle  condition 
index  for  60  mature  females  taken  in  the  fishery  (triangles)  and  87 
mature  females  taken  in  research  surveys  (circles I.  The  line  (not 
drawn  to  the  data  points  plotted)  represents  a  possible  relation  if 
losses  in  body  weight  were  equivalent  to  the  weight  of  the  spawn 
released,  computed  for  an  average  female  (130  mm)  where  the 
starting  point  is  her  potential  fecundity  of  3874  oocytes  and  ends 
with  0  eggs  and  a  mantle  condition  index  of  0.281  mg/mm2. 


possible,  but  it  seems  unlikely  for  the  nearly  exhausted 
L.  opalescens  that  are  near  the  end  of  their  life. 

Longevity  and  egg  deposition  rates 

Inferences  regarding  the  longevity  of  adult  spawning  L. 
opalescens  are  the  best  proxy  we  have  for  the  mortality 
rates  of  spawning  adults.  Previous  observers  (McGowen, 
1954;  Fields,  1965)  suggested  that  females  deposited  all 
their  eggs  in  one  night  and  death  soon  followed.  On  the 
other  hand,  it  is  unreasonable  to  expect  that  a  reduction 
of  609  in  body  weight  and  the  maturation  and  deposition 
of  up  to  78%  of  the  potential  fecundity  could  take  place 
in  24  hours.  Our  data  on  fecundity  and  mantle  condition 
show  an  initial  rapid  decline  in  the  number  of  oocytes 
followed  by  a  more  gradual  decline  (Fig.  7C),  indicating 
an  initial  period  of  intense  egg  laying  may  be  followed 
by  a  longer  one  where  fewer  eggs  are  deposited.  It  is  also 
important  to  recognize  that  ovaries  of  spawning  animals 
contain  a  wide  range  of  oocyte  sizes  (Fig.  6),  includ- 
ing many  small  unyolked  oocytes  (0.3-1  mm)  that  may 
mature  and  be  deposited  during  the  spawning  period. 
It  is  unlikely  that  all  these  processes  (body  resorption, 
dynamic  changes  in  rates  of  egg  deposition,  and  matura- 
tion of  small  unyolked  oocytes)  could  occur  in  one  24- 
hour  period.  Spawning  periods  longer  than  two  weeks 
also  seem  unlikely  because  mature  L.  opalescens  females 
may  require  extensive  feeding  periods  and  prolonged  ab- 
sences from  the  spawning  grounds;  these  behaviors  are 
inconsistent  with  our  energetic  analysis  in  the  preceding 
section.  Our  analysis  indicated  that  the  observed  reduc- 
tion in  eggs  can  be  fairly  well  explained  by  the  observed 
reduction  in  squid  dry  weight. 


Egg  deposition  rates  provide  another  way  to  infer  the 
longevity  of  spawning  squid.  The  best  evidence  for  the 
rate  of  egg  deposition  is  provided  by  females  judged,  on  the 
basis  of  their  high  mantle  condition  ( C^O.700  mg/mm2),  to 
be  new  recruits  to  the  spawning  grounds.  Considering  only 
those  new  recruits  that  have  ovulated  (postovulatory  fol- 
licles present  or  ova  in  the  oviduct ),  the  difference  between 
their  average  oocyte  standing  stock  (£y=2571)  and  their 
average  potential  fecundity  (£P=4020)  was  equivalent  to 
a  reduction  of  1449  oocytes  or  36%  of  their  potential  fecun- 
dity (Table  5).  If  the  difference  is  spawned  in  24  hours  or 
less,  then  36%  can  be  considered  as  an  average  for  the  first 
day  of  egg  deposition.  Instead  of  using  the  reduction  of  oo- 
cyte standing  stock,  one  could  consider  the  standing  stock 
of  ova  (ED )  to  be  equivalent  to  the  first  day  ( 24-hour  period ) 
of  spawning  in  these  new  recruits.  Their  average  ED  was 
1073  or  27%  of  their  potential  fecundity.  Thus  depending 
on  the  criteria,  the  first  day  of  spawning  might  be  27%  to 
36%  of  the  potential  fecundity.  We  prefer  36%  because  it  is 
unaffected  by  any  losses  due  to  egg  deposition. 

The  standing  stock  of  ova  (ED)  of  spawning  females 
with  lower  mantle  condition  (C<0.7  mg/mm2)  averaged 
9%  of  their  potential  fecundity.  If  the  average  ED  from 
these  females  is  a  crude  measure  of  daily  egg  deposition 
rates  after  the  first  day,  then  we  calculate  it  would  take 
seven  additional  days  [U00%-36%)/9%]  to  use  up  the 
remaining  potential  fecundity  or  a  total  spawning  period 
of  eight  days.  Eight  days  is  an  extreme  value  because 
an  adult  L.  opalescens  has  never  been  taken  with  zero 
oocytes.  The  minimum  residual  fecundity  was  22%  of  the 
potential  which  is  roughly  equivalent  to  about  two  days 
of  egg  deposition.  Thus,  six  days  may  be  a  better  guess 
of  the  maximum  longevity  of  spawning  L.  opalescens. 


318 


Fishery  Bulletin  102(2) 


Probably  very  few  females  would  be  expected  to  survive 
six  days  because  only  a  small  percentage  of  the  spawning 
population  (Table  4)  met  the  mantle  criteria  for  minimum 
residual  fecundity. 

In  summary,  our  best  guess  of  the  maximum  longevity 
of  squid  on  the  spawning  grounds  is  about  six  days.  Our 
best  description  of  daily  egg  deposition  is  a  rate  that  ends 
the  first  day  with  36%  of  the  potential  fecundity  deposited 
and  averages  about  9%  of  the  potential  per  day  over  the 
remaining  five  days  and  where  only  a  small  percentage 
of  the  females  live  to  deposit  78%  or  more  of  their  poten- 
tial fecundity. 

Egg  escapement 

We  examine  the  spawning  dynamics  of  Loligo  opalescens 
from  the  standpoint  of  possibly  using  fecundity  of  the 
catch  to  monitor  and  ultimately  regulate  escapement  of 
eggs  from  the  fishery.  The  key  variable  in  this  approach 
is  the  fraction  of  the  potential  fecundity  that  is  actually 
deposited  as  eggs  on  the  bottom  because  this  value  can  be 
directly  estimated  from  the  fecundity  of  the  catch.  Two 
other  important  parameters  are  the  daily  rate  of  total 
mortality  (2)  on  the  spawning  grounds  and  the  daily  rate 
of  egg  deposition  (y).  Neither  of  these  parameters  can  be 
directly  estimated  but  they  are  approximated  by  values 
that  are  most  consistent  with  our  observations  by  using 
a  model  (Eq.  1).  Our  observations  consist  of  the  fecundity 
of  the  catch  and  the  inferences  regarding  longevity  and 
egg  deposition,  presented  in  the  previous  section.  We  use 
our  approximations  for  egg  deposition  and  total  mortality 
in  a  second  model  (Eq.  12)  to  gain  an  idea  of  how  natural 
mortality  and  fishing  mortality  may  affect  egg  escape- 
ment. Lastly,  we  present  a  rapid  method  for  monitoring 


the  fecundity  of  the  catch  which  does  not  require  direct 
counting  of  oocytes  or  ova. 

Fraction  of  the  potential  fecundity  spawned  (Q5P)  In  a 
spawning  population  of  L.  opalescens,  the  mean  standing 
stock  of  oocytes  and  ova  (EYD),  when  expressed  as  a  frac- 
tion of  potential  fecundity,  is  equivalent  to  the  fraction 
of  the  potential  fecundity  of  the  population  that  remains 
in  the  spawners  (EYDIEP).  When  subtracted  from  one  (1- 
[EyD/EP] ),  the  difference  becomes  the  fraction  of  the  poten- 
tial fecundity  of  the  population  that  is  actually  spawned 
(Qsp).  For  this  interpretation  to  be  correct,  samples  must 
be  randomly  drawn  from  the  population  and  represent  all 
spawners  according  to  their  abundance  on  the  spawning 
grounds — from  the  newly  recruited  to  those  that  have 
been  spawning  for  extended  periods. 

Neither  the  females  taken  from  our  research  cruises  nor 
those  used  to  estimate  fecundity  from  the  landed  catch 
were  random  samples  of  the  spawning  population.  First, 
not  all  of  the  specimens  taken  during  the  two  research 
cruises  were  from  the  spawning  grounds.  Second,  the  60 
females  from  the  commercial  catch  were  not  randomly 
chosen  but  were  selected  to  represent  a  full  range  of  L 
and  C.  However,  by  weighting  our  fecundity  estimates  by 
a  random  sample  of  mantle  condition  from  the  fishery,  it 
was  possible  to  approximate  a  random  fecundity  sample  of 
spawners.  The  population  we  used  for  weighting  was  based 
on  the  mantle  condition  index  (C)  of  1275  randomly  taken 
specimens  from  the  commercial  catch  sampled  December 
1998  through  December  1999  (Table  4).  The  weighted 
and  unweighted  mean  standing  stocks  of  oocytes  and  ova 
(EYD)  were  similar  (Table  5),  indicating  that  our  previous 
selection  of  specimens  by  C  did  not  introduce  a  large  bias. 
For  the  unweighted  data,  EYD  was  2541  and  was  2599 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Lo/igo  opa/scens 


319 


Table  6 

Estimates  of  number  of  days  of  egg  deposition,  the  mean  number  of  eggs  deposited,  mean  standing  stocks  of  oocytes  and  ova 
remaining  in  female  L.  opalescens,  mean  number  of  eggs  deposited  at  the  end  of  the  first  night  (all  means  are  expressed  as  a 
fraction  of  the  potential  fecundity),  for  various  combinations  of  possible  egg  deposition  (u)  and  total  adult  mortality  (z)  rates. 
Model  provided  estimate  nearest  observed  data  when  z  =  0.45,  v  =  0.25,  and  tmax  =  8  days. 


Daily 
total 

mortality 
Cz) 


Daily 

egg 

deposition 

rate 

(v) 


Fraction  of 
potential 
fecundity 
deposited 

Qsp 

(Equation  li 


Fraction  of 
potential 
fecundity 

remaining 
in  females 

a-Qsp) 


Mean  number 

Fraction  of 

of  nights 

eggs  deposited 

of  egg 

at  the  end 

Days 

deposition 

of  the 

to  reach 

tQsp 

first  night 

78%  eggs 

(Equation  2) 

Q-e-") 

deposited 

2.45 

0.221 

6.057 

2.13 

0.362 

3.365 

1.88 

0.478 

2.329 

1.67 

0.221 

6.057 

1.48 

0.362 

3.365 

1.33 

0.478 

2.329 

1.08 

0.221 

6.057 

0.99 

0.362 

3.365 

0.91 

0.478 

2.329 

0.36 

6.0 

0.2 

0.2 

0.2 

0.45 

0.45 

0.45 

0.8 

0.8 

0.8 

Observed 


0.25 
0.45 
0.65 
0.25 
0.45 
0.65 
0.25 
0.45 
0.65 


0.458 
0.617 
0.706 
0.341 
0.486 
0.580 
0.237 
0.359 
0.447 
0.326  (SE  0.075) 


0.542 
0.383 
0.294 
0.659 
0.514 
0.420 
0.763 
0.641 
0.553 
0.674 


when  the  data  were  weighted  by  the  distribution  of  mantle 
conditions  in  the  catch.  The  mean  fraction  of  the  poten- 
tial fecundity  deposited  (QSP)  by  L.  opalescens  was  0.326 
(1-2599/3859).  That  much  of  the  fecundity  had  escaped 
(eggs  were  deposited)  before  the  market  squid  were  taken 
by  the  fishery  does  not  seem  unreasonable  because  22- 
36%  of  EP  may  be  deposited  during  the  first  day  of  spawn- 
ing. The  mean  QSP  is  an  important  index  because  it  mea- 
sures egg  escapement  as  a  fraction  of  potential  fecundity 
over  its  lifetime  (Eq.  1).  It  is  used  in  subsequent  sections 
to  identify  a  daily  total  mortality  rate  and  egg  escapement 
rate  for  the  average  female  in  the  population  that  best 
characterizes  the  sampled  L.  opalescens  population. 

Preferred  mortality  and  egg  deposition  rates  We  used 
Equation  1  to  evaluate  which  combination  of  a  range  of 
plausible  values  for  the  rates  of  daily  total  mortality  (z  of 
0.2,  0.45,  and  0.8)  and  daily  egg  deposition  (v  of  0.25,  0.45, 
and  0.65)  provides  an  estimate  closest  to  observed  Qsp 
(ESPIEP)  (Table  6).  The  combination  of  an  adult  daily  total 
mortality  (z)  rate  of  0.45,  a  daily  egg  deposition  (v)  rate  of 
0.25,  and  using  a  £max  of  8  days  gave  an  estimate  that  was 
most  consistent  with  the  observed  value  for  QSP  of  0.326 
(Table  6,  Fig.  10).  This  combination  of  rates  also  gave  an 
egg  depletion  of  78%  of  the  potential  fecundity  in  6  days 
which  was  consistent  with  our  best  guess  for  maximum 
longevity  and  maximum  fecundity.  On  the  other  hand,  the 
model  (using  1-e-"'  and  t=l)  predicts  that  about  22%  of  the 
potential  is  deposited  by  the  end  of  the  first  24  hours  (day  1) 
which  is  less  than  our  preferred  estimate  (36%)  based  on 
the  reduction  in  standing  stock  of  oocytes  but  is  closer  to  the 
one  based  on  the  standing  stock  of  ova  ( 27%).  A  possible  bio- 


logical explanation  for  the  difference  might  be  that  some  of 
the  ova  produced  during  the  first  day  of  deposition  might 
remain  in  the  oviduct  and  then  be  deposited  on  the  second 
day.  Regardless  of  the  uncertainties  regarding  the  fit  for 
the  initial  day  of  egg  deposition,  a  daily  total  mortality  rate 
of  0.45  and  daily  egg  deposition  rate  of  0.25  are  most  con- 
sistent with  the  field  data  known  at  the  present  time.  This 
means  that  the  average  spawning  period  is  very  short;  the 
average  female  lives  only  1.67  days  after  spawning  begins 
(ln(0.659)/-0.25;  Eq.  2).  It  is  interesting  that  1.67  days  for 
the  average  animal  is  not  a  radical  departure  from  Fields's 
(1965)  original  conclusion  of  a  single  night  of  spawning. 

Egg  escapement  from  the  fishery  In  L.  opalescens,  where 
the  fishery  targets  spawning  adults  that  die  after  spawn- 
ing, it  is  important  to  know  the  effect  of  fishing  mortality 
on  the  egg  escapement  rate  with  respect  to  the  lifetime 
fecundity  deposited,  RL,  t  (Eq.  12).  However,  not  all  terms 
in  Equation  12  are  observable  and  it  may  be  practical 
to  manage  the  fishery  by  monitoring  the  fraction  of  the 
potential  fecundity  that  is  deposited  on  the  bottom  (QSP= 
1-[EYD/EP]).  Nevertheless,  we  examined  the  potential 
effects  of  fishing  mortality  if)  on  the  egg  escapement  rate, 
Bf(mii,  when  natural  mortality  (m)  is  0.1,  0.25,  or  0.4,  and 
egg  deposition  (v)  is  0.25,  0.45,  or  0.65  (Fig.  11).  Because 
our  preferred  rates  from  the  previous  section  are  v  =  0.25 
and  2  =  0.45,  then  m  is  <0.45  with  fishing  because  z  =  m 
+  f.  If  we  use  v  =  0.25  and  set  daily  natural  mortality  rate 
high  (»i=0.4),  then /"is  0.05  andi?f  Jum  is  93%.  Doubling  the 
fishing  mortality  (to  0.1)  produces  an  absolute  difference  of 
6%  in  egg  escapement  (Fig.  11C).  Thus  at  a  high  m  of  0.4, 
escapement  is  relatively  insensitive  to  changes  in  daily 


320 


Fishery  Bulletin  102(2) 


fishing  mortality.  At  lower  natural  mortalities,  a 
change  in  fishing  mortality  has  a  greater  effect  on 
escapement.  At  m  =0.1  and  /'=  0.35  i?e,,malI  is  50%. 
Doubling  the  fishing  morality  to  0.7  Retnm  would 
be  33%,  producing  a  loss  of  17%  in  escapement 
(Fig.  11A).  Increasing  the  rate  of  daily  egg  deposi- 
tion (v)  from  our  preferred  value  of  0.25  to  0.65  also 
diminishes  the  effect  of  fishing  mortality  on  escape- 
ment but  the  effect  of  fishing  on  egg  escapement  is 
most  marked  at  the  low  natural  mortality  of  m  = 
0.1  and  is  relatively  minor  when  natural  mortality 
reaches  /?;  =  0.4.  Thus,  uncertainties  regarding  the 
true  initial  values  of  egg  deposition  seem  relatively 
unimportant  at  these  high  mortality  rates.  It  is 
important  to  remember  that  in  this  discussion  that 
we  are  discussing  daily  mortality  rates  that  last 
only  a  few  days  or  weeks  of  the  life  of  a  semelparous 
animal;  hence  the  rates  are  very  high  and  resemble 
the  typical  daily  mortality  rates  of  small  pelagic  fish 
eggs  (Alheit,  1993)  that  also  exist  for  short  periods. 

Cost  effective  methods  for  monitoring  fecundity     If 
egg  escapement  were  adopted  as  a  monitoring  and 
management  tool  for  the  market  squid  fishery,  a 
cost-effective  method  for  monitoring  fecundity  of 
L.  opalescens  would  be  needed.  A  direct  estima- 
tion of  the  standing  stock  of  oocytes  in  an  ovary  by  using 
a  microscope  and  video  system  (as  preformed  in  this 
study)  is  too  time  consuming  for  routine  monitoring  of  the 
fishery  because  it  takes  about  4  hours  per  specimen. 

Our  first  approach  for  an  indirect  estimator  was  to 
use  the  measurements  routinely  taken  by  CDF&G  staff 
who  sample  the  catch.  These  measurements  were  dorsal 
mantle  length,  mantle  condition  index,  and  an  oviduct 
classification  system  for  approximating  the  numbers  of 
ova.  To  estimate  the  oocyte  standing  stock  (EY)  of  the 
catch  females,  using  only  length  and  mantle  condition,  we 
fitted  a  nonlinear  model  to  the  data  for  all  squid  classed 


1.0 
0.9 
0.8 

z=0.45 
~         v=0.25 

0.78  of  potential 

fecundity  deposited^^      — 

0.7 

^     up  to  6.06  days 

0.6 

1  -  e-"^-^"^ 

0.5 

0.4 
0.3 

'm  67  n  ^41  mean  duration  of  spawning, 
(  -o  .  u.j  )  mean  fracti0n  eggS  deposited 

0.2 

0.1 

-/^^ 

0 

I        I        I 

1       1       1       1       1       1       1       1       1       1       1       1       1 

Elapsed  time  (/,  days) 

Figure  10 

The  cumulative  egg  release  curve  (solid  line)  and  the  density  func- 
tion of  longevity  on  the  spawning  grounds  of  adult  females  I  dashed 
line)  of  Loligo  opalescens  for  a  total  mortality  rate  (z)  of  0.45  and 
a  egg  release  rate  (v)  of  0.25.  The  plotted  solid  circle  represents 
mean  egg  deposition  estimated  by  the  model  as  a  proportion  of  the 
potential  fecundity  and  model  estimate  of  the  mean  duration  of 
the  spawning  period. 


as  mature  (spawning  individuals  and  pre-ovulatory)  in  our 
1998  research  survey  data  set.  This  yielded  the  equation 

£v=  220.453ell-99C  +  0-0079il,  (17) 

where  L  =  dorsal  mantle  length;  and 
C  =  mantle  condition  index. 

Equation  17  for  EY  explains  only  33%  of  the  variability 
within  the  survey  data  set  (n=90)  and  therefore  is  rather 
imprecise.  Using  this  model  we  estimated  EY  to  average 
about  2221  oocytes  in  the  ovaries  of  the  mature  females 


l.U 

m=0.10 

0.9 

r\\s 

0.8 

0.7 

\                  v=0.65 

W             v=0.45  

v  "••-..  N.        v=0.25  

0.6 
0.5 
0.4 

V^\ 

\^      ""••••■""""~~ 

0.3 

0.2 

1 

I      I      I      I      I      I      I      I      I 

0.0 


0.2 


0.4 


0.6 


1.0 


Fishing  mortality  (I) 


Figure  11 

The  egg  escapement  rate,  R,  ,  ( Eq.  12)  of  L.  opalescens  as  a  function  of  various  daily  natural  adult  mortality  rates  (in ),  daily  egg 
deposition  rates  (rO,  and  daily  fishing  mortality  rates  if).  In  each  panel,  the  solid  circle  indicates  the  /'value  for  preferred  values: 
2  =  0.45  and  v  =  0.25. 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Loltgo  opalscens 


321 


(n.=1275)  sampled  from  catch  during  the  period 
1998-99  (Fig.  12);  this  estimate  is  equivalent 
to  approximately  58%  of  the  potential  fecundity 
calculated  from  mean  length  (129  mm). 

Our  second  approach  was  to  estimate  total 
fecundity  (EYD,  standing  stock  of  oocytes  and 
ova)  indirectly  using  the  combined  formalin  wet 
weight  of  the  ovary  and  the  oviduct,  in  addition  to 
mantle  condition.  Combining  ovary  and  oviduct  in 
one  weight  is  more  efficient  than  weighing  them 
separately  because  much  less  time  is  required  for 
dissection.  Dorsal  mantle  length  was  also  con- 
sidered as  a  variable  but  it  was  not  significant. 
The  final  equation  for  the  total  standing  stock  of 
oocytes  and  ova  in  a  mature  female  squid  is 


EYD  =  378.28e12  33C  +  °  2447G  -  °-24CG> 


(18) 


where  C  =  mantle  condition  index;  and 

G  =  gonad  (ovary  and  oviduct)  weight. 


5000  r 


4000  - 


o     3000  - 


2000 


1000 


EY  =  220.453e<1  99C  +  000791.) 
for  L=  129  mm 


Mature  Females 

Collected  Dec.  1998-Dec.1999 

n=  1275 

Mean  C  (0.625)  ±2SE 


1 — 1 — 1 — r 


0.2 


0.3 


0.4 


1 — 1 — r~ 
0.5         0.6 


~~ 1 — 1 — r 

0.7         0.E 


~\ T~ 

0.9 


1 1 

1.0 


Mantle  condition  index  (C)  mg/mm2 


The  predicted  fecundity  related  well  to  the  observed 
with  a  pseudo  r2  of  0.60  (df=143).  We  also  used 
generalized  additive  models  to  estimate  fecundity 
(GAM,  pseudo  r2=0.64),  as  well  as  regression  on 
the  first  principal  component  which  explained  867r  of  the 
total  variation  (pseudo  r2=0.55).  Although  the  GAM  gave 
a  slightly  higher  pseudo  r2  than  the  parametric  nonlinear 
regression,  we  chose  the  later  for  easier  interpretation  and 
implementation.  A  pattern  existed  in  the  residuals  from  our 
model  (Fig.  13);  the  model  overestimated  some  fecundities 
at  high  mantle  condition  and  underestimated  fecundity 
at  low  mantle  condition.  This  pattern  in  the  residuals  is 
probably  related  to  the  differences  in  density  and  size  of 
oocytes  in  the  ovary.  Regardless  of  the  minor  problem  with 
the  residuals,  this  proxy  (Eq.  18)  for  the  standing  stock  of 
oocyte  and  ova  is  preferred  because  it  gives  a  much  more 
precise  estimate  at  the  minor  additional  cost  of  preserv- 
ing and  subsequently  determining  the  combined  weight 
of  ovary  and  oviduct.  Although  formalin  weight  of  ovary 
and  oviduct  are  not  presently  monitored  in  the  fishery,  it 
is  a  variable  that  could  be  added  to  fishery  protocols  at  a 
minor  increase  in  cost.  Another  benefit  of  this  more  pre- 
cise approach  using  EYD  is  that  oviduct  is  included  in  the 
estimate.  If  an  estimate  of  the  removal  of  fecundity  by  the 
fishery  is  needed,  ova  must  be  included.  Because  ova  are  not 
included  in  Equation  17,  to  add  them  requires  using  the  ovi- 
duct classification  system  (Table  1)  to  estimate  the  average 
number  of  ova — a  system  that  is  imprecise  but  cheap.  One 
could,  of  course,  use  Equation  17  for  EY  and  either  count  the 
ova  in  the  oviduct  or  weigh  the  oviduct,  but  that  would  take 
more  work  than  applying  Equation  18  for  EYD. 


Discussion 

Potential  fecundity 

Our  estimate  of  Loligo  opalescens  potential  fecundity  is 
based  on  a  regression  of  the  standing  stock  of  oocytes  on 


Figure  12 

Standing  stock  of  oocytes  in  the  ovary  (£>•)  as  a  function  of  mantle 
condition  index  (C)  for  a  129-mm  mature  female  L.  opalescens  as 
predicted  by  Equation  17  (equation  also  given  on  top  of  panel;  L  is 
dorsal  mantle  length).  Dashed  lines  are  ±2SE.  The  mean  EY  for  the 
females  taken  in  the  fishery  was  2221  oocytes. 


dorsal  mantle  length  for  mature  preovulatory  females 
having  yolked  oocytes  in  their  ovaries.  The  accuracy  of 
this  approach  depends  upon  the  assumption  that  these 
females  are  at  the  point  in  life  when  the  standing  stock 
of  oocytes  in  their  ovaries  is  equivalent  to  their  potential 
lifetime  fecundity.  This  key  assumption  would  not  hold  if 
some  of  the  mature  squid  classed  histologically  as  pre- 
ovulatory had  in  fact  spawned.  We  do  not  know  how  long 
postovulatory  follicles  are  distinguishable  from  atretic 
structures  in  the  ovary  of  L.  opalescens  and,  as  far  as  we 
know,  the  rate  of  degeneration  has  not  been  determined  for 
any  loliginid.  We  know  from  our  work  on  anchovy  (Hunter 
and  Goldberg,  1980;  Hunter  and  Macewicz,  1985a), 
although  it  is  not  a  cephalopod,  that  postovulatory  follicles 
are  distinguishable  from  atretic  structures  in  the  ovary  of 
anchovy  for  about  two  to  three  days  after  spawning  when 
the  water  temperature  is  about  16°C.  This  means  that  for 
undetected  spawning  to  occur  in  L.  opalescens,  the  inter- 
val between  ovulation  periods  would  likely  need  to  exceed 
three  days.  This  may  be  a  minimum  estimate  because  L. 
opalescens  spawn  at  lower  temperatures  (9-13  C,  Butler2) 
than  do  anchovy.  Definitely  a  laboratory  study  on  the 
rate  of  degeneration  is  necessary  because  postovulatory 
follicles  in  fish  degenerate  slower  at  lower  temperatures 
(Fitzhugh  and  Hettler,  1995).  In  addition  to  the  absence 
of  postovulatory  follicles,  the  oviduct  must  be  empty  for  a 
spawning  act  to  be  undetected.  Undetected  ovulation  and 
spawning  seems  unlikely  because  females  with  multiple 
stages  of  postovulatory  follicles  were  common  (87%  of  247 
mature  females),  females  with  only  old  postovulatory  fol- 
licles were  not  detected,  and  the  average  life  span  on  the 
spawning  grounds  may  only  be  a  few  days. 

Atretic  losses  of  oocytes  are  another  possible  bias  in 
estimating  potential  fecundity.  Atresia  (degeneration  and 
resorption  of  an  oocyte  and  its  follicle)  appears  to  be  a 


322 


Fishery  Bulletin  102(2) 


3000 

2000 

1000 

0 

CD 

> 

o 

-1000 

+ 

(0 

-2000 

/ 


*.. 


•  •  T*  .     I*    ••  •  * 


0.3  0.4 


0.5 


— i 1 1 1 — 

0.6  0.7  0.8  0.9 


1.C 


Mantle  condition  index  (mg/mm2) 


3000 1- 
2000- 
1000- 


0 


-1000 


-2000 


•  •  •  V  *     •  •  .  . 


L*J" ^ — ^*  *a   wm  ^ M • •    ■ 


0 


0.0006  r 


>.  0.0004  - 


0.0002 


~1 r 

2 


1 1 1 r 

6  8 

Gonad  weight  (g) 


10 


- 1 1 1 

12  14 


0.0b 


-2000 


-1000 


1000 


2000 


3000 


Residuals 


Figure  13 

Residual  plots  of  number  of  oocytes  and  ova  from  the  equation 
EYD  =  378.28e(2  33C  +  0  2447°-°  24CGl  (Eq.  18)  where  standing  stock  of  oocytes  and 
ova  (EYD )  are  predicted  from  mantle  condition  index  ( C)  and  gonad  weight  (G ). 
Bottom  panel  shows  probability  density  of  residuals. 


normal  part  of  ovarian  maturation  in  L.  oplaescens,  as  it 
is  the  case  for  L.  v.  reynaudii  (Melo  and  Sauer,  1998).  Our 
evidence  for  this  is  that  the  standing  stock  of  oocytes  in 
immature  female  L.  opalescens  declines  sharply  as  their 
ovaries  mature  (D  increases,  Fig.  7B).  Clearly  a  narrow 
window  of  opportunity  exists  for  an  unbiased  estimate 
of  the  potential  fecundity  of  L.  oplaescens.  If  the  count 
is  made  too  early  in  the  ovarian  maturation  process,  the 
count  will  either  be  low  because  extensive  primary  oogo- 
nia  production  may  be  still  be  occurring  (64-mm  female, 
Fig.  7A)  or  too  high  because  additional  oocytes  will  be 
absorbed  before  the  female  reaches  maturity.  If  the  count 


is  made  too  late,  it  will  be  impossible  to  find  a  female  that 
has  not  ovulated.  Our  selection  criteria  "presence  of  yolked 
oocytes"  (which  roughly  begins  at  a  oocyte  size  of  about 
1.1  mm)  filtered  out  the  very  high  counts  of  oocytes  associ- 
ated with  immature  ovaries. 

From  the  practical  standpoint,  dealing  with  atretic 
losses  that  may  continue  into  the  spawning  period  is  much 
less  important  for  L.  opalescens  than  for  L.  v.  reynaudii 
(Melo  and  Sauer,  1998;  Sauer  et  al.,  1999)  or  L.  pealeii 
(Maxwell  and  Hanlon,  2000).  In  these  squid,  where  the 
spawning  period  may  last  weeks  or  months,  atresia  may 
seriously  bias  potential  fecundity  estimates.  In  the  pres- 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Loligo  opolscens 


323 


ent  study  all  atretic  losses  would  be  attributed,  of  course, 
to  ovulation  and  spawning  but  the  chances  of  this  being  a 
major  error  seem  low.  Because  we  counted  atretic  as  well 
as  normal  oocytes,  atretic  losses  would  be  erroneously 
attributed  to  spawning  only  if  atresia  had  proceeded  to 
the  point  that  the  atretic  structure  could  not  be  identified 
as  that  of  an  oocyte  in  whole-mount  preparations  under  a 
light  microscope  ( 64x  power).  The  time  at  stage  for  atretic 
oocytes  in  L.  opalescens  ovaries,  as  well  as  other  squid,  is 
unknown.  The  duration  of  alpha-stage  atresia  of  yolked 
oocytes  in  anchovy  is  about  a  week  at  16°C  (Hunter  and 
Macewicz,  1985b)  and  we  suspect  for  the  larger  L.  opal- 
escens yolked  oocyte  that  the  alpha-stage  duration  may 
be  even  longer.  The  disappearance  of  unyolked  atretic  oo- 
cytes, as  an  oocyte-like  structure  that  would  be  counted, 
is  more  difficult  to  dismiss  because  so  little  is  known 
about  this  atretic  stage  and  its  duration.  If  our  estimate 
of  the  average  longevity  of  spawning  female  is  only  about 
1.67  days,  then  atretic  losses  of  even  small  unyolked  oo- 
cytes is  probably  not  an  important  bias.  It  would  be  useful 
if  a  way  could  be  found  to  estimate  oocyte  resorption  rates 
in  squid  although  it  may  be  very  difficult.  It  seems  more 
important  to  validate  our  preliminary  estimate  of  the 
average  longevity  of  spawning  squid,  because  if  true,  any 
concerns  regarding  atresia  could  be  dismissed. 

Mature  females  without  postovulatory  follicles  in  their 
ovaries  made  up  only  6%  of  the  247  females  examined 
histologically.  The  rarity  of  these  females  in  our  collections 
reduced  the  precision  of  our  potential  fecundity  estimate. 
Only  thirteen  of  the  fifteen  females  classed  as  a  mature 
preovulatory  female  were  usable  for  estimating  potential 
fecundity,  further  reducing  the  sample  size.  Such  a  small 
sample  size  not  only  results  in  a  low  precision  but  raises 
the  concern  that  the  sample  may  not  be  representative 
of  the  stock  as  a  whole.  The  fact  that  the  average  total 
fecundity  of  females  with  high  mantle  condition  from  the 
catch  was  close  to  the  predicted  value  based  on  the  thir- 
teen females,  indicates  that  the  latter  estimate  may  not  be 
biased.  Clearly  a  larger  sample  size  is  needed,  particularly 
if  egg  escapement  is  used  to  monitor  the  fishery.  It  would 
be  helpful,  in  obtaining  more  samples,  if  we  knew  the 
reason  for  the  apparent  rarity  of  mature  preovulatory  L. 
opalescens  females.  One  possibility  is  that  females  might 
pass  rapidly  from  the  initial  vitellogenesis  to  ovulation, 
perhaps  in  the  course  of  a  single  day  or  some  fraction  of  it, 
and  ovulation  might  begin  sometime  in  the  evening  when 
L.  opalescens  are  the  most  vulnerable  to  fishing.  Another 
possibility  is  that  mature  preovulatory  females  aggregate 
in  regions  not  heavily  fished  by  either  our  trawl  or  the 
fishery. 

Egg  escapement 

A  practical  suggestion  from  this  study  is  the  idea  of  man- 
aging spawning-ground  loliginid  fisheries  by  monitoring 
the  fecundity  of  the  catch  and  computing  the  fraction  of  the 
potential  fecundity  spawned.  Monitoring  the  escapement 
of  eggs  from  the  fishery  is  an  attractive  approach  for  Loligo 
opalescens  because  costs  are  moderate,  unlike  the  high 
cost  for  monitoring  egg  beds  that  cover  many  locations 


offshore  and  occur  at  any  time  of  the  year,  and  because 
traditional  fishery  assessment  models  are  difficult  to 
apply  or  inappropriate  at  the  present  time  (PFMC,  2002). 
To  proceed  with  escapement  fecundity  as  a  management 
tool,  it  would  be  necessary  to  set  a  target  level  for  egg 
escapement  and  to  relate  escapement  to  egg-per-recruit 
analysis  so  that  fishing  effort  could  be  adjusted  to  alter 
egg  escapement  rates.  Conceptual  work  along  these  lines 
has  been  completed  (Maxwell3) 

As  mentioned  earlier,  as  a  practical  matter  in  applying 
the  egg  escapement  method,  one  would  need  to  use  QSP, 
the  mean  fraction  of  the  potential  fecundity  escaping  (Eq. 
1),  as  a  proxy  for  the  more  comprehensive  and  more  use- 
ful measure  of  egg  escapement  Re  tmaz,  the  fraction  of  the 
expected  lifetime  fecundity  deposited  (Eq.  12).  Obviously, 
QSP  will  always  be  lower  than  Re-tm„  because  the  denomi- 
nator of  QSP  (the  fraction  ESPIEP)  is  potential  fecundity 
which  will  always  be  larger  than  the  denominator  for 
Retmal,  which  is  expected  lifetime  fecundity  (E).  Although 
quite  a  different  value,  QSP  is  a  useful  proxy  for  Re  tnai.  If 
natural  mortality  (m )  and  egg  deposition  rates  (v)  are  con- 
stant, changes  in  fishing  mortality  will  result  in  changes 
in  QSP  that  are  proportional  to  the  change  in  ReJmiI. 

However,  changes  will  not  be  proportional  if  either  v 
or  m  varies.  If  there  is  reason  to  believe  that  m  and  v  are 
varying  significantly,  the  use  of  QSP  as  a  proxy  for  Retm!Lll 
should  be  undertaken  with  caution. 

A  point  of  concern  in  applying  this  method  is  that  it  may 
be  difficult  to  substantially  change  escapement  of  eggs  by 
regulating  fishing  effort.  Our  model  indicated  that  egg  es- 
capement may  be  relatively  insensitive  to  changes  in  fish- 
ing mortality  if  natural  mortality  rates  are  as  high  as  we 
believe  them  to  be.  Of  equal  importance  to  management  is 
the  need  to  protect  egg  beds  from  damage  by  nets  and  to 
monitor  the  catch  to  prevent  any  change  that  might  result 
in  the  capture  of  significant  numbers  of  female  L.  opal- 
escens before  they  begin  to  deposit  eggs.  Thus  the  fraction 
of  the  catch  that  is  immature  females  must  be  monitored 
if  the  stock  is  managed  by  using  the  egg  escapement 
method.  For  simplicity,  our  calculations  of  escapement 
were  based  on  only  mature  females  because  immature 
females  were  only  2.6%  of  the  females  in  the  catch 
(1998-99)  and  their  inclusion  had  little  effect  on  param- 
eter estimates.  Egg  escapement  would  decrease  with  an 
increase  in  the  fraction  of  immature  in  the  catch.  As  none 
of  the  fecundity  of  a  captured  immature  female  escapes 
the  fishery,  a  relatively  small  increase  in  the  fraction 
of  immature  animals  in  the  catch  can  have  significant 
consequences. 

From  the  standpoint  of  fishery  management,  the  most 
important  unanswered  question  regarding  the  reproduc- 
tive biology  of  L.  opalescens  is  "how  long  do  they  remain  on 
the  spawning  grounds?"  or  the  equivalent  question  "what 


3  Maxwell,  M.  R..  L.  D.  Jacobson,  and  R.  Conser.  Unpubl. 
data.  Managing  squid  stocks  using  catch  fecundity  in  an 
eggs-per-recruit  model.  Southwest  Fisheries  Science  Center, 
National  Marine  Fisheries  Service,  8604  La  Jolla  Shores  Dr., 
La  Jolla,  CA  92037. 


324 


Fishery  Bulletin  102(2) 


is  the  daily  natural  mortality  of  the  spawners?"  Loligo 
opalescens  have  only  one  spawning  period  in  their  life 
time  (McGowan,  1954;  Fields,  1965;  Butler  et  al.,  1999) 
but  how  long  that  period  lasts  remains  unknown.  Melo 
and  Sauer  (1999)  concluded  that  the  spawning  period  of 
L.  v.  reynaudii  consisted  of  more  than  one  spawning  bout 
but  neither  the  number  of  bouts  nor  the  duration  of  each 
spawning  period  is  known.  In  a  laboratory  study  of  L. 
pealeii  (Maxwell  and  Hanlon,  2000),  the  number  of  bouts 
varied  from  one  to  ten,  the  interval  between  bouts  was 
highly  variable,  and  the  life  span  after  the  first  spawn- 
ing bout  was  from  3  to  50  days.  Our  best  guess  for  L. 
opalescens  under  fishing  conditions  was  an  average  life  on 
the  spawning  grounds  of  only  1.67  days  and  a  maximum 
longevity  of  about  6  days.  These  estimates  were  based  on  a 
simple  exponential  model,  constrained  by  various  proxies 
for  egg  deposition  rate,  longevity,  and  the  fraction  of  the 
potential  fecundity  in  the  catch  (the  only  directly  mea- 
sured value).  We  believe  that  two  of  our  estimates,  36%  of 
the  potential  fecundity  is  deposited  in  the  first  24  hours  of 
spawning  and  minimum  residual  fecundity  is  about  22% 
of  the  potential  fecundity,  are  on  relatively  firm  ground 
but  our  estimate  of  the  maximum  longevity  on  the  spawn- 
ing grounds  as  6  days  is  speculative.  New  information  on 
mortality  is  needed  because,  over  a  wide  range  of  daily 
mortality  rates,  our  model  yields  values  that  are  consis- 
tent with  observed  average  fraction  of  potential  fecundity 
in  the  catch.  Because  direct  measurement  of  mortality  on 
the  spawning  grounds  may  be  difficult,  it  may  be  useful 
to  develop  some  indirect  approaches.  For  example,  a  lab- 
oratory study  could  be  designed  to  generate  an  energy- 
based  model  that  converts  squid  mantle  tissue  loss  to 
deposited  eggs.  This  mantle-to-egg  conversion  rate  could 
be  used  to  assign  an  age  (time  elapsed  after  first  egg  depo- 
sition) to  modes  of  mantle  condition  from  fishery  samples. 
Mortality  could  then  be  computed  by  following  modes  of 
mantle  condition  through  time. 


Acknowledgments 

This  study  was  a  cooperative  project  between  the  Cali- 
fornia Department  of  Fish  &  Game  (CDF&G)  and  the 
National  Marine  Fisheries  Service  (NMFS)  from  start 
to  finish.  We  worked  closely  with  CDF&G  personnel 
throughout  the  study  with  port-sampling  data,  cruise  time, 
and  partial  financial  support  was  provided  by  CDF&G. 
We  worked  particularly  closely  with  M.  Yaremko,  A. 
Henry,  and  D.  Hanan  of  CDF&G.  J.  Welsh  assisted  in 
the  fecundity  work.  Others  that  contributed  include  J. 
Butler,  T.  Kudroschoff,  N.  Smith,  A.  Preti,  K.  Lazar, 
A.  Cossio,  and  at  sea  K.  Barsky,  T.  Bishop,  S  Charter, 
R.  Dotson,  D.  Fuller,  C.  Graff,  D.  Griffith,  P.  Hamdorf, 
A.  Hays,  B.  Horandy,  M.  Levy,  I.  Taniguchi,  J.  Ugoretz,  and 
L.  Zeidberg.  We  wish  to  thank  the  crews  of  the  research 
vessels  Jordan  and  Mako.  We  especially  wish  to  thank  diver 
J.  Hyde  who  observed,  collected,  and  photographed  squid  in 
La  Jolla  Canyon.  M.  Maxwell  and  two  anonymous  review- 
ers read  the  manuscript  and  provided  constructive  com- 
ments. R.  Allen  and  H.  Orr  improved  our  illustrations. 


Literature  cited 

Alheit,  J. 

1993.     Use  of  the  daily  egg  production  method  for  estima- 
ting biomass  of  clupeoid  fishes:  a  review  and  evaluation. 
Bull.  Mar.  Sci.  53:750-767. 
Arkhipkin,  A. 

1993.     Statolith  microstructure  and  maximum  age  of  Loligo 
gahi  (Myopsida:  Loliginidae)  on  the  Patagonian  shelf.     J. 
Mar.  Biol.  Assoc.  U.K.  73:979-982. 
Arkhipkin,  A.,  and  N.  Nekludova. 

1993.     Age,  growth  and  maturation  of  the  loliginid  squids 
Alloteuthis  africana  and  A.  subulata  on  the  west  African 
shelf.     J.  Mar.  Biol.  Ass.  UK.  73:949-961. 
Bower,  J.  R.,  and  Y.  Sakurai. 

1996.     Laboratory  observations  on  Todarodes  pacificus 
(Cephalopoda:  Ommastrephidae)  egg  masses.     Am.  Mala- 
cological  Bull.  13(l/2):65-71. 
Boyle,  P.  R.,  and  M.  A.  K.  Ngoile. 

1993.  Assessment  of  maturity  state  and  seasonality  of  repro- 
duction in  Loligo  forbesi  (Cephalopoda:  Loliginidae)  from 
Scottish  waters.  In  Recent  advances  in  fisheries  biology 
(T.  Okutani,  R.  K.  O'Dor,  and  T.  Kubodera,  eds.),  p.  37-48. 
Tokai  Univ.  Press,  Tokyo,  Japan. 

Boyle,  P.  R.,  G.  J.  Pierce,  and  L.  C.  Hastie. 

1995.     Flexible  reproductive  strategies  in  the  squid  Loligo 
forbesi.     Mar.  Biol.  121:501-508. 
Butler,  J.,  D.  Fuller,  and  M.  Yaremko. 

1999.  Age  and  growth  of  market  squid  iLoligo  opalescens) 
off  California  during  1998.  Calif.  Coop.  Oceanic  Fish. 
Invest.  Rep.  40:191-195. 

Coelho.  M.  L.,  J.  Quintela,  V.  Bettencourt,  G.  Olavo,  and 
H.  Villa. 

1994.  Population  structure,  maturation  patterns  and  fecun- 
dity of  the  squid  Loligo  vulgaris  from  southern  Portugal. 
Fish.  Res.  21:87-102. 

Collins,  M.  A.,  G.  M.  Burnell,  and  P.  G.  Rodhouse. 

1995.  Reproductive  strategies  of  male  and  female  Loligo 
forbesi  (Cephalopoda:  Loliginidae).  J.  Mar.  Biol.  Assoc. 
U.K.  75:621-634. 

Fields,  W.  G. 

1965.     The  structure,  development,  food  relations,  repro- 
duction, and  life  history  of  the  squid,  Loligo  opalescens 
Berry.     Calif.  Dep.  Fish  Game  Fish  Bull.  131,  108  p. 
Fitzhugh,  G.  R.,  and  W.F.  Hettler 

1995.     Temperature  influence  on  postovulatory  follicle  de- 
generation in  Atlantic  menhaden,  Brevoortia  tyrannus. 
Fish.  Bull.  93:568-572. 
Guerra,  A.,  and  F.  Rocha. 

1994.     The  life  history  of  Loligo  vulgaris  and  Loligo  for- 
besi (Cephalopoda:  Loliginidae)  in  Galician  waters  (NW 
Spain).     Fish.  Res.  21:43-69. 
Hatfield,  E.  M.  C. 

1991.  Post-recruit  growth  of  the  Patagonian  squid  Loligo 
gahi  (D'Orbigny).     Bull.  Mar.  Sci.  49(1-2)349-361. 

2000.  So  some  like  it  hot?  Temperature  as  a  possible  deter- 
minant of  variability  in  the  growth  of  the  Patagonian 
squid,  Loligo  gain  (Cephalopoda:  Loliginidae).  Fish.  Res. 
47:27-40. 

Hay,  D.  E.,  D.  N.  Outram,  B.  A.  McKeown,  and  M.  Hurlhurt. 

1987.     Ovarian  development  and  occyte  diameter  as  mat- 
uration criteria  in  Pacific  herring  {Clupea  harengus 
pallasi).     Can.  J.  Fish.  Aquat,  Sci.  44:1496-1502. 
Hunter,  J.  R..  and  S.  R.  Goldberg. 

1980.  Spawning  incidence  and  batch  fecundity  in  northern 
anchovy,  Engraulis  mordax.     Fish.  Bull.  77:641-652. 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Loligo  opalscens 


325 


Hunter,  J.  R.,  and  N.  C.  H.  Lo. 

1997.     The  daily  egg  production  method  of  biomass  estima- 
tion: some  problems  and  potential  improvements.     Ozeano- 
grafika  2:41-67. 
Hunter,  J.  R.,  N.  C.  H.  Lo,  and  R.  J.  H.  Leong. 

1985.  Batch  fecundity  in  multiple  spawning  fishes.  In  An 
egg  production  method  for  estimating  spawning  biomass  of 
pelagic  fish:  application  to  the  northern  anchovy  (Engrau- 
lis mordax)  (R.  Lasker,  ed.),  p.  67-77.  NOAA  Tech.  Rep. 
NMFS  36. 
Hunter,  J.  R.,  and  B.  J.  Macewicz. 

1985a.  Measurement  of  spawning  frequency  in  multiple 
spawning  fishes.  In  An  egg  production  method  for  esti- 
mating spawning  biomass  of  pelagic  fish:  application  to  the 
northern  anchovy  (Engraulis  mordax)  (R.  Lasker,  ed.),  p. 
79-94.  NOAA  Tech.  Rep.  NMFS  36. 
1985b.  Rates  of  atresia  in  the  ovary  of  captive  and  wild  north- 
ern anchovy,  Engraulis  mordax.  Fish.  Bull.  83:119-136. 
Hunter,  J.  R.,  B.  J.  Macewicz,  N.  C.  H.  Lo,  and  C.  A.  Kimbrell. 

1992.  Fecundity,  spawning,  and  maturity  of  female  Dover 
sole  Microstomas  pacificus,  with  an  evaluation  of  assump- 
tions and  precision.     Fish.  Bull.  90:101-128. 

Ikeda,  Y.,  Y.  Sakurai,  and  K.  Shimazaki. 

1993.  Maturation  process  of  the  Japanese  common  squid 
Todarodes  pacificus  in  captivity.  In  Recent  advances  in 
fisheries  biology  (T  Okutani,  R.  K.  O'Dor,  and  T.  Kubodera, 
eds.),  p.  179-187.     Tokai  Univ.  Press,  Tokyo,  Japan. 

Jackson,  G.  D. 

1993.  Seasonal  variation  in  reproductive  investment 
in  the  tropical  loliginid  squid  Loligo  chinensis  and  the 
small  tropical  sepioid  Idiosepius  pygmaeus.  Fish.  Bull. 
91:260-270. 

1994.  Statolith  age  estimates  of  the  loliginid  squid  Loligo 
opalescens  (Mollusca:  Cephalopoda):  corroboration  with 
culture  data.     Bull.  Mar.  Sci.  54:554-557. 

Jackson,  G.  D.,  A.  I.  Arkhipkin,  V.  A.  Bizikov,  and  R.  T.  Hanlon. 
1993.  Laboratory  and  field  corroboration  of  age  and  growth 
from  statoliths  and  gladii  of  the  loliginid  squid  Sepio- 
teuthis  lessoniana  (Mollusca:  Cephalopoda).  In  Recent 
advances  in  fisheries  biology  (T.  Okutani,  R.  K.  O'Dor, 
and  T.  Kubodera,  eds.),  p.  189-199.  Tokai  Univ.  Press, 
Tokyo,  Japan. 

Jackson,  G.  D.,  J.  W.  Forsythe,  R.  F.  Hixon,  and  R.  T.  Hanlon. 

1997.  Age,  growth,  and  maturation  of  Lolliguncula  brevis 
(Cephalopoda:  LoliginidaeHn  the  northwest  Gulf  of  Mexico 
with  a  comparison  of  length-frequency  versus  statolith  age 
analysis.     Can.  J.  Fish.  Aquat.  Sci.  54:2907-2919. 

Jackson,  G.  D.,  and  J.  Yeatman. 

1996.     Variation  in  size  and  age  at  maturity  in  Photololigo 

(Mollusca:  Cephalopoda)  from  the  northwest  shelf  of 

Australia.     Fish.  Bull.  94:59-65. 
Kjesbu,  O.  S.,  P.  R.  Witthames,  P.  Solemdal,  M.  G.  Walker. 

1998.  Temporal  variations  in  the  fecundity  of  Arcto-Norwe- 
giancod  {Gadus  morhua)  in  response  to  natural  changes  in 
food  and  temperature.    J.  Sea  Res.  40:  303-321. 

Knipe,  J.  H.,  and  R.  D.  Beeman. 

1978.  Histological  observations  on  oogenesis  in  Loligo 
opalescens.  In  Biological,  oceanographic,  and  acoustic 
aspects  of  the  market  squid,  Lolgio  opalescens  Berry  (C. 
W.  Recksiek  and  H.  W.  Frey,  eds.),  p.  23-33.  Calif.  Dep. 
Fish  Game  Fish  Bull.  169. 
Laptikhovsky,  V. 

2000.     Fecundity  of  the  squidLo/igo  vulgaris  Lamarck,  1798 


(Myopsida,  Loliginidae)  off  northwest  Africa.     Scientia 
Marina  64(3):275-278. 
Lopes,  S.  S.,  M.  L.  Coelho,  and  J.  P.  Andrade. 

1997.  Analysis  of  oocyte  development  and  potential  fecun- 
dity of  the  squid  Loligo  vulgaris  from  the  waters  of  south- 
ern Portugal.     J.  Mar.  Biol.  Assoc.  U.K.  77:903-906. 

Macewicz,  B.  J.,  and  J.  R.  Hunter. 

1994.  Fecundity  of  sablefish,  Anoplopoma  fimbria,  from 
Oregon  coastal  waters.  Calif.  Coop.  Oceanic  Fish.  Invest. 
Rep.  35:160-174. 

Maxwell,  M.  R.,  and  R.  T.  Hanlon. 

2000.     Female  reproductive  output  in  the  squid  Loligo  pea- 
led: multiple  egg  clutches  and  implications  for  a  spawning 
strategy.     Mar.  Ecol.  Prog.  Ser.  199:159-170. 
McGowan,  J.  A. 

1954.     Observations  on  the  sexual  behavior  and  spawning  of 
the  squid,  Loligo  opalescens,  at  La  Jolla,  California.     Calif. 
Fish  Game  Fish  Bull.  40(l):47-55. 
Melo,  Y  C,  and  W.  H.  H.  Sauer. 

1998.  Ovarian  atresia  in  cephalopods.  S.  Afr.  J.  Mar.  Sci. 
20:143-151. 

1999.  Confirmation  of  serial  spawning  in  the  chokka 
squid  Loligo  vulgaris  reynaudii  off  the  coast  of  South 
Africa.     Mar.  Biol.  135:307-313. 

Moltschaniwskyj,  N.  A.. 

1995.  Multiple  spawning  in  the  tropical  squid  Photolo- 
ligo sp.:  what  is  the  cost  in  somatic  growth?  Mar.  Biol. 
125:127-135. 

Moltschaniwskyj,  N.  A.,  and  J.  M.  Semmens. 

2000.  Limited  use  of  stored  energy  reserves  for  reproduc- 
tion by  the  tropical  loliginid  squid  Photololigo  sp.  J.  Zool. 
Lond.  251:307-313. 

Natsukari,  Y.,  and  N.  Komine. 

1992.  Age  and  growth  estimation  of  the  European  squid, 
Loligo  vulgaris,  based  on  statolith  microstructure.  J. 
Mar.  Biol.  Assoc.  U.K.  72:271-280. 

PFMC 

2002.  Report  of  the  Stock  Assessment  Review  (STAR) 
panel  for  market  squid.  Appendix  3.  In,  Status  of  the 
Pacific  Coast  coastal  pelagic  species  fishery  and  recom- 
mended acceptable  biological  catches:  stock  assessment 
and  fishery  evaluation  —  2002,  17  p.  Pacific  Fishery 
Management  Council.  Portland,  OR 

Quinn,  T.  J.,  and  R.  B.  Deriso. 

1999.  Quantitative  fish  dynamics.  Biological  resource 
Mmanagement  series,  542  p.  Oxford  Univ.  Press,  New 
York,  NY. 

Sauer,  W.  H.  H.,  Y.  C.  Melo,  and  W.  de  Wet. 

1999.  Fecundity  of  the  chokka  squid  Loligo  vulgaris  rey- 
naudii on  the  southeastern  coast  of  South  Africa.  Mar. 
Biol.  135:315-319. 

Semmens,  J.  M.,  and  N.  A.  Moltschaniwskyj. 

2000.  An  examination  of  variable  growth  in  the  loliginid 
squid  Sepioteuthis  lessoniana:  a  whole  animal  and  reduc- 
tionist approach.     Mar.  Ecol.  Prog.  Ser.  193:135-141. 

Woodhead,  A.  D. 

1960.     Nutrition  and  reproductive  capacity  in  fish.     Nutri- 
tion Society  Proceedings  19:23-28. 
Yang,  W.  T.,  R.  F.  Hixon,  P.  E.  Turk,  M.  E.  Krejci,  W.  H.  Hulet, 
and  R.  T.  Hanlon. 

1986.  Growth,  behavior,  and  sexual  maturation  of  the 
market  squid,  Loligo  opalescens,  cultured  through  the  life 
cycle.     Fish.  Bull.  84:771-798. 


326 


Fishery  Bulletin  102(2) 


Appendix  I 

Terms 

EP      potential  fecundity  (standing  stock  of  oocytes 
in  the  ovary  of  mature  females  prior  to  spawn- 
ing) 
ESPl      the  total  eggs  deposited  on  the  bottom  up  to 
time  t.  (t  in  days) 

Ec  total  number  of  eggs  deposited  by  mature 
females  in  the  catch  or  total  number  of  eggs 
escaped 

EM  total  number  of  eggs  deposited  by  mature 
females  prior  to  death  due  to  natural  mortal- 
ity 

EA  total  number  of  eggs  deposited  by  females  alive 
and  not  caught  by  fishery 

E  total  number  of  eggs  that  would  have  been 
spawned  during  a  squid's  lifetime  if  no  fishery 
existed 

EY      standing  stock  of  oocytes  in  the  ovary 

ED      standing  stock  of  ova  in  the  oviduct 

EYD  total  fecundity,  the  sum  of  both  the  number  of 
oocytes  in  the  ovary  and  ova  in  the  oviduct 

EYD  tlt  stocking  stock  of  oocytes  in  the  ovary  plus 
those  ova  in  the  oviduct  after  spawning  has 
begun  and  up  to  the  elapsed  time  tk,  where  tk 

ls  £  'max 

tmax  maximum  elapsed  time  with  the  time  0  being 
the  time  when  mature  females  are  about 
to  ovulate  or  total  elapsed  time  (in  days)  of 
spawners  on  the  spawning  ground 

ER  standing  stock  of  oocytes  remaining  in  ovary 
at  death 

m      daily  adult  natural  mortality  rate 

f     daily  fishing  mortality  rate  for  adults 

v      daily  egg  deposition  rate 

Qsr  i  =  EsPt  IEp  =  1-  ervt     fraction  of  potential  fecundity 

deposited  up  to  time  t 

e~zl      mortality  (survival)  curve 

JV„      number  of  mature  females  at  time  0 

Nr      total  number  of  spawners  in  the  catch 

Re  tk  egg  escapement  rate  =  ratio  of  eggs  deposited 
to  total  number  of  egg  which  would  be  spawned 
if  there  was  no  fishery,  at  a  given  elapsed  time 

Ret  egg  escapement  rate  up  to  the  maximum 

elapsed  time  Umax) 


Appendix  II 

For  any  elapsed  time  tk,  formulas  for  Ec,  EM,  EA  and  E: 


Ec  =  EpNJ 


EM  =  EPN0m 


1-e 


-lm+f)tk        ^_e-(m+f+v)t), 


m  +  f  m+f+v 


1-e 


-{m+f)tk        -.  _    -im+f+v)tk 


m  +  f  m+f+v 


EA  =  (EP-EYD.)Nk  =  AT0e-""+/  "*  EP  ( 1  -  e"1"' ) 


V  -mtJ  ,  m 

e         1 e 


E  =  EPN0 


The  derivation  for  EA  is  straight  forward  and  the  deriva- 
tions for  £r,  EM,  and  E,  similar  among  one  another,  are 
as  follows: 


where 


Ec=\  ESP,dC,  =  J  Ep~a-  <?"'''  )dC, 


=  Ep~N0jn-e-")e-'m+n'fdt, 


C.=N0-^—a-e-""+f") 
m  +  f 


is  the  number  of  removals  of  the  cohort  due  to  fishing  up 
to  time  t  (Quinn  and  Deriso,  1999). 


EM=\E~TldDnU=\Ep(\-e-«)dDnU 


EpNJ 


(m  +  f)(m  +  f  +  u) 


where 


D„,t=N0-—;(l- 


!!!_h  _<,-"»*/'") 


Macewicz  et  al.:  Fecundity,  egg  deposition,  and  mortality  of  Loligo  opa/scens 


327 


is  the  number  of  removals  of  the  cohort  due  to  natural 
mortality  up  to  time  t  (Quinn  and  Deriso,  1999)  and  E 
(when  no  fishing  takes  place): 


E=\  ESPjdDt  =  f  E^(  1  -  e"1"  )dDl 


=  EpN0\(l-e-l")e'm'mdt, 
o 

where  Dt  =  N0(l-e-mt)  is  the  number  of  removals  of  the 
cohort  due  to  natural  mortality  when  no  fishing  takes 
place. 


328 


Abstract— The  dusky  rockfish  (Se- 
bastes  ciliatus)  of  the  North  Pacific 
Ocean  has  been  considered  a  single 
variable  species  with  light  and  dark 
forms  distributed  in  deep  and  shal- 
low water,  respectively.  These  forms 
have  been  subjected  to  two  distinct 
fisheries  separately  managed  by  fed- 
eral and  state  agencies:  the  light  deep 
form  is  captured  in  the  offshore  trawl 
fishery;  the  dark  shallow  form,  in  the 
nearshore  jig  fishery.  The  forms  have 
been  commonly  recognized  as  the  light 
dusky  and  dark  dusky  rockfishes.  From 
morphological  evidence  correlated  with 
color  differences  in  some  400  speci- 
mens, we  recognize  two  species  cor- 
responding with  these  color  forms. 
Sebastes  ciliatus  (Tilesius)  is  the  dark 
shallow-water  species  found  in  depths 
of  5-160  m  in  the  western  Aleutian 
Islands  and  eastern  Bering  Sea  to 
British  Columbia.  The  name  Sebastes 
variabilis  (Pallas)  is  resurrected  from 
the  synonymy  of  S.  ciliatus  to  apply  to 
the  deeper  water  species  known  from 
depths  of  12-675  m  and  ranging  from 
Hokkaido,  Japan,  through  the  Aleu- 
tian Islands  and  eastern  Bering  Sea,  to 
Oregon.  Sebastes  ciliatus  is  uniformly 
dark  blue  to  black,  gradually  lightening 
on  the  ventrum,  with  a  jet  black  peri- 
toneum, a  smaller  symphyseal  knob, 
and  fewer  lateral-line  pores  compared 
to  S.  variabilis.  Sebastes  variabilis  is 
more  variable  in  body  color,  ranging 
from  light  yellow  to  a  more  usual  tan 
or  greenish  brown  to  a  nearly  uniform 
dark  dorsum,  but  it  invariably  has  a 
distinct  red  to  white  ventrum.  Syn- 
onymies, diagnoses,  descriptions,  and 
geographic  distributions  are  provided 
for  each  species. 


The  dusky  rockfishes  (Teleostei:  Scorpaeniformes) 
of  the  North  Pacific  Ocean:  resurrection  of 
Sebastes  variabilis  (Pallas,  1814)  and  a 
redescription  of  Sebastes  ciliatus  (Tilesius,  1813) 


James  W.  Orr 

Resource  Assessment  and  Conservation  Engineering  Division 

Alaska  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

7600  Sand  Point  Way  NE 

Seattle,  Washington  98115 

E-mail  address:  James  Orr@noaa.gov 


James  E.  Blackburn 

Alaska  Department  of  Fish  and  Game 

211  Mission  Road 

Kodiak,  Alaska  99615-9988 


Manuscript  approved  for  publication 
22  December  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:328-348(2004). 


Among  the  approximately  92  species 
of  Sebastes  found  in  the  North  Pacific, 
two  commercially  important  species 
long  identified  under  the  name  Sebastes 
ciliatus  have  been  taxonomically  prob- 
lematic. The  name  S.  ciliatus  (Tilesius, 
1813)  has  been  commonly  applied  to 
specimens  considered  to  represent  a 
single  variable  species  ranging  from 
northern  Japan  to  British  Columbia 
(Barsukov,  1964;  Westrheim,  1973; 
Shinohara  et  al.,  1994;  Mecklenburg  et 
al.,  2002),  and  the  name  S.  variabilis 
(Pallas,  1814)  has  been  treated  as  a 
junior  synonym  (Jordan  and  Gilbert, 
1881;  Eigenmann  and  Beeson,  1894; 
Jordan  and  Evermann,  1898;  Blanc 
and  Hureau,  1968).  Two  color  forms 
within  S.  ciliatus  have  been  reported 
and  hypothesized  to  be  distinct  species 
(Quast  and  Hall,  1972;  Eschmeyer  et 
al.,  1983;  Kessler,  1985;  Fig.  1).  The 
typically  light-colored  form,  commonly 
known  as  the  light  dusky  rockfish,  is 
often  found  in  large  aggregations  over 
the  outer  continental  shelf  and  upper 
slope  at  depths  down  to  675  m,  and 
less  frequently  in  nearshore  habitats. 
The  dark-blue  to  black  form,  commonly 
known  as  the  dark  dusky  rockfish,  is 
found  in  more  shallow  habitats  from 
nearshore  rocky  reefs  to  depths  no 
greater  than  160  m. 

These  forms  have  been  subjected  to 
two  distinct  fisheries  separately  man- 
aged by  U.S.  federal  and  Alaska  state 


agencies  since  1998.  The  light-colored 
deep  form  is  captured  in  the  offshore 
trawl  fishery  and  is  the  dominant  spe- 
cies of  the  pelagic  shelf  rockfish  fisher- 
ies complex  regulated  by  the  National 
Marine  Fisheries  Service  (NMFS). 
Specific  catch  limits  are  set  under  the 
designation  "dusky  rockfish."  The  oc- 
casional catch  of  the  dark  form  in  these 
offshore  waters  has  also  been  consid- 
ered "dusky  rockfish."  The  dark-colored 
shallow  form  is  found  commonly  in  the 
nearshore  jig  fishery  regulated  by  the 
Alaska  Department  of  Fish  and  Game. 
The  dark  form,  routinely  misidentified 
as  S.  melanops,  may  comprise  up  to  259c 
of  the  catch  in  the  "black  rockfish"  jig 
fishery  off  Kenai  Peninsula  (Clausen 
et  al.1)  and  is  managed  only  as  "other 
rockfish"  bycatch  within  the  fishery. 

Early  allozyme  analyses  (Tsuyuki 
et  al.,  1965,  1968)  indicated  signifi- 
cant genetic  differences  among  samples 
identified  as  S.  ciliatus.  A  more  detailed 
analysis  of  several  Sebastes  species 
(Seeb,  1986)  and  a  later  study  focused 


1  Clausen,  D.  M.,  C.  R.  Lunsford,  and 
J.  T.  Fujioka.  2002.  Pelagic  shelf 
rockfish.  In  Stock  assessment  and  fish- 
ery evaluation  report  for  the  groundfish 
resources  of  the  Gulf  of  Alaska  for  2002,  p. 
383-417.  North  Pacific  Fishery  Manage- 
ment Council,  605  W  4th  Ave,  Suite  306, 
Anchorage,  AK  99501. 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescription  of  Sebastes  ciliatus 


329 


Figure  1 

(A)  Sebastes  variabilis  (top),  UW  43494,  225.2  mm,  and  S.  ciliatus  (bottom),  UW  43493,  neotype,  266.4  mm, 
collected  at  37  and  25  m  depth,  respectively,  in  Lynn  Canal  near  Funter  Bay,  southeast  Alaska.  (Bl  Sebastes 
ciliatus  (top),  UW  45512,  235.2  mm,  and  S.  variabilis  (bottom),  UW  45511,  206.2  mm,  collected  syntopically 
at  67  m  depth  in  the  northern  Gulf  of  Alaska,  57.38061°N,  154.8009°W.  (C)  Sebastes  variabilis  (left),  UW 
43494,  150.6-225.2  mm,  and  S.  ciliatus  (right),  UW  43492,  153.7-241.1  mm,  collected  at  37  and  25  m  depth, 
respectively,  in  Lynn  Canal  near  Funter  Bay,  southeast  Alaska.  (D)  Sebastes  variabilis,  UW  43251,  390  mm 
(top)  and  410  mm  (bottom),  northern  Gulf  of  Alaska,  59.50446°N,  145.2262°W,  135  m  depth.  (E)  Sebastes 
mclanops  (top),  UW  43490,  and  S.  ciliatus  (bottom),  UW  43484,  313  mm,  collected  at  25  m  depth  in  Soapstone 
Cove,  southeast  Alaska. 


330 


Fishery  Bulletin  102(2) 


//,-m^'s'   <£/c£l%ytQ- ^Q  ft&*^/!£7<r?7is 7rOC/ zw 


B 


^/ 


£%tt2,.„  Sj 


• — S  .t_// eyes'    jfti. 


Figure  2 

( Ai  Pcrca  variabilis  Pallas,  MNHN  8670,  lectotype,  343.7  mm.  "mari  Americam  borealum."  (B)  Epinephelus  ciliatus  Tilesius, 
illustration  of  holotype  after  Tilesius  (1813),  specimen  presumed  lost. 


on  S.  ciliatus  (Westrheim  and  Seeb2)  concluded  that  the 
two  color  forms  were  distinct  sister  species.  Seeb's  recent 
work  with  microsatellite  DNA  data  has  revealed  discrete 
genetic  differences  between  the  two,  as  well  as  some  evi- 
dence for  infrequent  hybridization  (Seeb3).  Sequence  data 


2  Westrheim,  S.  J.,  and  L.  W.  Seeb.  1997.  Unpubl.  manu- 
script. Investigation  of  the  Sebastes  ciliatus  species  group. 
36  p.  Fisheries  and  Oceans  Canada,  Pacific  Biological  Sta- 
tion, Nanaimo,  BC,  Canada  V9R  5K6. 

3  Seeb,  L.  W.  2002.  Personal  commun.  Alaska  Department 
of  Fish  and  Game,  333  Raspberry  Road,  Anchorage,  AK  99518- 
1599. 


from  NADH  dehydrogenase  subunit  regions  of  the  mito- 
chondrial genome,  however,  have  not  revealed  differences 
between  the  two  forms  ( Lopez4,  Gray5),  nor  have  sequence 
data  from  other  work  on  closely  related  species  of  Sebastes 
(Bentzen  et  al.,  1998;  Sundt  and  Johansen,  1998;  Roques 
etal.,  2001). 


4  Lopez,  J.  A.  2000.  Personal  commun.  Iowa  State  Univ., 
Ames,  IA  50014. 

6  Gray,  A.  2000.  Personal  commun.  Fisheries  Division, 
School  of  Fisheries  and  Ocean  Sciences,  Univ.  Alaska  Fair- 
banks. 11120  Glacier  Highway,  Juneau,  AK  99801. 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescnption  of  Sebastes  ciliatus 


331 


In  this  study,  we  provide  morphological  evidence  from 
examination  of  about  400  specimens  collected  throughout 
the  geographic  and  bathymetric  range  of  the  species  to 
correlate  color  differences  with  meristic  and  shape  differ- 
ences. In  thus  recognizing  two  species,  S.  ciliatus  and  S. 
variabilis,  previously  referred  to  the  name  S.  ciliatus  (Tile- 
sius,  1813),  we  discuss  the  nomenclatural  consequences  of 
this  decision.  Both  species  were  originally  described  (as 
Epinephelus  ciliatus  Tilesius  and  Perca  variabilis  Pallas) 
on  the  basis  of  early  Russian  collections  from  along  the 
Aleutian  Islands  (Svetovidov,  1978,  1981).  The  type  series 
of  one  species  is  now  represented  by  a  single  extant  speci- 
men (Fig.  2A)  and  the  other  by  the  illustration  of  a  single, 
now  lost,  specimen  (Svetovidov,  1978,  1981;  Fig.  2B). 
Although  workers  since  the  turn  of  the  century  have  as- 
sociated the  name  S.  ciliatus  with  the  variably  light-col- 
ored species  (Jordan,  1896;  Jordan  and  Evermann,  1898; 
Barsukov,  1964;  Orr  et  al.,  1998,  2000;  Mecklenburg  et 
al.,  2002),  the  original  description  and  accompanying  il- 
lustration (Fig.  2B)  appear  to  describe  the  uniformly  dark 
species.  We  have  also  identified  the  remaining  syntype 
(Fig.  2A)  ofPeiea  variabilis  as  the  light  species.  Therefore, 
we  refer  the  dark,  shallow-water  species  (the  dark  rock- 
fish)  to  Sebastes  ciliatus  (Tilesius,  1813)  and  resurrect  the 
name  Sebastes  variabilis  (Pallas,  1814)  for  the  typically- 
light,  deeper-water  species  (the  dusky  rockfish). 


Methods  and  materials 

Counts  and  measurements  follow  Hubbs  and  Lagler 
(1958),  except  as  noted  below.  Unless  indicated  otherwise, 
standard  length  (SL)  is  used  throughout  and  was  always 
measured  from  the  tip  of  the  snout.  Depth  at  pelvic-fin  base 
was  measured  from  the  origin  of  the  dorsal  fin  to  the  base 
of  the  pelvic  fins  (at  the  articulation  of  the  pelvic-fin  spine); 
depth  at  anal-fin  origin,  from  the  base  of  the  last  dorsal- 
fin  spine  to  the  anal-fin  origin;  depth  at  anal-fin  insertion, 
from  dorsal-fin  insertion  to  anal-fin  insertion;  body  thick- 
ness, at  pectoral-fin  base;  head  thickness,  at  the  posterior 
orbital  rim;  prepelvic-  and  preanal-fm  length,  from  pelvic- 
fin  base  or  anal-fin  origin  to  the  tip  of  the  snout;  pelvic-fin 
to  anal-fin  length  from  pelvic-fin  base  to  anal-fin  origin; 
caudal  peduncle  dorsal  length  from  dorsal-fin  insertion  to 
caudal-fin  base;  caudal  peduncle  ventral  length  from  anal- 
fin  insertion  to  caudal-fin  base.  The  small  anterior  notch 
in  the  orbit  between  the  frontal  bone  and  lateral  ethmoid 
was  excluded  from  orbit  length  and  snout  length  measure- 
ments. Accessory  scales  are  small  scales  located  beyond 
the  posterior  field  of  major  scales.  The  swimbladder  mus- 
culature was  examined  after  dissection  according  to  the 
methods  of  Hallacher  (1974).  Institutional  abbreviations 
follow  Leviton  et  al.  (1985)  and  Leviton  and  Gibbs  ( 1988), 
as  modified  by  Poss  and  Collette  (1995). 

Individuals  were  identified  by  body  and  peritoneum  col- 
or (see  species  descriptions  below)  for  grouping  in  ANOVA 
and  ANCOVA,  as  well  as  for  labeling  individuals  in  graphs 
of  principal  components  analysis  scores.  Univariate  and 
multivariate  analyses  were  conducted  by  using  Statgraph- 
ics  Plus  4.1  (Manugistics,  Rockville,  MD)  and  Splus  2000 


(Mathsoft,  Inc.,  Seattle,  WA).  Differences  were  considered 
significant  at  P  <  0.05. 

Arcsine-transformed  morphometric  ratios  (with  SL  or 
head  length  as  denominator)  and  meristic  characters  were 
tested  to  meet  the  assumptions  of  normality  required  for 
ANOVA.  The  following  characters  exhibited  normal  distri- 
butions and  did  not  differ  significantly  in  variance  between 
species  and  were  subjected  to  ANOVA:  head  length,  orbit 
length,  snout  length,  interorbital  width,  suborbital  depth, 
gill-raker  length,  body  thickness,  pectoral-fin  base  width, 
pectoral-fin  ray  length,  caudal  peduncle  ventral  length, 
predorsal  length,  spinous  dorsal-fin  base,  soft  dorsal-fin 
base,  and  counts  of  lateral-line  pores  and  gill  rakers. 

For  morphometric  characters,  significant  differences 
were  also  identified  by  using  an  analysis  of  covariance 
(ANCOVA)  of  log-10-transformed  measurements  with  SL 
or  head  length  (HL)  as  covariates  when  assumptions  of 
normality  and  the  homogeneity  of  slopes  were  satisfied. 
The  ANCOVA  model  included  species  as  a  factor,  SL  or  HL 
as  a  covariate,  and  a  species/I  SL  or  HL)  interaction  (e.g., 
HL  =  C+Species+SL+(SpeeiesxSL)).  A  residual  analysis 
was  done  for  each  model  to  determine  the  appropriateness 
of  the  model.  Whenever  the  interaction  was  not  significant 
(at  the  5%  level),  a  reduced  model  was  used  by  dropping 
the  interaction  and  forcing  the  slopes  to  be  the  same 
(BD=C+Species  +SL).  This  removed  the  effect  of  SL  and 
HL  and  allowed  testing  for  significant  differences  between 
species.  The  following  morphometric  characters  met  the 
assumptions  required  for  ANCOVA:  head  length,  snout 
length,  interorbital  width,  gill-raker  length,  pectoral- 
fin  base  width,  pectoral-fin  ray  length,  caudal  peduncle 
ventral  length,  predorsal  length,  spinous  dorsal-fin-base 
length,  and  soft  dorsal-fin-base  length. 

On  a  dataset  of  specimens  with  all  characters,  sheared 
principal  components  analysis  (SPCA)  for  a  size-free 
analysis  (Bookstein  et  al.,  1985)  was  conducted  by  us- 
ing morphometric  characters,  and  a  standard  principal 
components  analysis  (PCA)  was  conducted  by  using  all 
meristic  characters.  Raw  morphometric  data  were  log- 
transformed  and  the  covariance  matrix  was  subjected 
to  SPCA,  as  was  the  correlation  matrix  of  raw  meristics. 
Differences  between  species  were  illustrated  by  plotting 
scores  of  sheared  PC2  against  sheared  PC3  and  sheared 
morphometric  PC2  against  the  standard  meristic  PCI. 
Separate  analyses  were  also  conducted  on  three  group- 
ings: 1)  each  species  by  depth,  2)  each  species  by  sex,  and 
3 )  shallow-water  populations  ofS.  ciliatus  and  S.  variabilis 
primarily  collected  in  the  vicinity  of  the  Triplet  Islands 
and  Monashka  Bay,  on  the  northeast  side  of  the  Kodiak 
Island  Archipelago,  and  the  vicinity  of  Lynn  Canal, 
Alaska.  Shallow  collections  were  defined  as  those  made  at 
less  than  50  m  depth,  and  deep  collections  were  taken  at 
depths  greater  than  50  m. 

These  plots  were  also  examined  for  groupings  indicative 
of  geographic  differences  in  body  shape  and  meristics. 
Geographic  areas  were  defined  as  follows:  British  Colum- 
bia, from  the  Straits  of  Juan  de  Fuca  to  Dixon  Entrance; 
southeast  Alaska,  from  Dixon  Entrance  to  Chatham 
Strait;  Gulf  of  Alaska,  from  Chatham  Strait  to  the  tip  of 
the  Alaska  Pennisula;  Aleutian  Islands  and  Bering  Sea, 


332 


Fishery  Bulletin  102(2) 


Table  1 

Proportional  morphometries  and  meristics  ofSebastes  cihatus  and  S.  variabilis  from  all  depths  and  regions.  Morphometric  data 
are  in  %SL  or  %HL.  X  =  statistically  significant  difference  at  0.05  level,  as  evaluated  by  ANOVA  and  ANCOVA.  when  appropri- 
ate; ns  =  not  statistically  significant  at  0.05  level,  n  =  number  offish  in  sample. 

S.ci 

liatus 

S.  varia 

nlis 

ANOVA      ANCOVA 

n 

Range 

Mean  ±SD 

n 

Range 

Mean  ±SD 

Meristics 

Dorsal-fin  spines 

138 

12-14 

13.0+0.2 

194 

13-14 

13.0  ±0.1 

Dorsal-fin  rays 

138 

13-17 

15.0  ±0.5 

194 

13-16 

15.0  ±0.4 

Anal-fin  rays 

139 

7-9 

7.9  ±0.4 

195 

7-9 

7.9  ±0.4 

Pectoral-fin  rays  (left) 

138 

17-19 

18.2  ±0.5 

194 

17-19 

18.0  ±0.3 

Pectoral-fin  rays  (right) 

137 

16-19 

18.2  ±0.4 

195 

16-19 

18.0  ±0.4 

Unbranched  pectoral-fin 
rays  ( left ) 

136 

8-10 

9.2  ±0.5 

195 

7-11 

9.1  ±0.4 

Unbranched  pectoral-fin 
rays  ( right  I 

108 

8-11 

9.2  ±0.5 

188 

7-11 

9.1  ±0.5 

Lateral-line  pores  (left) 

138 

39-50 

45.4  ±2.3 

188 

43-54 

48.5  ±1.9 

X 

Lateral-line  pores  (right) 

125 

40-54 

45.3  ±2.4 

172 

42-54 

48.5  ±1.9 

X 

Lateral-line  scales 

132 

44-60 

50.6  ±2.7 

177 

47-63 

52.8  ±2.7 

Gill  rakers 

137 

32-37 

34.8  ±1.1 

184 

32-37 

34.7  ±0.9 

X 

continued 

from  the  tip  of  the  Alaska  Peninsula  west  and  north  into 
the  Bering  Sea. 


Results 

Color 

Body  color  in  life  and  in  preservation  differs  consistently 
between  S.  ciliatus  and  S.  variabilis  (Fig.  1;  see  detailed 
description  below).  In  life,  S.  ciliatus  is  uniformly  bluish- 
black  to  gray,  with  slight  gradual  lightening  on  the  belly; 
the  peritoneum  is  invariably  jet  black.  In  contrast,  S. 
variabilis  varies  in  background  color  from  golden  yellow 
to  greenish  brown  to  dark  gray,  with  a  distinct  break 
between  the  darker  dorsum  and  the  invariably  white  to 
pink  ventrum,  particularly  at  the  base  of  the  anal  fin;  the 
peritoneum  is  gray  to  black.  In  S.  variabilis  preserved  for 
up  to  30  years,  the  distinct  break  along  the  ventrum  is 
retained  and  differs  from  the  uniformly  dark  preserved 
color  of  S.  ciliatus.  This  combination  of  characteristic 
body  and  peritoneum  color  was  used  initially  to  identify  in- 
dividuals as  either  S.  ciliatus  or  S.  variabilis  as  the  basis 
for  univariate  statistical  analyses. 

Meristic  characters 

Lateral-line  pore  and  gill-raker  counts  differed  signifi- 
cantly between  S.  ciliatus  and  S.  variabilis  from  all  depths 
and  regions,  S.  ciliatus  having  a  lower  range  and  mode  of 
counts  (Tables  1-3).  In  shallow  water,  only  lateral-line 
pore  counts  showed  significant  differences  (Table  4). 


Slight  clinal  variation  was  evident  for  lateral-line  pores 
in  S.  ciliatus  between  southeast  Alaska  collections  and 
northern  Gulf  of  Alaska  material  (Table  2).  In  the  PCA, 
counts  of  lateral-line  pores,  gill  rakers,  and  pectoral-fin 
rays  were  most  heavily  loaded  along  the  first  PC  axis, 
confirming  that  S.  ciliatus  has  typically  lower  lateral-line 
pore  and  gill-raker  counts  and  tends  to  have  a  higher  pec- 
toral-fin ray  count  (Tables  1-3,  5;  Fig.  3B). 

Morphometric  characters 

Among  morphometric  characters  meeting  statistical 
assumptions  for  ANOVA  or  ANCOVA,  head  length, 
interorbital  width,  suborbital  depth,  lower-jaw  length, 
gill-raker  length,  body  thickness,  pectoral-fin  base  width, 
predorsal  length,  and  soft-dorsal-fin-base  length  differed 
significantly  between  S.  ciliatus  and  S.  variabilis  across 
all  depths  and  regions  (Table  1).  Between  shallow-water 
S.  ciliatus  and  S.  variabilis,  all  the  above  characters, 
except  head  length,  interorbital  width,  and  predorsal 
length,  differed  significantly  (Table  4).  No  significant  dif- 
ferences were  found  in  analyses  within  species  by  depth 
or  sex. 

In  the  PCA  for  specimens  collected  across  all  regions 
and  depths,  clusters  of  S.  ciliatus  and  S.  variabilis  showed 
broad  overlap  and  only  slight  discrimination  among  indi- 
viduals along  the  PC2  axis  (Fig.  3A).  Principal  component 
2,  the  primary  shape  component,  described  1.8%  of  the 
total  variation,  and  PCI,  the  size  component  having  all 
loadings  positive,  described  96.4rr  of  the  variation.  Char- 
acters loading  most  heavily  along  the  PC2  axis  included 
suborbital  depth,  gill-raker  length,  orbit  length,  body 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescription  of  Sebastes  ciliatus 


333 


Table  1  (continued) 

S.  ciliatus 

S.  varia 

bilis 

ANOVA 

ANCOVA 

n 

Range 

Mean  ±SD 

n 

Range 

Mean  ±SD 

Morphometries 

Standard  length 

132 

83.8-340.0 

192 

77.7-430.8 

Head  length/SL 

113 

28.7-35.5 

32.9  ±1.2 

129 

28.1-36.2 

32.5  ±1.2 

X 

ns 

Orbit  length/HL 

111 

21.5-30.6 

25.6  ±2.0 

121 

20.6-33.5 

25.6  ±2.1 

ns 

Snout  length/HL 

111 

18.2-26.0 

21.4  ±1.6 

121 

17.1-27.1 

21.4  ±1.8 

ns 

ns 

Interorbital  width/HL 

111 

22.9-29.3 

25.9  ±1.3 

121 

22.5-30.4 

26.4  ±1.4 

X 

X 

Suborbital  depth/HL 

111 

4.1-7.8 

6.0  ±0.8 

121 

4.3-8.1 

6.2  ±0.7 

X 

Upper  jaw  length/HL 

111 

43.6-51.1 

47.4  ±1.5 

121 

42.7-54.0 

47.8  ±2.0 

Lower  jaw  length/HL 

105 

53.4-60.5 

56.5  ±1.7 

108 

52.8-62.7 

58.3  ±2.2 

X 

Gill  raker  length/HL 

98 

11.3-20.7 

15.0  ±1.6 

110 

11.6-19.9 

15.5  ±1.7 

ns 

X 

Depth  at  pelvic-fin  base/SL 

111 

32.5-42.7 

37.0  ±1.9 

121 

29.2-40.9 

36.2  ±1.7 

Depth  at  anal-fin  origin/SL 

111 

27.4-35.8 

31.2  ±1.7 

121 

26.7-35.4 

30.5  ±1.6 

Depth  at  anal-fin  insertion/SL 

109 

13.8-18.5 

15.9  ±1.0 

116 

13.1-18.0 

15.4  ±0.9 

Body  thickness/SL 

111 

14.9-22.3 

18.0  ±1.2 

120 

12.9-20.9 

17.3  ±1.5 

X 

Pectoral-fin  base  width/SL 

111 

9.5-11.9 

10.7  ±0.5 

121 

9.4-11.2 

10.2  ±0.4 

X 

X 

Pectoral-fin  ray  length/SL 

111 

24.6-31.8 

28.5  ±1.3 

121 

23.5-31.0 

28.2  ±1.4 

ns 

ns 

Pectoral-fin  length/SL 

111 

25.5-33.6 

29.8  ±1.5 

117 

24.2-35.1 

29.2  ±1.7 

Pelvic-fin  ray  length/SL 

111 

20.5-26.0 

22.7  ±1.1 

119 

19.2-29.2 

22.0  ±1.5 

Pelvic-fin  ray/Pelvic-fin  spine 

111 

52.4-67.4 

59.9  ±4.3 

121 

44.9-70.7 

59.9  ±5.5 

length 

Anal-fin  spine  I  length/SL 

83 

3.6-9.1 

5.2  ±0.9 

107 

3.3-8.8 

5.3  ±1.0 

Anal-fin  spine  II  length/SL 

84 

7.5-14.2 

10.6  ±1.2 

106 

5.8-13.6 

10.5  ±1.5 

Anal-fin  spine  III  length/SL 

84 

10.0-14.6 

12.4  ±1.1 

107 

9.5-15.6 

12.2  ±1.3 

Anal-fin  ray  1  length/SL 

83 

15.6-22.5 

19.0  ±1.2 

97 

15.1-21.0 

18.2  ±1.3 

Anal-fin  ray  2  length/SL 

84 

14.5-23.4 

20.2  ±1.2 

97 

15.3-23.1 

19.5  ±1.3 

Caudal-fin  length/SL 

60 

10.2-29.4 

21.7  ±2.7 

74 

15.4-26.9 

21.2  ±2.3 

Caudal  peduncle  depth/SL 

111 

9.3-13.7 

11.4+0.7 

121 

9.5-12.2 

10.9  ±0.5 

Caudal  peduncle  dorsal 

111 

11.9-16.2 

13.8+0.8 

121 

12.6-16.4 

14.1  ±0.8 

length/SL 

Caudal  peduncle  ventral 

111 

17.6-22.9 

20.5  ±1.0 

121 

17.1-24.3 

21.1  ±1.3 

ns 

ns 

length/SL 

Preanal  length/SL 

111 

60.0-77.4 

68.7  ±2.5 

121 

59.4-77.9 

68.1  ±2.6 

Pelvic-  to  anal-fin  length/SL 

105 

26.1-43.2 

32.5  ±3.1 

111 

25.9-41.7 

31.6  ±2.9 

Predorsal  length/SL 

111 

28.5-35.4 

32.1  ±1.3 

121 

28.1-35.2 

31.8  ±1.3 

X 

ns 

Spinous  dorsal-fin-base 

111 

32.1-43.6 

37.2  ±2.1 

121 

31.6-41.9 

37.0  ±2.1 

ns 

ns 

length/SL 

Soft  dorsal-fin-base  length/SL 

111 

21.9-30.8 

26.0  ±1.6 

121 

21.4-28.7 

24.8  ±1.6 

X 

X 

Anal-fin-base  length/SL 

111 

13.9-18.9 

16.5  ±1.0 

121 

13.1-18.4 

16.3  ±1.0 

Prepelvic-fin  length/SL 

111 

33.5-49.5 

39.7  ±2.6 

121 

34.0-46.5 

39.7  ±2.4 

thickness,  caudal-peduncle  dorsal  length,  and  upper-jaw 
length  (Table  6).  No  significant  regional  variation  was 
observed  within  the  overall  PCA. 

In  the  sheared  PCA  of  differences  in  shape  by  depth, 
both  species  showed  negligible  differences  within  broadly 
overlapping  clusters  of  individuals.  In  the  depth  analysis, 
loadings  along  the  PC2  axis  were  strongest  for  suborbital 
depth,  gill-raker  length,  and  orbit  length,  and  shallower 


individuals  tended  to  have  a  greater  suborbital  depth, 
longer  gill  rakers,  and  longer  orbit.  In  the  combined  mor- 
phometric  and  meristic  shallow-water  analysis,  slight 
differences  along  the  morphometric  PC2  axis  and  the 
meristic  PCI  axis  reflected  longer  gill  rakers  and  higher 
lateral-line  pore  counts  in  S.  variabilis.  No  differences 
were  found  in  the  PCA  comparing  sex  within  species 
(Tables  7-8). 


334 


Fishery  Bulletin  102(2) 


Table  2 

Counts  of  lateral-1 

ne  pores  and  gill 

rakers  for  Sebastes 

eiliatus 

and  S.  variabilis  by 

region. 

AI  oi 

BS 

=  Aleutian  Islands  or 

Bering  Sea;  GOA  = 

Gulf  of  Alaska;  SEAK 

=  Southeast  Alaska 

BC 

=  British  Columbia. 

Species 

Region 

Lateral-1 

ine  pores 

39 

40 

41 

42 

43 

44 

45 

46 

47 

48 

49 

50 

51 

52 

53 

54 

n 

Mean 

SD 

Sebastes  eiliatus 

AI  or  BS 

2 

3 

8 

4 

8 

1 

26 

44.73 

1.66 

GOA 

1 

1 

6 

7 

4 

11 

8 

5 

5 

3 

1 

52 

45.20 

2.22 

SEAK 

3 

3 

2 

2 

4 

2 

5 

3 

24 

46.75 

2.40 

Sebastes  variabilis 

AI  or  BS 

3 

4 

5 

1 

7 

4 

3 

27 

48.04 

1.87 

GOA 

5 

4 

14 

11 

15 

13 

8 

3 

1 

74 

48.68 

1.95 

SEAK 

1 

1 

3 

5 

9 

5 

14 

4 

2 

44 

48.75 

1.94 

BC 

1 

3 

1 

2 

3 

4 

1 

1 

16 

48.63 

2.25 

Region 

Gill  rakers 

32 

33 

34 

35 

36 

37 

n 

Mean 

SD 

Sebastes  eiliatus 

AI  or  BS 

GOA 

SEAK 

1 
2 
1 

9 

17 

1 

5 

23 

6 

10 

7 

13 

1 
2 
3 

1 

26 
51 
25 

34.04 
33.80 
34.77 

1.04 
0.87 
0.92 

Sebastes  variabilis 

AI  or  BS 
GOA 
SEAK 
BC 

3 

1 

1 
5 
4 
1 

10 

24 

18 

2 

8 
35 
15 

8 

2 
5 

8 
6 

1 
1 

25 

70 
46 
17 

34.22 
34.52 
34.61 
35.12 

1.13 
0.80 
0.89 
0.86 

Systematics 


Diagnosis 


Sebastes  eiliatus  (Tilesius,  1813) 
Dark  rockfish 

Figs.  1-4;  Tables  1-8 

Epinephelus  eiliatus  Tilesius,  1813:406,  pi.  16,  figs.  1-4 

(original  description,  one  specimen:  holotype  apparently 

lost,  sex  unknown,  approximately  413  mm  TL,  "Oceano 

orientali  Camtschatcam  et  Americam  alluenti"). 
Sebastichthys  eiliatus:  Jordan  and  Jouy,  1881:8  (in  part, 

new  combination). 
Sebastodes  eiliatus:  Jordan  and  Gilbert,  1883:658  (in  part, 

new  combination). 
Sebastostomus  eiliatus:  Eigenmann  and  Beeson,  1894:388 

(in  part,  new  combination). 
Sebastes  eiliatus:  Westrheim,  1973:1230  (in  part,  new 

combination). 
Sebastes  sp.  cf.  eiliatus:  Orr  et  al„  1998:26,  2000:26. 

Neotype 

UW  43493,  1(266.4  mm),  Lynn  Canal,  north  of  Funter 
Bay,  58.2467°N,  134.899°W,  25  m  depth,  13  July  1998. 

Material  examined 

A  total  of  140  specimens,  83.8-340.0  mm,  were  examined, 
including  the  neotype  above.  See  Appendix  for  catalog 
numbers  and  locality  data. 


A  species  of  Sebastes  with  the  following  combination  of 
character  states:  body  uniformly  black  to  dark  blue  or 
gray,  particularly  at  anal-fin  base  and  ventral  pectoral-fin 
rays;  peritoneum  jet  black;  symphyseal  knob  moderate  to 
strong;  extrinsic  swimbladder  muscle  with  anterior  fascia 
separating  sections  of  striated  muscles,  otherwise  of  type  I 
(a-z)  of  Hallacher  ( 1974 1;  lateral-line  pores  39-50,  lateral- 
line  scales  44-60;  pectoral-fin  (PI)  rays  16-19;  anal-fin  (A) 
rays  7-9;  dorsal-fin  (D)  rays  13-17;  vertebrae  28  (11-12  + 
16-17). 

Description 

D  XII-XIV.  13-17;  A  III,  7-9;  PI  16-19,  8-11  simple;  lat- 
eral-line pores  39-50(54),  scales  44-60;  gill  rakers  32-37 
( 10-11  +  22-27 );  vertebrae  28  (11-12  +  16-17).  Meristic  fre- 
quency and  statistical  data  are  presented  in  Tables  2-4. 

Morphometric  data  and  statistics  are  presented  in 
Tables  1  and  4.  Body  relatively  deep,  especially  at  nape, 
depth  at  pelvic-fin  base  32.5-42.7%  SL;  profile  of  dorsal 
margin  of  head  steep  from  snout  to  nape  above  anterodor- 
sal  margin  of  gill  slit,  flattening  to  dorsal-fin  origin;  mouth 
large,  with  posterior  end  of  maxilla  extending  between 
pupil  and  posterior  rim  of  orbit,  maxilla  length  43.6-51.1% 
HL;  symphyseal  knob  moderate  to  strong  and  having  blunt 
tip,  lower  jaw  length  53.4-60.5%  HL;  mandibular  pores  of 
moderate  size.  Cranial  spines  weak,  in  large  adults  cov- 
ered by  flesh,  head  smooth.  Nasal  spine  invariably  present; 
parietal  ridge  invariably  present  and  small  spine  typically 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescription  of  Sebastes  ciliatus 


335 


Table  3 

Counts  of  soft-dorsal-,  anal-,  and  pectoral-fin  rays  for  Sebastes  ciliatus  and  S.  variabilis  by  region.  AI  or  BS  =  Aleutian  Islands  or 
Bering  Sea;  GOA  =  Gulf  of  Alaska;  SEAK  =  Southeast  Alaska;  BC  =  British  Columbia,  n  =  number  offish  in  sample. 


Species 


Region 


Dorsal-fin  rays 


13 


14 


15 


16 


17 


Mean 


SD 


Sebastes  ciliatus 


Sebastes  variabilis 


AI  or  BS 

GOA 

SEAK 

AI  or  BS 

GOA 

SEAK 

BC 


2 

21 

3 

26 

15.04 

0.45 

4 

45 

2 

52 

14.94 

0.42 

5 

14 

4                1 

24 

15.04 

0.75 

25 

3 

28 

15.12 

0.33 

7 

54 

9 

70 

15.03 

0.49 

4 

40 

3 

47 

14.98 

0.39 

1 

16 

17 

14.94 

0.24 

Pectoral-fin  rays 

Region 


17 


18 


19 


Mean 


SD 


Sebastes  ciliatus 


Sebastes  variabilis 


AI  or  BS 

GOA 

SEAK 

AI  or  BS 

GOA 

SEAK 

BC 


1 

19 

44 

2 

16 

23 

6 

65 

1 

44 

15 

26 

52 
24 

27 
71 
47 
17 


18.19 
18.16 
18.17 
18.17 
17.93 
18.02 
18.12 


0.49 
0.37 
0.56 
0.38 
0.26 
0.25 
0.33 


Region 


Anal-fin  rays 


Mean 


SD 


Sebastes  ciliatus 


Sebastes  variabilis 


AIorBS 

GOA 

SEAK 

AI  or  BS 

GOA 

SEAK 

BC 


4 

21 

4 

47 

1 

20 

4 

19 

8 

63 

5 

41 

1 

15 

26 

7.88 

0.43 

51 

7.92 

0.27 

24 

8.08 

0.41 

27 

8.00 

0.50 

71 

7.88 

0.32 

46 

7.89 

0.31 

17 

8.00 

0.35 

present;  postocular  and  tympanic  spines  absent  or  obso- 
lete in  adults  (postocular  present  on  at  least  one  side  in 
23.2%  and  tympanic  present  on  at  least  one  side  of  37.7% 
of  specimens  examined),  most  often  present  in  juveniles. 
Interorbital  region  wide,  22.9-29.3%  HL,  strongly  convex; 
parietal  ridges  weak,  and  area  between  ridges  slightly 
convex;  preopercular  spines  5,  directed  posteroventrally; 
two  opercular  spines,  upper  spine  directed  posteriorly, 
lower  spine  directed  posteroventrally;  posttemporal  and 
supracleithral  spines  present;  lachrymal  spines  rounded, 
small;  dorsal  margin  of  opercle  nearly  horizontal;  lower 
margin  of  gill  cover  with  small  spines:  posteroventral  tip 
of  subopercle  and  anteroventral  tip  of  interopercle  rugose 
or  with  1-2  small  spines. 

Dorsal-fin  origin  above  anterodorsal  portion  of  gill  slit; 
dorsal  fin  continuous,  gradually  increasing  in  height  to 
spine  IV  and  decreasing  in  height  to  spine  XII;  spine  XIII 
much  larger,  forming  anterior  support  of  soft  dorsal  fin; 


membranes  of  spinous  dorsal  fin  moderately  incised,  less 
so  posteriorly;  soft  dorsal  fin  with  anterior  rays  longest, 
posterior  rays  gradually  shortening.  Anal-fin  spine  II 
shorter  than  III  (7.5-14.2  vs.  10.0-14.6%  SL),  anterior 
rays  longest  on  soft  rayed  portion  of  anal  fin,  posterior 
rays  gradually  shortening,  posterior  margin  perpendicu- 
lar to  body  axis  or  with  slight  posterior  slant,  anterior 
ray  tips  directly  ventral  to  or  forward  of  posterior  tips, 
anterior  tip  of  anal  fin  typically  rounded.  Pectoral  fins 
with  ray  10  longest,  extending  to  or  slightly  anterior  to 
vent,  fin-ray  length  24.6-31.8%  SL,  fin-base  to  ray-tip 
length  24.6-31.8%  SL;  fin-base  width  9.5-11.9%  SL.  Pel- 
vic fins  extend  about  60%  of  distance  from  pelvic-fin  base 
to  anal-fin  origin,  falling  well  short  of  vent,  ray  length 
20.5-26.0%  SL,  spine  length  52.4-67.4%,  ray  length. 
Caudal  fin  shallowly  emarginate,  length  10.2-29.4%  SL. 
Vent  positioned  below  dorsal-fin  spine  10,  1.6-5.8%  SL 
from  anal-fin  origin. 


336 


Fishery  Bulletin  102(2) 


Table  4 

Selected  proportional  morphometries  and  meristics  of  Sebastes  ciliatus  and  S.  variabilis  from  shallow-water  collections.  Mor- 
phometric  data  are  in  percent  standard  length  (SL)  or  head  length  (HL).  X  =  statistically  significant  difference  at  0.05  level,  as 
evaluated  by  ANOVA  and  ANCOVA,  when  appropriate;  ns  =  not  statistically  significant  at  0.05  level. 


S.  ciliatus 


S.  variabilis 


Range 


Mean  ±SD 


Range 


Mean  ±SD      ANOVA     ANCOVA 


Meristics  68 

Dorsal-fin  rays  13-16 

Anal-fin  rays  7-9 

Pectoral-fin  rays  (left)  17-19 

Lateral-line  pores  (left)  40-50 

Gill  rakers  32-37 

Morphometries 

Standard  length  83.8-331.0 

Head  length/SL  30.5-35.5 

Orbit  length/HL  21.5-30.6 

Snout  length/HL  18.2-26.0 

Interorbital  width/HL  22.9-28.2 
Suborbital  depth/HL  4.1-7.6 

Upper  jaw  length/HL  43.6-50.4 

Lower  jaw  length/HL  53.4-60.5 

Gill  raker  length/HL  11.8-20.7 

Depth  at  pelvic-fin  base/SL  34.0-40.8 

Depth  at  anal-fin  origin/SL  27.4-35.8 

Body  thickness/SL  14.9-21.3 
Pectoral-fin  base  width/SL  9.7-11.8 

Pectoral-fin  ray  length/SL  25.1-31.3 

Caudal  peduncle  depth/SL  9.8-13.7 

Caudal  peduncle  dorsal  length/SL  11.9-16.2 

Caudal  peduncle  ventral  length/SL  18.2-22.5 

Preanal  length/SL  62.5-77.4 

Predorsal  length/SL  29.3-35.4 

Spinous  dorsal-fin-base  length/SL  32.5-41.8 

Soft  dorsal-fin-base  length/SL  22.1-29.6 

Anal-fin-base  length/SL  14.5-18.9 

Prepelvic-fin  length/SL  36.4-49.5 


49 


15.0+0.5 
7.9  ±0.4 

18.1  ±0.5 

45.2  ±2.1 

34.3  ±1.1 


32.9  ±1.1 
25.6  ±2.1 
21.3  ±1.6 

25.6  ±1.2 
5.9  ±0.7 

47.3  ±1.5 

56.4  ±1.6 
15.2  ±1.5 

36.7  ±1.5 

31.2  ±1.7 
17.9  ±1.2 
10.6  ±0.5 
28.6  ±1.2 

11.5  ±0.7 
14.0  ±0.8 

20.6  ±1.0 

68.3  ±2.3 
32.2  ±1.2 
37.0  ±2.0 
26.2  ±1.4 
16.5  ±0.9 

39.4  ±2.3 


14-16 

15.0  ±0.4 

ns 

7-9 

8.0  ±0.3 

ns 

17-19 

18.0  ±0.2 

ns 

42-51 

47.9  ±2.0 

X 

32-36 

34.6+1.0 

ns 

83.3-363.0 

30.2-35.4 

32.7  ±1.0 

ns 

23.6-30.0 

26.4  ±1.6 

ns 

18.2-24.0 

20.9  ±1.4 

ns 

23.0-28.8 

25.8+1.1 

ns 

4.5-7.4 

6.0+0.6 

ns 

43.9-54.0 

47.3  ±2.0 

ns 

52.8-60.4 

56.6+1.8 

ns 

13.3-19.9 

16.1  ±1.6 

X 

33.7-38.6 

36.2+1.2 

ns 

27.0-31.7 

30.0  ±1.1 

ns 

12.9-20.9 

17.1  ±1.8 

X 

9.4-11.1 

10.2  ±0.4 

X 

24.8-31.0 

28.4  ±1.2 

ns 

9.7-11.8 

10.8  ±0.5 

ns 

12.6-15.3 

13.9  ±0.7 

ns 

17.1-24.3 

21.3  ±1.5 

X 

64.6-73.5 

68.2  ±2.1 

ns 

28.6-35.2 

31.8  ±1.4 

ns 

33.2-41.6 

36.6  ±1.8 

ns 

21.4-28.7 

25.0  ±1.6 

X 

13.6-18.4 

16.4  ±1.0 

ns 

36.3-46.5 

40.3  ±2.6 

ns 

ns 

ns 

ns 

ns 

ns 

X 

X 

ns 

ns 

ns 

ns 

ns 

ns 

ns 

X 

ns 

ns 

ns 

X 

ns 

ns 


Table  5 

Factor  loadings  for 

principal  component  (PC) 

analysis  of 

meristic  characters 

for  Sebastes  ciliatus  and  S 

variabilis 

collected  in  all  depths  and  regions. 

PCI 

PC2 

PC3 

Lateral-line  pores 

-0.6139 

-0.1000 

-0.4151 

Gill  raker 

-0.5993 

0.1360 

0.5217 

Dorsal-fin  rays 

-0.2731 

-0.5625 

0.4739 

Anal-fin  rays 

-0.0138 

-0.7389 

-0.3954 

Pectoral-fin  rays 

0.4349 

-0.3304 

0.4180 

Lateral  body  scales  with  many  (ca.  5-7)  accessory  scales 
in  posterior  field.  Maxilla  and  underside  of  mandible  com- 
pletely scaled;  suborbital  region  scaled;  branchiostegal 
rays  scaled. 

Gill  rakers  long,  ll.3-20.77r  HL.  and  slender  on  first 
arch,  longest  raker  in  joint  between  cerato-  and  hypobran- 
chials,  length  of  preceding  and  succeeding  rakers  progres- 
sively shorter;  rudiments  absent.  Pseudobranchs  37-38. 

Body  color  in  life  and  after  preservation  dark,  black  to 
gray,  lighter  in  deeper  water,  lightening  ventrally  on  belly 
and  occasional  reddening  from  base  of  pectoral  fin  to  base 
of  anal  fin,  uniformly  dark  from  soft  dorsal  fin  to  anal-fin 
base;  vague  darker  mottling  tapering  from  origin  of  soft 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescnption  of  Sebastes  ciliatus 


337 


0.34    ■ 
0.32    • 
0.30    - 
0.28   - 
0.26   - 
0.24    - 
0.22    - 
0.20   - 

A 

CO 

O 

Q_ 
C/3 

i 1 r 1 1                             i 

0.45 


0.5 


0.55 


0.6 
SPC2 


065 


0.7 


0.75 


O 

Q. 
W 


4    -| 

3    ■ 

2 

1 

0 
-1 
-2 
-3 
-4 
-5 
-3 


4 

3 

2 

1 

0   ■ 
-1 
-2 
-3 
-4 


-2.5 


-1.5 


-0.5 


0.5 
PC1 


1.5 


2.5 


3.5 


4.5 


0.45  0.5  0.55  0.6  0.65 

PC1  (meristic) 


0.7 


0.75 


Figure  3 

Plots  of  sheared  ( SPC )  and  standard  principal  component  <  PC )  scores  for 
morphometric  and  meristic  characters  for  Sebastes  ciliatus  (diamonds) 
and  S.  variabilis  (triangles).  (A)  Morphometric  characters  only,  (Bi 
meristic  characters  only,  and  ( C )  morphometric  ( SPC2 )  versus  meristic 
characters  (PCI). 


dorsal  fin  ventrally  and  forward  narrowing  across  lateral 
line,  faint  darker  mottling  also  present  farther  posterior 
at  soft  dorsal-fin  base.  Head  nearly  uniformly  dark,  two 
faint  bars  extending  from  orbit  to  preopercle,  a  faint  bar 
along  anterior  margin  of  maxilla.  Median  fins  uniformly 
dark  gray  to  black.  Pectoral  fins,  including  lower  rays,  gray 
to  black.  Pelvic  fins  dark.  Peritoneum  jet  black;  stomach, 
pyloric  caeca,  and  intestines  pale.  See  Figure  1  (A-C,  E), 
and  previously  published  color  figures  of  Kessler  (1985. 
"S.  ciliatus,  dark  dusky  rockfish"),  Kramer  and  O'Connell 
(1986;  "S.  ciliatus,  dark"),  Kramer  and  O'Connell  (1988, 
1995;  "S.  ciliatus,  Kodiak  specimen,"  and  "shallow  water 
specimen"),  Orr  et  al.  (1998,  2000;  "S.  sp.  cf.  ciliatus,  dark 
dusky  rockfish"),  Orr  and  Reuter  (2002;  "S.  ciliatus,  dark 
dusky"),  and  Mecklenburg  et  al.  (2002;  "S.  ciliatus,  dark 


phase").  Juveniles  in  life  (Fig.  1C)  similar  to  adults  in 
general  body  color,  often  brassy  on  breast  and  head. 

Largest  specimen  examined  340.0  mm  (425  mm  TL, 
412  mm  fork  length;  UW  46068).  Maximum  size  confirmed 
470  mm  fork  length  (RACE  Division6;  Orr,  personal  observ.). 

Distribution  and  natural  history 

The  range  of  Sebastes  ciliatus  based  on  material  exam- 
ined extends  from  the  western  Aleutian  Islands  and  east- 


6  RACE  (Resource  Assessment  and  Conservation  Engineering) 
Division.  2002.  Unpubl.  data  from  RACE  database.  Alaska 
Fisheries  Science  Center,  Natl.  Mar.  Fish.  Serv.,  NOAA,  7600 
Sand  Point  Way  NE,  Seattle,  WA  98115. 


338 


Fishery  Bulletin  102(2) 


Table  6 

Factor  loadings  for  shea 

■ed  principal  component  (SPC) 

analysis  of  morphometric 

characters  for  specimens  exam- 

ined  across  all  depths  and  regions 

for  Sebastes 

ciliatus 

and  S.  variabilis. 

PC  1 

SPC2 

SPC3 

Head  length 

0.1526 

0.1131 

0.0416 

Orbit  length 

0.1853 

0.2090 

0.0988 

Depth  at  anal-fin  origin 

0.2564 

0.1335 

-0.0136 

Snout  length 

0.2452 

0.1109 

0.0240 

Interorbital  width 

0.5753 

-0.8081 

-0.0485 

Suborbital  depth 

0.1905 

0.1481 

0.0675 

Upper  jaw  length 

0.1837 

0.1394 

0.0713 

Lower  jaw  length 

0.2332 

0.2321 

-0.9231 

Gill  raker  length 

0.1635 

0.0768 

0.0840 

Body  thickness 

0.2045 

0.1692 

0.1404 

Pectoral-fin  base 

0.2119 

0.1146 

0.1139 

Pectoral-fin  ray  length 

0.1551 

0.0848 

0.0105 

Caudal  peduncle  depth 

0.2113 

0.1436 

0.1757 

Caudal  peduncle 

0.1885 

0.1561 

0.0262 

dorsal  length 

Caudal  peduncle 

0.1601 

0.0829 

0.0773 

ventral  length 

Pre-anal-fin  length 

0.1313 

0.0732 

0.0800 

Predorsal-fin  length 

0.1516 

0.1063 

0.0456 

Spinous  dorsal-fin 

0.1549 

0.0976 

0.0504 

base  length 

Soft  dorsal-fin  base 

0.1602 

0.1271 

0.1493 

length 

Anal-fin  base  length 

0.1671 

0.1266 

0.1140 

Table  7 

Factor  loadings  for  sheared  principal  component  (SPCl 
analysis  of  morphometric  characters  for  shallow  water 
Sebastes  ciliatus  and  S.  variabilis. 

PCI 

SPC2 

SPC3 

Head  length 

0.1149 

0.0855 

0.0497 

Orbit  length 

0.1504 

0.1223 

0.0408 

Depth  at  anal-fin  origin 

0.1876 

0.1908 

0.1111 

Snout  length 

0.1586 

0.0886 

0.0687 

Interorbital  width 

0.2465 

0.1636 

-0.0394 

Suborbital  depth 

0.2428 

0.1156 

0.0024 

Upper  jaw  length 

0.5843 

-0.8015 

0.0042 

Lower  jaw  length 

0.1876 

0.1565 

0.0745 

Gill  raker  length 

0.2431 

0.1866 

-0.9189 

Body  thickness 

0.1782 

0.1386 

0.0734 

Pectoral-fin  base 

0.2103 

0.1889 

0.1819 

Pectoral-fin  ray  length 

0.2108 

0.1308 

0.1165 

Caudal  peduncle  depth 

0.1550 

0.1033 

0.0162 

Caudal  peduncle 
dorsal  length 

0.2126 

0.1552 

0.1930 

Caudal  peduncle 

0.1854 

0.1426 

-0.0224 

ventral  length 

Pre-anal-fin  length 

0.1520 

0.1098 

0.0381 

Predorsal-fin  length 

0.1499 

0.1124 

0.0603 

Spinous  dorsal-fin 
base  length 

0.1516 

0.0953 

0.0358 

Soft  dorsal-fin  base 

0.1642 

0.1395 

0.1171 

length 

Anal-fin  base  length 

0.1697 

0.1061 

0.1310 

ern  Bering  Sea,  through  the  Gulf  of  Alaska,  to  southeast 
Alaska.  Other  documented  records  extend  its  range  south 
to  Johnstone  Strait,  British  Columbia  (Peden  and  Wilson, 
1976;  Fig.  4).  It  is  common  throughout  its  range  in  shallow 
rocky  habitats,  and  our  material  was  collected  at  depths 
from  5  to  160  m,  its  total  recorded  depth  range. 

Sebastes  ciliatus  is  commonly  collected  with  S.  melanops 
by  trawl  and  hook-and-line  gear  in  shallow  waters,  where 
S.  ciliatus  is  commercially  fished  as  part  of  the  "black 
rockfish"  fishery  and  has  been  often  misidentified  as  S. 
melanops.  In  deeper  (>100  ml  trawls  in  Aleutian  and  Gulf 
of  Alaska  waters.  S.  ciliatus  is  commonly  found  in  asso- 
ciation with  S.  alutus  (Pacific  ocean  perch),  S.  polyspinis 
(northern  rockfish),  and  S.  variabilis  (dusky  rockfish).  Less 
frequently,  S.  uariegatus  (harlequin  rockfish),  S.  zacentrus 
(sharpchin  rockfish),  and  S.  proriger  (redstripe  rockfish) 
are  also  captured  with  S.  ciliatus.  A  large  (320  mm;  UW 
47417)  S.  ciliatus  was  found  in  the  stomach  of  a  Pacific  cod 
(Gadus  macrocephalus)  collected  in  the  Aleutian  Islands. 

Females  captured  in  summer  (May- July)  trawl  surveys 
are  most  often  ripe  with  eyed  larvae.  Near-term  females 
and  males  were  observed  in  July  in  shallow  waters  off 
southeast  Alaska  in  contrast  to  individuals  of  S.  variabi- 


Table  8 

Factor  loadings  for 

principal 

component  (PC)  analysis  of 

meristic  characters 

for  shallow  water 

Sebastes 

ciliatus 

and  S.  variabilis. 

PCI 

PC2 

Lateral-line  pores 

-0.5933 

0.1677 

Gill  rakers 

-0.4274 

-0.6851 

Dorsal-fin  rays 

-0.5742 

-0.0777 

Anal-fin  rays 

-0.3667 

0.6191 

Pectoral-fin  rays 

0.0325 

-0.3364 

lis,  which  were  all  immature  at  this  time  (Orr,  personal 
observ.). 

Etymology 

The  specific  name  ciliatus  is  derived  from  the  Latin  word 
"cilium"  for  "eyelid"  or  "eyelash"  and  alludes  to  the  numer- 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescnption  of  Sebastes  ciliatus 


339 


Figure  4 

Distribution  of  Sebastes  ciliatus  based  on  material  examined  (open  circles)  and  recent  National  Marine  Fisheries 
Service  survey  data  (closed  circles)  for  the  years  1999  to  2002.  Each  symbol  may  represent  more  than  one  capture. 


ous  accessory  scales  (similar  to  fringing  eyelashes)  that 
are  found  on  the  posterior  field  of  the  larger  scales  in  most 
species  of  Sebastes  (Tilesius,  1813). 

Remarks 

Tilesius  (1813)  based  his  description  of  Epinephelus  cil- 
iatus on  a  single  specimen  collected  in  the  North  Pacific 
"bordering  Kamchatka  and  America,"  probably  during 
the  Krusenstern  expedition  of  1803-06  (Bauchot  et  al., 
1997;  Svetovidov,  1978,  1981;  Pietsch,  1995).  Although  the 
illustration  of  the  specimen  was  published  (Tilesius,  1813; 
Fig.  2B),  the  specimen  itself  has  since  been  lost,  probably 
before  the  transfer  of  the  Kunstkammer  collection  to  the 
Zoological  Museum  of  the  Academy  of  Sciences,  St.  Peters- 
burg ( Svetovidov,  1978, 1981).  Because  S.  ciliatus  may  easily 
be  confused  with  other  dark-colored  Sebastes  and  S.  variabi- 
lis, we  have  herein  designated  UW  43493,  collected  in  Lynn 
Canal  of  southeast  Alaska,  as  the  neotype  of  S.  ciliatus. 

The  illustration  of  the  holotype  of  E.  ciliatus  Tilesius 
(1813)  depicts  a  uniformly  dark  individual  of  Sebastes,  and 
most  of  its  reported  meristics  and  other  characters  are 
consistent  with  both  S.  ciliatus  and  S.  variabilis.  However, 
its  lateral-line  pore  count  is  low  at  43,  and  although  falling 
well  within  the  range  found  in  the  material  examined  of 
S.  ciliatus,  the  count  is  represented  in  only  one  individual 
of  S.  variabilis  examined.  Along  with  its  low  lateral-line 


pore  count,  a  moderate  symphyseal  knob  is  illustrated, 
similar  to  that  of  S.  ciliatus,  excluding  its  identification 
as  S.  melanops,  a  common  and  similarly  colored  Sebastes 
found  within  the  geographic  range  of  S.  ciliatus. 

The  anal-fin  posterior  margin  of  the  specimen  illus- 
trated shows  a  moderate  posterior  slant,  and  tips  of  the 
posteriormost  rays  extend  well  past  those  of  the  anterior 
rays.  Sebastes  ciliatus  may  have  an  anal  fin  with  a  slight 
posterior  slant,  unlike  S.  variabilis,  but  it  is  never  as 
pronounced  as  the  illustration  indicates.  However,  this 
character  is  not  found  in  any  other  dark-colored  species  of 
Sebastes  presently  known  from  the  Aleutian  Islands  and 
northern  Gulf  of  Alaska  west  of  Kodiak  Island.  Sebastes 
entomelas  has  an  anal  fin  with  a  strong  posterior  slant  to 
its  posterior  margin,  but  the  northernmost  record  of  this 
species  is  Kodiak  Island  (Allen  and  Smith,  1988;  Love, 
2002 ;  Mecklenburg  et  al.,  2002 )  where  it  is  rare  ( RACE  Di- 
vision6). Sebastes  entomelas  also  has  a  much  higher  count 
of  lateral-line  pores  (50-60;  Love  et  al.,  2002). 

One  syntype  of  Perca  variabilis  was  sent  by  Martin  H. 
K.  Lichtenstein  (1780-1857),  the  director  of  the  Berlin 
Zoological  Museum  in  1813,  to  Georges  Cuvier  at  the  Mu- 
seum National  d'Histoire  Naturelle  in  Paris,  and  has  been 
preserved  as  MNHN  8670  (Svetovidov,  1981;  Fig.  2 A). 
Although  originally  from  the  collections  of  Carl  Heinrich 
Merck  (1761-1799;  Svetovidov,  1981;  Blanc  and  Hureau, 
1968;  Bauchot  and  Desoutter,  1986)  and  thus  contempora- 


340 


Fishery  Bulletin  102(2) 


neous  with  Tilesius's  material,  this  specimen  probably  did 
not  serve  as  the  example  for  Tilesius's  (1813)  illustration. 
The  illustration  is  of  the  left  side  of  a  whole  fish,  whereas 
MNHN  8670  is  the  dried  skin  and  head  of  the  right  side. 
Counts  and  measurements  taken  from  the  original  descrip- 
tion and  compared  with  the  specimen  indicate  that  it  is 
improbable  that  the  left  side  of  this  individual  was  the 
subject  of  the  illustration.  The  counts  provided  by  Tilesius 
(1813)  include  D  XIII,  (soft  rays  not  given);  A  III,  8;  PI  18; 
lateral-line  pores  43.  MNHN  8670  differs  in  counts  of  anal- 
fin  rays  (9)  and  in  lateral-line  pores  (49).  Although  the  sym- 
physeal  knob  is  reduced  and  the  anal-fin  margin  is  strongly 
slanted  posteriorly  in  the  Tilesius  illustration,  MNHN  8670 
has  a  strong  symphyseal  knob  and  a  perpendicular  anal-fin 
margin  with  a  distinctly  pointed  tip  (Fig.  2,  A  and  B). 

Sebastes  variabilis  (Pallas,  1814) 
Dusky  rockfish 

Figs.  1-3,  5;  Tables  1-8 

Perca  variabilis  Pallas,  1814:241  (original  description, 

three?  specimens;  lectotype  hereby  designated,  MNHN 

8670,  dried  skin,  sex  unknown,  343.7  mm.  "mari  Ameri- 

cam  borealum";  other  syntypes  apparently  lost). 
Sebastes  variabilis:  Cuvier,  in  Cuvier  and  Valenciennes, 

1829:547  (new  combination). 
Sebastichthys  ciliatus:  Jordan  and  Jouy,  1881:8  (in  part, 

new  combination). 
Sebastodes  ciliatus:  Jordan  and  Gilbert,  1883:658  (in  part, 

new  combination). 
Sebastosto/nus  ciliatus:  Eigenmann  and  Beeson,  1894:388 

(in  part,  new  combination). 
Sebastes  ciliatus:  Westrheim,  1973:1230  (in  part,  new 

combination). 

Material  examined 

A  total  of  253  specimens,  48.0-430.8  mm,  including  the 
lectotype  listed  above,  were  examined.  See  Appendix  for 
additional  catalog  numbers  and  locality  data. 

Diagnosis 

A  species  of  Sebastes  with  the  following  combination  of 
character  states:  body  light  yellow  to  greenish  brown  to 
gray,  typically  greenish  brown,  with  orange  flecks  vari- 
ously present  on  sides,  particularly  light  ventrally  above 
anal-fin  base  and  on  ventral  pectoral-fin  rays;  peritoneum 
light  gray  to  jet  black;  symphyseal  knob  strong;  extrin- 
sic swimbladder  muscle  with  a  single  section  of  striated 
muscle,  lacking  anterior  fascia,  otherwise  of  type  I  (a-z) 
of  Hallacher  (1974);  lateral-line  pores  43-54,  lateral-line 
scales  47-63;  pectoral-fin  rays  16-19;  anal-fin  rays  7-9; 
dorsal-fin  rays  13-16;  vertebrae  28-29  (11-12  +  16-18). 

Description 

D  XIII-XIV,  13-16;  A  III,  7-9;  PI  16-19,  7-11  simple; 
lateral-line  pores  43-54,  scales  47-63;  gill  rakers  32-37 


110-11  +  22-26);  vertebrae  28-29  ( 11-12  +  16-18)  (one  of 
ten  specimens  with  27  vertebrae,  with  one  caudal  vertebra 
bearing  two  neural  and  two  haemal  spines).  Meristic  fre- 
quency and  statistical  data  are  presented  in  Tables  2-4. 

Morphometric  data  and  statistics  are  presented  in 
Tables  1  and  4.  Body  relatively  deep,  especially  at  nape, 
depth  at  pelvic-fin  base  29. 2-40. 91  SL:  profile  of  dorsal 
margin  of  head  steep  to  nape  above  anterodorsal  margin 
of  gill  slit,  flattening  to  dorsal-fin  origin.  Mouth  large, 
with  posterior  end  of  maxilla  extending  beyond  pupil  to  or 
beyond  posterior  rim  of  orbit,  maxilla  length  42.7-54.01 
HL;  symphyseal  knob  strong  with  blunt  tip,  lower-jaw 
length  52.8-62.71  HL;  mandibular  pores  of  moderate 
size.  Cranial  spines  weak,  in  large  adults  covered  by  flesh, 
head  smooth.  Nasal  spines  invariably  present;  parietal 
ridge  invariably  present  and  small  spine  typically  pres- 
ent; postocular  and  tympanic  spines  typically  absent  or 
obsolete  in  adults  (weak  postocular  spines  present  on  at 
least  one  side  in  29.71  and  weak  tympanic  spines  present 
on  at  least  one  side  in  50.61  of  specimens  examined)  are 
typically  present  in  juveniles.  Interorbital  region  wide, 
22.5-30.41  HL,  strongly  convex;  parietal  ridges  weak, 
and  area  between  ridges  slightly  convex;  preopercular 
spines  5,  directed  posteroventrally;  two  opercular  spines, 
upper  spine  directed  posteriorly,  lower  spine  directed 
posteroventrally;  posttemporal  and  supracleithral  spines 
present;  lachrymal  spines  rounded,  small;  dorsal  margin 
of  opercle  nearly  horizontal;  lower  margin  of  gill  cover 
with  small  spines:  posteroventral  tip  of  subopercle  and 
anteroventral  tip  of  interopercle  rugose  or  with  1  or  2 
small  spines. 

Dorsal-fin  origin  above  anterodorsal  portion  of  gill  slit; 
dorsal  fin  continuous,  gradually  increasing  in  height  to 
spine  IV  or  V  and  decreasing  in  height  to  spine  XII;  spine 
XIII  much  larger,  forming  anterior  support  of  soft  dorsal 
fin;  membranes  of  spinous  dorsal  fin  moderately  incised, 
less  so  posteriorly;  soft  dorsal  fin  with  anterior  rays  lon- 
gest, posterior  rays  gradually  shortening.  Anal-fin  spine 
II  shorter  than  III  (5.8-13.6  vs.  9.5-15.61  SL),  anterior 
rays  longest  on  soft  rayed  portion  of  anal  fin,  posterior  rays 
gradually  shortening,  posterior  margin  perpendicular 
to  body  axis  or  with  slight  posterior  slant,  anterior  ray 
tips  directly  ventral  to  posterior  tips,  anterior  tip  of  anal 
fin  typically  pointed.  Pectoral  fins  with  ray  10  longest, 
extending  to  or  slightly  anterior  to  vent,  fin-ray  length 
23.5-31.01  SL,  fin  base  to  ray  tip  length  24.2-35.11  SL; 
fin-base  width  9.4-11.21  SL.  Pelvic  fins  extend  about  601 
of  distance  from  pelvic-fin  base  to  anal-fin  origin,  falling 
well  short  of  vent,  ray  length  19.2-29.21  SL,  spine  length 
44.9-70.71  ray  length.  Caudal  fin  slightly  emarginate, 
length  15.4-26.91  SL.  Vent  positioned  below  dorsal-fin 
spine  10,  2.2-7.01  SL  from  anal-fin  origin. 

Lateral  body  scales  with  many  (ca.  5-7)  accessory  scales 
in  posterior  field.  Maxilla  and  underside  of  mandible  com- 
pletely scaled;  suborbital  region  scaled;  branchiostegal 
rays  scaled. 

Gill  rakers  long,  11.6-19.91  HL,  and  slender  on  first 
arch,  longest  raker  in  joint  between  cerato-  and  hypobran- 
chials,  length  of  preceding  and  succeeding  rakers  progres- 
sively shorter;  rudiments  absent.  Pseudobranchs  36-38. 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescription  of  Sebastes  aliatus 


341 


Figure  5 

Distribution  of  Sebastes  variabilis  based  on  materia]  examined  (open  circles!  and  recent  National  Marine  Fisheries 
Service  survey  data  (closed  circles)  for  the  years  1999  to  2002.  Each  symbol  may  represent  more  than  one  capture. 


Body  color  in  life  variable  (Fig.  ID);  adults  typically 
light,  greenish-tan  (Fig.  IB),  often  darker  gray  dorsally 
(Fig.  ID),  rarely  lighter  yellow  overall  (Fig.  ID);  invari- 
ably lightening  ventrally  to  pinkish-white  on  head,  belly, 
anal-fin  base,  and  caudal  peduncle;  a  clear  demarcation 
between  darker  dorsum  and  light  ventrum  above  anal-fin 
base;  vague  darker  mottling  tapering  from  origin  of  soft 
dorsal-fin  ventrally  and  forward  narrowing  across  lateral 
line,  faint  darker  mottling  also  present  farther  posterior  at 
soft  dorsal-fin  base,  mottling  most  evident  in  tan  individu- 
als; brown  to  orange  "flecks"  present  on  sides  of  body  on 
posterior  fields  of  scales,  appearing  as  darker  speckling  in 
juveniles.  Head  similar  in  background  color  to  body,  two 
prominent  bars  extending  from  orbit  to  preopercle,  a  prom- 
inent bar  along  anterior  margin  of  maxilla  in  darker  indi- 
viduals (these  bars  obsolete  in  light  individuals).  Median 
fins  and  pelvic  fins  uniformly  gray,  lighter  in  light-bodied 
individuals.  Pectoral  fins  brown  to  grayish  pink;  lower 
rays  pink.  Peritoneum  light  gray  to  jet  black,  typically 
dark  gray;  stomach,  pyloric  caeca,  and  intestines  pale. 
See  Figure  1  ( A-D)  and  previously  published  color  figures 
of  Kessler  (1985;  "Sebastes  sp.,  light  dusky  rockfish"), 
Kramer  and  O'Connell  (1986;  "S.  ciliatus,  light"),  Kramer 
and  O'Connell  (1988,  1995;  "S.  ciliatus,  light  specimen"), 
Orr  et  al.  (1998,  2000;  "S.  ciliatus,  light  dusky  rockfish"), 
Orr  and  Reuter  (2002;  "light  dusky"),  Mecklenburg  et  al. 
(2002;  "light  phase"). 


Juveniles  in  life  (Fig.  1C)  lighter  than  adults,  with  dor- 
sum light-brown  to  tan,  background  covered  with  orange- 
brown  speckles,  often  with  distinct  dark  band  at  base  of 
soft  dorsal  fin;  head  brassy;  ventrum  pink  on  lower  jaw. 
breast,  and  base  of  anal  fin,  lightening  to  white  on  belly. 

Largest  specimen  examined  430.8  mm  (527.7  mm  fork 
length  [FL],  541.3  mm  TL;  UW  44253).  Maximum  size 
reported  590  mm  FL  (RACE  Division6). 

Distribution  and  natural  history 

Sebastes  variabilis  is  recorded  from  a  single  specimen  off 
Hokkaido,  Japan  (Shinohara  et  al.,  1994),  and  from  other 
specimens  collected  from  the  east  coast  of  Kamchatka 
to  Cape  Ol'utorskii  (at  60°N)  in  the  western  Bering  Sea. 
along  the  Aleutian  Islands  to  60°N  in  the  eastern  Bering 
Sea,  through  the  Gulf  of  Alaska  south  to  Johnstone  Strait, 
British  Columbia  (Peden  and  Wilson,  1976;  Richards  and 
Westrheim,  1988;  Fig.  5),  and  to  central  Oregon  (based  on 
a  recently  collected  single  specimen  [UW  46575] ).  The  ear- 
lier record  of  Schultz  (1936)  and  Alverson  and  Welander 
( 1952 )  from  Washington  at  Neah  Bay  was  reidentified  by 
Westrheim  (1968)  as  S.  entomelas. 

Although  the  depth  of  collection  for  material  examined 
ranges  from  6  to  370  m,  and  the  species  is  recorded  at 
depths  to  675  m,  large  adults  are  commonly  found  along 
the  edge  of  the  continental  shelf  at  depths  of  100-300  m, 


342 


Fishery  Bulletin  102(2) 


where  the  species  is  the  target  of  commercial  fisheries  in 
the  Gulf  of  Alaska.  During  trawl  surveys,  it  is  most  com- 
monly associated  with  S.  alutus  and  S.  polyspinis,  and 
at  greater  depths  with  S.  aleutianus  (rougheye  rockfish) 
throughout  its  range  in  Alaskan  waters  (Reuter,  1999; 
Ackley  and  Heifetz,  2001). 

Females  and  males  captured  during  summer  ( May- July ) 
trawl  surveys  ranged  widely  in  maturity  state.  Occasional 
ripe  females  were  observed,  although  most  females  were 
maturing  (Orr,  personal  observ. ).  A  high  percentage  of 
females  caught  in  trawl  surveys  during  early  April  off 
southeast  Alaska  were  releasing  larvae,  indicating  that 
parturition  occurs  in  the  spring  (Lunsford7).  During  July 
in  shallower  waters  (ca.  40  m)  of  southeast  Alaska,  all  S. 
variabilis  collected  were  immature. 

Etymology 

The  specific  name  variabilis  is  presumed  to  be  a  reference 
by  Pallas  (1814)  to  the  wide  range  of  body  color  in  the 
species. 

Remarks 

Pallas  ( 1814 )  described  Perca  variabilis  from  at  least  three 
specimens  probably  collected  by  Merck  during  the  1786-94 
Billings  expedition  to  the  Russian  Far  East,  including  the 
Aleutian  Islands,  eastern  Bering  Sea,  and  northern  Gulf 
of  Alaska  (Schmidt,  1950;  Svetovidov,  1978,  1981;  Pierce, 
1990).  One  specimen  was  more  completely  described  and 
used  by  him  to  obtain  a  set  of  counts  and  measurements. 
The  other  specimens  were  used  to  describe  variation  in 
the  species,  as  in  the  following  excerpt  translated  by  the 
authors  from  the  Latin  text  of  Pallas  (1814):  "Body  colored 
according  to  life  and  sex,  varied,  sometimes  dark  blue, 
belly  white,  fins  blackish;  female  red  below;  those  older 
wholly  red  or  even  purplish...."  Pallas  (1814)  ultimately 
based  the  name  P.  variabilis  on  the  supposed  variability 
in  color  in  this  species. 

Jordan  and  Evermann  (1898)  examined  an  individual 
from  the  Pallas  collection,  recognized  by  him  as  the  "sum- 
mer variety"  of  P.  variabilis  (ZMB  8145).  They  identified 
this  "summer  variety"  as  Sebastes  aleutianus  Jordan  and 
Evermann  (Jordan,  1884,  1885;  Jordan  and  Evermann, 
1898),  a  species  easily  distinguished  from  both  S.  varia- 
bilis and  S.  ciliatus  by  its  full  complement  of  eight  pairs 
of  strong  cranial  spines.  These  specimens  have  since  been 
lost,  probably  during  the  destruction  of  the  Berlin  Zoo- 
logical Museum  during  World  War  II  (Paepke  and  Fricke, 
1992).  Although  Jordan  and  Gilbert  (1883)  wrote  that 
S.  proriger  was  also  confounded  with  S.  ciliatus,  Jordan 
(1885)  and  Jordan  and  Evermann  (1898)  corrected  this 
statement,  noting  that  only  S.  ciliatus  and  S.  aleutianus 
were  included  within  the  material  described  as  E.  ciliatus 
and  P.  variabilis  by  Tilesius  and  Pallas. 


7  Lunsford,  C.  2002.  Personal  commun.  National  Marine 
Fisheries  Service,  Aukc  Bay  Laboratory,  Alaska  Fisheries 
Science  Center,  11305  Glacier  Highway,  Juneau,  AK  99801- 
8626. 


Although  MNHN  8670  (Fig.  2A)  is  from  the  Pallas  col- 
lection (Svetovidov,  1978,  1981),  it  is  apparently  not  the 
specimen  used  for  the  complete  description.  In  his  original 
account,  Pallas  (1814)  listed  the  following  meristic  data 
(modified  to  standard  notation):  D  XIII,  15;  A  III,  7;  PI  17 
(8  simple).  Although  the  dorsal-fin  ray  count  is  identical 
with  that  of  the  MNHN  8670  specimen,  both  anal-  and 
pectoral-fin  ray  counts  differ.  All  elements  are  well  pre- 
served and  easily  counted. 

Comparisons  of  proportions  are  more  difficult  to  inter- 
pret because  measurements  had  not  been  standardized 
at  the  time  of  the  original  description.  However,  of  those 
measurements  that  can  be  readily  compared,  the  following 
significant  differences  were  found,  providing  additional 
evidence  that  this  individual  was  not  the  specimen  used 
for  the  primary  description:  total  length  (391.6  mm  in 
Pallas  [1814;  "longitudo  majoris  speciminis"]  vs.  414.0  mm 
taken  from  MNHN  8670),  head  length  (101.6  mm  ["capi- 
tis a  summa  maxilla  ad  operculum  angulum"]  vs.  either 
96.6  mm  [standard  head  length]  or  110.2  mm  [head  length 
to  tip  of  lower  jaw]).  The  specimen  used  by  Pallas  for  the 
detailed  meristics  and  morphometries  is  presumed  lost 
(Svetovidov,  1978). 

Ayres  ( 1854)  misidentified  S.  melanops  from  the  vicinity 
of  San  Francisco  Bay  as  S.  variabilis.  Giinther  (I860)  and 
Ayres  (1862,  1863)  placed  S.  variabilis  of  Ayres  into  the 
synonymy  of  S.  melanops. 

Comparisons 

Sebastes  variabilis  is  most  similar  to  S.  ciliatus;  the  latter 
is  distinguished  by  its  uniformly  dark-blue  to  black  color. 
Sebastes  ciliatus  is  invariably  dark  at  the  base  of  the  anal 
fin  and  on  the  lower  pectoral  rays,  areas  of  lighter  color  in 
S.  variabilis  even  in  those  individuals  that  have  an  overall 
dark  body.  The  peritoneum  of  S.  ciliatus  is  always  jet  black, 
unlike  the  usual  gray  peritoneum  of  S.  variabilis,  which 
however  may  often  be  dark  or  occasionally  jet  black.  In 
combination  with  these  color  differences,  a  low  count  of 
39-42  lateral-line  pores  will  distinguish  S.  ciliatus  from 
S.  variabilis,  although  the  total  range  of  counts  overlaps 
considerably. 

The  extrinsic  morphological  features  of  the  swimblad- 
der  of  both  S.  ciliatus  and  S.  variabilis  are  of  type  I  (a-z) 
of  Hallacher  (1974)  in  which  the  anterior  muscle  mass 
originates  from  the  occipital  region  of  the  neurocranium. 
attaches  to  the  pectoral  girdle  near  the  insertion  of  Baude- 
lot's  ligament,  passes  between  the  epineural  and  pleural 
ribs  of  vertebrae  3  and  4,  passes  ventral  to  the  pleural  rib 
of  vertebrae  5,  and  continues  posteriorly  as  three  tendons 
that  insert  on  the  pleural  ribs  of  vertebrae  8,  9,  and  10. 
In  S.  ciliatus  and  not  S.  variabilis,  the  anterior  striated 
muscle  mass  is  separated  into  two  sections  by  a  thin  fascia, 
similar  to  the  condition  reported  in  S.  paucispinis  alone 
among  species  of  Sebastes  (Hallacher,  1974).  The  morpho- 
logical features  of  the  complex  differ  significantly  in  S. 
paucispinis,  however,  in  that  the  striated  muscle  does  not 
attach  to  the  pectoral  girdle  but  bypasses  it  to  insert  by  a 
single  tendon  into  the  posterior  portion  of  the  swimblad- 
der.  Only  five  specimens  each  of  S.  ciliatus  and  S.  varia- 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescription  of  Sebastes  ciliatus 


343 


bills  were  dissected  in  the  present  study  to  examine  these 
muscles.  Additional  material  should  be  examined  to  assess 
the  intraspecific  variability  and  systematic  significance  of 
this  character  complex. 

Typical  habitats  of  these  two  species  also  differ.  Adult 
S.  ciliatus  are  found  in  nearshore  shallow  habitats  at 
maximum  depths  of  160  m  and  are  abundant  in  protected 
coves  on  the  outer  coast  of  Alaska.  Sebastes  variabilis,  in 
contrast,  is  found  along  the  continental  shelf  margin  at 
depths  to  675  m.  However,  adult  S.  variabilis  have  also 
been  collected  in  nearshore  waters  as  shallow  as  40  m  ( UW 
43494).  In  areas  of  sympatry,  such  as  the  inside  waters 
in  Lynn  Canal  of  southeast  Alaska  and  Monashka  Bay  of 
Kodiak  Island,  S.  variabilis  is  found  at  greater  depths  in 
stronger  current,  whereas  S.  ciliatus  is  found,  often  with 
S.  melanops,  among  kelp  (Macrocystis)  on  rocky  ledges 
(Blackburn  and  Orr,  personal  observ. ). 

Other  uniformly  dark  colored  species  of  Sebastes,  such 
as  S.  melanops  and  S.  mystinus,  may  also  be  confused 
with  S.  ciliatus  and  darker  individuals  of  S.  variabilis,  al- 
though both  may  be  distinguished  on  the  basis  of  color  and 
morphological  features.  The  body  of  S.  melanops  is  dark 
bluish-black,  has  black  speckling  on  the  dorsum  and  lat- 
eral surfaces,  and  a  distinctly  white  ventrum  (in  contrast 
with  the  slightly  lighter  ventrum  of  S.  ciliatus  (Fig.  IE]). 
Unlike  S.  ciliatus,  in  which  the  peritoneum  is  jet  black,  S. 
melanops  has  a  white  peritoneum.  In  S.  melanops,  five  or 
six  faint  light  blotches  slightly  larger  than  the  orbit  are 
present  on  the  dorsum  about  midway  between  the  lateral 
line  and  the  dorsal-fin  base.  These  blotches  are  especially 
prominent  underwater,  and  in  Alaska  easily  distinguish 
S.  melanops  from  both  S.  variabilis  and  S.  ciliatus,  which 
lack  blotches  (Lauth8;  see  color  figures  of  Love,  2002,  and 
Stewart  and  Love,  2002).  The  symphyseal  knob  in  S. 
melanops  is  obsolete,  consisting  only  of  a  fleshy  pad  at  the 
tip  of  the  mandible,  unlike  the  distinct  bony  knob  of  S. 
ciliatus  and  S.  variabilis.  Mandibular  pores  of  S.  melanops 
are  obsolete,  as  compared  with  the  larger,  readily  appar- 
ent pores  of  S.  ciliatus  and  S.  variabilis.  Vertebral  counts 
also  differ,  from  28-29  in  S.  ciliatus  and  S.  variabilis  to  26 
in  S.  melanops.  Sebastes  melanops  ranges  from  southern 
California  to  Atka  Island  in  the  Aleutian  Islands  (Meck- 
lenburg et  al.,  2002)  and  the  southern  Bering  Sea,  where 
its  presence  is  documented  by  a  single  recent  collection 
(UW  47037).  Most  previous  reports  from  the  Bering  Sea 
may  be  of  S.  ciliatus. 

Sebastes  mystinus  is  also  similar  to  S.  ciliatus  but  may 
be  distinguished  by  the  four  distinct  dark  bars  across  its 
head  and  nape  contrasting  with  its  general  body  color  of 
light  mottled  bluish-gray.  The  mouth  of  S.  mystinus  is 
smaller  than  that  of  S.  ciliatus,  and  the  maxilla  extends 
only  to  the  middle  of  the  pupil  rather  than  to  the  posterior 
portion  of  the  orbit  as  in  S.  ciliatus.  Like  S.  melanops  and 
most  Pacific  Sebastes,  S.  mystinus  has  26-27  vertebrae, 
compared  to  the  28-29  vertebrae  of  S.  ciliatus  and  S. 


s  Lauth,  R.  R.  1998.  Personal  commun.  Resource  Assess- 
ment and  Conservation  Engineering  Division,  Alaska  Fisheries 
Science  Center,  Natl.  Mar.  Fish.  Serv.,  NOAA,  7600  Sand  Point 
Way  NE,  Seattle,  WA  98115. 


variabilis.  Sebastes  mystinus  has  been  recorded  from  Sitka 
Harbor,  Alaska  (Kramer  and  O'Connell,  1995),  to  Punta 
Santo  Tomas,  northern  Baja  California  (Hobson,  2002). 
Earlier  reports  from  Kodiak  Island,  the  Aleutian  Islands, 
and  Bering  Sea  are  undocumented  ( Quast  and  Hall.  1972 ; 
Kramer  and  O'Connell,  1995;  Mecklenburg  et  al.,  2002) 
and  probably  refer  to  S.  ciliatus. 

Sebastes  polyspmis  is  commonly  caught  in  trawls  and 
may  be  confused  with  S.  variabilis,  especially  when  pre- 
served. It  can  be  distinguished  from  S.  variabilis  by  its 
modal  count  of  14  dorsal-fin  spines  and  light  (pink  or  white 
when  live)  oblique  band  across  the  lower  rays  of  the  pecto- 
ral fin,  which  remains  prominent  when  recently  preserved. 
In  life,  the  overall  color  of  S.  polyspinis  is  reddish-orange 
to  pink,  overlaid  with  gray-green  mottling  and  fine  green 
spots.  Evermann  and  Goldsborough  (1907)  considered  the 
then  undescribed  S.  polyspinis  within  the  range  of  varia- 
tion of  "S.  ciliatus"  because  at  least  one  lot  (USNM  6243) 
was  misidentified  by  them  as  "S.  ciliatus."  They  probably 
also  confused  S.  melanops  with  S.  variabilis,  or  possibly  S. 
polyspinis,  describing  the  color  in  life  of  S.  melanops  from 
Alaska  as  "olive-brown,  blotched  with  dirty  red."  Sebastes 
melanops  never  has  a  trace  of  red,  whereas  the  most  com- 
mon color  pattern  of  S.  variabilis  could  be  adequately 
described  by  this  phrase. 

In  the  western  Pacific,  two  species,  the  dark-colored  S. 
taczanowskii  and  the  light-colored  S.  schlegelll,  may  be 
confused  with  S.  ciliatus  and  S.  variabilis,  respectively. 
Both  may  be  distinguished  from  S.  ciliatus  and  S.  varia- 
bilis by  modal  counts  of  pectoral-fin  rays  (15  in  S.  tacza- 
nowskii and  17  in  S.  schlegelii  vs.  18  in  both  S.  ciliatus  and 
S.  variabilis)  and  vertebrae  (26  in  both  S.  taczanowskii 
and  S.  schlegelii  vs.  28-29  in  S.  ciliatus  and  S.  variabilis). 
Sebastes  schlegelii  may  also  be  distinguished  by  its  typi- 
cal dorsal-fin  spine  count  of  12  (vs.  13  in  S.  ciliatus  and  S. 
variabilis). 

Implications  for  fisheries  management 

The  dusky  rockfish  (S.  variabilis)  and  the  dark  rockfish 
(S.  ciliatus)  have  been  subjected  to  two  distinct  fisheries 
separately  managed  by  U.S.  federal  and  Alaska  state 
agencies:  S.  variabilis  is  captured  in  the  offshore  trawl 
fishery;  S.  ciliatus,  in  the  nearshore  jig  fishery.  Although 
the  offshore  fisheries  for  dusky  rockfish  only  inciden- 
tally catch  the  dark  rockfish  and  are  managed  for  dusky 
rockfish,  the  nearshore  fishery  is  not  managed  for  dark 
rockfish,  and  instead  the  species  has  been  routinely  mis- 
identified  as  black  rockfish  (S.  melanops).  Sebastes  ciliatus 
has  been  found  to  comprise  up  to  25<7<r  of  the  catch  in  the 
"black  rockfish"  jig  fishery  of  the  northern  Gulf  of  Alaska 
(Clausen  et  al.1). 

Several  differences  in  biologically  significant  param- 
eters were  evident  from  specimens  examined  in  our  study 
and  from  observations  in  survey  data.  The  two  species 
are  typically  found  in  different  habitats,  attain  different 
maximum  sizes,  and  show  differences  in  reproductive 
seasonality.  Recognizing  the  two  as  distinct  species  is  the 
first  step  towards  establishing  a  biologically  based,  spe- 
cies-specific management  scheme. 


344 


Fishery  Bulletin  102(2) 


Acknowledgments 

We  appreciate  the  early  efforts  of  J.  Geil  and  J.  Westrheim 
to  encourage  this  research  through  spirited  discussion  and 
by  providing  unpublished  data,  courtesy  of  Fisheries  and 
Oceans  Canada,  Pacific  Biological  Station.  We  thank  D. 
Clausen,  C.  Lunsford,  T.  W.  Pietsch,  D.  E.  Stevenson,  and 
M.  E.  Wilkins  for  overall  reviews  and  K.  Mier  for  a  statisti- 
cal review.  The  care  and  transportation  of  specimens  from 
Alaska  is  not  a  small  task  and  we  thank  the  many  collec- 
tors who  made  this  project  possible:  fisheries  observers  W. 
Benecki,  M.  Brown,  K.  Barber,  T.  Droz,  J.  Peeples,  and  J. 
Wiersema;  personnel  of  the  AFSC's  RACE  Division  W.  C. 
Flerx,  R.  Harrison,  R.  R.  Lauth,  G.  R.  Hoff,  M.  Martin,  N. 
W.  Raring,  J.  S.  Stark,  H.  Zenger,  and  M.  Zimmermann. 
For  support  of  nearshore  field  work  in  southeast  Alaska, 
we  thank  J.  Fujioka,  D.  Clausen,  and  B.  Wing.  The  Uni- 
versity of  Washington  Fish  Collection  archived  most  of 
the  fresh  material  examined  for  this  study  and  we  appreci- 
ate the  support  of  the  curator  (T.  W.  Pietsch)  and  collec- 
tions managers  (K.  E.  Pearson  and  B.  W.  Urbain).  We  also 
thank  the  following  curators  and  collections  managers 
and  their  institutions  for  hosting  visits  for  examination 
of  material  or  for  the  loan  of  specimens:  B.  Wing  ( ABL). 
K.  Sendall  (BCPM),  W.  Eschmeyer,  D.  Catania  (CAS), 
B.  Sheiko  (KIE).  J.-C.  Hureau,  G.  Duhamel,  P.  Pruvost 
(MNHN),  D.  F.  Markle  (OSU),  and  G.  D.  Johnson.  S. 
Jewett,  and  S.  J.  Raredon  (USNM). 


Literature  cited 

Ackley.  D.  R..  and  J.  Heifetz. 

2001.     Fishing  practices  under  maximum  retainable  bycatch 
rates  in  Alaska's  groundfish  fisheries.     Alaska  Fish.  Res. 
Bull.  8(11:22-44. 
Allen.  M.  J.,  and  G.  B.  Smith. 

1988.     Atlas  and  zoogeography  of  common  fishes  in  the 
Bering  Sea  and  northeastern  Pacific.     NOAA  Tech.  Rep. 
NMFS66,  151  p. 
Alverson.  D.  L..  and  A.  D.  Welander. 

1952.     Notes  on  the  scorpaenid  fishes  of  Washington  and 
adjacent  areas,  with  a  key  for  their  identification.     Copeia 
1952:138-143. 
Ayres,  W.  O. 

1854.  [Descriptions  of  new  species  of  California  fishes.] 
Proc.  Calif.  Acad.  Sci.  1:1-22. 

1862.  (Descriptions  of  fishes  believed  to  be  new.)  Proc. 
Calif.  Acad.  Sci.  2:209-218. 

1863.  Notes  on  the  sebastoid  fishes  occurring  on  the  coast  of 
( 'alifornia,  U.S.A.     Proc.  Zool.  Soc.  London  26:390-402. 

Barsukov,  V.  V. 

1964.  Key  to  the  fishes  of  the  family  Scorpaenidae.  Soviet 
Fisheries  Investigations  in  the  northeast  Pacific.  So- 
viet Fisheries  Investigations  in  the  northeast  Pacific 
53:226-262. 

Bauchot,  M.-L.,  J.  Daget,  and  R.  Bauchot. 

1997.  Ichthyology  in  France  at  the  beginning  of  the  19th 
century:  the  histoire  naturelle  des  poissons  of  Cuvier 
I 1769-1832)  and  Valenciennes  (1794-1865),  p.  27-80.  In 
Collection  building  in  ichthyology  and  herpetology  (T.  W. 
Pietsch  and  W.  D.  Anderson  Jr.,  eds.),  593  p.  Am.  Soc. 
Ichthyol.  Herpetol.  Spec.  Publ.  3. 


Bauchot,  M.-L.,  and  M.  Desoutter. 

1986.  Catalogue  critique  des  types  de  poissons  du  Museum 
national  d'Histoire  naturelle.  Sous-ordre  des  Percoidei 
l families  des  Apogonidae,  Centrachidae,  Centropomidae, 
Dinolestidae.  Glaucosomatidae,  Grammatidae.  Kuhliidae. 
Percidae,  Percichthyidae,  Plesiopidae.  Priacanthidae. 
Pseudochromidae.  Teraponidaei.  Bull.  Mus.  natl.  Hist, 
nat.,  Paris.  4e  ser.,  8,  sect.  A.  4  suppl.:  51-130. 
Bentzen,  P.,  J.  M.  Wright,  L.  T.  Bryden.  M.  Sargent,  and 
K.  C.  T.  Zwanenburg. 

1998.     Tandem  repeat  polymorphism  and  heteroplasmy  in 
the  mitochondrial  control  region  of  redfishes  {Sebastes: 
Scorpaenidae l.     J.  Hered.  89:1-7. 
Blanc,  M.,  and  J.  C.  Hureau. 

1968.     Catalogue  critique  des  types  de  poissons  du  Museum 
national  d'Histoire  naturelle  (poissons  a  joues  cuirassees), 
70  p.     Publ.  Diverces  Mus.  Natl.  Hist.  Nat.  23. 
Bookstein,  F.  L.,  B.  Chernoff,  R.  L.  Elder,  J.  M.  Humphries  Jr., 
G.  R.  Smith,  and  R.  E.  Strauss. 

1985.     Morphometries  in  evolutionary  biology,  177  p.     Acad. 
Nat.  Sci.  Spec.  Publ.  15,  Philadelphia,  PA. 
Cuvier,  G.,  and  A.  Valenciennes. 

1829.     Histoire  naturelle  des  poissons.  Vol.  4,  518  p.     Stras- 
bourg. 
Eigenmann,  C.  H.,  and  C.  H.  Beeson. 

1894.     A  revision  of  the  fishes  of  the  subfamily  Sebastinae 
of  the  Pacific  coast  of  America.     Proc.  U.S.  Nat.  Mus. 
17:375-407. 
Eschmeyer,  W.  N.,  E.  S.  Herald,  and  H.  Hammann 

1983.     A  field  guide  to  Pacific  coast  fishes  of  North  America. 
Houghton  Mifflin  Company.  Boston,  MA,  336  p. 
Evermann,  B.  W..  and  E.  L.  Goldsborough. 

1907.     The  fishes  of  Alaska.     Bull.  Bur.  Fish.  26:219-360. 
Gunther,  A.  C.  G. 

1860.     Catalogue  of  the  fishes  in  the  British  Museum. 
Catalogue  of  the  acanthopterygian  fishes  in  the  collection 
of  the  British  Museum.  Gobiidae  to  Notacanthi.     Cat. 
Fishes  3:1-586. 
Hallacher,  L.  E. 

1974.     The  comparative  morphology  of  extrinsic  swimblad- 
der  musculature  in  the  scorpionfish  genus  Sebastes  (Pisces: 
Scorpaenidae).     Proc.  Cal.  Acad.  Sci.  40:59-86. 
Hobson,  T. 

2002.     Sebastes  mystinus.     In  Rockfishes  of  the  northeast 
Pacific  (M.  Love,  M.  Yoklavich,  and  L.  Thorsteinson,  eds.), 
p.  215-218.     Univ.  California  Press,  Los  Angeles.  CA. 
Hubbs,  C.  L..  and  K.  F.  Lagler. 

1958.     Fishes  of  the  Great  Lakes  region,  revised  ed.     Cran- 
brook  Inst.  Sci.  Bull.  26,  213  p. 
Jordan,  D.  S. 

1883.     Notes  on  American  fishes  preserved  in  the  museums 
at  Berlin.  London,  Paris,  and  Copenhagen.     Proc.  Acad. 
Nat.  Sci.  Phila.  1884:281-293. 
1885.     A  catalogue  of  fishes  known  to  inhabit  the  waters  of 
North  America,  north  of  the  Tropic  of  Cancer,  with  notes  on 
the  species  discovered  in  1883  and  1884.     Rep.  U.S.  Fish 
Coram.  13:789-973. 
1896.     Notes  on  fishes,  little  known  or  new  to  science.     Proc. 
Calif.  Acad.  Sci.  6:201-244. 
Jordan,  D.  S.,  and  B.  W.  Evermann. 

1898.  The  fishes  of  North  and  Middle  America:  a  descrip- 
tive catalogue  of  the  species  offish-like  vertebrates  found 
in  the  waters  of  North  America,  north  of  the  Isthmus 
of  Panama.  Part  II.  Bull.  U.S.  Natl.  Mus.  47:i-xxx, 
1241-2183. 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescnption  of  Sebastes  ciliatus 


345 


Jordan,  D.  S.,  and  C.  H.  Gilbert. 

1881.     Description  of  Sebastichthys  mystinus.     Proc.  U.S. 

Natl.  Mus.  4:70-72. 
1883.     Synopsis  of  the  fishes  of  North  America.     Bull.  U.S. 
Natl.  Mus.  16:i-liv  +  1-1018. 
Jordan,  D.  S.,  and  P.  L.  Jouy. 

1881.     Checklist  of  duplicates  of  fishes  from  the  Pacific  coast 
of  North  America,  distributed  by  the  Smithsonian  Institu- 
tion in  behalf  of  the  U.S.  National  Museum.     Proc.  U.S. 
Natl.  Mus.  4:8. 
Kessler,  D.  W. 

1985.  Alaska's  saltwater  fishes  and  other  sea  life,  358  p. 
Alaska  Northwest  Publishing  Co.,  Anchorage,  AK. 

Kramer,  D.  E.,  and  V.  O'Connell. 

1986.  Guide  to  northeast  Pacific  rockfishes  genera  Sebastes 
and  Sebastolobus.  Alaska  Sea  Grant,  Mar.  Adv.  Bull.  25, 
78  p. 

1988.  Guide  to  northeast  Pacific  rockfishes  genera  Sebastes 
and  Sebastolobus.  [1st  revision.]  Alaska  Sea  Grant.  Mar. 
Adv.  Bull.  25,  78  p. 
1995.  Guide  to  northeast  Pacific  rockfishes  genera  Sebastes 
and  Sebastolobus.  [2nd  revision.]  Alaska  Sea  Grant,  Mar. 
Adv.  Bull.  25,  78  p. 
Leviton,  A.  E.,  and  R.  H.  Gibbs  Jr. 

1988.     Standards  in  herpetology  and  ichthyology.  Standard 
symbolic  codes  for  institutional  resource  collections  in 
herpetology  and  ichthyology.  Suppl.  no.  1:  additions  and 
corrections.     Copeia  1988:280-282. 
Leviton,  A.  E.,  R.  H.  Gibbs  Jr.,  E.  Heal,  and  C.  E.  Dawson. 

1985.     Standards  in  herpetology  and  ichthyology:  part  I.  Stan- 
dard symbolic  codes  for  institutional  resource  collections  in 
herpetology  and  ichthyology.     Copeia  1985:802-832. 
Love,  M.  S. 

2002.     Sebastes  entomelas.     In  Rockfishes  of  the  northeast 
Pacific  (M.  Love,  M.  Yoklavich,  and  L.  Thorsteinson,  eds.), 
p.  172-174.     Univ.  California  Press,  Los  Angeles,  CA. 
Love,  M.  S.,  M.  Yoklavich,  and  L.  Thorsteinson  (eds.). 

2002.     The  rockfishes  of  the  northeast  Pacific,  405  p.     Univ. 
California  Press,  Los  Angeles,  CA. 
Mecklenburg,  C.  W.,  T  A.  Mecklenburg,  and 
L.  K.  Thorsteinson. 
2002.    Fishes  of  Alaska,  1037  p.    Am.Fish.Soc.,Bethesda,MD. 
Orr.  J.  W.,  M.  A.  Brown,  and  D.  C.  Baker. 

1998.  Guide  to  rockfishes  (Scorpaenidae)  of  the  genera 
Sebastes,  Sebastolobus,  and  Adelosebastes  of  the  northeast 
Pacific  Ocean.  NOAA  Tech.  Memo.  NMFS-AFSC-95, 46  p. 
2000.  Guide  to  rockfishes  (Scorpaenidae)  of  the  genera 
Sebastes,  Sebastolobus,  and  Adelosebastes  of  the  northeast 
Pacific  Ocean,  2nd  ed.  NOAA  Tech.  Memo.  NMFS-AFSC- 
117.  48  p. 
Orr,  J.  W„  and  R.  Reuter. 

2002.     Sebastes  ciliatus.    In  Rockfishes  of  the  northeast 
Pacific  (M.  Love,  M.  Yoklavich,  andL.  Thorsteinson,  eds.), 
p.  151-153.     LTniv.  California  Press,  Los  Angeles,  CA. 
Paepke,  H.-J.,  and  R.  Fricke. 

1992.     Critical  catalog  of  the  types  of  the  fish  collection  of  the 
Zoological  Museum  Berlin.  Part  4:  Scorpaeniformes.     Mitt. 
Zool.  Mus.  Berl.  68:267-293. 
Pallas,  P.  S. 

1814.  Zoographia  Rosso-Asiatica,  sistens  omnium  anima- 
lium  in  extenso  Imperio  Rossico  et  adjacentibus  maribus 
observatorum  recensionem.  domocilia,  mores  et  descriptio- 
nes,  anatomen  atque  iconem  plurimorum.  Petrop..  Acad. 
Sci.,  vol.  3,  428  p. 
Peden,  A.  E.,  and  D.  E.  Wilson. 

1976.     Distribution  of  intertidal  and  subtidal  fishes  of  north- 


ern British  Columbia  and  southeastern  Alaska.     Syesis 
9:221-248. 
Pierce,  R.  A. 

1990.     Russian  America:  a  biographical  dictionary.     Alaska 
Historical  Commission  Studies  in  History,  No.  132,  555 
p.     Limestone  Press,  Fairbanks,  AK. 
Pietsch.  T  W.  fed.) 

1995.  Historical  portrait  of  the  progress  of  ichthyology, 
from  its  origins  to  our  own  time,  by  Georges  Cuvier,  366 
p.  Edited  and  annotated  by  T.  W.  Pietsch,  translated  from 
the  French  by  A.  J.  Simpson.  Johns  Hopkins  Univ.  Press, 
Baltimore,  MD. 
Poss,  S.,  and  B.  B.  Collette. 

1995.     Second  survey  offish  collections  in  the  LTnited  States 
and  Canada.     Copeia  1995:48-70. 
Quast,  J.  C.  and  E.  L.  Hall. 

1972.     List  of  fishes  of  Alaska  and  adjacent  waters  with 
a  guide  to  some  of  their  literature.     NOAA  Tech.  Rep. 
NMFS  SSRF-658,  47  p. 
Reuter,  R.  F. 

1999.     Describing  dusky  rockfish  (Sebastes  ciliatus)  habitat 
in  the  Gulf  of  Alaska  using  historical  data.  Unpubl.  MS 
thesis,  83  p.     California  State  Univ.,  Hayward,  CA. 
Richards,  L.  J.,  and  S.  J.  Westrheim. 

1988.     Southern  range  extension  of  the  dusky  rockfish. 
Sebastes  ciliatus,  in  British  Columbia.     Can.  Field-Nat. 
102:251-253. 
Roques.  S.,  J.-M.  Sevigny,  and  L.  Bernatchez. 

2001.  Evidence  for  broadscale  introgressive  hybridization 
between  two  redfish  (genus  Sebastes)  in  the  North-west 
Atlantic:  a  rare  marine  example.     Mol.  Ecol.  10:149-165. 

Schmidt,  P.  Y 

1950.  Fishes  of  the  Sea  of  Okhotsk.  Academy  of  Sciences 
of  the  USSR,  Transactions  of  the  Pacific  Committee,  vol. 
6,  Moscow-Leningrad,  392  p.  [In  Russian,  1965  transl. 
by  the  Israel  Prog.  Sci.  Trans.  Available  from  the  National 
Technical  Information  Service,  Springfield,  VA.] 
Schultz.  L.  P. 

1936.     Keys  to  the  fishes  of  Washington,  Oregon  and 
closely  adjoining  regions.     Univ.  Washington  Publ.  Biol. 
2:103-228. 
Seeb.  L. 

1986.     Biochemical  systematics  and  evolution  of  the  scor- 
paenid  genus  Sebastes.     Ph.D.  diss.,  176  p.     Univ.  Wash- 
ington. Seattle,  WA. 
Shinohara.  G.,  M.  Yabe,  and  T.  Honma. 

1994.     Occurrence  of  the  scorpaenid  fish,  Sebastes  ciliatus, 
from  the  Pacific  coast  of  Hokkaido,  Japan.     Bull.  Biogeogr. 
Soc.  Japan  49:61-64. 
Stewart,  E.,  and  M.  Love. 

2002.  Sebastes  melanops.  In  Rockfishes  of  the  northeast 
Pacific  (M.  Love,  M.  Yoklavich,  and  L.  Thorsteinson,  eds.), 
p.  204-207.     Univ.  California  Press.  Los  Angeles,  CA. 

Sundt  R.  C.  and  T.  Johansen. 

1998.     Low  level  of  interspecific  DNA  sequence  variation 

of  the  mitochondrial  16S  rDNA  in  north  Atlantic  redfish 

Sebastes  (Pisces,  Scorpaenidae).     Sarsia  83:449-452. 
Svetovidov,  A.  N. 

1978.     The  types  of  the  fish  species  described  by  P.  S.  Pallas 

in  "Zoographi  Rosso-Asiatica"  (with  a  historical  account  of 

publication  of  this  book).     Nauka,  Leningrad,  35  p.     [In 

Russian.] 
1981.     The  Pallas  fish  collection  and  the  Zoographia 

Rosso-Asiatica:  an  historical  account.     Arch.  Nat.  Hist. 

10:45-64. 


346 


Fishery  Bulletin  102(2) 


Tilesius,  W.  G.  von. 

1813.     Iconum  et  descriptionum  piscium  Camtschaticorum 
continuatio  tertia  tentamen  monographiae  generis  Agoni 
Blochiani  sistens.     Mem.  Acad.  Imp.  Sci.  St.  Petersb. 
4:406-478. 
Tsuyuki.  H.,  E.  Roberts,  and  W.  E.  Vanstone. 

1965.     Comparative  zone  electropherograms  of  muscle 
myogens  and  blood  hemoglobins  of  marine  and  fresh- 
water vertebrates  and  their  application  to  biochemical 
systematics.     J.  Fish.  Res.  Board  Canada  22:203-213. 
Tsuyuki,  H.,  E.  Roberts,  R.  H.  Lowes,  W.  Hadaway,  and 
S.  J.  Westrheim. 

1968.     Contribution  of  protein  electrophoresis  to  rockfish 
iScorpaenidae)  systematics.     J.  Fish.  Res.  Board  Canada 
25:2477-2501. 
Westrheim,  S.  J. 

1968.  First  records  of  three  rockfish  species  (Sebastodes 
aurora,  S.  ciliatus,  and  Sebastolobus  altivelis)  from  waters 
off  British  Columbia.  J.  Fish.  Res.  Board  Canada  25: 
2509-2513. 
1973.  Preliminary  information  on  the  systematics,  distri- 
bution, and  abundance  of  the  dusky  rockfish,  Sebastes 
ciliatus.     J.  Fish.  Res.  Board  Can.  30:1230-1234. 


Appendix 

Material  examined 

Sebastes  ciliatus  Bering  Sea:  UW  45488,  1(329.7  mm), 
56.65°N,  167.8167°W,  100  m  depth,  5  February  2001; 
UW  22474,  1(221.7  mm),  59.25°N,  177.8167°W,  160  m 
depth,  1987.  Aleutian  Islands:  UW  43447,  1(290  mm), 
51.2534°N,  179.2689°E,  165  m  depth,  26  July  1997;  UW 
45588,  2(240-380  rami,  Aleutian  Is.,  FV  Vesteraalen, 
1997;  ABL  69-13,  2(228.8-250.9  mm),  Amchitka  I., 
Constantine  Harbor,  51.4°N,  179.3667°E,  6  September 
1968;  UW  46489  (UK  T002383),  1(279.7  mm),  51.2707°N, 
179.2118°E,  98  m  depth,  26  July  1997;  UW  45498, 
1(318.5  mm),  51.9921°N,  174.1031°W,  93  m  depth,  6  July 
1997;  UW  43039,  4(83.8-206.5  mm),  Amchitka  I.,  Con- 
stantine Harbor,  off  Kirilof,  7  December  1961;  UW  43423, 
2(295.1-315.9  mm),  Amchitka  I.,  51.3833°N,  178.9°E,  88 
m  depth,  20  April  1993;  UW  43436,  12(272-315  mm), 
51.7552°N,  175.6726°E,  94  m  depth,  1  August  1997;  UW 
14416,  1(183  mm),  Amchitka  I.,  Constantine  Harbor, 
5  September  1955;  UW  43458,  5(263.9-288.6  mm), 
51.2707°N,  179.2118°E,  98  m  depth,  26  July  1997;  UW 
4766,  1(189  mm),  Atka  I.,  Atka  village,  18  August  1938; 
UW  46483,  2(230-253  mm),  51.9730°N,  176.0841°E, 
78  m  depth,  20  July  2000;  UW  46484,  2(275-293  mm), 
51.9638°N,  176.0288°E,  64  m  depth,  20  July  2000;  UW 
47417,  1(320  mm),  Aleutian  Is.,  recovered  from  stomach 
of  Gadus  macrocephalus.  Gulf  of  Alaska:  UW  43242, 
6(320-390  mm),  53.7345°N,  165.5425°W,  89  m  depth, 
25  June  1996;  UW  43272,  5(274-325.5  mm),  55.093°N, 
157.8048°W,  76  m  depth,  10  June  1996;  UW  43420, 
1(334.7  mm),  55.9042°N,  157.0751°W,  101  m  depth,  18 
June  1996;  UW  45508,  2(242.4-292.6  mm),  55.1056°N, 
157.9594°W,  78  m  depth,  2  June  1999;  UW  45512, 


1(235.2  mm),  57.3806°N,  154.8009°W,  67  m  depth,  7  June 
1999;  UW  45509,  1(298  mm),  54.3215°N,  161.8107°W, 
72  m  depth,  23  May  1999;  UW  47412.  1(323.7  mm), 
52.8953°N,  168.297°W,  99  m  depth,  21  May  2001;  UW 
46488,  1(270  mm),  56.3686°N,  154.0495°W,  69  m  depth,  24 
June  2001;  UW 46584, 1(295  mm),  55.1056°N,  157.9594°W, 
78  m  depth,  2  June  1999;  UW  46485,  1(320.1  mm), 
55.01862°N,  157.882°W,  76  m  depth,  8  June  2001.  Kodiak 
Island  area:  UW 43059, 1(200.7  mm).  Kodiakl..  Monashka 
Bay,  SW  of  Trenton  Pt.,  57.8367°N,  152.4°W,  28  July  1977; 
UW  47289,  2(  169-203.7  mm),  Kodiak  I.,  Monashka  Bay,  5 
m  depth,  57.8383°N,  152.4283°W,  5  m  depth,  31  July  1982; 
UW  44035,  2(216.7-274.5  mm),  Kodiak  I.,  Monashka 
Bay,  57.8367°N,  152.4°W,  18  m  depth,  2  August  1982;  UW 
44036,  11(178.2-263.7  mm),  Kodiak  I.,  Monashka  Bay, 
57.8383°N,  152.4283°W,  12  m  depth,  3  August  1982;  UW 
44045,  19(140.6-307.8  mm),  Kodiak  I.,  Monashka  Bay, 
57.8383°N,  152.4283°W,  15  m  depth,  12  August  1982;  UW 
44049,  1(291.6  mm),  Kodiak  I.,  Monashka  Bay,  57.8383°N, 
152.4283°W,  12  m  depth,  11  September  1982;  UW  44050, 
1(331  mm),  NE  of  Kodiak  I.,  Triplets  Is.,  57.98°N,  152.48°W. 
18  m  depth.  1  July  1983;  UW  44051,  1(319.7  mm).  Kodiak 
I.,  Monashka  Bay,  57.8367°N,  152.4°W,  6  m  depth,  14 
October  1982.  Southeast  Alaska:  ABL  60-7,  1(164  mm), 
Kuiu  I.,  Washington  Bay,  ca.  5  mi  W  of  Petersburg.  2  June 
1960;  ABL  62-105,  1(224.5  mm),  Little  Port  Walter,  SE 
tip  of  Baranoff  I.;  ABL  68-295,  1(214.1  mm),  NW  tip  of 
Lincoln  I.,  ca.  5  mi  N  of  Pt.  Retreat,  Lynn  Canal,  7.5  m 
depth,  4  July  1968;  ABL  87-15,  1(  111.8  mm),  Port  Althorp, 
58.1317°N,  136.3333°W,  126-146  m  depth,  18  July  1982; 
UW  43484,  17(202.8-315  mm),  Cross  Sound,  Chicha- 
gof  I.,  Soapstone  Cove,  58.FN,  136.5°W,  25  m  depth,  11 
July  1998;  UW  43485,  19(124.4-264.7  mm).  Lisianski 
Strait,  57.925°N,  136.288°W,  10  m  depth,  12  July  1998; 
UW  43492,  3(153.7-241.1  mm),  Lynn  Canal,  Funter  Bay, 
58.2467°N,  134.8983°W.  25  m  depth,  13  July  1998;  UW 
22426, 2(134.8-169.4  mm),  Alexander  Archipelago,  Biorka 
Channel,  S.  of  Sitka,  33  m  depth,  21  February  1982. 

Sebastes  variabilis  Japan:  HUMZ  125816,  1(334.0  mm), 
Pacific  coast  off  Kushiro,  Hokkaido,  42.6833°N,  144.7667°E, 
200  m  depth,  2  March  1993.  Bering  Sea:  KIE  1277, 
1(320.3  mm),  east  coast  of  Kamchatka  off  Cape  OTutorskii, 
60.4667°N,  171.75°E,  370  m  depth,  18  August  1994;  KIE 
1409,  2(334.5-344.7  mm).  Commander  Is.,  Bering  I., 
55°N,  166.05°E,  90-120  m  depth,  30  April  1996;  KIE 
uncat,  1(324.5  mm),  Karaginskiy  Bay,  1  August  1998;  UW 
43500,  1(314.5  mm),  55.3179°N,  167.5503°W,  147  m  depth. 
3  July  1998;  UW  43499,  1(370  mm),  56.7°N,  163.4°W,  77 
m  depth,  18  June  1998;  UW  43498,  3(330-340  mm). 
54.4763°N,  159.6925°W,  4  June  1996;  UW  40308, 
1(180  mm),  55.36°N,  163.44°W,  84  m  depth.  7  May  1990; 
UW  40311,  1(187  mm),  55.55°N,  163.75°W,  84  m  depth,  3 
May  1990;  UW  44166,  1(339.3  mm),  Alaska,  "2-1-92"  at 
1800  hours;  UW  44182,  2(258.2-266  mm).  FV  Yukon 
Challenger,  haul  105,  8  March  1993;  UW  44253, 
1(430.8  mm),  Aleutian  Is.,  52.3°N,  173.8-174.7°W,  106-113 
mdepth,  lOApril  199LUW 44255, 1(375.7  mm),56.4192°N. 
152.853°W,  182  m  depth,  24  March  1990;  UW  44261, 
1(331.1  mm),  57.3333°N.  151.4167°W,  128  m  depth;  UW 


Orr  and  Blackburn:  Resurrection  of  Sebastes  variabilis  and  redescription  of  Sebastes  ciliatus 


347 


47411,  1(360  mm),  Bering  Sea,  winter  2001,  P.  J.  Sullivan. 
Aleutian  Islands:  UW  43480,  1(370  mm),  51.2522°N, 
179.199°E,  173  m  depth,  20  July  1997;  UW  43460, 
1(340  mm),  Aleutian  Is.,  FV  Dominator,  summer  1997; 
UW  43438,  5(330-370  mm),  51.9252°N,  176.3817°W, 
122  m  depth,  23  July  1997;  UW  43455  (KU  T002038), 
1(275  mm),  51.2707°N,  179.2118°E,  98  m  depth,  26  July 
1997;  UW  45499,  2(327-330  mm),  51.9921°N,  174.1031°W, 
93  m  depth;  UW  45632,  3(83.3-126.1  mm),  Amchitka  I., 
NE  of  Sand  Beach  Cove,  51.5°N,  179°E,  36  m  depth,  20 
August  1971;  UW  45460,  1(98.7  mm),  54.0386°N, 
166.6406°W,  85  m  depth,  22  May  2000;  UW  43441, 
2(380-381  mm),  54.17902°N,  166.3255°W,  240  m  depth, 

12  June  1997;  UW  43442,  1(355  mm),  54.0386°N, 
166.6572°W,  90  m  depth,  13  June  1997;  UW  43445, 
2(315-340  mm).  54.3773°N,  165.608°W,  90  m  depth,  11 
June  1997;  UW  43459  ( KU  T002038),  1(264.2  mm), 
51.2707°N,  179.2118°E,  98  m  depth,  26  July  1997;  UW 
43461,  1(330  mm),  53.6905°N,  167.2648°W,  112  m  depth. 

13  June  1997;  UW  43416,  1(290  mm),  54.796°N. 
163.2772°W,  89  m  depth,  1  June  1996;  UW  43443. 
1(230  mm),  52.8589°N,  172.4586°E,  146  m  depth.  5  August 
1997;  UW  43444,  1(220  mm),  52.8589°N,  172.4586°E,  146 
m  depth,  5  August  1997;  UW  45588,  2(350-350  mm), 
Aleutian  Is.,  FV  Vesteraalen,  summer  1997;  UW  46482, 
4(175-225  mm),  52.8280°N.  168.9904°W,  44  m  depth,  20 
May  2001.  Gulf  of  Alaska:  UW  43204,  7(350-420  mm). 
55.4327°N,  158.9439°W,  155  m  depth.  8  June  1996;  UW 
43200,  6(320-390  mm),  55.2924°N,  156.6652°W,  114  m 
depth,  12  June  1996;  UW  43214,  1(190  mm),  57.3175°N, 
154.8356°W,  82  m  depth,  21  June  1996;  UW  43212, 
1(300  mm),  54.9351°N,  157.4668°W,  153  m  depth,  12  June 
1996;  UW 43201, 8(290-380  mm),  54.1125°N,  161.7306°W, 
111  m  depth,  1  June  1996;  UW  43203,  8(305-405  mm), 
54.1125°N,  161.7306°W,  111  m  depth,  1  June  1996;  UW 
43211,  2(380-410  mm),  54.6806°N,  158.9407°W,  95  m 
depth,  8  June  1996;  UW  43213,  1(345  mm).  53.9849°N, 
163.2663°W,  108  m  depth,  30  May  1996;  UW  43416, 
1(290  mm),  54.7960°N,  163.2772°W,  89  m  depth,  1  June 
1996;  UW  43417,  1(241.4  mm),  55.2898°N,  158.3123°W. 
130  m  depth.  10  June  1996;  UW  43377,  1(356.4  mm), 
55.6441°N,  134.9706°W,  202  m  depth,  26  July  1996;  UW 
45587,  4(  325.3-376.2  mm),  55.4351°N,  156.5458°W,  167  m 
depth,  15  June  1996;  UW  44123,  1(85.3  mm),  57.2128°N. 
152.7898°W,  135  m  depth,  24  October  1997;  KU  T3178. 
1(348.7  mm).  58.8191°N,  140.3303°W,  185  m  depth,  14 
July  1999;  KU  T003215,  1(324.3  mm),  58.8191°N, 
140.3303°W,  185  m  depth,  14  July  1999;  KU  T003216, 
1(373.9  mm).58.8191°N,  140.3303°W,  185mdepth,  14  July 
1999;  USNM  32014,  1(240.1  mm),  Tolstoi  Bay,  October 
1882;  UW  45477,  2(351.5-352.8  mm),  59.1787°N, 
149.1194°W,  157  m  depth,  29  June  1999;  UW  45510, 
4(335-370  mm),  58.9658°N,  148.1749°W,  251  m  depth,  14 
July  1996;  UW  45511,  3(206.2-213.4  mm),  57.3806°N, 
154.8009°W,  67  m  depth,  7  June  1999;  UW  46487, 
1(333.1  mm),  52.8953°N,  168.297°W,  99  m  depth,  21  May 
2001;  UW  43427, 1(350  mm),  54.2758°N,  161.4326°W,  122 
m  depth,  3  June  1996;  UW  43428,  1(315  mm),  55.9042°N, 
157.0751°W,  101  m  depth,  18  June  1996;  UW  43466, 
1(380  mm),  59.4469°N,  140.4849°W,  226  m  depth,  30  July 


1993;  UW  43471,  1(371.7  mm),  58.0895°N,  150.5977°W, 
141  m  depth,  5  August  1993:  UW  43473,  1(342  mm), 
59.4469°N,  140.4849°W,  226  m  depth,  18  July  1996;  UW 
22475,  2(240-270  mm),  54.0167°N,  160.8°W,  170  m  depth, 
4  November  1981;  UW  40912,  1(213.6  mm).  Prince  Wil- 
liam Sound,  60.5658°N,  147.5866°W,  70  m  depth,  2  Octo- 
ber 1989;  UW  43214, 1(  187.9  mm),  57.3175°N,  154.8356°W. 
82  m  depth,  21  June  1996;  UW  43251,  8(320-420  mm), 
59.5045°N,  145.2262°W,  135  m  depth.  17  July  1996. 
Kodiak  Island  area:  ABL  66-890, 1(103  mm),  Marmot  Bay, 
Kodiak  I.,  57.9333°N,  152.1167°W,  1964;  UW  44052, 
1(215.5  mm),  Kodiak;  UW  44032,  1(93.4  mm),  Cook  Inlet, 
Kachemak  Bay,  59.6°N,  151. 3°W,  "<50  m"  depth,  8  Septem- 
ber 1981;  UW  44033,  2(95.7-112  mm),  Cook  Inlet,  Kache- 
mak Bay,  59.6°N,  151. 3°W,  "<50  m"  depth,  10  October 
1981;  UW  47148,  5(95.5-112  mm).  Cook  Inlet,  Kachemak 
Bay,  59.6°N,  151. 3°W.  10  October  1981;  UW  44034, 
1(363  mm),  NE  of  Kodiak  I.,  Triplets  Is.,  57.98°N, 
152.48°W,  20-24  m  depth,  1  July  1982;  UW  44037. 
1(335  mm),  E  of  Kodiak  I.,  57.72-57.87°N.  151.8-152.2°W, 
crab  pot,  18  August  1982;  UW  44038,  2(275.3-310  mm), 
57.975°N,  151.8433°W,  144  m  depth,  19  August  1982;  UW 

44039,  3(266.7-300.4  mm).  Kodiak  I.,  Monashka  Bay, 
57.8367°N,  152. 4°W.  15  m  depth.  31  August  1982;  UW 

44040.  5(269.7-303.7  mm),  Kodiak  I.,  Monashka  Bay, 
57.8367°N,  152.4°W,  20  m  depth,  14  October  1982;  UW 
44041, 1(281.1  mm),  NE  of  Kodiak  I..  Triplets  Is.,  57.98°N. 
152.48°W,  20  m  depth,  2  July  1983;  UW  43381, 1(340  mm), 
Triplets  Is.,  hook  and  line,  57.98°N,  152.48°W,  34  m  depth, 
15  July  1993;  UW  44042,  2(199.5-215.8  mm),  Shelikof 
Strait  off  mouth  of  Uyak  Bay,  57.7°N,  153.92°W,  100  m 
depth,  2  April  1984;  UW  44043,  12(231.3-271.3  mm), 
Shelikof  Strait;  UW  44044,  2(229.6-257.8  mm),  Kodiak  I., 
Monashka  Bay,  jig,  57.8367°N,  152.4°W,  15  m  depth.  12 
August  1982;  UW  44046,  1(302  mm),  E  of  Kodiak  I., 
58.5217°N,  151.3333°W,  154  m  depth,  21  August  1982;  UW 
44047,  1(321.7  mm).  E  of  Kodiak  I.,  58.85°N,  151.8167°W, 
113  m  depth,  21  August  1982;  UW  44048,  1(318.3  mm). 
Kodiak  I.,  Monashka  Bay,  57.8367°N,  152.4°W,  18  m 
depth,  11  September  1982;  UW  46486,  1(278.5  mm), 
56.6941°N.  151.9115°W,  59  m  depth,  27  June  2001;  UW 
47362,  3(122-219  mm),  56.3686°N,  154.0495°W,  69  m 
depth,  24  June  2001.  Southeast  Alaska:  UW  43495, 
1(262  mm),  Lynn  Canal,  N  of  Funter  Bay.  58.03°N, 
134.8967°W,  40  m  depth,  13  July  1998;  ABL  66-156, 
1(269.4  mm),  Barlow  Cove,  19  mi  NW  of  Juneau,  58.3197°N, 
134.8967°W,  12  February  1967;  ABL  68-301, 1(249.8  mm), 
Lynn  Canal,  off  reef  at  N  end  of  Little  I.,  ca.  9  mi  N  of  Pt. 
Retreat,  58.5417°N,  135.0433°W,  3  August  1968;  ABL  69- 
116.  2(77.7-136.3  mm),  Chichagof  I.,  Ogden  Passage 
between  Khaz  Bay  and  Portlock  Harbor,  57.6333°N, 
136.1617°W.  10  September  1969;  ABL  69-122. 
2(82-82.1  mm),  Chichagof  I.,  Icy  Strait  off  SE  end  of 
Pleasant  I.,  58.3333°N,  135.6333°W;  ABL  70-103, 
1(173.2  mm),  Gastineau  Channel,  Marmion  I.,  ca.  9  mi.  S 
of  Juneau,  58.2°N,  134.2533°W,  10  July  1970;  UW  43494, 
11  (150.7-225.2  mm),  Lynn  Canal,  Funter  Bay,  58.2467°N, 
134.8988°W,  37  m  depth,  13  July  1998;  UW  44117, 
26(  153.8-272.7  mm),  Funter  Bay,  58.2467°N,  134.8983°N, 
40  m  depth,  13  July  1998;  UW  44118,  5(151.9-190.0  mm), 


348 


Fishery  Bulletin  102(2) 


Funter  Bay,  58.2467°N,  134.8983°W,  40  m  in  depth,  13 
July  1998;  UW  46526,  12(48.0-93.5  mm),  Sitka  Sound, 
Middle  I.,  57.1°N,  135.45°W,  4  June  2001;  UW  48866, 
3(78.1-156.7  mm),  Alexander  Archipelago,  Biorka  Chan- 
nel, S  of  Sitka,  33  m  depth,  21  February  1982.  British 
Columbia:  BCPM  974-0623-001,  5(342-367  mm),  Hecate 
Strait,  Moresby  Gully,  RV  G.  B.  Reed,  52.31°N,  130.4867°W, 
185-199  m  depth,  14  September  1974;  BCPM  974-416, 
1(154.7  mm),  Dundas  I.,  Brundige  Inlet,  reef  at  entrance 
to  E  arm,  1-5  m  depth;  BCPM  974-419,  1(124.8  mm), 
Dundas  Is.,  Brundige  Inlet,  just  N  of  island  at  entrance  to 
E  arm,  54.6043°N,  130.8612°W,  6-15  m  depth,  19  June 
1974;  BCPM  974-434,  2(108.3-113.8  mm),  mouth  of  Brun- 
dige Inlet  E  shore,  8-12  m  depth;  BCPM  974-447, 
2(  168.9-175.6  mm),  Welcome  Harbour  channel,  54.0225°N, 
130.6133°W,  6-12  m  depth,  2  July  1974;  BCPM  974-467, 
2(127.9-162.1  mm),  off  Parkin  Islets  (E  side),  54.6261°N, 
130.4639°W,  6-8  m  depth,  12  July  1974;  BCPM  974-468, 
1(120  mm),  offNtipofBirniels.,  54.6045°N,  130.4508°W, 
6-12  m  depth,  13  July  1974;  BCPM  974-485,  1(  129.9  mm). 
off  island  in  Griffith  Harbor,  53.6011°N,  130.5486°W,  5-8 
m  depth,  20  July  1974;  BCPM  974-489, 2(145.8-151.5  mm), 
Safa  Islets,  54.7733°N,  130.6067°W,  6-18  m  depth,  28  July 


1974.  Oregon:  UW  46575,  1(355  mm),  44.4°N,  124.783°W. 
265  m  depth,  17  May  2002. 

Significant  comparative  material  examined 

Sebastes  melanops  USNM  342,  syntypes.  Cape  Flattery, 
Washington,  and  Astoria,  Oregon;  UW  47037,  1(350  mm), 
Bering  Sea,  55.2333°N,  164.65°W,  2  February  2002;  UW 
43490,  3(195-230  mm),  Cross  Sound,  Chichagof  I.,  Soap- 
stone  Cove,  58.1°N,  136. 5°W,  25  m  depth,  11  July  1998; 
UW  47288,  2(163-182  mm),  Kodiak  L,  Monashka  Bay, 
5  m  depth,  57.8383°N,  152.4283°W,  5  m  depth,  2  August 
1982. 

Sebastes  mystinus  USNM  27031,  syntype,  1(346  mm), 
Monterey,  California;  USNM  27085,  syntype,  1(269  mm). 
Monterey,  California:  USNM  26971,  syntype,  1(212  mm). 
Monterey,  California. 

Sebastes  polyspinis  USNM  60243,  2(75.3-141.3  mm), 
Alaska,  Chignik  Bay,  13.6  km  S  of  Tuliumnit  Point,  56°N, 
RV  Alabatross,  Sta.  4285,  57-108  m  depth,  10  August 
1903. 


349 


Abstract— The  dynamics  of  the  sur- 
vival of  recruiting  fish  are  analyzed  as 
evolving  random  processes  of  aggrega- 
tion and  mortality.  The  analyses  draw 
on  recent  advances  in  the  physics  of 
complex  networks  and,  in  particular, 
the  scale-free  degree  distribution  aris- 
ing from  growing  random  networks 
with  preferential  attachment  of  links 
to  nodes.  In  this  study  simulations 
were  conducted  in  which  recruiting 
fish  1)  were  subjected  to  mortality  by 
using  alternative  mortality  encounter 
models  and  2)  aggregated  according 
to  random  encounters  (two  schools 
randomly  encountering  one  another 
join  into  a  single  school )  or  preferential 
attachment  (the  probability  of  a  suc- 
cessful aggregation  of  two  schools  is 
proportional  to  the  school  sizes).  The 
simulations  started  from  either  a  "dis- 
aggregated" (all  schools  comprised  a 
single  fish)  or  an  aggregated  initial  con- 
dition. Results  showed  the  transition  of 
the  school-size  distribution  with  pref- 
erential attachment  evolying  toward 
a  scale-free  school  size  distribution, 
whereas  random  attachment  evolved 
toward  an  exponential  distribution. 
Preferential  attachment  strategies 
performed  better  than  random  attach- 
ment strategies  in  terms  of  recruit- 
ment survival  at  time  when  mortal- 
ity encounters  were  weighted  toward 
schools  rather  than  to  individual  fish. 
Mathematical  models  were  developed 
whose  solutions  (either  analytic  or 
numerical)  mimicked  the  simulation 
results.  The  resulting  models  included 
both  Beverton-Holt  and  Ricker-like 
recruitment,  which  predict  recruitment 
as  a  function  of  initial  mean  school  size 
as  well  as  initial  stock  size.  Results 
suggest  that  school-size  distributions 
during  recruitment  may  provide  infor- 
mation on  recruitment  processes.  The 
models  also  provide  a  template  for 
expanding  both  theoretical  and  empiri- 
cal recruitment  research. 


Recruitment  as  an  evolving  random  process 
of  aggregation  and  mortality 


Joseph  E.  Powers 

Southeast  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

75  Virginia  Beach  Drive 

Miami,  FL  33149 

E-mail  address:  loseph  powers@noaa.gov 


Manuscript  approved  for  publication 
10  December  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:349-365  (2004). 


The  study  of  recruitment  processes 
has  traditionally  addressed  mortal- 
ity (predation  and  starvation)  and 
the  effects  of  patchiness  on  mortality 
(Vlymen,  1977;  Beyer  and  Laurence, 
1980;  Hunter,  1984;  Rothschild,  1986); 
hence  the  importance  of  aggregation 
and  mortality  in  recruitment  processes 
of  marine  fish  populations  has  long 
been  noted.  Ecological  processes  of 
starvation,  growth,  and  predation  of 
larval  fish,  coupled  with  oceanographic 
factors  show  the  inherent  variability  in 
these  processes  (Koslow,  1992;  Mertz 
and  Myers,  1994,  1995;  Pepin,  1991; 
Rickman  et  al.,  2000;  Comyns  et  al., 
2003).  In  particular  Rickman  et  al. 
(2000)  have  indicated  the  importance 
of  the  magnitude  of  fecundity  in  the 
variability  of  egg  and  larval  mortal- 
ity. Indeed,  Koslow  ( 1992 )  argued  that 
fecundity  and  the  associated  variability 
in  egg  and  larval  mortality  will  limit 
our  ability  to  determine  stock-recruit- 
ment relationships. 

Stock-recruitment  models  have  gen- 
erally emphasized  the  static  results 
of  recruitment  processes  rather  than 
the  dynamics  themselves.  Indeed,  al- 
though the  classic  stock-recruitment 
models  such  as  the  Beverton-Holt  and 
Ricker  have  been  related  to  microscale 
processes  (Beverton  and  Holt,  1957; 
Ricker,  1958;  Paulik,  1973;  Harris, 
1975),  the  dynamics  at  those  scales 
were  not  explored,  primarily  because 
there  was  not  a  theoretical  basis  for  do- 
ing so  ( Rothschild,  1986 ).  Nevertheless, 
there  is  a  need  to  develop  a  theoretical 
understanding  of  small-scale  inter- 
action processes  during  recruitment, 
particularly  as  they  relate  to  group 
formation. 


Group-formation  ( aggregation  of  fish 
into  schools),  schooling  (shoaling)  be- 
havior, and  the  evolutionary  motivations 
for  formation  of  schools  continue  to  be 
important  research  topics  (Pitcher  and 
Parrish,  1993;  Landa,  1998).  Schooling 
behavior  has  variously  been  attributed 
to  predator-avoidance,  predator-attack 
dilution,  and  hydrodynamic  and  forag- 
ing advantages  (see  Pitcher  and  Par- 
rish, 1993,  for  a  review).  One  of  the  first 
models  for  school  formation  was  that  of 
Anderson  ( 19S1 )  in  which  he  empirically 
observed  skewed  distributions  in  which 
small  schools  were  more  prevalent  than 
larger  ones.  Subsequently,  Bonabeau 
and  Dagorn  (1995),  Gueron  and  Levin 
(1995),  Niwa  (1998),  and  Bonabeau  et 
al.  (1999),  developed  group-size  distri- 
bution models.  In  particular,  Bonabeau 
et  al.  (1999)  in  comparing  group-size 
distributions  of  tunas,  sardinella,  and 
buffalo  suggested  that  power-law  dis- 
tributions may  be  quite  generic.  Niwa 
(1998)  noted  that  Anderson's  original 
model  allowed  for  power-law  distribu- 
tions. Power-laws  are  termed  scale-free 
because  they  exhibit  no  intrinsic  scale. 
Similarly,  existence  of  a  power-law  is 
often  referred  to  as  "scaling." 

Recently,  power-law  distributions 
have  arisen  in  studies  of  the  physics 
of  small-world  and  evolving  networks 
(for  example  the  world  wide  web,  ac- 
tor collaborations,  scientific  citations 
[Barabasi  and  Albert, 1999],  biological 
cellular  networks  [Fell  and  Wagner, 
2000],  and  ecosystem  structure  [Sole 
and  Montoya,  2001]).  In  particular, 
Barabasi  and  Albert  (1999)  demon- 
strated that  a  randomly  evolving  net- 
work would  result  in  a  scale-free  degree 
distribution  if  the  network  is  growing 


350 


Fishery  Bulletin  102(2) 


(the  number  of  nodes  is  increasing)  and  if  the  new  nodes 
were  linked  to  existing  nodes  by  preferential  attachment. 
Preferential  attachment  (or  the  "rich-get-richer"  phenom- 
enon) occurs  when  a  new  node  is  linked  to  an  existing  node 
with  a  probability  proportional  to  the  number  of  links  al- 
ready attached  to  that  node.  More  formally,  the  Barabasi 
and  Albert  model  is  created  by  adding  a  new  node  at  each 
time  step  and  by  randomly  linking  it  to  m  existing  nodes 
proportional  to  the  number  of  links  at  the  existing  nodes. 
After  a  large  number  of  time  steps,  the  probability  of  a  node 
having  k  links  (the  degree  distribution)  scales  as  a  power- 
law  P(k)~k  '-',  where  y  =  3,  independent  of  m.  The  Barabasi 
and  Albert  result  differs  from  the  classic  random  network 
model  of  Erdos  and  Renyi  ( 1960 )  in  which  nodes  are  linked 
randomly  to  existing  nodes,  leading  to  P(k)~exp(-Xk ).  Sub- 
sequent research  has  expanded  on  the  Barabasi  and  Albert 
model  to  examine  aging,  removal  and  rewiring  of  nodes, 
removal  of  links,  fitness  and  attractiveness  of  nodes,  and 
local  modifications  to  preferential  attachment  (see  Albert 
and  Barabasi,  2002,  for  a  review  of  these  developments) 

The  generic  occurrence  of  scale-free  school-size  distribu- 
tions suggest  that  modeling  of  aggregation  and  mortality 
processes  using  the  analogy  of  random  networks  may  be 
fruitful.  The  approach  may  provide  insight  into  recruit- 
ment dynamics  and  a  theoretical  basis  for  further  inves- 
tigation. This  study  attempts  to  do  that  and  is  organized 
in  the  following  manner.  First,  a  simulation  model  of  the 
recruitment  process  is  developed  in  which  aggregation  and 
mortality  occur  based  upon  some  simple  rules  of  prefer- 
ential attachment  and  random  attachment.  Attachment 
rules  are  presented  as  metaphors  for  more  complex  behav- 
iors. Next,  analytical  models  are  created  that  mimic  the 
simulations,  and  results  of  the  simulations  and  analytical 
models  are  compared.  Finally,  the  implications  for  existing 
stock-recruitment  models  and  investigation  of  recruitment 
processes  are  discussed. 


ity  is  the  removal  of  nodes  (fish)  and,  if  there  are  no  more 
fish  in  the  school,  then  the  removal  of  schools.  A  simulation 
model  with  simple  rules  of  mortality  and  aggregation  was 
created  to  examine  the  dynamics  of  these  processes. 

The  simulation  model  followed  individual  fish  and 
schools  through  a  recruitment  period,  i.e.,  the  passage 
of  time  until  an  arbitrary  time  of  recruitment.  During  a 
recruitment  period  fish  and  schools  undergo  encounters 
of  mortality  and  aggregation.  Starting  at  time  r=0  with  S 
fish,  iV,_0  schools  and  kt  t=0  fish  in  school  i  U=l,2,  . . .  ,  N0), 
simulations  were  conducted  by  randomly  generating  an 
encounter  event  (mortality  or  aggregation!.  If  the  event 
was  a  mortality,  then  a  school  was  randomly  selected  by 
using  the  appropriate  mortality  rate  model  (;?*.  discussed 
below).  If  the  size  of  that  school  was  greater  than  one, 
then  that  size  was  reduced  by  one.  If  the  school  size  was 
equal  to  one,  then  the  number  of  schools  was  reduced  by 
one  and  this  school  was  eliminated  from  the  list. 

If  the  event  was  an  aggregation,  then  two  distinct 
schools  were  randomly  selected  by  using  the  appropriate 
aggregation  rate  model  (w,  also  discussed  below).  The  two 
schools  were  combined,  leaving  one  school  whose  size  was 
the  sum  of  the  two  original  ones  and  one  fewer  total  num- 
ber of  schools.  The  probability  of  an  event  being  a  mortality 
was  ml(m+w)  and  the  converse  probability  of  an  aggrega- 
tion was  l-m/(m+w).  Time  increments  of  each  event  were 
computed  using  At=m~l  for  mortality  events  and  (mw)~l 
for  aggregation  events.  Results  at  time  t  were  collated  into 
the  number  offish  surviving  to  time  /  (denoted  by  R,  >,  the 
number  of  schools,  Nr  the  school  size  distribution,  Pt(k), 
and  the  average  school  size,  k r  Note  thati?,  =Ntkr  Simula- 
tions were  run  until  there  were  no  fish  left. 

Encounter  rates  The  encounter  rates,  m  and  w,  were 
based  upon  random  movements  in  statistical  mechanics 
(Tolman,  1979)  in  which  the  encounter  rate  (£/)  of  objects 
of  type  i  with  objects  of  type 7  is  described  by 


Methods 


U=(G,  +  GO  Dp,  {v?  +  v/)"3, 


(1) 


Simulation  of  individuals  in  ecology  and  population 
dynamics  (individual-based  models)  have  become  increas- 
ingly popular  (McCauley  et  al.,  1993).  However,  it  is  often 
difficult  to  understand  the  dynamics  of  large  individu- 
ally based  models  (Pascual  and  Levin,  1999).  Thus,  it  is 
important  to  obtain  models  that  describe  dynamics  of 
groups  that  incorporate  individual  behavior  (Flierl  et 
al.,  1999).  The  models  that  are  developed  here  include  an 
individually  based  model  (simulation  model)  and  an  ana- 
lytical model  that  describes  "mean-field"  dynamics  of  the 
individuals  behavior. 

Simulation  model 

The  recruiting  fish  of  a  year  class  may  be  modeled  as  a 
network  offish  in  which  a  fish  "links"  to  other  fish  to  form 
schools.  (Note  that  in  this  context  it  is  assumed  that  a 
"school"  includes  aggregations  consisting  of  a  single  fish). 
Thus,  the  process  of  aggregation  is  a  process  of  adding 
links  to  nodes  (aggregation  of  schools).  Similarly,  mortal- 


where  G,  =  the  size  of  the  detection  space  at  which  object 
detects  object  type./'; 
Z),  =  the  density  of  objects  of  type  i;  and 
u,    =  the  velocity  (in  three-dimensional  space)  at 
which  object  i  moves  in  the  environment. 

For  these  simulations  the  G  parameters  were  scaled  to 
one  and  the  velocity  parameters  (v's)  were  collapsed  into 
two  encounter  rates:  ,11  for  mortality  encounters  (scaled  to 
unity)  and  a  for  aggregation  encounters. 

Mortality  rate  In  the  simulations,  mortality  of  fish  is 
perpetrated  by  mortality  agents.  If  the  mortality  agents 
randomly  encounter  schools  of  fish,  then  the  probability 
of  a  successful  mortality  (the  removal  of  a  fish  from  the 
system)  is  proportional  to  the  school  size  k.  Under  these 
conditions  Equation  1  reduces  to  Equation  2  with 


G=G, 


■UV?  +V3.  )<={!■■ 


(2) 


m  =  2^ENk, 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


351 


where  E  =  the  density  of  mortality  agents;  and 

k   =  the  encounter  rate  of  fish  with  mortality 
agents. 

Note  that  on  average  Equation  2  reduces  to  m  =  2iiElNlkl= 
2iiEt  Rt  =  -dRIdt.  Hence,  if  the  density  of  mortality  agents 
is  constant  throughout  the  recruitment  period,  then  mor- 
tality is  density  independent  and  mortality  is  proportional 
to  abundance.  An  alternative  interpretation  of  Equation  2 
is  that  the  mortality  agents  randomly  encounter  fish  and 
that  all  encounters  result  in  a  successful  mortality.  The 
mortality  model  (Eq.  2)  will  be  referred  to  as  mdl  (for  den- 
sity-independent). It  is  not  expected  that  mdi  is  the  most 
realistic,  but  rather  it  provides  a  basis  for  comparison. 

A  second  mortality  alternative  is  where  mortality  agents 
randomly  encounter  schools,  whereupon  they  always  per- 
petrate a  successful  mortality:  mN  =  2iiEtNt.  This  model, 
like  mdi,  assumes  that  the  density  of  mortality  agents  are 
constant  throughout  the  recruitment  period. 

For  purposes  of  simulation,  the  density  of  mortality 
agents  at  the  onset  of  the  recruitment  process  was  speci- 
fied to  be  unity  (£0=1).  For  the  two  mortality  models,  mdi 
and  mN,  this  meant  that  £=1  throughout  a  simulation. 

More  realistic  density-dependent  mortality  models  are 
immediately  suggested.  The  first  is  a  density-dependent 
model  in  which  the  ratio  of  mortality-agent  density  to  the 
number  of  schools  remains  constant  throughout  the  re- 
cruitment period,  i.e.,  EtINt  remains  constant  throughout 
the  recruitment  period.  This  leads  to  mdN  =  2f.iNzk,  where 
EtINt  was  set  equal  to  one.  In  this  model  the  ratio  of  mor- 
tality agents  to  schools  is  constant,  agents  and  schools 
randomly  encounter  one  another,  and  the  probability  of  a 
successful  mortality  (given  there  is  an  encounter)  is  pro- 
portional to  the  number  offish  that  are  in  the  school  that 
is  encountered  (mortality  success  is  related  preferentially 
toward  larger  schools). 

A  second  density-dependent  model  is  where  the  mortal- 
ity agent  density  is  proportional  to  the  number  offish  (Etl 
Rt  is  a  constant  set  equal  to  one,  mdR=2iiR2=2iiN2k2).  In 
this  model  the  ratio  of  mortality  agents  to  the  number 
of  fish  in  the  population  is  constant;  agents  and  schools 
randomly  encounter  one  another;  and  the  probability  of  a 
successful  mortality  (given  there  is  an  encounter)  is  pro- 
portional to  the  number  offish  that  are  in  the  school  that 
is  encountered  (mortality  success  related  preferentially 
toward  larger  schools).  Another  interpretation  of  this 
model  is  that  agents  randomly  encounter  fish,  at  which 
time  the  fish  suffers  mortality  at  a  probability  independent 
of  school-size  characteristics. 

A  third  density-dependent  model  depicts  mortality- 
agent  density  proportional  to  school  size  (Etlkt  is  a  con- 
stant set  equal  to  one,  mdk=2\.i  Nk'1).  In  this  model  the 
ratio  of  mortality  agents  to  mean  school  size  is  constant, 
agents  and  schools  randomly  encounter  one  another,  and 
the  probability  of  a  successful  mortality  (given  that  there 
is  an  encounter)  is  proportional  to  the  number  offish  that 
are  in  the  school  that  is  encountered.  Another  interpreta- 
tion of  this  model  is  that  agent  density  is  proportional  to 
the  number  of  schools,  agents  encounter  schools  prefer- 
entially according  to  school  size,  and  the  probability  of  a 


successful  mortality  (given  that  there  is  an  encounter)  is 
proportional  to  the  number  of  fish  that  are  in  the  school 
that  is  encountered. 

Subsequently  it  will  be  shown  that  the  first  density- 
dependent  model  is  related  to  a  Ricker-like  stock-recruit- 
ment model  and  the  second  model  is  exactly  equivalent 
to  a  Beverton-Holt  model.  Definitions  of  the  mortality 
models  are  summarized  in  Table  1.  Note  that  in  the  ini- 
tial applications  of  these  mortality  models,  it  is  assumed 
that  a  mortality  encounter  results  in  mortality  of  one 
fish.  More  detailed  mortality  models  in  which  a  number 
of  fish  greater  than  one  are  removed  by  mortality  may  be 
implemented  in  the  future.  Clearly,  these  would  be  more 
biologically  realistic  in  many  instances.  However,  the 
emphasis  of  this  study  is  on  the  possible  scaling  behavior 
of  school-size  distributions.  Barabasi  and  Albert  (1999) 
showed  that  the  scaling  behavior  of  a  growing  random 
network  is  independent  of  the  number  of  randomly  se- 
lected links  at  each  time  step.  With  this  analogy,  simple 
increases  in  mortality  per  encounter  are  not  expected  to 
alter  the  scaling  behavior  of  the  school-size  distributions. 
Therefore,  the  one-fish-per-mortality-encounter  approach 
was  used  in  these  initial  simulations. 

Aggregation  rate 

Similar  to  mortality-rate  encounters,  aggregations  were 
investigated  as  1)  random  attachment  of  two  unique 
schools  (wN=2aNiN-D)  and  2)  preferential  attachment  of 
two  unique  schools  i  and./'  (w  =2aN(N-\)klkJ;  [Table  1]). 
Note,  the  trivial  alternative  where  there  was  no  attach- 
ment, (a=0),  results  in  equivalence  between  the  mortal- 
ity models  mdN,  and  mdR\  whereas  mdl  becomes  a  simple 
proportional  mortality  rate.  Thus,  results  of  models  with 
o=0  are  uninteresting  in  the  context  of  this  study  and  are 
not  presented. 

Initial  conditions  Each  simulation  was  conducted  with 
one  of  two  alternative  initial  conditions.  The  first  alterna- 
tive was  one  of  complete  disaggregation  in  which  simula- 
tions were  initiated  with  S  fish,  S  schools,  and  one  fish  in 
each  school  (NQ=S,  kQ=l).  The  second  alternative  initial 
condition  was  constructed  from  the  population  dynamics 
of  a  typical  fish  population.  The  main  assertion  of  this 
alternative  is  that  the  eggs  or  larval  fish  produced  by 
one  female  during  spawning  constitutes  one  school  at  the 
onset  of  the  recruitment  process.  Thus,  the  fecundity  per 
female  at  age  is  a  measure  of  initial  school  size  and  the 
abundance  of  females  at  age  is  a  measure  of  the  frequency 
of  schools  of  that  size.  More  precisely,  the  initial  condition 
was  constructed  for  a  population  of  females  greater  than 
five  years  of  age  (age  of  maturity),  where  their  fecundity 
at  age,  F '  is  proportional  to  weight  at  age  determined 
from  a  von  Bertalanffy  growth  equation  with  parameters 
K=0.2  and  L^  =  10,  and  an  allometric  parameter  of  3: 
(F  =1000 [(l-exp(-age( 0.2)))] 3).  Abundance  at  age,  Aage, 
was  computed  with  an  instantaneous  mortality  rate  of  0.2: 
[A  =Zexp(-0.2(age-5))].  The  scalarXwas  obtained  from 
the  approximate  solution  to  S  =2Fage  Aage,  where  F  and  A 
were  integer  values  and  S  was  the  initial  number  of  fish 


352 


Fishery  Bulletin  102(2) 


Table  1 

Summary  of  definitons  of  the  mortality  models  used  in  this  study. 

Model 

Definition 

Mortality  rates': 

mdi  =  2ftNk 
m  ix  =  2fiN2k 
mdR  =  2fiN2kk 
mdk=2ijNkk 

mN  =  2iiN 

density-independent 

density-dependent,  mortality  agents  proportional  to  A'' 
density-dependent,  mortality  agents  proportional  to  R 
density-dependent,  mortality  agents  proportional  to  k 
random  encounters  with  schools 

Aggregation  rates: 

wN  =  2ctN(N-l) 

random  encounters  with  schools 

wpa  =  2aN(N-l)klkJ 

preferential  attachment  of  schools  i  andj 

Initial  conditions: 

Disaggregated 
Aggregated 

N0  =  S,k0=l 

(see  text  and  Table  2) 

Mean  field  equivalents  usee 

in  analytical  model  (see 

text): 

mdi  =  2fiNk              m(/v  = 

2^iNk2 

mJR  =  2fiN2k2 

mM  =  2iiNk2             wpa  = 

2NiN-l)k2 

Key  to  figures  of  simulation 

results 

Figure  1:  disaggregated 

md,              Wpa 

a  =  10-6 

S  =  106 

Figure  2:  disaggregated 

mJ,                 WN 

«  =  io-6 

S=106 

Figure  3:  aggregated 

"hi,                  ™pa 

a  =  IO"6 

S  =  108 

Figure  4:  aggregated 

"IJN               Wpa 

a  =1.5xl0-6 

S  =  2  x  106 

Figure  5:  aggregated 

mdN           wN 

«=  1.5xl0-6 

S  =  2  x  106 

1  In  all  simulations,  ft  was  set  equal  to  1. 

Table  2 

The  aggreg 

ated  in 

itial  school-size 

distribution 

,  when  S  = 

1,000,000 

.  Per  capita  female  fecundity  at 

age  is  a  measure  of  school  size, 

num 

ber  of  female; 

at 

age  is  a  measure 

of  freq 

jency  of  sc 

hools.  See 

text  for  details 

of  computation. 

School  size 

Freq 

.  of  schools 

Freq.  x  size 

School  size 

Freq 

of  schools 

Freq.  x  size 

252 

348 

87,696 

857 

47 

40.279 

341 

284 

96,844 

882 

38 

33.516 

427 

233 

99,491 

903 

31 

27,993 

508 

190 

96,520 

920 

25 

23.000 

581 

156 

90,636 

934 

21 

19,614 

596 

1 

596 

946 

17 

16,082 

646 

128 

82,688 

955 

14 

1.3,370 

703 

104 

73,112 

963 

11 

10,593 

751 

85 

63,835 

970 

9 

8730 

793 

70 

55,510 

975 

7 

6825 

828 

57 

47,196 

979 

6 

5874 

Sum 

of  freq. 

x  size  = 

S  = 

=  1,000.000. 

of  a  simulation.  Then  one  school  of  an  appropriate  magni- 
tude, M,  was  added  such  that  the  M  +^Fll/,vA„.,,,  was  exactly 
equal  to  S.  Note  that  under  this  construction  the  school 
sizes  in  the  distribution  do  not  vary  with  S  (except  for  the 
one  school  of  size  M),  whereas  the  frequency  of  schools  by 
size  do.  An  example  of  the  initial  distribution  with  the  use 
of  this  construction  is  given  in  Table  2. 


Analytical  models 

Analytical  models  of  aggregation  and  recruitment  are 
presented,  where  the  models  are  developed  from  first  prin- 
ciples and  the  parameters  have  an  interpretation  in  the 
physics  and  biology  of  the  recruitment  process.  Hopefully, 
the  nature  of  the  parameters  can  guide  model  selection, 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


353 


and  the  estimates  may  provide  a  theoretical  framework  for 
empirical  research  on  recruitment  processes. 

Noting  that  R,=Ntkt,  the  recruitment  dynamics  depicted 
in  the  simulations  may  be  modeled  by  using  Equations 
3-6  in  which  recruitment  is  dependent  on  the  particular 
mortality  and  aggregation  models  that  are  chosen  (m  and 
w;  Table  1): 

dR,l  =  -m  =  d(N,k,  "/dt=(dkl  Idf'N,  +(dN,  I  dt"k,        (3) 


dNJdt 


-2aNt{Nt-l)\l 


dN,  I  dt  =  -w  -  mxPu 


dk,/dt  =  mft,  [( k,  - 1)1  ( N,  - 1) 
-< m  -  m^Pv  )l  N,+  wkt  /<JV,_i ) 


(4) 


(5) 


dPkl  Idt  =  ( mk+lPk+u  -  mkPKt  )INt-  wkP,!t  I N,       k  >  1 

i-l 

+  wYJP,Pk-,INl  (6) 

=  m2PuINt  - miPua-Pu  >/<  AT,  -1)  k  =  l 

-w1PuINl  , 

where  Ph ,  =  the  proportion  of  schools  with  k  fish  in  them 
at  time  t . 

Also,  mk  and  wk  denote  encounter  rates  appropriate  to 
schools  of  size  k,  whereas  unsubscripted  m  and  w  denote 
mean  field  dynamics  and,  thus,  the  kt  t's  are  replaced  by 
kt's  (see  Table  1). 

The  first  term  in  Equation  4  denotes  the  reduction  in 
number  of  schools  due  to  aggregation  events;  the  second 
term  denotes  a  reduction  due  to  mortality  events  on 
schools  with  one  fish  in  them.  Similarly,  the  first  term  in 
the  mean  school-size  equation  ( Eq.  5 )  describes  the  change 
in  mean  school  size  due  to  mortality  events  on  schools 
of  size  equal  to  one;  the  second  term  is  due  to  mortality 
events  on  schools  of  size  greater  than  one;  and  the  third 
term  is  due  to  aggregation  events.  Finally,  the  first  term 
in  Equation  6  describes  the  change  in  probability  of  school 
size  k  due  to  mortality;  the  second  term  describes  loss  due 
to  aggregation;  and  the  third  describes  gain  due  to  aggre- 
gation. Of  particular  importance  is  Pl  t:  when  Px ,  is  zero, 
the  loss  of  schools  occurs  only  due  to  aggregation.  When 
Pj ,  is  positive,  then  loss  of  schools  is  accelerated  due  to 
mortality  (Eq.  4). 

The  goal  is  to  obtain  solutions  to  Equations  3-6  as 
functions  of  a,  ju,  and  the  initial  conditions.  If  one  can  be 
assured  that  there  will  not  be  a  school  composed  of  one 
fish  during  a  particular  recruitment  period  (Pj  ,=0),  then 
Equation  6  is  eliminated,  the  Pj  ( terms  drop  out  of  Equa- 
tions 4  and  5,  and  a  numerical  or  analytical  solution  to 
the  differential  equations  can  be  obtained,  which  is  com- 
putationally feasible  for  use  in  fitting  to  stock-recruitment 
data.  For  example,  when  there  is  preferential  aggregation 
(w  )  and  mortality  agents  are  proportional  to  schools 
(mdN),  the  equations  reduce  to 


dktldt-- 


ldN 


I N,  +  w    k,( Nj_-1)=  -2/jNtk,  +  2aNtkt3 


Analytical  solutions  were  obtained  for  some  of  the  mor- 
tality and  aggregation  models  when  P1(=0  throughout 
the  recruitment  process  (Appendix  1).  In  particular  for 

m„ 


!Randwpa- 


Rt  =  SHl+2iitS)  (7) 

Nt=NQ+(a/n"[S-S/e%*  %itS"]  (8) 


^=^/[l+2^S  +  2ctfS^]. 


(9) 


which  is  the  Beverton-Holt  stock-recruitment  model 
expanded  to  include  equations  for  the  number  of  schools 
and  the  mean  school  size.  Interestingly,  Equation  9  indi- 
cates that  monitoring  the  school-size  distribution  two  or 
more  times  during  a  recruitment  procession  would  yield 
estimates  of  the  stock-recruitment  parameters  without 
having  direct  measures  of  the  number  of  surviving  fish. 
Equation  7  predicts  recruitment  by  using  one  parameter. 
j.1. ,  the  rate  of  mortality  encounters  during  the  recruitment 
period.  However,  spawning  stock  biomass  is  often  used  as  a 
surrogate  for  the  number  of  initial  stock,  S.  Thus,  another 
parameter  is  needed  to  convert  spawning  stock  biomass 
to  S  in  Equation  7.  In  that  case  the  recruitment  model 
becomes  Rt  =  aS/(l+2utaS),  where  a  is  another  parameter 
related  to  fecundity.  The  additional  parameter  will  be 
needed  for  all  the  models  discussed  here,  if  spawning  stock 
biomass  is  the  measure  of  initial  stock. 

The  assumption  that  Plt=0  for  all  t  of  a  recruitment 
period  may  not  be  justified  in  all  situations.  An  approxi- 
mation was  developed  (Appendix  2)  to  be  applied  when 
the  initial  conditions  are  disaggregated  and  when  there 
is  preferential  attachment.  In  this  circumstance,  the  dif- 
ferential equation  (Eq.  6)  when  k  =  l  is  replaced  by 


dPu/dt  =  -wPu/N,  +  m  ( 1  -  Pu)INt. 


(10) 


Results 


Simulations 

Several  hundred  simulations  were  conducted  under  vari- 
ous initial  stock  sizes  (S),  aggregation  parameters  (o),  ini- 
tial aggregation  conditions,  and  mortality  and  aggregation 
models  (m  and  w).  An  example  set  of  results  are  presented 
in  Figures  1-5  (a  key  to  figures  is  in  Table  1). 

A  typical  example  of  the  evolution  of  the  school-size 
distribution  is  given  in  Figure  1  for  the  disaggregated 
initial  condition,  a=10"6,  S=106,  mortality  model  mdi  and 
aggregation  model  w  .  In  this  example  both  the  mortality 
and  aggregation  models  exhibit  preferential  attachment, 
and  the  school-size  distribution  approaches  scale-free  be- 
havior P(k)~k->,  although  y  evolves  over  time.  Eventually, 
a  so-called  "giant  cluster"  is  formed  by  the  aggregation 
process,  in  which  all  the  fish  attach  to  one  school.  This  has 


354 


Fishery  Bulletin  102(2) 


1 ,000.000 


o  f =  0.1 1 

+  f=0.40 
»  f=0.95 


A  O 


V 

\ 


sw 


luminal  iih  i 


School  size 


-r   1,000,000 


--    5,000,00 


( 

Figure  1 

Simulated  dynamics  of  school-size  distributions  with  mdl  as  the  mortality  model 
and  w  as  the  aggregation  model.  This  simulation  started  with  disaggregated 
initial  conditions  (JVn  =  Sl,  where  S=106.  The  aggregation  parameter  was  oc=10-6. 
The  top  panel  shows  school-size  distributions  (in  log-log  scale)  at  selected  times 
(/).  The  lower  panel  gives  the  mean  school  size  (kbar)  and  school  abundance  (AM 
versus  time. 


been  shown  to  be  an  analog  of  Bose-Einstein  condensation 
(Bianconi  and  Barabasi,  2001;  Dorogovtsev  and  Mendes, 
2002)  and  gelation  (Krapivsky  et  al.,  2000).  Greater  mix- 
ing rates  Cot's)  and  larger  densities  (N's)  accelerate  the 
aggregation  process  and  the  formation  of  the  giant  cluster. 
The  average  size,  k,  increases  over  time  from  the  disag- 
gregated initial  condition  until  a  giant  cluster  is  formed. 
The  number  of  schools  declines  over  time  because  of  both 
aggregation  and  the  mortality  of  fish  in  schools  that  only 
have  one  fish  in  them. 

When  there  is  random  aggregation  beginning  from  a 
disaggregated  initial  condition  (a=10-6,  S=10K,  mdi,  wN  ; 


Fig.  2),  the  school-size  distribution  exhibits  exponential 
behavior  P(k)~exp(-/Jt),  with  A  evolving  over  time.  This 
is  equivalent  to  the  Erdos  and  Renyi  (1960)  results  for 
random  graphs.  A  comparison  of  Figure  2  with  Figure  1 
shows  the  difference  between  preferential  attachment  and 
random  attachment,  i.e.,  the  difference  between  scale-free 
and  exponential  models. 

Aggregated  initial  conditions  (Figs.  3-5)  result  in  a 
transition  from  the  initial  distribution  to  either  scale- 
free  or  exponential  distribution.  During  the  transition, 
the  size  of  the  smallest  school  gradually  becomes  smaller 
until  there  is  a  finite  probability  of  schools  with  one  fish  in 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


355 


1.000.000    - 

o 

+ 

A 

+ 

+• 

o    ( =  0.03 
+    (=0.33 
a   (=0.62 

>. 
o 

c 

B-    1,000   - 

1  - 

A 
O 

+ 
A 

O 

+ 
A 

O 

+ 
A 

-t- 
A 

+ 
A 

+ 

A 

* — I 1 

School  size 


J3 


-r    1.000,000 


Figure  2 

Simulated  dynamics  of  school-size  distributions  with  mdl  as  the  mortality 
model  and  wN  as  the  aggregation  model.  This  simulation  started  with  disag- 
gregated initial  conditions  (N0  =  S),  where  S=106.  The  aggregation  parameter 
was  a=10~6.  The  top  panel  shows  school-size  distributions  (frequency  in  log)  at 
selected  times  (().  The  lower  panel  gives  the  mean  school  size  (kbar)  and  school 
abundance  (JV)  versus  time. 


them.  At  this  point  the  reduction  in  the  number  of  schools 
is  accelerated  because  of  the  mortality  of  fish  that  are  in 
"schools"  in  which  they  are  the  only  member,  and  because 
of  the  loss  of  schools  attributed  to  aggregation. 

Model  comparisons 

Numerical  integration  of  Equations  3-5  matched  the  sim- 
ulation results  (Fig.  6,  when  P1(=0),  indicating  that  the 
mathematical  model  describes  the  simulation  dynamics. 
The  numerical  techniques  are  sufficiently  efficient  to  be 
used  in  a  curve-fitting  context.  Evaluations  of  the  approxi- 
mation (Appendix  2)  indicate  that  the  approximation  may 
be  useful  for  predictions  of  recruitment,  when  compared 


with  the  simulations.  However,  the  components  of  recruit- 
ment, kt  and  Nr  were  biased  (Fig.  7).  Further  research  is 
needed  to  develop  estimates  of  P1 1  and,  more  generally, 
P(k)  under  other  models  and  initial  conditions. 

Recruitment  was  compared  between  mortality  models 
and  aggregation  models  (Fig.  8).  If  the  mortality  model 
was  either  mdl  or  mdR,  then  the  mortality  rate  was  not  af- 
fected by  the  school-size  distribution:  random  attachment 
and  preferential  attachment  perform  equally  as  well  in 
terms  of  survival  at  a  given  time.  But  if  mortality  encoun- 
ters proportional  to  school  density  (.mdN)  were  imposed, 
then  there  were  better  survival  rates  with  preferential  at- 
tachment than  with  random  attachment  (Fig.  8,  A  and  B). 
Conversely,  mortality  encounters  proportional  to  school 


356 


Fishery  Bulletin  102(2) 


o  f=0 
+  f  =  0.35 
A  f=  1.50 


1.000 

School  size 


T  2.000 


Figure  3 

Simulated  dynamics  of  school-size  distributions  using  mdl  as  the  mortal- 
ity model  and  w  as  the  aggregation  model.  This  simulation  started  with 
aggregated  initial  conditions  (S  =  106).  The  aggregation  parameter  was 
a=10~6.  The  top  panel  shows  school-size  distributions  lin  log-log  scale)  at 
selected  times  (t).  The  lower  panel  gives  the  mean  school  size  Uibar)  and 
school  abundance  IN)  versus  time. 


size  (mdk)  led  to  poorer  survival  with  preferential  attach- 
ment (Fig.  8,  C  and  D). 


Discussion 

Koslow  (1992),  Rickman  et  al.  (2000),  and  others  have 
commented  on  the  inherent  variability  in  stock-recruit- 
ment data  and  the  limited  predictive  power  of  determin- 
istic stock-recruitment  models.  Therefore,  there  is  no 
expectation  that  one  could  select  the  models  described 
here  over  other  stock-recruitment  models  on  the  basis  of 
fits  to  data.  Although  the  aggregation-mortality  models 


may  be  fitted  to  stock-recruitment  data,  the  real  useful- 
ness is  as  a  guide  to  selection  of  stock-recruitment  models 
used  in  management,  as  a  mechanism  for  integrating 
research  on  recruitment  processes,  and  as  a  tool  for  explor- 
ing the  structure  of  recruitment  variability. 

The  aggregation-mortality  models  were  introduced 
by  using  an  analogy  with  evolving  random  networks 
( Barabasi  and  Albert,  1999 )  and  were  shown  to  be  analyti- 
cally equivalent  (Appendix  2).  Modeled  fish  are  subjected 
to  competing  forces  of  organization  (aggregation)  and  decay 
(mortality),  as  in  a  network  in  which  links  to  nodes  in  the 
network  are  created,  destroyed,  and  rewired  (Albert  and 
Barabasi.  2002).  An  important  finding  of  Barabasi  and 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


357 


o  f=  0 
+  t=5t 

a  f=  2f 


School  size 


0002 
[ 


Figure  4 

Simulated  dynamics  of  school-size  distributions  using  mJX  as  the  mortal- 
ity model  and  w  as  the  aggregation  model.  This  simulation  started  with 
aggregated  initial  conditions  (S=2xl06).  The  aggregation  parameter  was 
o=1.5x  10"6.  The  top  panel  shows  school-size  distributions  (in  log-log  scale) 
at  selected  times  (t).  The  lower  panel  gives  the  mean  school  size  (kbar)  and 
school  abundance  (N)  versus  time. 


Albert  (1999)  was  that  scaling  of  the  aggregate-size  dis- 
tribution was  dependent  on  the  type  of  aggregation,  spe- 
cifically preferential  attachment.  Bonabeau  and  Dagorn 
noted  the  generic  occurrence  of  scaling  of  aggregation 
distributions  in  nature  (Bonabeau  and  Dagorn,  1995)  and 
this  scaling  of  aggregation  distributions  motivated  the 
development  of  the  models  presented  here. 

The  emphasis  of  the  aggregation  models  was  on  prefer- 
ential attachment  and  on  comparison  of  model  results  with 
results  for  models  with  random  attachment  strategies.  The 
preferential  attachment  rule  used  in  the  simulations  was 
that  aggregation  rates  were  proportional  to  the  size  of  the 
school  encountered.  But,  what  is  meant  by  preferential 


attachment  and  does  preferential  attachment  occur  in 
nature?  Clearly,  a  fish,  school  or  mortality  agent  has  no 
global  knowledge  of  the  proportional  size  of  a  school  that 
is  encountered.  However,  preferential  attachment  in  these 
models  is  a  metaphor  for  aggregation  strategies  that  are 
weighted  toward  larger  school  sizes.  Indeed,  studies  of 
networks  have  shown  that  attachment  may  be  proportional 
to  a  power  of  school  size  and  still  produce  scale-free  prop- 
erties (Albert  and  Barabasi,  2002).  Also,  network  studies 
have  shown  that  scale-free  distributions  occur  when  a 
wide  number  of  attachment  criteria  are  included,  such 
as  the  "fitness"  of  the  object  being  encountered  and  the 
attractiveness  of  local  conditions  (Bianconi  and  Barabasi, 


358 


Fishery  Bulletin  102(2) 


1,000  i 

0 

0 

O 

O 

0 

o 

o    (=0 

+    f=5f  -6 

a   f=  1f  -6 

o 
o 
o 
o 

>, 

0 

<D 

o 

cr 

o 

CD 

* 

o 

D 

3^j* 

o 

cr 

TMi1 

o 

CD 

o 

3 

'^MhK* 

o 

LL 

A           +             AK         +       ++     HtH- 

o 
o 

0 

A   +                A           +     + 

A                                            * 

+                   OL               +  -H- 
« A Htt- 

e . 

400 

Icsooz  size 


1,000    -| 


Figure  5 

Simulated  dynamics  of  school-size  distributions  using  mdN  as  the  mortal- 
ity model  and  wN  as  the  aggregation  model.  This  simulation  started  with 
aggregated  initial  conditions  (S=2xl06).  The  aggregation  parameter  was 
ot=1.5x  10~6.  The  top  panel  shows  school-size  distributions  (in  log-log  scalei 
at  selected  times  it).  The  lower  panel  gives  the  mean  school  size  ikbar)  and 
school  abundance  (N )  versus  time. 


2001;  Calderelli  et  al„  2002;  Vazquez,  2003).  Biological 
concepts  of  fitness,  feeding  behavior,  predator-avoidance 
behavior,  and  habitat  suitability  appear  to  fall  within 
the  attachment  criteria  examined  in  physics  literature. 
Oceanographic  stability  (Myers  and  Pepin,  1994),  assorta- 
tive  schooling  by  color  patterns  (Crook,  1999),  chemosen- 
sory  stimuli  (Quinn  and  Busack,  1985),  and  larval  fitness 
indices  from  RNA/DNA  ratios  (Pepin,  1991;  Suneetha  et 
al.,  1999)  may  be  mechanisms  that  directly  or  indirectly 
influence  aggregation  size  and,  thus,  distribution. 

The  geometry  of  the  school  size  itself  may  be  sufficient  to 
produce  preferential  attachment  behavior,  as  well.  In  the 


models  of  this  study,  the  detection  spaces  (G,  +Gf  in  Equa- 
tion 1)  were  set  to  unity  and  assumed  to  be  independent  of 
school  size.  However,  the  detection  space  may  be  related 
to  school  size.  For  example,  if  a  school  of  one  fish  has  a 
spheroid  detection  space  around  itself  with  radius  equal 
to  1,  then  using  the  geometry  of  an  aggregation  of  /<■  fish, 
the  detection  space  of  the  aggregate  would  be  proportional 
to  kv:\  Alternatively,  if  the  detection  space  were  a  two-di- 
mensional circle  with  a  radius  of  1,  then  the  aggregate's 
detection  space  would  be  proportional  to  &"'-'.  Substituting 
size-dependent  detection  spaces  into  the  random  mortality 
and  aggregation  models  would  be  sufficient  to  induce  pref- 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


359 


B 


1.000,000  2,000.000 


3,000.000    -[ 


2,000.000  0 

Stock  size  (no.  of  fish) 


1,000,000 


■  Mathematical  model        •  Simulation  model 


Figure  6 

Stock-recruitment  relationships  from  the  mathematical  models  (Eqs.  3-5,  aggre- 
gated initial  conditions!  compared  with  simulation  results:  I  Ai  density-independent 
mortality  (mc/l I  and  preferential  attachment  (u'l  evaluated  at  4=1,  r<=10".  ii  =  l;  (B) 
density-dependent  mortality  proportional  to  fish  imdR  I  and  preferential  attachment 
(w  )  evaluated  at  t=10~5,  a=3  x  10  5,  ,n  =  l;  ( C )  density-dependent  mortality  propor- 
tional to  schools  (rndN)  and  preferential  attachment  iw  )  evaluated  at  t=5x  10~4, 
a=1.5x  10~6,  f(=l;  and  (D)  density-dependent  mortality  proportional  to  school  size 
(mdk)  and  preferential  attachment  (w    )  evaluated  at  ^=10-3,  a=2x  10_li,  ,11  =  1. 


erential  interaction  even  when  encounters  are  random: 
schools  are  randomly  encountered,  but  the  encounter  event 
itself  is  weighted  toward  larger  schools.  Thus,  the  shape  of 
the  detection  space  may  be  another  mechanism  by  which 
preferential  attachment  may  be  exhibited. 

In  the  models  presented,  it  is  blithely  assumed  that 
mortality  is  caused  by  undefined  mortality  agents.  How- 
ever, most  larval  recruitment  research  has  been  directed 
at  starvation  and  predation  as  determinants  of  recruit- 
ment variability  (Lasker,  1975;  Hunter,  1984;  Bailey  and 
Houde,  1989;  Chambers  and  Trippel,  1997,  for  example). 
The  mortality  models  used  here  clearly  fit  within  the  pre- 
dation paradigm:  mortality  from  predation  results  from 


encounters  with  mortality  agents  of  specific  density  and 
size.  Whereas,  mortality  from  starvation  ensues  from  a 
lack  of  encounters  with  prey  agents  of  sufficient  density 
and  size.  In  certain  situations  starvation  processes  might 
be  aptly  described  by  the  predation-encounter  approach 
used  in  this  study.  However,  further  research  is  needed 
to  evaluate  their  appropriateness  and  to  develop  alterna- 
tive modifications  to  Equations  3-6.  A  mechanism  to  do 
this  may  be  the  inclusion  of  fragmentation  of  schools  into 
the  models.  In  the  models  as  they  are  now  characterized, 
new  schools  are  not  created,  the  number  of  schools  only 
becomes  smaller  through  either  aggregation  or  through 
mortality  on  schools  of  a  single  fish.  Fragmentation  might 


360 


Fishery  Bulletin  102(2) 


1 ,000,000  2,000,000 


2.000,000  0 

Stock  size  (no.  offish) 


Mathematical  model        •  Simulation  model 


Figure  7 

Stock-recruitment  relationships  determined  from  the  mathematical  models  (Eqs. 
4,  5,  and  10,  disaggregated  initial  conditions)  compared  with  simulation  models. 
(A  and  Bl  Recruitment  at  r=l  with  mortality  encounters  proportional  to  school  size 
(mdk)  at  a=5xl0~"  and  ,n  =  l;  A  is  recruitment  and  B  is  the  mean  school  size.  iC  and 
D)  Recruitment  at  t=10~5  with  mortality  encounters  proportional  to  school  density 
im/vi  at  o=0.2  and  ii=l;  C  is  recruitment  and  D  is  mean  school  size. 


occur  due  to  secondary  effects  of  mortality  encounters, 
as  well  as  other  factors  such  as  starvation.  For  example, 
Sogard  and  Olla  (1997)  have  shown  predation-risk  and 
hunger  to  be  related  to  group  cohesion. 

The  formation  of  a  giant  cluster  (a  single  school  en- 
compassing all  the  fish)  is  an  important  feature  of  the 
attachment  process.  The  simulations  showed  that  with 
preferential  attachment  the  recruitment  process  passes 
through  a  phase  where  the  size  distribution  is  scale  free, 
then  a  critical  point  is  reached  where  a  giant  cluster  is 
being  formed,  i.e.  a  single  school  begins  to  attract  all  the 
fish.  Research  on  complex  networks  has  shown  the  condi- 
tions for  formation  of  the  giant  cluster  (Aiello  et  al.,  2000; 
Albert  and  Barabasi,  2002).  This  should  be  investigated 
for  the  school  aggregation  models  because  it  is  likely  that 
the  mortality  models  used  in  the  present  study  would  no 
longer  be  appropriate  once  the  giant  cluster  is  formed.  In- 
deed in  some  fish  stocks,  schools  may  aggregate  into  giant 
clusters  on  a  local  scale  and  then  aggregation  may  stop  for 


reasons  such  as  juveniles  entering  a  benthic  phase.  The 
resulting  distribution  of  school  sizes  may  be  the  cluster 
distribution  across  benthic  habitats.  Spatial  limitations 
of  aggregation  are  an  important  feature  of  individually 
based  models  (Pascual  and  Levin,  1999).  Again,  this  may 
be  an  important  area  for  research. 

What  is  the  benefit  of  preferential  attachment?  If  mor- 
tality encounters  are  proportional  to  school  density,  then 
recruitment  survival  rates  are  improved  when  there  are 
fewer  schools  for  a  given  number  of  fish,  i.e.  when  prefer- 
ential attachment  is  employed  rather  than  random  attach- 
ment (Fig.  8).  Perhaps,  preferential  attachment  strategies 
are  a  useful  evolutionary  hedge  against  uncertainty  in 
the  nature  of  the  mortality  dynamics.  Conversely,  when 
mortality  encounters  are  proportional  to  school  size,  then 
better  survival  is  achieved  when  schools  are  smaller,  i.e. 
with  random  attachment  (Fig.  8).  If  mortality  by  preda- 
tors is  related  to  larger  schools,  or  if  attainment  of  prey  is 
inversely  related  to  larger  schools,  then  more  solitary  life 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


361 


A  Mortality  encounters  proportional  to  school 
r 


density;  aggregated  initial  conditons 


Mortality  encounters  proportional  to  school 


80.000  ■ 

-D  density,  disaggregated  initial  conditons 

1 

0 , 1 

C  Mortality  encounters  proportional  to  school 
size;  aggregated  initial  conditons 


o  if 

0 


D  Mortality  encounters  proportional  to  school 
size;  disaggregated  initial  conditons 


2,000.000 

Stock  size  (no.  of  fish) 


Random  attachment 


Preferential  attachment 


Figure  8 

Stock-recruitment  results  with  preferential  attachment  comparer]  with  random  attach- 
ment (w  versus  wN).  (A)  Recruitment  at  r=l  with  mortality  encounters  proportional  to 
school  density  imdN)  at  a=l(H'  and  ,«=5xl0  4  with  an  aggregated  initial  condition;  (Bl 
recruitment  at  r=10~5  with  mortality  encounters  proportional  to  school  density  lmdN)  at 
a=0.2  and  ji=1  with  a  disaggregated  initial  condition;  (C)  recruitment  at  t=l  with  mortal- 
ity encounters  proportional  to  school  size  (.mdk)  at  a=10~9  and  jt<=5xl0~41  with  an  aggre- 
gated initial  condition;  (Dl  recruitment  at  t=l  with  mortality  encounters  proportional  to 
school  size  l»!iM.)  at  a=5x  10~7  and  ti=l  with  a  disaggregated  initial  condition. 


history  strategies  may  evolve.  Perhaps,  the  random  ag- 
gregation model  would  be  most  effective  for  solitary  preda- 
tory fish  when  their  mortality  is  imposed  by  a  mdk-type 
model.  For  fish,  this  may  be  more  likely  to  occur  at  later 
life  stages  than  at  recruitment.  If  mortality  encounters 
are  proportional  to  fish  (mrlR),  then  results  are  intermedi- 
ate and  preferential  attachment  and  random  attachment 
perform  equally  as  well. 

The  density-dependent  mortality  models  implicitly  in- 
corporate a  predator-prey  interaction.  Alternative  preda- 
tor-prey interactions  examined  were  those  in  which  preda- 
tor density  was  proportional  to  fish,  to  schools,  or  to  the 
number  offish  within  a  school  (school  size)  throughout  a 
recruitment  period.  In  reality  mortality  is  perpetrated  by 
a  variety  of  agents  at  many  different  scales.  Some  agents 
act  at  the  scale  of  the  population  (Nk  ),  some  at  the  scale  of 


schools  (N  ),  some  at  the  scale  of  mean  school  size  (k ),  and 
some  at  the  scale  of  a  local  school  (£■).  The  mixture  of  preda- 
tory agents  and  their  densities  can  cause  various  kinds  of 
dynamics  including  oscillatory,  chaotic,  and  stable  behav- 
ior (Wilson  1996,  Pascual  and  Levin  1999).  Therefore,  it  is 
unlikely  that  the  models  in  this  study,  in  which  predator- 
prey  ratios  are  constant,  would  be  predictive  of  anything 
other  than  average  behavior  during  recruitment.  However, 
the  analytical  approach  allows  changes  in  the  scale  of 
predator-prey  interaction  over  time.  We  can  model  this  as 
ml=2uNtaktb,  where  a  and  b  are  dynamic  (time-dependent) 
and,  perhaps,  correlated.  Although  we  may  wish  to  use 
the  Beverton-Holt  model  (a=b=2)  or  the  Ricker-like  model 
(a=2,  b=l)  as  a  representation  of  average  dynamics,  it 
remains  that  recruitment  variability  will  be  influenced  by 
the  dynamics  of  the  exponents,  a  and  b.  Numerical  evalua- 


362 


Fishery  Bulletin  102(2) 


tion  of  the  differential  equations  by  using  random  variates 
at  each  time  step  may  be  a  mechanism  to  evaluate  how 
the  variability  of  a  and  b  within  a  recruitment  period  are 
translated  into  the  variability  structure  around  a  stock- 
recruitment  relationship. 

The  model  formulations  used  in  the  present  study  have 
been  characterized  from  the  underlying  physical  pro- 
cesses. By  doing  so,  research  may  be  directed  at  empirical 
and  experimental  measurement  of  specific  stock-recruit- 
ment parameters,  which  opens  the  models  to  testing 
and  verification.  Additionally,  results  indicate  that  the 
school-size  distribution  contains  a  rich  source  of  informa- 
tion on  the  mortality  and  aggregation  processes  and  that 
monitoring  of  the  distribution  during  recruitment  could 
be  useful  for  understanding  recruitment  variability  and 
model  structure. 


Acknowledgments 

I  would  like  to  thank  the  reviewers  for  their  constructive 
comments  and  the  National  Marine  Fisheries  Service  for 
allowing  me  the  opportunity  to  conduct  this  research. 


Literature  cited 

Aiello,  W.  F.,  F.  Chung,  and  L.  Lu. 

2000.  A  random  graph  model  for  massive  graphs.     Proc 
32nd  ACM  Symp.  Theor.  Comp.,  p.  171-180. 

Albert.  R..  and  A.  Barabasi. 

2002.     Statistical  mechanics  of  complex  networks.     Rev. 
Mod.  Phys.  74(l):47-98. 
Anderson,  J.  J. 

1981.     A  stochastic  model  for  the  size  of  fish  schools.     Fish. 
Bull.  79:315-323. 
Bailey,  K.  M.,  and  E.  D.  Houde. 

1989.     Predation  on  eggs  and  larvae  of  marine  fishes  and  the 
recruitment  problem.     Adv.  Mar.  Biol.  25:1-83. 
Barabasi,  A.,  and  R.  Albert. 

1999.     Emergence  of  scaling  in  random  networks.     Science 
286:5098-512. 
Bianconi,  G.,  and  A.  Barabasi. 

2001.  Bose-Einstein  condensation  in  complex  networks. 
Phys.  Rev.  Lett.  86:5632-5635. 

Beverton,  R.  J.  H.,  and  S.  J.  Holt. 

1957.     On  the  dynamics  of  exploited  fish  populations.     Fish. 
Invest.  Lond.  Ser.  2,  19  p. 
Beyer,  J.  E.,  and  G.  C.  Laurence. 

1980.     A  stochastic  model  of  larval  fish  growth.     Ecol. 
Model.  8:109-132. 
Bonabeau,  E.,  and  L.  Dagorn. 

1995.     Possible  universality  in  the  size  distribution  of  fish 
schools.     Phys.  Rev.  E  51:R52220-R52223. 
Bonabeau,  E.,  L.  Dagorn,  and  P.  Freon. 

1999.     Scaling  in  animal  group-size  distributions.     Proc. 
Natl.  Acad.  Sci.  96:4472-4477. 
Caldarelli,  G.,  A.  Capocci,  P.  De  Los  Rios,  and  M.A.  Munoz. 

2002.  Scale-free  networks  from  varying  vertex  intrinsic 
fitness.     Phys.  Rev.  Lett.  89,  258702. 

Chambers,  C,  and  E.  A.  Trippel. 

1997.     Early  life  history  and  recruitment  in  fish  populations. 
632  p.     Chapman  and  Hall,  London. 


Comyns,  B.  H.,  R.  F.  Shaw,  and  J.  Lyczkowski-Shultz. 

2003.  Small-scale  spatial  and  temporal  variability  in 
growth  and  mortality  of  fish  larvae  in  the  subtropical 
northcentral  Gulf  of  Mexico:  implications  for  assessing 
recruitment  success.     Fish.  Bull.  101:10-21. 

Crook,  A.  C. 

1999.  Quantitative  evidence  for  assortative  schooling  in  a 
coral  reef  fish.     Mar.  Ecol.  Prog.  Ser.  176:17-23. 

Dorogovtsev,  S.  N„  and  J.  F.  F.  Mendes. 

2000.  Scaling  behaviour  of  developing  and  decaying  net- 
works.    Europhys.  Lett.  52:33-39 

2002.     Evolution  of  networks.     Adv.  Phys.  51:1079-1187. 
Erdos,  P.,  and  A.  Renyi. 

1960.     On  evolution  of  random  graphs.     Publ.  Math.  Inst. 
Hung.  Acad.  Sci.  5:17-61. 
Fell,  D.  A.,  and  A.  Wagner. 

2000.     The  small  world  of  metabolism.     Nat.  Biotech. 
18:1121-1122. 
Flierl,  G.,  D.  Grunbaum,  S.A.  Levin,  and  D.  Olson. 

1999.  From  individuals  to  aggregations:  the  interplay  be- 
tween behavior  and  physics.     J.  Theor.  Biol.  196:397-454. 

Gueron.  S.,  and  S.  A.  Levin. 

1995.     The  dynamics  of  group  formation.     Math.  Biosci. 
128:243-264. 
Harris,  J.  G.  K. 

1975.     The  effect  of  density-dependent  mortality  on  the 
shape  of  the  stock  and  recruitment  curve.     J.  Cons.  Int. 
Explor.  Mer  36:144-149. 
Hunter,  J.  R. 

1984.  Inferences  regarding  predation  on  the  early  life 
stages  of  cod  and  other  fishes.  In  The  propagation  of  cod. 
Gadus  morhua  L.  (E.  Dahl,  D.  S.  Danielssen,  E.  Moksness, 
and  P.  Solemdan,  eds.),  p.  533-562.  Flodevigen  Rap- 
portser  1984. 
Koslow,  J.  A. 

1992.  Fecundity  and  the  stock-recruitment  relationship. 
Can.  J.  Fish.  Aquat.  Sci.  49:210-217. 

Krapivsky,  P.  L.,  S.  Redner,  and  F.  Leyvraz. 

2000.  Connectivity  of  growing  random  networks.  Phys. 
Rev.  Lett.  85:4629-4632. 

Landa.  J.  T. 

1998.     Bioeconomics  of  schooling  fishes:  selfish  fish,  quasi- 
free  riders,  and  other  fishy  tales.     Environ.  Biol.  Fish. 
53:353-364. 
Lasker,  R. 

1975.     Field  criteria  for  survival  of  anchovy  larvae:  the 
relation  between  inshore  chlorophyll  maximum  layers  and 
successful  first  feeding.     Fish.  Bull.  73:453-462. 
McCauley,  E..  W.  G.  Wilson,  and  A.M.  de  Roos. 

1993.  Dynamics  of  age-structured  and  spatially  structured 
predator-prey  interactions:  individual-based  models  and 
population-level  formulations.  Am.  Naturalist  142:412- 
442. 

Mertz,  G.,  and  R.  A.  Myers. 

1994.  Match/mismatch  predictions  of  spawning  dura- 
tion versus  recruitment  variability.  Fish.  Oceanogr. 
3(4):236-245. 

1995.  Estimating  the  predictability  of  recruitment.  Fish. 
Bull.  93:657-665. 

Myers,  R.  A.,  and  P.  Pepin. 

1994.     Recruitment  variability  and  oceanographic  stability. 
Fish.  Oceanogr.  3(4):246-255. 
Niwa.  H.-S. 

1998.  School  size  statistics  offish.  J.  Theor.  Biol.  195:351- 
361. 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


363 


Pascual,  M.,  and  S.  A.  Levin. 

1999.  From  individuals  to  population  densities:  search- 
ing for  the  intermediate  scale  of  nontrivial  determinism. 
Ecology  80(71:2225-2236. 

Paulik,  G.  J. 

1973.     Studies  of  the  possible  form  of  the  stock-recruitment 
curve.     Rapp.  P.-v.  Reun.  Cons.  Int.  Explor.  Mer  164: 
303-315. 
Pepin,  P. 

1991.     Effect  of  temperature  and  size  on  development, 
mortality,  and  survival  rates  of  early  life  history  stages  of 
marine  fish.     Can.  J.  Fish.  Aquat.  Sci.  48:503-518. 
Pitcher,  T.  J.,  and  J.  K.  Parrish. 

1993.     Functions  of  shoaling  behaviour  in  teleosts.     In  Be- 
haviour of  teleost  fishes  (T.  J.  Pitcher,  ed.l.  p.  363-439. 
Chapman  and  Hall,  London. 
Quinn,  T.  P.,  and  C.  A.  Busack. 

1985.  Chemosensory  recognotion  of  siblings  in  juvenile  coho 
salmon  Oncorhynchus  kisutch.     Anim.  Behav.  33:51-56. 

Ricker,  W.  E. 

1958.     Handbook  of  computations  for  biological  statistics  of 
fish  populations,  119  p.     Fish.  Res.  Board  Can.  Bull. 
Rickman,  S.  J.,  N.  K.  Dulvy,  S.  Jennings,  and  J.  D.  Reynolds. 

2000.  Recruitment  variation  related  to  fecundity  in  marine 
fishes.     Can.  J.  Fish.  Aquat.  Sci.  57:116-124. 

Rothschild,  B.  J. 

1986.  Dynamics  of  marine  fish  populations.  269  p.  Har- 
vard Univ.  Press,  Cambridge.  MA. 

Sogard.  S.  M„  and  B.  L.  Olla. 

1997.  The  influence  of  hunger  and  predation  risk  on 
group  cohesion  in  a  pelagic  fish,  walleye  pollock  Theragra 
chalcogramma .     Environ.  Biol.  Fish.  50:405-413. 

Sole,  R.  V.,  and  J.  M.  Montoya. 

2001.  Complexity  and  fragility  in  ecological  networks. 
Proc.  R.  Soc.  London  B.  268:2039-2045. 

Suneetha,  K.-B.,  A.  Folkvord,  and  A.  Johannessen. 

1999.     Responsiveness  of  selected  condition  measures  of 
herring,  Clupea  harengus,  larvae  to  starvation  in  rela- 
tion to  ontogeny  and  temperature.     Environ.  Biol.  Fish. 
54:191-204. 
Tolman,  R.  C. 

1979.     The  principles  of  statistical  mechanics,  661  p.     Dover 
Pubis.,  New  York,  NY. 
Vazquez,  A. 

2003.     Growing  networks  with  local  rules:  preferential 
attachment,  clustering  hierarchy  and  degree  correla- 
tions.    Phys.  Rev.  E  67:056104. 
Vlymen,  W.  J. 

1977.     A  mathematical  model  of  the  relationship  between 
larval  anchovy  iEngraulis  mordax)  growth,  prey  micro- 
distribution,  and  larval  behavior.     Environ.  Biol.  Fish. 
2:211-233. 
Wilson,  W.  G. 

1996.  Lotka's  game  in  predator-prey  theory:  linking  popu- 
lations to  individuals.     Theor.  Pop.  Biol.  50:368-393. 


Mortality  proportional  to  fish:  mdR 
Preferential  aggregation:  w 


Rt 


l  +  2,uSt 


Nt=£  +  ^S 


kt=R,/N, 


ko     fi{        1  +  2[iSt 


l  +  2/iSr  +  2osSV 


Mortality  proportional  to  fish:  mdR 
Random  aggregation:  wN 


"*      l  +  2/iS< 

s 

1  '      S-lS-Ve"2* 

k, 

=  Rt/Nt  =  S-aS-^-2°' 

Density-independent:  mdl 
Preferential  aggregation:  wpa 


R,  =  Se-2'" 


Art=  J.  _.£_£(!_  e-4/« 


^      2/1 
k,=R,/N,-- 


v-2"' 


1_S*o   «q_c-4/i/) 

2    n 


Density-independent:  mdi 
Random  aggregation:  wN 


Nt  = 


R,  =  Se"2'* 
S 


kt=Rt/Nt  =  e~2f"  (S-iS-^e-'2*) 


Appendix  1 

Analytical  solutions  to  Equations  3-5  for  selected  mortal- 
ity and  aggregation  models.  Solutions  assume  that  Plt  =  0 
for  all  t  evaluated  and  that  the  number  of  schools  is  large. 
No  analytical  solutions  were  found  for  (mdN,wpa),  (mdN, 


Wpa>'OT(mo 


•wpa>- 


Mortality  proportional  to  schools:  mdN 
Random  aggregation:  wN 


Rt  =Se 


-2(1/  J feo_ 

|s-(S- v? 


-2a* 


364 


Fishery  Bulletin  102(2) 


N, 


S-(S-k0)e' 


k,  =RtINt  = 


k^e** 


[S-(S-kf))e-2,"\' 


Mortality  proportional  to  school  size:  mdk 
Random  aggregation:  wN 


and  Albert  (1999).  Dorogovtsev  and  Mendes  (2000)  and 
Albert  and  Barabasi  (2002). 

When  the  aggregation  model  is  preferential  attachment 
(w  )  (ignoring  for  the  moment  the  nonstationarity  of  N 
and  P),  then  the  partial  differential  of  a  school  of  size  klt 
with  respect  to  Rt  has  been  shown  by  Dorogovtsev  and 
Mendes  (2000)  to  asymptotically  be 


dkuim,  =  P,ik,,IR,), 


(AD 


where  ft,  is  the  net  rate  of  decay  per  each  mortality  event, 
i.e., 


R, 


l  +  2pSi-^(S-Ml-e~2°*] 
a 


fit 


l-wpalm. 


With  specific-mortality  models,  fi,  is 


(A2) 


Nt  = 


k,=R,/N,=- 


S-iS-k0)e' 


S-iS-kpte 


-2at 


l  +  2^t-^S-k0)[l-e~ 
a 


Random  mortality  encounters:  mN 
Random  aggregation:  wN 

V  i„  \  S  <J 


Rl=S-£-]n\Z-(e""-l)  +  l\ 
a      [ko 


mdl:    p,  =  l-ial  ii)iN,  -Dk,  =  l-ial  /j)R, 
mdN:  p,  =  l-(a/ n)(Nt-ljktl Nt  =  l-ial  n)k, 
mdR:  p,  =  l-ial  f.t)iN,  -DIN,  =l-(a//i) 
mdk:   P,  =  l-(a/fi)(Nt  -l)  =  l-(a/p)P,  Ik,, 

where  the  approximations  on  the  right  assume  that  the 
number  of  schools  is  large.  The  first  term  of  (A2)  denotes 
the  removal  of  a  fish  proportional  to  school  size  for  a  mor- 
tality event;  the  second  term  denotes  aggregation  events 
proportional  to  school  size.  If  ft,  is  independent  of  time 
ift,=ft),  then  Dorogovtsev  and  Mendes  (2000)  showed  that 
under  continuum  conditions 


Pi  k ) 


7  =  1  +  1/ /J. 


(A3) 


W,=- 


2ftf 


S-{S-ko)e 


k,=R,IN, 


Appendix  2 

Characteristics  of  school-size  distribution  under 
preferential  attachment 

Much  of  the  recent  literature  on  evolving  complex  net- 
works has  been  directed  at  determining  the  degree  distri- 
bution, i.e.,  the  probability  P(k)  of  a  node  having  k  links 
(Albert  and  Barabasi,  2002).  When  the  network  grows 
or  declines  proportional  to  k  or  when  links  are  rewired 
to  be  proportional  to  k,  then  P(k>  can  be  determined  by 
using  continuum  theory  (Dorogovtsev  and  Mendes,  2000; 
Albert  and  Barabasi,  2002)  leading  to  scale-free  degree 
distributions.  Therefore,  when  preferential  attachment 
and  nonrandom  mortality  are  used,  then  the  model  may  be 
couched  as  a  scale-free  network  in  the  manner  of  Barabasi 


Equation  A3  is  equivalent  to  the  results  of  Dorogovtsev 
and  Mendes  (2000),  Krapivsky  et  al.  (2000),  and  Albert 
and  Barabasi  (2002)  and  suggest  that  ft,  may  be  a  useful 
approximation  for  determining  the  power-law  tail  of  the 
school-size  distribution  (Appendix  Fig.  1). 

The  simulation  results  showed  the  dynamics  of  Pkt. 
When  the  aggregated  initial  condition  was  imposed,  at 
the  start  of  the  simulations  there  were  no  schools  with 
only  one  fish  in  them  (P,,  =  0).  Eventually,  as  the  number 
of  schools  and  fish  declined,  P„  became  positive.  Finally, 
as  the  distribution  became  scale-free,  -i)Px,l Hk  became 
negative  and  remained  so  throughout  the  remainder  of  the 
simulation  or  until  a  single  giant  cluster  was  formed  (Ap- 
pendix Fig.  1).  Conversely,  if  the  initial  conditions  began 
with  schools  being  disaggregated,  then  dPuldk  began  as 
a  negative  number  and  remained  so  until  either  a  giant 
cluster  formed  or  there  were  no  more  fish  remaining. 

An  approximation  is  suggested  by  the  above  results 
for  circumstances  when  the  initial  conditions  are  disag- 
gregated and  when  there  is  preferential  attachment:  the 
differential  equation  dP,. ,  Idt  when  k  =  1  ( Eq.  6 )  is  replaced 

by 


dPuldt 


-irPy,IN,  +  ;?Ml-Pu)/JV,. 


(A4) 


Powers:  Recruitment  as  an  evolving  random  process  of  aggregation  and  mortality 


365 


1  1 ,000 

School  size  (no.  of  fish) 

Appendix  Figure  1 

School-size  distribution  at  selected  times.  (A)  School-size 
distribution  of  a  simulation  starting  with  a  disaggregated 
initial  condition,  S=106,  n=10_6,  where  mortality  is  density- 
independent  (»!(/l)  and  there  is  preferential  attachment  (w^). 
(B)  Distribution  starting  with  an  aggregated  initial  condition, 
S=2xl06,  «=1.5xl0~6,  where  mortality  is  density-dependent 
proportional  to  schools  imdN)  and  with  preferential  attachment 
(w  a).  The  dotted  lines  are  the  predictions  of  y  =  1  +  1//3  from 
Equation  A2,  horizontally  offset  for  viewing. 


366 


Abstract— We  examined  the  diets  and 
habitat  shift  of  juvenile  red  snapper 
(Lutjanus  campechanus)  in  the  north- 
east Gulf  of  Mexico.  Fish  were  col- 
lected from  open  sand-mud  habitat 
(little  to  no  relief),  and  artificial  reef 
habitat  (1-m3  concrete  or  PVC  blocks), 
from  June  1993  through  December 
1994.  In  1994,  fish  settled  over  open 
habitat  from  June  to  September,  as 
shown  by  trawl  collections,  then  began 
shifting  to  reef  habitat  —  a  shift  that 
was  almost  completed  by  December  as 
observed  by  SCUBA  visual  surveys. 
Stomachs  were  examined  from  1639 
red  snapper  that  ranged  in  size  from 
18.0  to  280.0  mm  SL.  Of  these,  850 
fish  had  empty  stomachs,  and  346  fish 
from  open  habitat  and  443  fish  from 
reef  habitat  contained  prey.  Prey  were 
identified  to  the  lowest  possible  taxon 
and  quantified  by  volumetric  measure- 
ment. Specific  volume  of  particular 
prey  taxa  were  calculated  by  dividing 
prey  volume  by  individual  fish  weight. 
Red  snapper  shifted  diets  with  increas- 
ing size.  Small  red  snapper  (<60  mm 
SL)  fed  mostly  on  chaetognaths.  cope- 
pods,  shrimp,  and  squid.  Large  red 
snapper  (60-280  mm  SL)  shifted  feed- 
ing to  fish  prey,  greater  amounts  of 
squid  and  crabs,  and  continued  feeding 
on  shrimp.  We  compared  red  snapper 
diets  for  overlapping  size  classes  (70- 
160  mm  SL)  offish  that  were  collected 
from  both  habitats  (Bray-Curtis  dis- 
similarity index  and  multidimensional 
scaling  analysis).  Red  snapper  diets 
separated  by  habitat  type  rather  than 
fish  size  for  the  size  ranges  that  over- 
lapped habitats.  These  diet  shifts  were 
attributed  to  feeding  more  on  reef  prey 
than  on  open-water  prey.  Thus,  the 
shift  in  habitat  shown  by  juvenile  red 
snapper  was  reflected  in  their  diet  and 
suggested  differential  habitat  values 
based  not  just  on  predation  refuge  but 
food  resources  as  well. 


Diet  shifts  of  juvenile  red  snapper 

(Lutjanus  campechanus) 

with  changes  in  habitat  and  fish  size 


Stephen  T.  Szedimayer 

Marine  Fish  Laboratory 

Department  of  Fisheries 

Auburn  University 

8300  State  Highway  104 

Fairhope,  Alabama  36532 

E-mail  address:  sszedlmas'acesag  auburn.edu 

Jason  D.  Lee 

Barry  Vittor  &  Associates 
8060  Cottage  Hill  Rd. 
Mobile,  Alabama  36695 


Manuscript  approved  lor  publication 
4  November  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
..1  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:366-375  (200  1 1 


Larval  red  snapper  (Lutjanus  cam- 
pechanus) spend  approximately  26  days 
in  the  plankton,  prior  to  metamorpho- 
sis and  first  appearance  on  benthic 
substrate.  For  the  most  part  the  fish 
settle  on  open  substrate,  where  peaks 
in  recruitment  are  observed  in  August 
and  September,  after  which  they  may 
move  to  more  structured  habitat  some- 
time within  the  first  year  ( Szedimayer 
and  Conti,  1999).  The  apparent  advan- 
tage of  this  habitat  shift  would  be 
increased  food  resources  and  protec- 
tion from  predators.  To  help  clarify  the 
value  of  increased  food  resources  on 
reef  habitats,  comparisons  of  diets  from 
the  two  habitats  are  necessary.  Also, 
because  many  fish  species  shift  diets 
with  increasing  size  (Sedberry  and 
Cuellar,  1993;  Burke,  1995;  Rooker, 
1995;  Lowe  et  al.,  1996),  we  need  to 
distinguish  possible  ontogenetic  diet 
differences  from  shifts  that  are  due  to 
habitat. 

Previous  red  snapper  diet  studies 
have  focused  on  larger  individuals  and 
on  small  sample  sizes  for  fish  <250  mm 
SL  (Camber,  1955;  Moseley,  1966; 
Bradley  and  Bryan,  1975).  Camber 
(1955)  described  the  diets  of  15  "small 
red  snapper"  from  Campeche  Banks, 
and  reported  that  14  of  the  15  stom- 
achs contained  small  penaeid  shrimps. 
Moseley  ( 1966)  described  the  diets  of  45 
"juvenile  red  snapper"  collected  off  the 
coasts  of  Texas,  and  28  off  Louisiana. 
Louisiana  fish  fed  on  fishes,  shrimps. 


detritus,  and  stomatopods,  and  Texas 
fish  fed  on  shrimps,  crabs,  and  mysid 
shrimps. 

Perhaps  the  most  comprehensive  red 
snapper  diet  study  to  date  has  been 
that  of  Bradley  and  Bryan  ( 1975)  which 
described  the  diets,  by  season,  of  trawl- 
collected  (open  sand-mud  habitat)  and 
hook-and-line  reef  "rough  bottom  ar- 
eas" fish  off  the  Texas  coast.  They  de- 
scribed the  diets  of  258  open-habitat 
and  190  reef  red  snapper  and  found 
that  juvenile  red  snapper  (25-325  mm 
FL)  were  dependent  on  shrimp,  crabs, 
and  other  crustaceans  and  that  adults 
(325-845  mm  FL)  were  dependent  on 
fish,  crabs,  and  other  crustaceans. 
They  described  a  change  in  juvenile 
red  snapper  diet  as  fish  size  increased, 
"young  red  snapper  depend  almost 
exclusively  upon  invertebrates,"  and 
showed  a  gradual  increase  in  verte- 
brate prey  with  growth.  However,  they 
did  not  separate  out  the  proportions  of 
their  "juvenile"  red  snapper  that  were 
collected  from  reef  versus  open  habi- 
tat. Thus,  the  shift  from  open  to  reef 
habitat  is  still  poorly  understood.  If  and 
when  this  shift  occurs  and  whether  this 
shift  is  accompanied  with  a  diet  shift 
that  is  independent  of  fish-size  effects 
needs  to  be  defined. 

The  purpose  of  the  present  study- 
is  to  describe  the  diet  of  red  snapper 
off  the  coast  of  Alabama  —  from  the 
juvenile  stage  (just  after  settlement  I  to 
one-year  old  fish.  We  examined  overall 


Szedlmayer  and  Lee:  Diet  shifts  of  Lutjanus  campechanus 


367 


ontogenetic  shifts  in  red  snapper  diet  with  increasing  size 
and  possible  changes  in  diet  with  habitat  shifts  from  open 
substrate  to  structured  habitat  (artificial  reefs). 


Materials  and  methods 

Red  snapper  were  collected  from  open-flat  substrate  (sand 
and  mud)  and  reef  habitats  (artificial  reefs;  Fig.  1).  The 
open  habitat  was  located  approximately  6  km  south  of 
Mobile  Bay,  Alabama  (30'06'N,  88°03'W),  and  ranged 
in  depth  from  12  to  20  m.  Previous  studies  showed  very 
high  concentrations  of  age- 0  red  snapper  from  these  areas 
(Szedlmayer  and  Shipp,  1994;  Szedlmayer  and  Conti, 
1999 ).  The  artificial  reef  habitats  were  located  in  the  Hugh 
Swingle  artificial  reef  area,  approximately  20  km  south 
of  Mobile  Bay,  AL,  and  ranged  in  depth  from  18  to  23  m 
(Szedlmayer  and  Shipp,  1994;  Szedlmayer,  1997). 

We  collected  fish  from  open  substrate  by  trawl  (7.62-m 
head  rope,  2.54-cm  mesh,  2-mm  codend  mesh).  Samples 
were  taken  every  two  weeks  from  June  to  December 
1994;  however,  time  between  samples  was  longer  in  the 
winter  because  of  poor  weather.  Each  trawl  was  fished 
for  10  min,  and  all  age-0  and  age-1  red  snapper  collected 
were  placed  on  ice,  returned  to  the  laboratory,  and  frozen 
for  later  analysis.  Bottom  dissolved  oxygen,  salinity,  and 
temperature  were  sampled  with  a  Hydrolab  Surveyor  II  at 
each  location  (Szedlmayer  and  Conti,  1999). 

Prior  to  diet  analysis,  red  snapper  were  thawed,  weighed 
to  the  nearest  0.1  g,  and  measured  to  the  nearest  0.1  mm 
SL.  The  whole  fish  was  preserved  in  10%  formalin  if  SL 
was  <50  mm,  whereas  for  larger  fish,  stomachs  were  re- 
moved and  preserved.  After  48  hours  in  formalin,  stomach 
samples  were  transferred  to  75%  isopropyl  alcohol. 

Concrete  block  and  PVC  artificial  reefs  (1  m3)  were 
placed  in  the  Hugh  Swingle  reef  area  in  August  1992 
and  July  1993  (Szedlmayer,  1997).  "Reef  is  used  here  for 
defining  these  artificial  habitats.  Reefs  were  not  sampled 
for  a  minimum  of  3  months  after  placement.  Red  snapper 
were  collected  from  June  1993  through  December  1994. 
Fish  were  collected  from  these  reefs  by  SCUBA  divers 
first  placing  a  drop  net  (3.0  m  radius,  1.3  cm  square  mesh) 
over  the  reef  and  then  releasing  rotenone  into  the  enclosed 
area.  Reef  fish  were  placed  on  ice  in  the  field  and  trans- 
ported back  to  the  laboratory.  Approximately  12-18  h 
after  collection  all  reef  fish  were  weighed  to  the  nearest 
0.1  g  and  measured  to  the  nearest  1.0  mm.  Stomachs  were 
fixed  in  10%  formalin,  and  after  24  h  transferred  to  75%' 
isopropyl  alcohol.  Red  snapper  size  classes  were  also  esti- 
mated by  SCUBA  visual  surveys  in  July  and  August  1994. 
On  each  visual  survey,  divers  counted  red  snapper  by  50- 
mm  size  intervals.  Bottom  dissolved  oxygen,  salinity,  and 
temperature  were  sampled  with  a  Hydrolab  Surveyor  II 
during  each  survey. 

All  stomachs  were  dissected  and  contents  placed  in  petri 
dishes.  All  prey  were  counted  and  identified  to  the  lowest 
possible  taxon.  Volume  was  calculated  by  using  an  adapta- 
tion of  the  method  described  by  Hellawell  and  Able  (1971). 
Each  prey  taxon  from  each  stomach  was  placed  into  a  glass 
well  of  a  known  depth.  A  cover  slide  was  placed  on  the  well, 


Mobile 
Bay 


X 


If 


FL 


Gulf  of  Mexico 


O 


Gulf  of  Mexico 


O 


o 


o 


30.00  N 


□    □ 


DnaDnr 


0  5  10  Kilometers 


80.00  W 


Figure  1 

Collection  sites  for  red  snapper  ( Lutjanus  campechanus)  in  the 
northern  Gulf  of  Mexico.  Open  circles  are  open  habitat  trawl 
sites,  and  gray  squares  are  1-m3  concrete  or  PVC  artificial  reefs. 


depressing  the  prey  taxon  to  a  known  depth  (e.g.,  1  mm). 
The  prey  were  video  taped  with  a  high-8  Sony  camera 
and  images  were  digitized  with  Image  Pro  2.0  software 
(Media  Cybernetics.  Silver  Spring,  MD).  Image  size  was 
calibrated  to  0.01  mm  by  a  stage  micrometer.  The  surface 
area  of  each  preparation  was  measured  by  using  Image 
Pro  software.  Volume  was  calculated  by  multiplying  the 
surface  area  by  the  known  depth.  Specific  volumes  for  par- 
ticular prey  taxa  were  calculated  by  dividing  prey  volume 
by  individual  fish  weight  (mnv'Vfish  wt  g).  Comparisons  of 
diet  shift  by  increasing  fish  size  were  made  by  grouping 
prey  taxa  into  ten  prey  groups  and  by  calculating  specific 
volume  for  10-mm-size  intervals  of  red  snapper. 

A  dissimilarity  index  (Bray-Curtis)  was  calculated 
from  specific  volumes  of  individual  prey  taxa,  for  overlap- 
ping size  classes  of  red  snapper  both  within  and  between 
habitats:  Bray-Curtis  =  IW^-Yj/IiY^+Y^),  where  Y  = 
specific  volume  of  jth  species,  and  j  and  k  are  the  samples 
being  compared  (Field  et  al.,  1982).  The  dissimilarities 
were  then  used  in  a  multidimensional  scaling  analysis 
(MDS;  Schiffman  et  al.,  1981).  The  MDS  provided  a  two- 
dimensional  "map"  of  the  distances  between  samples  (fish 


368 


Fishery  Bulletin  102(2) 


6 
4 
2 
0 
8 
£  6 
!     4 

O 

0  2 

1  0 

O 

E     4 

E 

g     2 

!  o 

o 

6       6 

C 

yj      4 

0_ 

o  2 
0 
6 
4 
2 
0 


Trawls  -  1 1 
28  Jun-1  Jul 


-d 


Reefs  =  5 
25  May-2  Jun 


ELi 


Trawls  =  31 
10-19  Jul 


Visual 

Reefs  =  13 

1-25  Jul 


L 


Trawls  =  32 
31  Jul-12  Aug 


Visual 

Reefs  =  ! 

9  Aug 


J_L 


Trawls  =  26  Reefs  =  2 

•23  Aug  23  Aug 


1  id1 

^t         Trawlj 


Trawls  =  32 

9  SeP  Reefs  =  3 

6  Sep 


JTlhffl 


Trawls  =  25 
25-28  Sep 


Reefs  =  4 
28  Sep-6  Oct 


II 


Trawls  =  25 
12-21  Oct 


Reefs  =  4 
19-24  Oct 


Jktft^f^n 


Trawls  =  26 
30  Oct-9  Nov 


Reefs  =  3,  9-18  Nov 


^bjMkffl 


Trawls  =  15 
12-16  Dec 


Reefs  =  6 
5-8  Dec 


IT>^ 


Trawls  =  26 
9-12  Jan 


£\ 


Reefs  =  3 
25Mar-15  Apr 


100  200  100 

Standard  length  (mm) 


200 


Figure  2 

Movement  patterns  for  age-0  red  snapper  iLutjanus  campechanus)  from  the 
northern  Gulf  of  Mexico  in  1994.  Black  bars  represent  trawl  samples,  grey  bars 
represent  reef  drop-net  samples,  and  white  bars  represent  SCUBA  visual  surveys 
of  concrete  reefs. 


size  and  habitat  type)  in  Euclidian  space  based  on  the 
Bray-Curtis  index.  Thus,  comparisons  of  red  snapper  diets 
were  based  on  all  prey  taxa,  yet  independent  of  capture 
habitat  and  fish  size. 


Results 

In  the  sampling  areas  during  the  summer  and  fall  of  1994, 
salinity  ranged  from  30  to  35  ppt.  Dissolved  oxygen  was 
7  ppm  in  the  early  summer,  decreased  to  3  ppm  in  July  and 
August,  and  increased  to  7  ppm  in  the  fall.  Temperature 
was  22°C  in  June,  increased  to  28°C  in  late  August,  then 
dropped  to  just  below  20°C  by  December.  No  significant 
differences  were  detected  between  trawl  and  reef  sites  for 
these  environmental  measures  U-test,  Ps0.05). 

Red  snapper  showed  a  clear  shift  in  habitat  during  their 
first  few  months  of  life  (Fig.  2).  Fish  first  recruited  to  open 


habitat  at  the  end  of  June,  at  sizes  <40  mm  SL.  Fish  con- 
tinued to  recruit  to  open  habitat  until  early  September,  at 
which  time  they  were  larger  ( 30  to  100  mm  SL )  and  began 
shifting  to  more  structured  habitat.  By  mid-October  most 
age-0  fish  had  moved  to  reef  habitat.  During  the  initial 
settlement  no  new  recruits  were  collected  or  visually  ob- 
served on  the  artificial  habitats  (Fig.  2).  Overall,  only  red 
snapper  <160  mm  SL  were  collected  from  open  habitat, 
whereas  only  red  snapper  >70  mm  SL  were  collected  from 
reef  habitat.  Size  overlapped  from  70.0  to  160  mm  SL  be- 
tween habitats  (Fig.  3). 

A  total  of  1639  red  snapper  stomachs  were  analyzed: 
570  from  open  substrate  and  1069  from  reef  habitat.  Prey 
were  found  in  789  (48'<  )  of  the  total  stomachs  examined, 
346  (61%)  from  the  open  habitat  and  443  (41%)  from  the 
reef  habitat  (Fig.  3).  Trawl-collected  red  snapper  were 
mostly  collected  from  site  one,  but  sample  sizes  were  also 
large  (>30  with  prey)  at  two  other  sites  (Table  1).  Total  red 


Szedlmayer  and  Lee:  Diet  shifts  of  Lut/anus  campechanus 


369 


Open  Habitat  n=346 
Reef  Habitat  n=443 
Empty  n=850 


QD=- 


■i~~ r    i    T    i     i 
60      80     100    120    140    160    180   200   220   240   260   280   300 
Size  class  (mm  SL) 

Figure  3 

A  comparison  of  red  snapper  {Lutjanus  campechanus)  length  frequencies  between 
open  and  reef  habitats  in  the  northern  Gulf  of  Mexico.  Gray  bars  =  empty  stomachs 
from  both  habitats. 


Table  1 

Number  of  red  snapper  (Lutjanus  campechanus)  stom- 
achs sampled  in),  and  number  of  stomachs  containing 
prey  from  open  and  reef  habitat  in  the  northeast  Gulf  of 
Mexico. 

Open  trawl  sites 

Reef  habitats 

n                   n  with  prey 

n 

n  with  prey 

356                    223 

108 

53 

45                        21 

17 

5 

75                       58 

198 

115 

57                       33 

55 

31 

37                       11 

249 

71 

50 

23 

14 

1 

89 

45 

11 

5 

209 

74 

35 

10 

22 

4 

12 

6 

snapper  collected  from  the  reefs  varied  by  site  (from  11  to 
249  fish),  but  large  samples  were  collected  from  at  least  6 
different  reefs  (Table  1).  Large  sample  sizes  were  collected 
during  most  months  over  open  habitat,  with  the  exception 
of  November  1994  (n=12),  and  for  most  months  (6  out 


Table  2 

Number  of  red  snapper  (Lutjanus  campechanus)  stomachs 
sampled  (re),  and  number  containing  prey,  by  month  and 
year,  from  open  and  reef  habitat  in  the  northeast  Gulf  of 

Mexico. 

Open 

habitat 

Reef  habitat 

Month 

r 

with 

Month 

n  with 

and  year 

n 

prey 

and  year 

n 

prey 

Jul    1994 

56 

43 

Jun 

1993 

94 

50 

Aug  1994 

169 

109 

Oct 

1993 

370 

169 

Sep  1994 

187 

98 

May 

1994 

141 

37 

Oct  1994 

97 

52 

Jun 

1994 

46 

37 

Nov  1994 

16 

12 

Aug 

1994 

41 

8 

Dec  1994 

45 

32 

Sep 

1994 

155 

86 

Oct 

1994 

76 

28 

Nov 

1994 

65 

12 

Dec 

1994 

81 

16 

of  9)  from  reef  sites  (Table  2).  Only  red  snapper  stomachs 
containing  prey  were  used  in  our  analyses. 

Red  snapper  diets  showed  55  different  prey  identi- 
fied to  the  lowest  possible  taxon.  In  general,  red  snap- 
per diets  were  dominated  by  fish  (43%),  squid  (29.5%), 
shrimp  (16.4%),  and  crabs  (4.4%;  Table  3).  Specifically, 
the  "shrimp"  group  included  Mysidacea  (mysid  shrimps), 
Stomatopoda  (mantis  shrimps),  Penaeidea  (penaeid 


370 


Fishery  Bulletin  102(2) 


Table  3 

Specific  volume  (mnvVfi 

sh  weight  g)  for  prey 

taxa  from  red 

snapper  (Lutjanus  compel 

ha n us  1.  %  =  percent 

specific-voli 

lme  of  total 

volume,  Habitat  =  prey 

habitat.  General  prey  groups  are  noted  in  quotation  marks,  unid.  =  unidentified. 

Prey  taxa 

Total  volume 

Percent 

Lowest  taxon 

Specific  volume 

Percent 

Habitat 

Osteichthyes  "fish" 

5408.2 

43.5 

unid.  fish 

3465.9 

27.9 

Halichoeres  spp. 

650.4 

5.2 

reef 

Blenniidae 

279.2 

2.2 

reef 

Serranidae 

278.1 

2.2 

reef 

Serranus  subligarius 

240.8 

1.9 

reef 

Centropristis  ocyurus 

207.3 

1.7 

reef 

Engraulidae 

117.9 

0.9 

open 

Ophichthidae 

100.6 

0.8 

open 

Cynoglossidae 

35.2 

0.3 

open 

Triglidae 

20.8 

0.2 

open 

Ophichthus  sp. 

10.8 

0.1 

open 

Cephalopoda  "squid" 

3665.6 

29.5 

Loliginidae 

3665.6 

29.5 

open 

Natantia  "shrimp" 

2033.7 

16.4 

unid.  shrimp 

544.6 

4.4 

Sicyoninae 

359.6 

2.9 

reef 

Hippolytidae 

345.7 

2.8 

reef 

Penaeidae 

264.5 

2.1 

open 

Alpheidae 

131.1 

1.1 

reef 

Sergestidae 

24.2 

0.2 

open 

Luciferinae 

22.6 

0.2 

open 

Ogyrididae 

8.8 

0.1 

open 

Stomatopoda  "shrimp" 

Squillidae 

221.8 

1.8 

open 

Mysidacea  "shrimp" 

Mysidacea 

109.8 

0.9 

open 

Reptantia  "crabs" 

550.8 

4.4 

Portunidae 
unid.  crab 

302.0 
143.0 

2.4 
1.2 

mixed 

Diogeninae 

51.6 

0.4 

open 

Leucosiidae 

20.7 

0.2 

reef 

Xanthidae 

16.7 

0.1 

reef 

Porcellanidae 

7.3 

0.1 

reef 

Chaetognatha 

199.6 

1.6 

Sagitta  spp. 

199.6 

1.6 

open 

Polychaeta 

130.1 

1.0 

Polycheata 

75.4 

0.6 

mixed 

Polychaeta 

Onuphidae 

34.0 

0.3 

open 

Maldanidae 

19.9 

0.2 

open 

Ascidiacea  "tunicate" 

121.0 

1.0 

Ascidiacea 

121.0 

1.0 

reef 

Calanoida  "copepod" 

118.2 

1.0 

Calanoida 

113.3 

0.9 

open 

Octopodidae 

93.6 

0.8 

Octopus  sp. 

93.6 

0.8 

reef 

unid. 

79.5 

0.6 

unid. 

79.5 

0.6 

Amphipoda 

13.8 

0.1 

Amphipoda 

9.4 

0.1 

mixed 

Ostracoda 

6.1 

0.0 

Ostracoda 

6.1 

0.0 

open 

shrimps),  and  Caridea  (caridean  shrimps).  In  addition, 
all  Squillidae  were  probably  Squilla  empusa,  according 
to  Hopkins  et  al.,  (1987).  Among  fish,  many  were  uniden- 
tified due  to  digestion,  but  if  proportions  of  unidentified 
fish  are  similar  to  identified  fish,  then  dominant  fish 
prey  included  Halichoeres  spp.,  (5.2c/r ),  Blenniidae  (2.2%), 
and  Serranidae  (2.2%).  Two  prey  fish  were  identified  to 
species:  Serranus  subligarius  (1.9%),  and  Centropristis 
ocyurus  (1.7%). 


Among  the  squid  taxon,  one  genus  dominated:  Lolli- 
guncula  spp.,  (29.59!  I,  but  all  squid  were  either  L.  brevis 
or  Loligo  pealeii  (Hopkins  et  al.,  1987).  Among  shrimp, 
dominant  taxa  included  Sicyoninae  (2.9%),  Hippolytidae 
(2.8%),  Penaeidae  (2.1%),  Squillidae  (1.8%),  and  Alpheidae 
(1.1%).  Among  crabs,  dominant  taxa  were  mostly  Por- 
tunidae (2.49; ).  Other  groups  showing  greater  than  1.0% 
included  Chaetognatha  [Sagitta  sp.  1.69J ),  and  Ascidiacea 
or  tunicates  (1.0%;  Table  3). 


Szedlmayer  and  Lee:  Diet  shifts  of  Lutjanus  campechanus 


371 


16  48  45  41    37  41   33  35  37  59  67   37  40  37  41    28  28  35  29  13    15  15    5     3 


0         20       40       60        80       100      120      140      160      180     200     220     240 

Size  class  (mm  SL) 

Figure  4 

Stomach  contents  by  specific-volume  for  ten  higher  taxonomic  groups 
over  10-mm  size  classes  of  red  snapper  {Lutjanus  campechanus)  from 
both  open  and  reef  habitats  in  the  northern  Gulf  of  Mexico.  Numbers 
on  the  upper  axis  are  the  number  of  red  snapper  that  contained  prey 
for  each  respective  size  class. 


Red  snapper  shifted  diets  with  increasing  size.  For  red 
snapper  <60.0  mm  SL,  diets  were  dominated  by  shrimp, 
chaetognaths,  squid,  and  copepods.  Large  red  snapper 
(60-280  mm  SL)  shifted  to  feeding  on  fish  prey,  greater 
amounts  of  squid  and  crabs,  and  continued  feeding  on 
shrimp  (Fig.  4). 

The  diets  of  juvenile  red  snapper  changed  as  they 
moved  from  open  to  reef  habitats.  Fish  collected  had 
overlapping  sizes  of  70.0  to  160.0  mm  SL  from  both  open 
and  reef  habitats,  and  the  MDS  analysis  for  this  size 
range  showed  a  clear  separation  of  diets  between  the  two 
habitats  (Fig.  5).  Two  points  that  were  outliers  (R75,  T155 1 
were  biased  because  they  represented  only  one  fish  each, 
and  the  third  outlier  (R85)  was  difficult  to  explain. 

The  clear  separation  of  red  snapper  diets  shown  by  the 
MDS  analysis  can  be  attributed  to  several  prey  shifts  that 
accompanied  habitat  shifts.  For  prey  crabs,  open-habitat 
red  snapper  diets  were  dominated  by  Xanthidae,  and 
smaller  amounts  of  Paguridae,  Portunidae,  Diogeninae, 
and  Pinnotheridae  (Fig.  6),  whereas  diets  of  red  snapper 
from  reef  habitats  shifted  to  a  dominance  by  Portunidae 
and  Diogeninae  ( Fig.  7).  For  prey  shrimp,  open  habitat  red 
snapper  diets  were  dominated  by  Penaeidae  and  Mysida- 
cea  (Fig.  8),  whereas  diets  from  reef  habitats  shifted  to  a 
dominance  of  Sicyoninae,  Hippolytidae,  Alpheidae,  and 
Squillidae  (Fig.  9).  For  prey  fish,  open-habitat  red  snap- 
per diets  were  dominated  by  Engraulidae  (although  most 
were  unidentified;  Fig.  10),  whereas  diets  from  reef  habitat 
clearly  reflected  prey  fish  from  reef  habitats  and  included 
Blenniidae,  Serranidae,  and  three  prey  fish  identified  to 
genera,  Centropristis  spp,  Halichoeres  spp.,  and  Sen-anus 
spp.  (Fig.  11). 


-3-2-1  0  1  2  3 

X(unitless) 

Figure  5 

Multidimensional  scaling  of  diets  for  red  snapper  (Lutja- 
nus campechanus)  based  on  the  Bray-Curtis  dissimilarity 
index  computed  for  specific  volume  of  prey  taxa  both  within 
and  between  habitats  for  overlapping  size  classes  (70.0  to 
159.9  mm  SL).  The  letter  and  number  accompanying  each 
point  indicates  the  habitat  and  size  class  that  each  point 
represents  (e.g.,  T  =  trawl,  R  =  reef,  75  =  75  mm  SL  size 
class).  Circles  were  drawn  by  hand.  Axes  are  unitless. 


Discussion 

The  present  study  provides  a  substantial  sample  size 
(ft  =  1639)  for  red  snapper  diet  analysis  and  a  relatively 


372 


Fishery  Bulletin  102(2) 


16    48  45  41    37   41    32   30  20  15 


6      3     3 


"I — I 1 r 


Crabs 
Open  habitat 


n 


T 1 1 1 1 — 

UNIDcrab 
l  l  Paguridae 
Xanthidae 
Portunidae 
Diogeninae 
I        I  Pinnotheridae 


a 


j=u 


t — T — T — T — T — T — T — T — ~l 1 1 f — T 

0         20        40        60        80       100      120      140 
Size  class  (mm) 

Figure  6 

Crab  prey  from  open  habitat.  Stomach  contents  by  spe- 
cific volume  over  10-mm  size  classes  of  red  snapper  iLut- 
janus  campechanus)  from  the  northern  Gulf  of  Mexico. 
Numbers  on  the  upper  axis  are  the  number  of  red  snapper 
that  contained  prey  for  each  respective  size  class. 


16  48   45  41    37   41    32  30   20  15     8     6     3     3 


0  4 
0 


1 1 1 ' 1 — 

UNID  shrimp 
Lucifennae 
Mysidacea 
I  I  Ogyndidae 
^^1  Penaeidae 
Sergestidae 
Sicyoninae 
I        I  Squillidae 


"Shrimp" 
Open  habitat 


- 


II 


1 


I 


I 


I 


nW 


20        40        60        80        100      120      140 
Size  class  (mm) 


Figure  8 

"Shrimp"  prey  from  open  habitat.  Stomach  contents  by 
specific  volume  over  10-mm  size  classes  of  red  snap- 
per (Lutjanus  campechanus)  from  the  northern  Gulf  of 
Mexico.  Numbers  on  the  upper  axis  are  the  number  of  red 
snapper  that  contained  prey  for  each  respective  size  class. 


5  17  44  59  31  37  34  40  28  28  35  29  13  15  15    5    3 

5     3  - 
m 

E 

I      I      I      I      1      I      1      I      I      1      I      I      I 

1        1  UNIDcrab 
Crabs                  1        1  Xanthidae 
Reef  habitat      ^m  Portunidae 
I        I  Diogeninae 
1        1  Pinnotheridae 
fSSg  Porcellanidae 
|::::::|  Leucosiidae 

i    i   i 

^E, 

"5     2- 
> 

<u 
n. 

Specific  volume 
i 

i-i 

_ 

lR.il 

1 

1 

■ 

r-] 

II 

u            1       1       1       1       1       1       1       1       1       1       1       1       1       1       1       1 

80      100     120     140     160     180     200     220     240 

Size  class  (mm) 

Figure  7 

Crab  prey  from  reef  habitat.  Stomach  contents  by  specific 

volume  over  10-mm  size  classes  of  red  snapper  (.Lutjanus 

campechanus)  from  the  northern  Gulf  of  Mexico.  Num- 

bers on  the  upper  axis  are  the  number  of  red  snapper  that 

contained  prey  for  each  respective  size  class. 

4  - 


5  17  44  59  31  37  34  40  28  28  35  29  13  15  15  5  3 

— ' 1 ' — I — 


"T" 


"T" 


"Shrimp" 
Reef  habitat 


I  I  Alpheidae 

I  I  UNID  shrimp 

I  I  Hippolytidae 

l  l  Sergestidae 

I  I  Sicyoninae 

I  I  Squillidae 


80      100     120     140     160     180    200    220    240 
Size  class  (mm) 

Figure  9 

"Shrimp"  prey  from  reef  habitat.  Stomach  contents  by  spe- 
cific volume  over  10-mm  size  classes  of  red  snapper  iLut- 
janus  campechanus)  from  the  northern  Gulf  of  Mexico. 
Numbers  on  the  upper  axis  are  the  number  of  red  snapper 
that  contained  prey  for  each  respective  size  class. 


Szedlmayer  and  Lee:  Diet  shifts  of  Lut/anus  campechanus 


373 


20 


15 


10 


16  48   45  41    37  41    32  30   20  15     8     6 


■      I      ■      I      ■      I 

Fish 

Open  habitat 

.     I        I  Cynoglossidae 
|^H  Engraulidae 
1        1  Ophichthidae 

I  1  Serranidae 

II  Synodontidae 
1         1  UNIDfish 

1        1  Triglidae 

i 

■, 

1 

i 
i 

20 


40        60        80 
Size  class  (mm) 


100      120      140 


Figure  10 

Fish  prey  from  open  habitat.  Stomach  contents  by  specific 
volume  over  10-mm  size  classes  of  red  snapper  (Lut/anus 
campechanus)  from  the  northern  Gulf  of  Mexico.  Num- 
bers on  the  upper  axis  are  the  number  of  red  snapper  that 
contained  prey  for  each  respective  size  class. 


1b  - 

— 

14  - 

CT 

- 

13  - 

<1> 

12  - 

< 

11  - 

</) 

~ 

10  - 

E 

q  - 

r 

8  - 

r, 

7  - 

>, 

Cl 

b  - 

<D 

R  - 

t- 

3 

4  - 

5   17  44  59  31  37  34  40  28  28  35  29  13  15  15   5    3 


3  - 


Fish 
Reef 

Habitat 


Blennndae 

Centropristis 

Halichoeres 

Ophichthidae 

Serranidae 

Serranus 

Triglidae 

unid 


HI 


o 


80      100     120     140     160     180     200     220     240 
Size  class  (mm) 

Figure  11 

Fish  prey  from  reef  habitat.  Stomach  contents  by  spe- 
cific volume  over  10-mm  size  classes  of  red  snapper 
(Lutjanus  campechanus)  from  the  northern  Gulf  of 
Mexico.  Numbers  on  the  upper  axis  are  the  number  of  red 
snapper  that  contained  prey  for  each  respective  size  class. 


high  percentage  of  stomachs  with  food  (48%)  compared  to 
past  studies.  Rooker  ( 1995 )  also  showed  a  high  percentage 
(69%;  312  out  of  449  stomachs)  of  schoolmaster  snapper 
(Lutjanus  apodus)  contained  prey,  when  fish  were  col- 
lected from  depths  similar  to  those  of  the  our  study  ( 1  to  27 
ml.  The  higher  percentage  of  stomachs  with  prey  found  in 
our  study  compared  to  past  studies  of  red  snapper  ( Stea- 
rns, 1884;  Camber.1955;  Moseley,  1966)  may  be  due  to  the 
shallower  depths  sampled  (18  m;  DeMartini  et  al.,  1996). 
Juvenile  red  snapper  showed  feeding  patterns  similar 
to  many  other  marine  fishes.  After  settlement,  from  ap- 
proximately 20  to  60  mm  SL,  they  showed  a  wide-ranging 
diet  that  included  shrimp,  copepods,  chaetognaths,  and 
squid.  Prey  fish  were  also  found  in  the  stomachs  of  the 
smallest  red  snapper  collected  (15-20  mm  SL)  but  were 
not  a  dominant  component.  Sweatman  (1993)  reported 
similar  results  for  the  snapper  Lutjanus  quinquelineatus, 
ranging  from  24  to  29  mm  SL,  i.e.,  piscivorous  in  the  first 
few  days  after  settlement.  Above  60  mm  SL,  fish  prey 
tended  to  dominate  specific  volume,  but  not  by  feeding  less 
on  shrimp  because  shrimp  continued  to  be  an  important 
prey.  Squid  became  another  dominant  component  of  red 
snapper  diet  at  about  100  mm  SL  and  also  continued  as  an 
important  prey  up  to  240  mm  SL.  Unfortunately,  sample 
size  was  reduced  above  230  mm  SL,  and  it  was  difficult 
to  estimate  if  squid  and  fish  continued  as  dominant  prey 
components  above  these  size  classes.  Sedberry  and  Cuellar 
( 1993 )  reported  a  similar  shift  in  diets  of  reef-associated 
vermilion  snapper  (Rhomboplites  aurorubens).  This  spe- 


cies shifted  from  small  crustaceans  to  fishes  and  cepha- 
lopods  over  a  size  range  similar  to  that  of  red  snapper  in 
the  present  study.  Moseley  ( 1966 )  reported  a  "slow  transi- 
tion from  zooplankton  to  macro  animals  for  red  snapper 
sizes  between  40  and  90  mm" — a  transition  that  probably 
included  fish  prey  that  he  did  not  specifically  identify. 
Bradley  and  Bryan  (1975),  showed  a  shift  in  juvenile  red 
snapper  diets  with  size  (25-325  mm  FL).  Their  smallest 
red  snapper  keyed  on  invertebrates,  then  showed  a  sharp 
increase  in  dependency  upon  prey  fish  above  175  mm 
FL,  when  Batrachoididae  (toadfish)  became  a  dominant 
component.  These  shifts  in  diet  are  important  in  helping 
to  identify  fish  habitat  and  are  potentially  key  aspects  of 
early  survival. 

Red  snapper  showed  two  major  habitat  shifts  in  their 
first  year.  Juvenile  red  snapper  first  settled  from  the 
plankton  to  benthic  substrate  near  20  mm  SL  ( Szedlmayer 
and  Conti,  1999 ).  The  present  study  showed  a  second  shift 
from  open  habitat  to  reef  habitat  starting  at  about  70  mm 
SL  (Fig.  3).  No  fish  smaller  than  70  mm  were  collected 
from  the  reefs,  and  smaller  red  snapper  were  rarely  ob- 
served on  these  reefs  from  SCUBA  visual  surveys.  No  fish 
larger  than  160  mm  SL  were  caught  from  the  open  habitat 
but  were  present  on  the  reefs.  This  finding  suggested  that 
red  snapper  had  shifted  to  reef  habitat  by  160  mm  SL  but 
also  may  have  avoided  trawl  gear  as  described  earlier  for 
age-0  red  snapper  (Bradley  and  Bryan,  1975)  and  age-0 
summer  flounder  (Paralichthys  dentatus)  (Szedlmayer 
and  Able,  1993).  However,  no  large  (150-300  mm)  red 


374 


Fishery  Bulletin  102(2) 


snapper  were  observed  over  open  habitat  by  a  SCUBA 
visual  survey  despite  our  observations  that  red  snapper 
are  attracted  to  SCUBA  divers.  Thus  we  suggest  that  a 
shift  in  habitat  was  more  likely  the  cause  of  this  absence 
than  trawl  avoidance. 

The  distinct  diet  shift  as  red  snapper  changed  habitats 
was  independent  of  increasing  size  and  suggested  that 
different  benthic  habitats  play  a  critical  role  in  the  early 
life  history  of  this  species.  This  separation  was  completely 
independent  of  "a  priori"  knowledge  of  sample  location  and 
fish  size.  For  example,  the  MDS  analysis  showed  almost 
complete  separation  based  on  habitat  rather  then  fish  size 
(Fig.  5).  These  differences  between  open  and  reef  habitat 
were  readily  apparent  when  prey  taxa  were  separated 
into  lower  taxonomic  categories.  For  example,  fishes  such 
as  Halichoeres  spp.,  Serranus  spp.,  and  Centr-op/istis  spp., 
were  found  only  in  the  diets  of  reef-collected  red  snapper. 
These  species  are  closely  tied  to  reef  structure  (Nelson 
and  Bortone,  1996).  Prey  shrimp  also  showed  distinct 
differences  in  red  snapper  diets  between  habitats.  Over 
open  habitat,  Mysidacea,  Penaeidae,  and  Sergestidae  were 
important  components.  After  the  shift  to  reef  habitat, 
Mysidacea  were  absent  and  Penaeidae  and  Sergestidae 
were  greatly  reduced,  and  Sicyoninae,  Hippolytidae,  and 
Alpheidae  became  the  dominant  shrimp  components. 
The  latter  are  all  families  typically  associated  with  reef 
habitats  (Chance,  1970;  Pequegnat  and  Heard,  1979). 
One  exception  was  the  increased  feeding  on  Squillidae,  an 
open  habitat  crustacean,  at  the  largest  size  classes  of  this 
study  (220-250  mm  SL;  Fig.  9).  For  crabs,  the  separation 
was  not  as  clear,  because  of  the  dominance  of  Portunidae, 
which  can  be  assigned  to  both  open  and  reef  habitats. 
However,  increases  in  reef  crabs  were  still  apparent  with 
habitat  shift,  i.e.,  Diogeninae,  Porcellanidae,  and  Leucosi- 
idae  can  all  be  considered  reef  prey.  Although  Bradley  and 
Bryan  (1975)  pooled  "juvenile"  red  snapper  over  open  and 
reef  habitats,  they  did  show  a  marked  increase  in  fish  prey 
above  175  mm  FL.  This  increase  was  almost  exclusively 
due  to  Batrachoididae  or  toadfishes,  which  are  typically 
found  in  reef  habitat.  We  did  not  observe  any  toadfish  prey 
in  our  juvenile  red  snapper  collections,  but  its  presence  in 
this  earlier  study  is  consistent  with  present  findings  show- 
ing a  shift  to  feeding  on  reef-habitat  prey. 

Red  snapper  diet  shifted  to  greater  percentages  of  reef- 
prey  with  movement  to  reef  habitat,  but  with  this  shift 
they  also  continued  feeding  on  other  prey.  This  flexibility 
in  feeding  habits  allows  red  snapper  to  take  advantage 
of  prey  from  wide-ranging  habitats.  Similar  diet  shifts 
related  to  habitat  shifts  have  been  shown  in  schoolmaster 
snapper,  (L.  apodus)  (Rooker,  1995).  The  schoolmaster 
snapper  shifted  from  nearshore  mangroves  to  coral  reef 
habitats  near  70  mm  SL;  diets  offish  s70.0  mm  SL  were 
dominated  by  crustaceans,  particularly  amphipods  and 
crabs.  Fish  >70.0  mm  SL  fed  on  fishes  and  to  a  lesser  ex- 
tent crabs,  shrimps,  and  stomatopods.  Similar  diet  shifts 
were  also  shown  for  several  fish  species  of  Puget  Sound. 
For  example  in  pile  perch  {Rhacochilus  vacca),  striped 
seaperch  iEmbiotoca  lateralis),  and  quillback  rockfish 
(Sebastes  maliger),  the  smallest  juveniles  preyed  on  open- 
habitat  plankton  and  benthic  fauna,  and  medium-size 


and  larger  fish  (>121  mm)  of  all  three  species  shifted  their 
diets  to  include  reef-associated  prey  ( Hueckel  and  Stayton, 
1982).  However,  at  larger  sizes  these  three  species  were 
not  totally  dependent  on  reef-associated  prey. 

We  have  examined  red  snapper  diets  based  on  specific 
volume  of  food.  Although  many  other  studies  have  used 
an  index  of  relative  importance  (IRI;  Pinkas  et  al.,  1971: 
Cortes,  1997),  we  were  specifically  interested  in  the  nutri- 
tional value  of  particular  prey,  and  prey  separation  into 
open-habitat  or  reef-habitat.  With  IRIs  these  separations 
would  be  more  difficult  to  define,  e.g.,  pelagic  prey  with 
high  numbers  might  be  considered  more  important,  but 
actually  provide  little  nutritional  value  to  red  snapper 
diets  (Macdonald  and  Green,  1983).  Future  studies  on  the 
effects  of  red  snapper  predation  on  prey  distributions  may 
be  better  suited  for  using  IRIs. 

In  summary,  red  snapper  diets  from  open  habitat 
showed  prey  taxa  associated  with  open  sand-mud  sub- 
strate and  the  planktonic  environment.  Open-habitat  prey, 
such  as  chaetognaths,  are  known  to  be  pelagic  as  well  as 
benthic,  as  are  sergestid  shrimp,  calanoid  copepods,  my- 
sids,  and  stomatopods  (Williams,  1968;  Manning,  1969; 
Gosner,  1978;  Stuck  et  al.,  1979;  Alldredge  and  King, 
1985;  Lindquist  et  al.,  1994 ).  Red  snapper  shifted  diets  to 
reef-associated  prey  with  their  habitat  shift,  and  this  diet 
shift  was  independent  offish  size.  These  diet  shifts  were 
clearly  apparent  for  both  fish  and  shrimp  prey  but  less 
so  for  crab  prey.  As  shown  with  marine  fish  species  from 
Puget  sound,  red  snapper  diets  from  reef  habitat  were  not 
restricted  to  reef-associated  prey.  For  example,  squids 
were  an  important  prey  over  both  open  and  reef  habitats 
in  the  present  study  and  our  findings  agree  with  those 
of  Bradley  and  Bryan  (1975).  The  squids  Loligo  sp.,  and 
Lolliguncula  sp.  are  both  abundant  in  nearshore  coastal 
waters  and  are  not  typically  associated  with  reef  structure 
(Gosner  1978;  Laughlin  and  Livingston,  1982;  Hopkins 
et  al.,  1987).  Availability  and  ease  in  capture  could  be  a 
key  as  to  why  squid  are  important  for  red  snapper  over 
size  ranges  of  40  to  240  mm  SL.  This  flexibility  in  feeding 
habits  allows  red  snapper  to  take  advantage  of  prey  from 
wide-ranging  habitats,  but  clear  shifts  to  additional  reef 
prey  supports  the  hypothesis  that  reef  structure  provides 
new  prey  resources. 


Acknowledgments 

We  thank  Joseph  Conti,  Kori  M.  Heaps,  and  Frank  S. 
Rikard  for  help  in  field  collections  and  invertebrate 
identification.  This  study  was  funded  by  NOAA,  NMFS. 
MARFIN  grant  number  USDC-NA47FF0018-0.  This  is  a 
contribution  of  the  Alabama  Agricultural  Station. 


Literature  cited 

Alldredge,  A.  L.,  and  J.  M.  King. 

1985.  The  distance  demersal  zooplankton  migrate  above 
the  benthos:  implications  for  predation.  Mar.  Biol. 
84:253-260. 


Szedlmayer  and  Lee:  Diet  shifts  of  Lut/anus  campechanus 


375 


Bradley,  E.,  and  C.  E.  Bryan. 

1975.  Life  history  and  fishery  of  the  red  snapper  {Lutja- 
nus campechanus)  in  the  northwestern  Gulf  of  Mexico: 
1970-1974.  In  Proceedings  of  the  twenty-seventh  Gulf 
and  Caribbean  Fisheries  Institute,  p.  77-106.  Univ. 
Miami,  Miami.  FL. 

Burke,  J.  S. 

1995.  Role  of  feeding  and  prey  distribution  of  summer 
and  southern  flounder  in  selection  of  estuarine  nursery 
habitats.     J.  Fish  Biol.  47:355-366. 

Camber,  C.  I. 

1955.  A  survey  of  the  red  snapper  fishery  of  the  Gulf 
of  Mexico,  with  special  reference  to  the  Campeche 
Banks.  Florida  State  Board  of  Conservation  Technical 
Series  12:1-64. 

Chance,  F  A.,  Jr. 

1970.  A  new  shrimp  of  the  genus  Lysmata  (Decapoda: 
Hippolytidae)  from  the  western  Atlantic.  Crustaceana 
19:59-66. 

Cortes,  E. 

1997.  A  critical  review  of  methods  of  studying  fish  feeding 
based  on  analysis  of  stomach  contents;  application  to  elas- 
mobranch  fishes.     Can.  J.  Fish.  Aquat.  Sci.  54:726-738. 

DeMartini.  E.  E..  F.  A.  Parrish,  and  D.  M.  Ellis. 

1996.  Barotrauma-associated  regurgitation  of  food:  impli- 
cations for  diet  studies  of  Hawaiian  pink  snapper.  Pristi- 
pomoides  filamentosus  (family  Lutjanidaei.  Fish.  Bull. 
94:250-256. 

Field,  J.  G..  K.  R.  Clarke,  and  R.  M.  Warwick. 

1982.     A  practical  strategy  for  analyzing  multispecies  dis- 
tribution patterns.     Mar.  Ecol.  Prog.  Ser.  8:37-52. 
Gosner,  K.  L. 

1978.     A  field  guide  to  the  Atlantic  seashore,  329  p.     Hough- 
ton Mifflin  Company,  Boston,  MA. 
Hellawell.  J.  M..  and  R.  Able. 

1971.  A  rapid  volumetric  method  for  the  analysis  of  the 
food  of  fishes.     J.  Fish  Biol.  3:29-37. 

Hopkins.  T.  S.,  J.  F  Valentine,  and  L.  B.  Lutz. 

1987.     An  illustrated  guide  with  key  to  selected  benthic 
invertebrate  fauna  of  the  Northern  Gulf  of  Mexico.     Mis- 
sissippi Alabama  Sea  Grant  Publ.  087(10).  Biloxi.  MS. 
Hueckel.  G.  S.,  and  R.  L.  Stayton. 

1982.     Fish  foraging  on  an  artificial  reef  in  Puget  Sound, 
Washington.     Mar.  Fish.  Rev.  44:  38-44. 
Laughlin,  R.  A.,  and  R.  J.  Livingston. 

1982.     Environmental  and  trophic  determinants  of  the 
spatial/temporal  distribution  of  the  brief  squid  {Lolligun- 
cula  brevis)  in  the  Apalachicola  estuary  (North  Florida, 
USA).     Bull.  Mar.  Sci.  32:489-497. 
Lindquist.  D.  G..  L.  B.  Cahoon,  I.  E.  Clavijo,  M.  H.  Posey. 
S.  K.  Bolden,  L.  A.  Pike,  S.  W.  Burk,  and  P.  A.  Cardullo. 

1994.     Reef  fish  stomach  contents  and  prey  abundance  on  a 
reef  and  sand  substrata  associated  with  adjacent  artificial 
and  natural  reefs  in  Onslow  Bay,  North  Carolina.     Bull. 
Mar.  Sci.  55:308-318. 
Lowe,  C.  G„  B.  M.  Wetherbee,  G.  L.  Crow,  and  A.  L.  Tester. 

1996.  Ontogenetic  dietary  shifts  and  feeding  behavior  of 
the  tiger  shark,  Galeocerdo  cuvier,  in  Hawaiian  waters. 
Environ.  Biol.  Fish.  47:203-211. 


Macdonald,  J.  S.,  and  R.  H.  Green. 

1983.     Redundancy  of  variables  used  to  describe  importance 
of  prey  species  in  fish  diets.     Can.  J.  Fish.  Aquat.  Sci. 
40:635-637. 
Manning.  R.  B. 

1969.     Stomatopod  Crustacea  of  the  Western  Atlantic,  380 
p.     Univ.  Miami  Press.  Miami,  FL. 
Moseley,  F.  N. 

1966.     Biology  of  the  red  snapper.  Lutjanus  aya  Bloch,  of 
the  northwestern  Gulf  of  Mexico.     Publ.  Inst.  Mar.  Sci. 
Univ.  Texas  11:90-101. 
Nelson.  B.  D.,  and  S.  A.  Bortone. 

1996.  Feeding  guilds  among  artificial-reef  fishes  in  the 
northern  Gulf  of  Mexico.     Gulf  Mex.  Sci.  2:66-80. 

Pequegnat,  L.  H..  and  R.  W.  Heard. 

1979.     Synalpheus  agelas,  new  species  of  snapping  shrimp 
from  the  Gulf  of  Mexico  and  Bahama  Islands  (Decapoda: 
Alpheidae).     Bull.  Mar.  Sci.  29:110-116. 
Pinkas,  L.,  M.  S.  Oliphant,  I.  L.  R.  Iverson. 

1971.     Food  habits  of  albacore.  bluefin  tuna,  and  bonito  in 
California  waters.     Calif.  Fish  Game  152:1-105. 
Rooker,  J.  R. 

1995.     Feeding  ecology  of  the  schoolmaster  snapper,  Lut- 
janus apodus  iWalbum).  from  southwestern  Puerto 
Rico.     Bull.  Mar.  Sci.  56:881-894. 
Schiffman,  S.  S.,  M.  L.  Reynolds,  and  F  W.  Young. 

1981.     Introduction  to  multidimensional  scaling,  413  p.     Ac- 
ademic Press  Inc..  New  York,  NY. 
Sedberry.  G.  R.,  and  N.  Cuellar. 

1993.     Planktonic  and  benthic  feeding  by  the  reef  associated 
vermilion  snapper.  Rhomboplites  aurorubens  (Teleostei, 
Lutjanidae).     Fish.  Bull.  91:699-709. 
Stearns,  S. 

1884.     On  the  position  and  character  of  the  fishing  grounds 
of  the  Gulf  of  Mexico.     Bull.  U.S.  Fish.  Comm.  4:289-290. 
Stuck,  K.  C,  H.  M.  Perry,  and  R.  W.  Heard. 

1979.     An  annotated  key  to  the  Mysidacea  of  the  north  cen- 
tral Gulf  of  Mexico.     Gulf  Res.  Rep.  6:  225-238. 
Sweatman,  H.  P.  A. 

1993.     Tropical  snapper  (Lutjanidae)  that  is  piscivorous  at 
settlement.     Copeia  1993:1137-1139. 
Szedlmayer.  S.  T. 

1997.  LHtrasonic  telemetry  of  red  snapper,  Lutjanus 
campechanus,  at  artificial  reef  sites  in  the  northeast  Gulf 
of  Mexico.     Copeia  1997:846-850. 

Szedlmayer,  S.  T,  and  K.  W.  Able. 

1993.  Ultrasonic  telemetry  of  age-0  summer  flounder.  Para- 
lichthys  dentatus.  movements  in  a  southern  New  Jersey 
estuary.     Copeia  1993:728-736. 

Szedlmayer,  S.  T.,  and  J.  Conti. 

1999.     Nursery  habitats,  growth  rates,  and  seasonality  of 

age-0  red  snapper.  Lutjanus  campechanus,  in  the  northeast 

Gulf  of  Mexico.     Fish.  Bull.  97:626-635. 
Szedlmayer,  S.  T,  and  R.  L.  Shipp. 

1994.  Movement  and  growth  of  red  snapper.  Lutjanus 
campechanus,  from  an  artificial  reef  area  in  the  northeast- 
ern Gulf  of  Mexico.     Bull.  Mar.  Sci.  55:  887-896. 

Williams,  A.  B. 

1968.  Substrate  as  a  factor  in  shrimp  distribution.  Lim- 
nol.  Oceanog.  3:283-290. 


376 


Abstract— We  measured  growth  and 
movements  of  individually  marked 
free-ranging  juvenile  white  shrimp 
[Litopenaeus  setiferus)  in  tidal  creek 
subsystems  of  the  Duplin  River, 
Sapelo  Island.  Georgia.  Over  a  period 
of  two  years,  15.974  juvenile  shrimp 
(40-80  mm  TL)  were  marked  inter- 
nally with  uniquely  coded  microwire 
tags  and  released  in  the  shallow  upper 
reaches  of  four  salt  marsh  tidal  creeks. 
Subsequent  samples  were  taken  every 
3-6  days  from  channel  segments 
arranged  at  200-m  intervals  along 
transects  extending  from  the  upper  to 
lower  reach  of  each  tidal  creek.  These 
collections  included  201,384  juvenile 
shrimp,  of  which  184  were  marked 
recaptures.  Recaptured  shrimp  were  at 
large  an  average  of  3-4  weeks  (range: 
2-99  days)  and  were  recovered  a  mean 
distance  of  <0.4  km  from  where  they 
were  initially  marked.  Mean  residence 
times  in  the  creek  subsystems  ranged 
from  15.2  to  25.5  days  and  were  esti- 
mated from  exponential  decay  func- 
tions describing  the  proportions  of 
marked  individuals  recaptured  with 
increasing  days  at  large.  Residence 
time  was  not  significantly  correlated 
with  creek  length  (Pearson  =  -0.316, 
P=  0.684  I,  but  there  was  suggestive 
evidence  of  positive  associations  with 
either  intertidal  (Pearson  r  =  0.867, 
P=0.133)  or  subtidal  (Pearson  /-=0.946, 
P=0.054)  drainage  area.  Daily  mean 
specific  growth  rates  averaged 
0.009  to  0.013  among  creeks;  mean 
absolute  growth  rates  ranged  from 
0.56-0.84  mm/d,  and  were  lower  than 
those  previously  reported  for  juvenile 
penaeids  in  estuaries  of  the  southeast- 
ern United  States.  Mean  individual 
growth  rates  were  not  significantly 
different  between  years  (/-test,  P>0.30) 
but  varied  significantly  during  the 
season,  tending  to  be  greater  in  July 
than  November.  Growth  rates  were 
size-dependent,  and  temporal  changes 
in  size  distributions  rather  than  tem- 
poral variation  in  physical  environ- 
mental factors  may  have  accounted  for 
seasonal  differences  in  growth.  Growth 
rates  differed  between  creeks  in  1999 
U-test,  P<0.015),  but  not  in  1998  (r-test, 
P>0.5).  We  suggest  that  spatial  varia- 
tion in  landscape  structure  associated 
with  access  to  intertidal  resources  may 
have  accounted  for  this  apparent  inter- 
annual  difference  in  growth  response. 


Manuscript  approved  for  publication 
30  September  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:376-388(20011. 


Individual  growth  rates  and  movement  of 
juvenile  white  shrimp  (Litopenaeus  setiferus) 
in  a  tidal  marsh  nursery* 

Stacey  Webb 

Florida  Department  of  Environmental  Protection 

Water  Quality  Standards  and  Special  Proiects  Program 

2600  Blair  Stone  Road,  M.S.  3560 

Tallahassee,  Floida  32399 

E-mail  address:  stacey  fekervcudep  state. fl. us 

Ronald  T.  Kneib 

UGA  Marine  Institute 
Sapelo  Island,  Georgia  31327 


In  2001,  shrimp  became  the  most  popu- 
lar seafood  in  the  United  States  when 
per  capita  consumption  (1.54  kg)  sur- 
passed that  of  canned  tuna  (1.32  kg) 
for  the  first  time  in  recorded  history 
(NOAA1).  Although  77%  of  the  catch 
is  from  the  Gulf  of  Mexico,  commercial 
fisheries  in  Atlantic  coastal  states  of 
the  southeastern  United  States  also 
depend  heavily  on  penaeid  shrimp  pop- 
ulations. Of  the  three  most  common 
estuarine-dependent  penaeid  species 
(Litopenaeus  setiferus,  Farfantepenaeus 
aztecus,  and  F.  duorarum)2  harvested 
in  the  South  Atlantic  Bight,  white 
shrimp  Litopenaeus  setiferus  domi- 
nate, comprising  >70%  of  the  catch  in 
the  region  (North  Carolina  to  the  east 
coast  of  Florida)  and  75-87%  in  South 
Carolina  and  Georgia  (NMFS:I). 

Concerns  over  the  possibility  of  de- 
pleting the  resource  as  early  as  the 
1930s  led  to  intensive  studies  of  the 
life  cycle  (Lindner  and  Anderson,  1956; 
Williams,  1984).  The  white  shrimp  has 
an  annual  life  cycle  that  can  be  divid- 
ed into  offshore  (oceanic)  and  inshore 
(estuarine)  phases.  Adults  spawn  in 
Atlantic  waters  in  spring  and  the  post- 
larvae  migrate  into  estuaries,  aided  by 
flood  tides  and  wind-generated  currents 
(Lindner  and  Anderson,  1956;  Wenner 
et  al.,  1998).  Postlarvae  penetrate  into 
the  shallow  upper  reaches  of  the  nurs- 
ery habitat  where  juveniles  achieve  a 
substantial  portion  of  their  adult  body 
mass  before  moving  into  deeper  creeks, 
rivers,  and  sounds  where  they  approach 
maturity  and  emigrate  seaward  to 


spawning  areas  (Muncy,  1984;  Wil- 
liams, 1984). 

Given  the  commercial  importance 
and  early  interest  in  this  species,  sur- 
prisingly little  research  has  focused  on 
the  juvenile  stages  within  tidal  marsh 
nursery  habitats  ( Minello  and  Zimmer- 
man, 1985;  Zein-Eldin  and  Renaud. 
1986;  Knudsen  et  al.,  1996;  McTigue 
and  Zimmerman,  1998).  Seasonal  mi- 
grations and  ontogenetic  movements 
of  white  shrimp  between  coastal  ocean 
spawning  grounds  and  estuarine  nurs- 
eries are  well  known  ( Dall  et  al..  1990 1. 
as  are  the  sometimes  extensive  migra- 
tions of  adult  shrimp  along  the  Atlantic 
coast,  primarily  to  the  south  during 
fall  and  early  winter,  and  northward  in 
late  winter  and  early  spring  (Lindner 
and  Anderson,  1956;  Shipman,  1983). 
Within  the  estuary,  juvenile  white 


■Contribution  921  of  the  Univerisity  of 
Georgia  Marine  Institute,  Sapelo  Island, 
GA. 

1  NOAA  (National  Oceanic  and  Atmo- 
spheric Administration.  2002.  Shrim 
p  overtakes  canned  tuna  as  top  U.S.  sea- 
food. Website:  http://www.noaanews. 
noaa.gov/stories/s970.htm.  I  Accessed: 
28  August  2002.] 

-  These  species  were  all  previously  included 
in  the  genus  Penaeus,  but  the  subgenera 
were  elevated  to  genera  by  Perez-Farfante 
and  Kensley  (1997). 

:!  NMFS  (National  Marine  Fisheries 
Service).  2002.  Unpubl.  data.  Web- 
site: http://www.st.nmfs.gov/stl/commer- 
cial/index.html.  lAccessed  29  August 
2002.1 


Webb  and  Kneib:  Individual  growth  rates  and  movement  of  Litopenaeus  setiferus  in  a  tidal  marsh  nursery 


377 


shrimp,  in  contrast  to  other  penaeid  species,  are  found 
across  a  wider  range  of  environmental  conditions  and 
habitats  (Kutkuhn,  1966)  and  often  make  tidal  excursions 
between  subtidal  and  vegetated  intertidal  habitats  to  for- 
age (Mayer,  1985;  Kneib.  1995;  2000).  However,  relatively 
little  is  known  about  movements  within  subtidal  creeks  of 
the  primary  nursery  areas,  and  the  degree  to  which  indi- 
viduals exhibit  fidelity  to  a  particular  tidal  creek  drainage 
system  is  unknown. 

Direct  measurement  of  juvenile  shrimp  growth  rates 
within  the  nursery  have  also  been  rare.  Most  growth 
estimates  for  free-ranging  juvenile  shrimp  are  based  on 
analyses  of  size-frequency  data,  which  can  be  misleading 
(Loesch,  1965).  Shrimp  grow  rapidly  while  in  the  estua- 
rine  nursery  throughout  the  summer  and  early  fall,  and 
juveniles  approach  adult,  or  commercially  harvestable  size 
within  2-4  months  after  immigration  to  the  estuary 
(Kutkuhn,  1966;  Williams,  1984).  Mean  absolute  growth 
rates  of  0.7-1.1  mm/d  are  commonly  reported  for  many 
penaeids  (Dall  et  al.,  1990).  However,  growth  studies  are 
difficult  to  compare  because  the  rate  of  growth  may  vary 
between  years  and  among  seasons,  as  well  as  with  size, 
age,  and  sex  of  individuals  (Perez-Farfante,  1969).  Growth 
estimates  for  Litopenaeus  setiferus  range  widely,  from  10  to 
65  mm/month  (Williams,  1984).  Previous  estimates  were 
based  on  a  variety  of  approaches  including  experimental 
studies  in  aquaria  and  ponds  (Pearson,  1939;  Johnson  and 
Fielding,  1956),  size  distributions  from  tagging  studies  of 
adults  (Lindner  and  Anderson,  1956),  length-frequency 
distributions  of  juveniles  in  field  samples  (Gunter,  1950; 
Williams,  1955;  Loesch,  1965;  Harris,  1974;  Mayer,  1985), 
mark-recapture  of  uniform  size  ranges  of  subadults  and 
adults  (Klima,  1974),  and  mark-recapture  of  shrimp  in 
marsh  ponds  (Knudsen  et  al.,  1996).  Many  estimates  of 
growth  for  small  (<80  mm  TL)  juvenile  L.  setiferus  have 
been  extrapolated  from  mark-recapture  studies  of  larger 
(>100  mm  TL)  individuals  (Lindner  and  Anderson,  1956; 
Harris,  1974;  Klima,  1974).  However,  there  is  a  paucity 
of  empirical  data  on  growth  rates  of  small,  free-ranging 
juvenile  white  shrimp  within  natural  estuarine  nursery 
habitats.  The  purpose  of  the  present  study  was  to  provide 
reliable  data  on  growth  and  movements  of  individual  ju- 
venile white  shrimp  within  a  natural  estuarine  nursery 
environment  and  to  initiate  an  assessment  of  spatial 
variation  in  habitat  quality  in  relation  to  tidal  marsh 
landscape  structure. 

Recent  innovations  in  tagging  techniques  have  pro- 
duced an  effective  way  to  obtain  information  on  individual 
organisms  through  the  use  of  sequentially  numbered  bi- 
nary-coded microwire  tags  (Northwest  Marine  Technol- 
ogy. Inc.  Shaw  Island,  WA).  Microwire  tags  were  first  used 
in  tagging  experiments  by  Jefferts  et  al.  (1963)  and  have 
since  been  used  successfully  to  tag  a  variety  of  crustaceans 
including  prawns  (Prentice  and  Rensel,  1977),  crayfish 
(Isely  and  Eversole,  1998),  blue  crabs  (van  Montfrans  et 
al.,  1986;  Fitz  and  Weigert,  1991),  and  lobsters  (Krouse 
and  Nutting,  1990;  Uglem  and  Grimsen,  1995).  Results 
of  these  studies  and  others  generally  show  that  tag  reten- 
tion rates  are  high  and  tagging  has  little  effect  on  the 
growth  or  survival  of  the  fishes  and  crustaceans  in  which 


microwire  tags  have  been  used.  In  a  laboratory  study 
involving  240  juvenile  white  shrimp,  Kneib  and  Huggler 
(2001)  confirmed  that  tag  retention  was  high  (-98%), 
growth  rates  between  tagged  and  control  individuals  were 
not  significantly  different,  and  the  best  location  (based 
on  tag  retention  and  survival)  for  tag  injection  was  in  the 
muscle  tissue  of  the  abdomen.  This  type  of  tag  allows  for 
identification  of  individuals  because  each  tag  is  etched 
with  a  unique  number  encoded  in  binary  form.  In  addi- 
tion, the  tag  is  completely  internal  and  inconspicuous,  thus 
eliminating  problems  associated  with  external  tags  (e.g., 
streamer-type  tags)  that  might  interfere  with  molting  or 
increase  predation  risk  (Garcia  and  LeReste,  1981;  van 
Montfrans  et  al.,  1986;  Isely  and  Eversole,  1998). 


Materials  and  methods 

Study  area 

All  samples  were  collected  from  four  tidal  creek  subsys- 
tems associated  with  the  Duplin  River  on  the  west  side  of 
Sapelo  Island,  Georgia.  The  Duplin  River  tidal  drainage 
( -11  km2 )  includes  almost  10  km-  of  tidal  salt  marsh  that  is 
inundated  twice  daily  by  unequal  tides  with  a  mean  range 
of  2.1  m  (Wadsworth,  1980).  Smooth  cordgrass  iSpartina 
alterniflora)  is  the  dominant  vegetation  in  the  intertidal 
marshes  of  this  area.  Seasonal  water  temperatures  aver- 
age between  10°C  and  30°C,  and  salinity  is  characteristi- 
cally polyhaline,  ranging  from  15  to  30  ppt  (Kneib,  1995). 
Freshwater  flow  into  the  system  is  intermittent  and 
originates  largely  from  local  upland  runoff  and  indirect 
flows  by  several  interconnected  tidal  channels  from  the 
Altamaha  River  about  8  km  to  the  southwest  (Ragotzkie 
and  Bryson.  1955). 

Tidal  creeks  included  in  this  study  were  Post  Office 
Creek  (PO)  and  Stacey  Creek  (SO  in  1998,  and  the  East 
and  West  forks  (EF.  WF,  respectively)  of  the  upper  Duplin 
River  in  1999  (Fig.  1).  Logistical  constraints  precluded 
sampling  shrimp  populations  from  more  than  two  creek 
systems  within  the  same  year,  and  different  pairs  of 
creeks  were  chosen  in  each  of  the  two  years  to  broaden  the 
spatial  coverage  of  the  study.  High-resolution  black  and 
white  photographs  (1:16000  scale)  from  an  aerial  survey  of 
the  region  in  December  1989  were  used  to  measure  broad- 
scale  structural  characteristics  of  the  creek  systems,  in- 
cluding areal  extent  of  the  intertidal  and  subtidal  portions 
of  each  drainage.  The  metrics  and  methods  of  extracting 
the  information  from  the  photographs  are  fully  described 
elsewhere  (see  Webb  and  Kneib,  2002). 

Field  sampling 

Shrimp  were  collected  by  cast  net  along  the  shallow  (<1  m 
depth)  edges  of  the  subtidal  portion  of  each  creek  system 
during  low  tide.  Preliminary  studies  showed  that  1.52-m 
diameter  nets  with  ca.  1-cm  mesh  size  collected  the  range 
of  juvenile  shrimp  sizes  (40-80  mm)  targeted  for  mark- 
ing in  this  study.  All  samples  were  collected  within 
2-3  hours  of  low  tide  to  ensure  that  the  shrimp  popula- 


378 


Fishery  Bulletin  102(2) 


Figure  1 

Map  of  the  salt  marsh  estuary  in  the  vicinity  of  Sapelo  Island,  Geor- 
gia, showing  locations  of  the  tidal  creek  subsystems  within  the  tidal 
drainage  of  the  Duplin  River. 


tion  was  restricted  to  the  tidal  creek  channels  and  had  no 
refuge  in  the  intertidal  vegetation.  A  series  of  stations,  at 
intervals  of  approximately  200  m,  was  established  along 
the  length  of  the  subtidal  portion  of  each  creek  from  the 
upper  reaches  to  the  mouth,  so  that  the  number  of  stations 
within  a  creek  depended  on  the  navigable  length  of  the 
subtidal  channel.  There  were  13  stations  in  PO,  11  in  SC, 
9  in  EF,  and  7  in  WF.  Salinity,  water  temperature,  and 
dissolved  oxygen  were  measured  near  the  surface  (<1  m 
depth)  at  the  mouth  of  the  tidal  creek  on  each  day  of 
sampling  by  using  a  YSI  model  85  meter  ( YSI,  Inc.  Yellow 
Springs,  OH).  Juvenile  shrimp  were  marked  with  uniquely 
coded  microwire  tags  (1.1  mm  longx0.25-mm  diameter, 
Northwest  Marine  Technology  [NMT],  Inc.  Shaw  Island, 
WA),  which  were  injected  into  the  muscle  tissue  of  the 
first  abdominal  segment.  We  used  a  hand-held  multishot 
injector  (NMT)  that  was  designed  to  cut,  magnetize,  and 
inject  sequentially  coded  tags  from  a  continuous  stain- 
less-steel wire  spool.  Each  tag  was  etched  with  six  lines  of 
binary  code  that  could  be  read  under  a  microscope  (25x) 
and  translated  into  a  set  of  numbered  coordinates.  Only 
three  of  the  coded  lines  were  required  to  identify  a  unique 
individual.  A  master  line  contained  a  distinguishing 
sequence  code  that  was  necessary  to  properly  interpret 
codes  on  data  lines  designated  D3  and  D4.  The  numeri- 


cal values  associated  with  these  coded  lines  were  entered 
into  a  sequential  tag  conversion  computer  program  (GR 
[Growth  Rate],  version  1.3,  Northwest  Marine  Technology. 
Inc.  Shaw  Island,  WA)  that  output  the  unique  tag  number 
corresponding  to  those  coordinates. 

A  reference  tag  was  archived  for  every  shrimp  marked  in 
order  that  the  code  on  either  side  of  the  tag  injected  into  a 
shrimp  was  known  with  certainty.  This  was  necessary  to 
ensure  positive  identification  of  recaptured  individuals  be- 
cause the  injector  was  designed  only  to  cut  tags  to  a  known 
length  (1.1  mm)  and  did  not  distinguish  between  the  be- 
ginning and  end  of  sequential  codes  and  often  cut  tags  that 
included  a  portion  of  two  adjacent  codes.  Prior  to  their  re- 
lease, marked  shrimp  were  passed  across  a  magnetometer 
(NMT)  which  signaled  the  presence  of  the  ferromagnetic 
tag  with  an  audible  tone  and  flashing  light.  All  shrimp 
collected  after  these  marking  sessions  were  scanned  in 
the  same  manner  and  when  a  tag  was  detected,  it  was 
removed  from  the  recaptured  shrimp,  cleaned  and  read 
under  the  microscope.  The  two  reference  tags  bracketing 
the  recovered  tag  were  then  located  in  the  archive  set  to 
determine  the  date,  location,  and  initial  length  at  release 
of  the  marked  shrimp.  Thus  the  growth  rate,  time  at  large, 
and  distance  between  points  of  release  and  recapture  could 
be  determined  with  certainty  for  individual  shrimp. 


Webb  and  Kneib:  Individual  growth  rates  and  movement  of  Litopenaeus  setiferus  in  a  tidal  marsh  nursery 


379 


Shrimp  were  marked  and  released  only  in  the  upper 
reaches  of  each  tidal  creek.  During  the  marking  process, 
small  batches  of  shrimp  (<50  individuals)  were  collected 
and  held  in  insulated  plastic  coolers.  Water  in  the  coolers 
was  exchanged  each  time  a  new  batch  of  shrimp  was  col- 
lected. Only  active  individuals,  40-80  mm  total  length 
(TL,  tip  of  the  rostrum  to  end  of  the  telson)  and  in  appar- 
ently good  condition,  were  candidates  for  marking.  The 
marking  process  required  a  minimum  field  crew  of  two 
researchers.  One  measured  shrimp  and  recorded  data, 
while  the  other  injected  tags  and  released  marked  shrimp 
at  the  edge  of  the  tidal  creek. 

We  attempted  to  mark  1000  shrimp  during  a  3-day  pe- 
riod in  each  tidal  creek  near  the  beginning  of  every  month 
from  July  to  November  in  both  years.  It  was  not  possible 
to  tag  shrimp  in  both  creeks  simultaneously;  therefore 
shrimp  were  marked  and  released  in  week  1  during  the 
first  week  of  the  month,  then  in  creek  2  during  the  second 
week  of  the  month.  The  remainder  of  the  month  was  spent 
collecting  marked  shrimp  (Fig.  2).  Inclement  weather 
and  mechanical  problems  with  the  tagging  equipment 
sometimes  prevented  us  from  achieving  the  goal  of  tag- 
ging 1000  shrimp  per  creek  within  the  first  week  of  each 
month.  In  August  1998,  the  tag  injector  malfunctioned  on 
the  first  day  of  tagging  in  SC  and  was  unavailable  for  sev- 
eral weeks  while  it  was  being  refurbished.  Consequently, 
sampling  was  suspended  in  SC  during  that  time.  Although 
the  same  problem  occurred  while  we  attempted  to  mark 
shrimp  from  the  EF  in  September  1999,  we  continued 
sampling  in  an  attempt  to  recapture  shrimp  tagged  in 
previous  months. 

A  total  of  6  sampling  events  after  marking  were  planned 
in  each  creek  per  month  (Fig.  2).  A  sample  consisted  of  the 
combined  contents  of  10  haphazard  casts  of  the  net  along 
the  edge  of  the  tidal  creek  within  each  station  per  sampling 
date.  Shrimp  populations  were  usually  sampled  at  3-day 
intervals  for  21  days  beginning  from  the  midpoint  of  the 
marking  period.  A  consistent  exception  was  the  second 
sample  in  the  series,  which  occurred  at  a  6-day  interval  to 
accommodate  the  marking  effort  in  the  second  creek  and 
to  keep  the  sampling  effort  consistent  in  all  creeks.  Inclem- 
ent weather  interrupted  the  sampling  schedule  on  occasion 
and  when  unfavorable  conditions  persisted  for  more  than  3 
days,  some  of  the  planned  sampling  events  after  marking 
were  cancelled;  some  months  were  represented  by  fewer 
than  6  sampling  events.  Sampling  was  terminated  when, 
as  a  result  of  normal  seasonal  emigration  from  the  nursery 
areas,  shrimp  densities  declined  to  the  point  that  they  could 
no  longer  be  consistently  collected  from  the  tidal  creeks  by 
cast  net  (19  November  1998  and  21  November  1999). 

Catches  of  marked  shrimp  from  each  station  were  re- 
tained in  separate  plastic  bags,  placed  on  ice,  and  trans- 
ported to  the  laboratory.  A  subsample  of  shrimp  from  each 
station  (every  tenth  individual)  was  measured  (TL,  mini 
to  construct  size  distributions.  If  a  sample  included  fewer 
than  100  shrimp,  all  were  measured.  Sex  was  not  deter- 
mined. All  individuals  were  scanned  for  the  presence  of 
tags  and  when  a  marked  shrimp  was  detected,  it  was  mea- 
sured (TL,  mm)  before  the  tag  was  removed  and  stored  in  a 
plastic  vial  for  reading  at  a  later  date.  For  each  recapture. 


Marking  — 

•  •• 

• 

coo 
o 

o 

•  Creek  1 
O  Creek  2 

Postmarking 

OOO 

sampling 

o 

10  15  20  25 

Days  of  the  month 


30 


Figure  2 

Schedule  for  monthly  marking  and  postmarking  field 
schedule  for  the  juvenile  white  shrimp  study. 


we  recorded  date,  creek,  station  of  recapture  (i.e.,  distance 
from  original  release  site)  and  size  offish. 

Daily  instantaneous  (specific)  growth  rates  (mm/[mm/ 
d] )  were  calculated  as 

[(In  L2- In  Lj)/t], 

where  L2  =  total  length  (mm)  of  an  individual  on  the  date 
of  recapture; 
L1  =  initial  total  length  (mm)  on  the  date  of  tag- 
ging; and 
t  =  number  of  days  at  large. 

Daily  absolute  growth  rates  (mm/d)  also  were  calculated 
(L2-L1/t)  to  facilitate  comparison  with  estimates  from 
previous  studies.  Displacement  (distance  moved)  was 
determined  by  comparing  the  location  of  recapture  with 
the  original  location  at  marking.  Residence  time  within  a 
tidal  creek  was  determined  from  a  plot  of  the  proportion  of 
recaptured  individuals  against  time-at -large  for  each  creek 
system.  First,  using  the  Regression  Wizard  in  the  com- 
puter program  SigmaPlot®  (version  8.0,  SPSS,  Inc.  Chicago, 
ID,  we  fitted  the  data  to  an  exponential  decay  function: 

(y=a  x  e~bt) , 

where  y  =  the  proportion  of  total  recaptures; 

t  =  time  at  large;  and 
a  and  b  =  the  estimated  parameters. 

Constraints  imposed  on  the  fit  were  a  =  l  (because  the 
proportion  of  total  recaptures  could  not  exceed  1)  and 
fe>0  (because  this  was  an  exponential  decay  function). 
Mean  residence  time  for  shrimp  in  each  creek  was  then 
estimated  from  the  area  below  the  fitted  curves  describing 
the  proportion  of  recaptures  with  time  at  large.  This  was 
calculated  with  the  macro  function  "area  below  curves" 
included  in  the  "Toolbox"  menu  selection  of  SigmaPlot® 
(vers.  8.0,  SPSS,  Inc.  Chicago,  ID  which  uses  the  trapezoi- 
dal rule  to  estimate  the  area  under  curves. 

Statistical  analyses 

Most  of  the  data  analyses  used  statistical  procedures  in 
version  8.0  of  the  computer  software  package  Systat® 


380 


Fishery  Bulletin  102(2) 


(SPSS,  Inc.  Chicago,  ID.  When  parametric  tests  were 
performed,  residuals  were  analyzed  to  determine  whether 
the  data  met  the  required  assumptions  (Sokal  and  Rohlf, 
1995).  Levene's  test  was  used  to  evaluate  conformity  to 
the  assumption  of  variance  homogeneity  among  groups. 
When  this  assumption  was  violated,  the  data  were  trans- 
formed and  retested.  If  the  assumptions  were  still  not  met, 
then  an  appropriate  nonparametric  test  was  applied  (e.g., 
Kruskal-Wallis  one-way  analysis  of  variance ).  Two  sample 
/-tests  were  used  to  compare  spatial  and  temporal  varia- 
tion in  water  temperature,  salinity,  and  dissolved  oxygen 
between  creeks  within  a  year  and  between  years.  August 
was  omitted  in  comparisons  of  data  between  creeks  in 
1998,  and  between  years  because  sampling  in  SC  was 
suspended  during  August  1998.  Regression  analyses  were 
performed  to  determine  whether  there  were  significant 
linear  relationships  between  initial  shrimp  length  and 
growth  rates  within  each  tidal  creek.  One-way  ANOVA 
(controlling  for  the  covariate  initial  length)  was  used  to 
test  for  differences  in  growth  rates  between  creeks  within 
each  year.  A  similar  approach  (controlling  for  initial  size) 
was  used  to  test  for  monthly  (seasonal)  differences  in 
growth  rate  within  years.  If  growth  rates  did  not  differ 
significantly  between  creeks,  the  data  were  pooled  within 
year,  otherwise  creeks  were  treated  separately.  Only  indi- 
viduals at  large  for  a  month  or  less  (to  ensure  that  growth 
was  representative  of  individual  months)  were  included 
in  the  analyses. 

Shrimp  at  large  for  fewer  than  3  days  were  excluded 
from  the  statistical  analyses  to  reduce  certain  antici- 
pated biases  associated  with  estimating  growth  rates. 
These  included  1)  measurement  error  (assumed  to  be  at 
least  1  mm),  which  would  likely  represent  a  substantial 
proportion  of  the  growth  rate  estimate  when  absolute 
change  in  size  was  small;  2)  increments  of  growth  associ- 
ated with  molting  (Dall  et  al.,  1990),  which  could  either 
underestimate  growth  for  shrimp  that  had  been  at  large 
for  a  short  time  or  had  not  molted  since  they  were  tagged 
or  overestimate  growth  if  shrimp  were  recaptured  shortly 
after  the  first  molt  following  marking;  and  3)  size-specific 
growth,  where  shrimp  marked  at  a  relatively  small  size 
and  smaller  shrimp  exhibit  a  higher  relative  growth,  so 
that  early  recaptures  could  represent  larger  than  average 
growth  rates. 


Results 

Physical  parameters 

Average  water  temperature,  salinity,  and  dissolved  oxygen 
(measured  at  the  mouth  of  each  tidal  creek)  were  similar 
between  creeks  within  years  (see  Table  1  in  Webb  and 
Kneib,  2002).  In  1998,  temperature  ranged  from  18.9  to 
33.4°C,  salinity  from  18.2  to  28.0  ppt,  and  dissolved  oxygen 
from  1.4  to  11.3  mg/L.  In  1999,  temperature  ranged  from 
15.0  to  33.4°C,  salinity  from  23.9  to  32.5  ppt,  and  dis- 
solved oxygen  from  0.8  to  7.1  mg/L.  Temperature  followed 
expected  seasonal  patterns  each  year;  mean  values  were 
highest  in  summer  and  declined  toward  autumn.  Results 


of  /-tests  with  separate  variance  estimates  showed  no 
significant  differences  between  years  in  either  mean  tem- 
perature (  =  0.14,  df=134.0,  P=0.80)  or  dissolved  oxygen 
(£=1.82,  df=115.9,  P=0.07)  but  mean  salinity  was  sig- 
nificantly (£=11.63,  df=122.7,  P<0.01)  higher  in  1999  (28.1 
ppt)  than  in  1998  (24.8  ppt).  Cumulative  rainfall  was  83.9 
cm/yr  in  1998  and  82.9  cm/yr  in  1999  (Garbisch4).  These 
values  were  indicative  of  drought  conditions  because  they 
were  well  below  the  long-term  mean  annual  precipitation 
value  of  ca.  132  cm/yr  reported  for  Sapelo  Island  between 
May  1957  and  March  2003  (Southeast  Regional  Climate 
Center5). 

Growth 

Shrimp  collections  during  recapture  efforts  ranged  from 
20,077  to  78,724  individuals,  but  the  proportion  of  marked 
individuals  recaptured  was  low  in  both  years,  averaging 
just  over  1%  (Table  1).  However,  the  recaptures  included 
184  individuals  for  which  growth  rates  and  net  movements 
within  the  nursery  were  known  precisely. 

Daily  absolute  growth  rates  of  individuals,  which 
ranged  from  0.25  to  2.5  mm,  averaged  0.86,  0.78,  0.84,  and 
0.61  mm  at  PO,  SC,  WF  and  EF,  respectively.  The  mean 
values  are  on  the  low  end  of  the  range  reported  in  previ- 
ous studies  of  juvenile  Litopenaeus  setiferus  with  other 
methods  and  conducted  in  different  locations  (Table  2). 
Daily  specific  growth  rates  were  size-dependent  in  both 
years.  Negative  linear  relationships  between  growth  rate 
and  initial  size  (i.e.,  smaller  shrimp  grew  relatively  faster) 
was  the  prevalent  trend  in  all  creeks  (Fig.  3).  No  sig- 
nificant difference  (£=1.19,  df=74,  P=0.237)  in  growth  was 
detected  between  PO  and  SC,  where  mean  (±SD)  specific 
(instantaneous)  daily  growth  rates  were  0.014  ±0.006  and 
0.012  ±0.007,  respectively.  In  1999,  shrimp  exhibited  sig- 
nificantly U=2.12,  df=56,  P=0.038)  higher  mean  specific 
growth  rates  in  the  WF  ( 0.014  ±0.008)  compared  to  the  EF 
(0.010  ±0.006)  of  the  Duplin  River.  The  physical  environ- 
ment was  similar  at  these  sites  (Webb  and  Kneib,  2002), 
and  there  was  no  significant  difference  (£=1.43,  df=81, 
P=0.156)  in  the  mean  final  sizes  of  shrimp  recaptured 
from  these  sites.  However,  the  mean  (±SD)  initial  size  of 
marked  shrimp  at  EF  (61.3  ±8.3)  was  significantly  U=2.20, 
df=81,  P=0.031)  larger  than  at  WF  (56.0  ±10.9);  therefore 
a  lower  specific  growth  rate  was  to  be  expected  at  EF. 

On  a  finer  temporal  scale,  seasonal  variation  in  growth 
rates  occurred  in  both  years,  more  rapid  growth  early  in 
the  season,  and  a  general  increase  in  the  mean  size  of  in- 
dividuals as  the  season  progressed  were  evident  i  Fig.  4). 
The  earlier  observation  that  specific  growth  rate  declined 
with  size  (Fig.  3)  opens  the  possibility  that  seasonal 
variation  in  growth  rates  could  be  explained  simply  by 


4  Garbisch,  -J.  Unpubl.  data.  Univ.  Georgia  Marine  Institute 
Flume  Dock  Monitoring  Station,  NOAA,  Sapelo  Island  Nat  ional 
Est  ua  vine  Research  Reserve.  Univ.  Georgia  Marine  Institute. 
Sapelo  Island,  GA  31327. 

5  Southeast  Regional  Climate  Center.  Unpubl.  data.  Website: 
http://water.dnr.state.sc.us/water/climate/sercc/elimateinfo/ 
Instmical/historicaLga.html.     [Accessed  21  November  20031. 


Webb  and  Kneib:  Individual  growth  rates  and  movement  of  Litopenaeus  setiferus  in  a  tidal  marsh  nursery 


381 


004 


0.03 


0.02 


0.01 


%      0.1 


EF 


r2  =  0.24 
P=0.06 


40  50 


70         80 


0.04 


PO 


r2  =  0.47 

0.03 

• 
• 

P=<0.001 

0.02 

• 

£••• 

• 
• 

• 

«>v 

•  •  • 

--*V 

0.01 

••. 

• 

0.00 

, 

■% . . . 

30       40        50       60       70        80       90 


30       40        50       60       70        80       90 


0.04 
0.03 
0.02 
0.01 
0.00 


sc 


r2  =  0.16 
P=0.04 


30       40        50       60       70        80       90 


Initial  size  (TL,  mm) 

Figure  3 

Scatter  plots  and  linear  regression  results  for  the  relationships  between  individual  specific 
growth  rates  and  initial  sizes  of  recaptured  juvenile  shrimp  in  each  tidal  creek  subsystem: 
EF  =  East  Fork  of  Duplin  River,  WF  =  West  Fork  of  Duplin  River,  PO  =  Post  Office  Creek, 
SC  =  Stacey  Creek. 


Table  1 

Monthly  summary  of  the  number  of  shrimp  tagged. 

number  collected 

in  subsequent 

sampling,  and  the 

number  of  tagged 

shrimp 

recaptured  in  each  tidal  creek. 

Collection  site  and  sampling  process 

July 

August 

September 

October 

November 

Total 

Post  Office  Creek 

Number  tagged  and  released 

1004 

1077 

1025 

1000 

779 

4885 

Number  collected  after  tagging 

7477 

13.851 

4730 

8384 

1298 

35,740 

Number  of  tags  recaptured 

13 

16 

16 

22 

1 

68 

Stacey  Creek 

Number  tagged  and  released 

719 

91 

804 

862 

623 

3099 

Number  collected  after  tagging 

5877 

0 

5466 

6376 

2358 

20,077 

Number  of  tags  recaptured 

2 

0 

12 

16 

3 

33 

East  Fork  of  Duplin  River 

Number  tagged  and  released 

812 

1000 

0 

1024 

1003 

3839 

Number  collected  after  tagging 

23,080 

14,490 

11,020 

14,804 

3449 

66,843 

Number  of  tags  recaptured 

7 

11 

0 

4 

4 

26 

West  Fork  of  Duplin  River 

Number  tagged  and  released 

1008 

1000 

447 

696 

1000 

4151 

Number  collected  after  tagging 

32,130 

17,926 

10,954 

10,600 

7114 

78,724 

Number  of  tags  recaptured 

27 

16 

3 

4 

7 

57 

382 


Fishery  Bulletin  102(2) 


Table  2 

Summary  of  estimated  mean 
if  reported  in  other  units. 

daily  absolute  growth  rates 

for  juvenile  Litopenaeus  setiferus.  Growth  rates  were  converted  to  mm/d 

Reference 

Location 

Growth  rate 
(mm/d) 

Method  and  notes 

Gunter,  1950 

Gulf  of  Mexico.  Texas 

0.8-1.3 

Size  frequency  in  field  samples,  juveniles  28-100  mm 

Williams,  1955 

coastal  North  Carolina 

1.2 

Size  frequency  in  field  samples,  progression  of 
maximum  sizes  of  juveniles,  32-117  mm 

Johnson  and  Fielding, 

1956 

Florida 

1.3 

Pond  culture,  juveniles 

Lindner  and  Anderson 

1956 

South  Atlantic  Bight 
and  Gulf  of  Mexico 

1.0-1.3 

Extrapolated  for  juveniles  40-80  mm  from  Walford 
plot  results  using  field  mark-recapture  (disc  tags)  data 
for  individuals  70-205  mm 

Loesch,  1965 

Mobile  Bay,  Alabama 

0.6-1.0 
2.2 

Size  frequency  from  spring  and  summer  field  samples; 
progression  of  maximum  sizes  of  juveniles  50-135  mm 

juveniles  15-70  mm 

Klima,  1974 

Galveston  Bay,  Texas 

1.4-1.8 

Extrapolated  for  juveniles  40-80  mm  from  Walford 
plot  results  determined  from  field  mark-recaptured 
(stain-injected)  subadults  (117  mm) 

coastal  Louisiana 

1.0-1.3 

Extrapolated  for  juveniles  40-80  mm  from  Walford 
plot  results  determined  from  field  mark-recaptured 
(stain-injected)  subadults  (120  mm) 

Mayer,  1985' 

Sapelo  Island,  Georgia 

0.9-1.5 

Estimated  from  modal  size-frequency  data  for 
juveniles  20-120  mm 

Knudsonetal..  1996 

coastal  Louisiana 

0.3-0.7 

Mark-recapture  (injected  pigments)  of  juveniles 
45-58  mm  (initial  size)  from  coastal  marsh  ponds 

This  study 

Sapelo  Island,  Georgia 

0.6-0.9 

Monthly  mark-recapture  (coded  ferromagnetic  tags) 
of  juveniles  40-80  mm  (initial  size)  from  subtidal 
creeks 

'  Mean  growth  rates  reported  ir 
derived  directly  from  the  data 

Table  3  of  Mayer  1 1985 )  were  inconsistent  with  cohort  data  in  Figure  8  of  that  thesis;  rates  reported  here  were 
Doints  shown  in  Figure  8  of  Mayer's  thesis. 

changes  in  the  average  size  of  shrimp  within  the  nursery 
over  time. 

We  tested  this  hypothesis  by  comparing  mean  growth 
rates  among  months  after  controlling  for  initial  length  as 
a  covariate.  For  these  analyses,  the  1998  data  from  PO  and 
SC  were  pooled  because  there  was  no  evidence  of  a  differ- 
ence in  growth  rates  between  these  two  creeks;  the  1999 
data  from  EF  and  WF  were  analyzed  separately  because 
mean  growth  rates  differed  between  these  two  systems. 
After  removing  the  effect  of  initial  size,  there  was  no  sig- 
nificant difference  among  months  in  1998,  nor  in  1999  at 
EF,  but  significant  differences  in  mean  growth  remained 
detectable  among  months  at  WF  (Table  3).  The  findings 
from  WF  also  were  unusual  in  that  the  covariate  (initial 
length)  was  not  a  significant  factor  in  the  analysis.  Post 
hoc  multiple  comparisons  (Bonferroni,  experiment-wise 
<*=0.05)  of  mean  growth  rates  among  months  (without 
accounting  for  the  covariate)  indicated  that  the  specific 
growth  rate  in  July  (0.021)  was  significantly  greater  than 
thai  in  the  other  months  (0.007  to  0.011).  This  was  the 


only  statistically  significant  evidence  of  seasonal  varia- 
tion in  growth  apparently  not  associated  with  shrimp  size 
distributions. 

With  respect  to  spatial  variation  in  growth  rates  of  ju- 
venile shrimp,  the  most  notable  observation  in  this  study 
was  the  relatively  low  mean  growth  rate  observed  at  EF 
compared  to  the  other  sites.  This  difference  could  have 
resulted  from  the  larger  mean  initial  size  of  individuals 
tagged  at  EF  (61.3  mm)  compared  with  those  at  WF  (56.0) 
in  1999.  However,  a  similar  difference  in  mean  initial  sizes 
of  marked  shrimp  between  tidal  creek  subsystems  (SC, 
64.2  mm;  PO,  59.6  mm)  in  the  previous  year  did  not  result 
in  a  significant  difference  in  growth  rates.  When  we  con- 
sidered the  structural  characteristics  of  each  tidal  creek  at 
a  landscape  level,  the  EF  subsystem  had  the  largest  tidal 
drainage  area  (119.5  ha. )  compared  to  the  other  sites  ( 58.6 
to  104.9  ha.),  but  proportionally  less  of  that  area  was  inter- 
tidal  drainage.  There  was  a  stronger  correlation  between 
mean  growth  rate  (pooled  across  all  individuals  within  a 
creek)  and  the  proportion  of  the  drainage  area  that  was 


Webb  and  Kneib:  Individual  growth  rates  and  movement  of  Litopenaeus  setiferus  in  a  tidal  marsh  nursery 


383 


Table  3 

Summary  of  ANOVA  results  for  the  effects  of  month  on  specific  growth  rate  of  Litopenaeus  setiferus  after  controlling  for  the 
covariate  initial  length.  Only  individuals  at  large  between  3  and  32  days  were  included  in  the  analyses.  PO  =  Post  Office  Creek; 
SC  =  Stacey  Creek;  EF  =  East  Fork  of  Duplin  River;  and  WF  =  West  Fork  of  Duplin  River.  Prob.  =  probability. 

1998  (PO  and  SC) 

EF 

WF 

Source                                                    df 

F-value             Prob. 

df 

F-value           Prob. 

df 

F-value 

Prob. 

Covariate  (initial  size)                         1 
Month                                                      4 
Error                                                     70 

30.77               <0.01 
1.25                0.30 

1 

3 

11 

1.42               0.26 
0.74               0.55 

1 

4 

36 

1.42 
4.32 

0.24 
<0.01 

Oct 

(39) 


Nov 
(15) 


Month 


Figure  4 

Mean  daily  specific  growth  rates  and  mean  initial  size  of 
marked  and  recaptured  juvenile  white  shrimp  by  month. 
Only  individuals  at  large  for  3  to  32  days  were  included. 
Error  bars  are  2  SEs  and  the  number  of  observations  are 
given  in  parentheses  below  each  month. 


intertidal  (Fig.  5A)  than  there  was  between  growth  and 
mean  initial  size  (Fig.  5B)  at  the  landscape  level.  There 
was  almost  no  correlation  between  proportion  of  drainage 
area  that  was  intertidal  and  initial  mean  size  of  marked 
shrimp  (Pearson  r=-0.18,  P=1.0). 

Residence  time  and  movement  of  marked  shrimp 

Recaptured  shrimp  were  at  large  for  up  to  99  days,  but 
mean  residence  time  for  individuals  marked  in  all  four 
tidal  creeks  was  between  15  and  26  days  (Fig.  6).  Mean 
residence  time  was  greatest  at  EF  and  least  at  SC.  During 
their  time-at-large,  net  displacement  (distance  between 
mark  and  recapture  sites)  of  the  marked  individuals 
ranged  from  0  to  3000  m  ,  but  averaged  258-373  m  in  all 
creeks.  There  was  no  evidence  of  a  significant  relationship 
between  time-at-large  and  mean  net  displacement  (linear 
regression  F=1.48;  df=l,45;  P=0.23),  but  movement  was 
slightly  related  to  shrimp  size,  and  larger  individuals 
showed  greater  displacement  (Fig.  7).  Variation  in  resi- 
dence time  among  creek  subsystems  was  not  significantly 


0.016 

A 

0.014  • 

Pearson  r=  0.090                       wp  # 
P=0.10                         po«/- 

0.012  ■ 

^^^        •  SC 

CO 

^      0.010  ■ 

^-^ 

g 

EF  * 

S      0.008  I 

£               0.84           0.86           0.88          0.90           0.92 

%_                         Intertidal  area/total  drainage  area 

™       0.016 

B 

J       0.014 
0.012 

A  WF 

""""■ •— ^*  PO 

~~-C^_^^        SC  A 

0.010 

Pearson  r=  -0.55 

P  =  0.45                                 A£F 

56            58            60            62            64 

Mean  initial  size  (TL,  mm) 

Figure  5 

Correlations  between  mean  daily  specific  growth 

rate  in  each  tidal  creek  (PO=Post  Office,  SC  =  Stacey 

Creek,  EF=East  Fork  of  Duplin  River,  WF=West 

Fork  of  Duplin  River).  (A)  The  proportion  of  the  tidal 

drainage  area  that  is  intertidal;  IB)  mean  initial  size 

of  marked  shrimp. 

correlated  with  length  of  the  creek  mainstem  (Pearson 
r=-0.32,  P=0.684),  but  there  was  evidence  of  positive 
associations  with  the  amount  of  intertidal  (Pearson 
r=0.87,  P=0.133)  and  subtidal  (Pearson  r=0.95,  P=0.054) 
drainage  areas  within  each  subsystem. 

Most  shrimp  (939r)  were  recaptured  in  the  same  tidal 
creek  subsystem  in  which  they  were  originally  marked, 
but  there  was  some  evidence  of  movement  among  creeks 
and  between  the  subtidal  and  intertidal  components  of 
the  shrimp  nursery  within  creeks.  The  marked  individual 
at  large  for  the  longest  time  (99  days)  was  recaptured 
at  the  same  station  where  it  was  originally  marked  (net 


384 


Fishery  Bulletin  102(2) 


Mean  residence  time 
(area  under  each  curve) 

Post  Office  Creek  (20.4  days) 
Stacey  Creek  (15.2  days) 


100 


Mean  residence  time 

(area  under  each  curve): 

i  East  Fork  (25.5  days) 
i  West  Fork  (20.4  days) 


40  60 

Days  at  large 


100 


Figure  6 

Estimates  of  mean  residence  times  of  marked  shrimp 
in  (A)  tidal  creeks  sampled  in  1998  and  ( B )  creeks  sam- 
pled in  1999.  Estimates  are  based  on  the  area  under  the 
curves  describing  the  proportion  of  recaptured  marked 
shrimp  in  each  creek  system  that  still  remained  to  be 
captured  after  the  indicated  number  of  days-at-large. 


displacements 0  m).  In  contrast,  one  shrimp  marked  at 
PO  demonstrated  a  net  displacement  of  3  km  when  it  was 
recaptured  at  SC  after  61  days  at  large.  Nine  shrimp  (at 
large  from  18  to  49  days)  marked  at  WF  were  recaptured 
at  EF.  and  two  (at  large  19  and  45  days)  tagged  at  EF  were 
recaptured  at  WF.  It  was  not  possible  to  determine  pre- 
cisely when  these  shrimp  moved  out  of  the  creek  in  which 
they  were  tagged  or  how  long  they  were  present  in  the 
creek  subsystem  where  they  were  ultimately  recaptured. 
For  the  growth  rate  analyses,  it  was  assumed  that  most 
growth  occurred  while  the  shrimp  were  in  the  creek  and 
where  individuals  were  marked.  The  mean  (±SD)  final 
size  (mm,  TL)  of  individuals  that  moved  between  creek 
subsystems  was  significantly  (separate  variance  estimate 
<=2.62,  df=16.9,  P=0.018)  larger  (78.9  ±7.4)  than  that  of 
the  group  tagged  and  recaptured  in  the  same  subsystem 
(71.3  ±13.1);  the  initial  mean  size  of  the  two  groups  was 
nearly  identical  (57.5  ±10.8  and  57.7  ±10.4,  respectively). 
Two  shrimp  (at  large  7  and  17  days)  tagged  at  EF  were 
recaptured  at  high  tide  in  flume  weirs  located  25  m  into 


1000 
800 
600 
400 
200 


F=7.10,  df  =  1.37 
P  =  0.01,  r2  =  0.16 


40        50        60        70        80        90        100      110 
Final  size  (TL.  mm) 

Figure  7 

The  effect  of  shrimp  size  at  recapture  on  mean  dis- 
tance between  mark  and  recapture  locations  (dis- 
placement!. Summary  results  from  the  linear  regres- 
sion ANOVA  performed  on  the  data  are  shown.  Values 
of  mean  displacement  were  based  on  data  from  2-11 
individuals  within  each  size. 


the  interior  of  the  intertidal  marsh  drained  by  that  tidal 
creek  subsystem.  The  flume  weir  samples  were  part  of  an 
ongoing  study  (Kneib,  unpubl.  data)  to  determine  nekton 
use  of  the  intertidal  marsh  surface  (see  Kneib,  1991,  1997; 
Kneib  and  Wagner,  1994). 


Discussion 

Growth 

Mean  growth  rates  of  juvenile  white  shrimp  measured 
in  this  study  (0.6-0.9  mm/d)  were  near  the  lower  end  of 
the  range  of  estimates  previously  reported  for  juvenile 
white  shrimp  along  the  U.S.  Atlantic  and  Gulf  coasts 
(Table  2).  The  principal  difference  between  the  present 
and  previous  studies  is  that  the  values  presented  in  this 
study  were  based  on  direct  measurements  of  free-rang- 
ing individual  juvenile  shrimp  rather  than  on  extrapola- 
tions from  batch  mark-recaptures  of  larger  individuals 
or  changes  in  modal  size  frequencies.  The  open  nature  of 
estuarine  ecosystems,  prolonged  seasonal  recruitment  to 
the  nursery,  and  ontogenetic  differences  in  mortality  and 
movement  all  may  confound  the  interpretation  of  size-fre- 
quency data  (Haywood  and  Staples,  1993).  Given  that  our 
growth  values  were  based  on  actual  changes  in  the  size 
of  individuals  rather  than  estimated  from  the  apparent 
growth  trajectories  of  cohorts,  we  are  confident  that  the 
mean  growth  rates  reported  here  accurately  reflect  those 
of  free-ranging  juvenile  white  shrimp  (40-80  mm  TL)  in 
the  polyhaline  portion  of  the  tidal  marsh  nursery  habitat 
of  coastal  Georgia. 

Temporal  differences  in  observed  growth  rates  in  this 
study  may  have  resulted  from  either  variation  in  environ- 
mental conditions  or  spatial  variation  in  habitat  quality. 


Webb  and  Kneib:  Individual  growth  rates  and  movement  of  Litopenaeus  setiferus  in  a  tidal  marsh  nursery 


385 


Penaeids  are  most  abundant  in  tidal  marsh  nurseries 
when  physical  conditions  (eg.,  temperature  and  salinity) 
appear  optimal  for  their  growth  and  survival  (Zein-Eldin 
and  Renaud,  1986),  but  environmental  variability  is 
characteristic  of  most  estuaries  and  therefore  is  an  obvi- 
ous starting  point  for  explaining  observed  differences  in 
shrimp  growth  among  sites  or  times.  Salinity  was  the 
only  environmental  factor  we  measured  that  showed  a 
significant  difference  between  years  but  could  not  be  as- 
sociated with  any  interannual  difference  in  mean  growth 
rates. 

Temperature  may  affect  the  growth  and  estuarine  dis- 
tribution of  juvenile  penaeids  more  than  salinity  (Vetter, 
1983),  and  interactions  between  salinity  and  temperature 
may  have  even  greater  effects  than  variation  in  either  fac- 
tor alone  (Zein-Eldin  and  Renaud,  1986).  Mean  tempera- 
tures throughout  our  study  period  (with  the  exception  of 
November)  in  both  years  were  largely  within  the  optimum 
range  for  growth  of  white  shrimp  which,  in  the  laboratory, 
was  reported  to  be  between  25°and32.5°C  (Zein-Eldin  and 
Griffith,  1969).  Higher  temperatures  generally  contrib- 
ute to  faster  growth  in  young  penaeids  (Perez-Farfante. 
1969;  Muncy,  1984),  and  therefore  it  seems  reasonable  to 
expect  seasonal  variation  in  temperature  to  be  reflected 
in  growth  rates.  However,  this  interpretation  is  con- 
founded by  the  fact  that  growth  rates  also  are  size  depen- 
dent (Fig.  3,  Table  3)  and  that  increases  in  mean  size  of 
juvenile  white  shrimp  (Fig.  4)  occurred  while  tempera- 
ture in  the  nursery  habitat  was  decreasing  from  the  July 
maxima.  It  seems  likely  that  growth  rates  of  juvenile 
white  shrimp  were  robust  over  the  relatively  narrow  range 
of  seasonal  variation  in  temperature  and  salinity  observed 
in  the  present  study. 

Alternatively,  differences  in  growth  between  certain 
sites  could  be  the  result  of  spatial  variation  in  habitat 
quality.  This  variation  need  not  be  a  function  of  water 
quality,  but  rather  a  function  of  some  structural  aspect 
of  the  nursery  habitat.  There  was  a  strong  correlation  be- 
tween mean  growth  rates  and  the  proportion  of  tidal  creek 
drainage  area  that  was  intertidal.  Only  four  creek  sub- 
systems were  examined  in  ours  study,  and  we  recognize 
that  this  is  an  insufficient  sample  size  to  justify  anything 
more  than  a  suggestive  hypothesis.  However,  evidence  of 
relationships  between  the  amount  of  intertidal  habitat 
and  penaeid  shrimp  production  (Turner,  1977,  1992),  as 
well  as  the  amount  of  intertidal  creek  edge  and  juvenile 
shrimp  abundance  in  adjacent  subtidal  creeks  (Webb  and 
Kneib,  2002),  supports  the  contention  that  intertidal  ac- 
cessibility is  an  important  component  of  nursery  habitat 
quality  for  juvenile  white  shrimp.  We  propose  that  the 
ratio  between  intertidal  and  shallow  subtidal  habitat  may 
be  a  key  feature  of  white  shrimp  nursery  habitat  quality. 
When  tidally  inundated,  the  intertidal  portion  of  marsh 
creek  drainage  systems  is  used  extensively  by  juvenile 
white  shrimp  (Kneib,  1995,  2000),  most  likely  as  a  rich 
foraging  area  (Kneib,  1997),  and  the  shallow  subtidal 
portion  functions  as  a  low  tide  refuge  and  corridor  for  the 
seasonal  migration  of  postlarvae  and  subadults  between 
the  open  estuary  and  coastal  ocean  spawning  grounds 
and  the  juvenile  nursery  (Kneib,  1997,  2000). 


Movement  and  residence  time 

Understanding  the  causes  of  broad-scale  migration  of 
penaeids  has  obvious  implications  for  predicting  com- 
mercial catches  and  therefore  these  causes  have  been  the 
focus  of  research  on  shrimp  movements  for  decades  ( Perez- 
Farfante,  1969;  Muncy,  1984).  However,  finer-scale  move- 
ments, which  may  affect  growth  and  survival  of  juvenile 
shrimp  within  the  estuary,  are  not  as  well  known.  Emigra- 
tion of  white  shrimp  from  estuaries  is  determined  by  size, 
maturity,  and  environmental  conditions  (Muncy,  1984), 
and  size  plays  a  principal  role  (Dall  et  al.,  1990).  In  the 
South  Atlantic  Bight,  larger  white  shrimp  (>100  mm  TL) 
begin  emigrating  from  the  nursery  to  commercial  fishing 
areas  in  the  nearshore  coastal  ocean  in  August  (Lindner 
and  Anderson  1956,  Shipman,  1983).  We  collected  few 
shrimp  >100  mm  in  the  tidal  marsh  creeks,  which  is  con- 
sistent with  previous  observations  of  ontogenetic  migra- 
tion to  deeper  waters.  According  to  growth  rates  measured 
in  this  study,  a  shrimp  of  40  mm  TL  would  become  large 
enough  to  emigrate  from  the  estuary  to  the  coastal  ocean 
in  2-3  months  (i.e.,  a  shrimp  tagged  at  40  mm  TL  could 
reach  85-108  mm  TL  in  2.5  months). 

The  presence  of  high  densities  of  small  juvenile  white 
shrimp  in  the  upper  reaches  of  Georgia's  tidal  marsh 
creeks  (Harris,  1974;  Hackney  and  Burbanck,  1976;  Webb 
and  Kneib,  2002)  has  supported  the  contention  advanced 
by  Weinstein  ( 1979)  that  these  areas  are  primary  nurser- 
ies for  juvenile  fish  and  shellfish.  However,  it  has  been 
unclear  whether  these  aggregations  represent  stable  res- 
ident populations  or  are  composed  of  tidal  transients  that 
constantly  migrate  among  creek  subsystems  within  the 
broader  estuarine  nursery.  Young  shrimp  are  known  to 
move  short  distances  to  avoid  unfavorable  physiochemical 
conditions  (Hackney  and  Burbanck,  1976;  Dall  etal.,  1990) 
and  routinely  make  tidally  mediated  excursions  between 
subtidal  and  intertidal  portions  of  the  nursery  to  forage 
or  escape  predators  (Kneib,  1995,  1997).  Our  findings 
showed  that  juvenile  white  shrimp  also  tended  to  remain 
resident  in  the  upper  reaches  of  tidal  creeks  where  they 
were  originally  tagged  until  attaining  a  size  ( 80-100  mm ) 
at  which  they  normally  begin  to  emigrate  from  the  nursery 
(Perez-Farfante,  1969). 

Although  there  was  some  movement  between  tidal  creek 
subsystems,  the  high  level  of  site  fidelity  demonstrated 
by  juvenile  white  shrimp  was  remarkable  given  the  open- 
ness and  degree  of  tidal  flux  in  the  Duplin  River  system 
(mean  tide  range=2.1  m).  Data  from  the  chemical  analysis 
of  shrimp  tissue  composition  also  suggest  limited  move- 
ments of  juvenile  penaeids  within  estuarine  nurseries. 
Using  the  stable  isotopes  of  carbon  and  nitrogen  from  mus- 
cle tissues  of  pink  shrimp  (Farfantepenaeus  duorarum), 
Fry  et  al.  (1999)  traced  shrimp  movements  within  and 
between  seagrass  and  mangrove  habitats  of  southwestern 
Florida.  They  found  distinct  differences  among  individu- 
als sampled  from  similar  inshore  habitat  types  separated 
by  small  (3-5  km)  open  water  distances,  indicating  that 
individuals  remained  "resident"  in  specific  portions  of  the 
estuary  at  least  for  several  weeks.  Noting  a  similar  study 
in  coastal  Louisiana,  Fry  et  al.  (2003)  suggested  that 


386 


Fishery  Bulletin  102(2) 


small  juvenile  brown  shrimp  (Farfantepenaeus  aztecus) 
are  more  transient  in  suboptimal  habitat  (open  bays  and 
deeper  channels)  and  exhibit  less  movement  upon  reach- 
ing optimal  habitat  (ponds  and  shallow  channels). 

The  only  study  with  which  we  can  directly  compare  our 
findings  on  residence  time  and  movements  was  conducted 
by  Knudsen  et  al.  (1996)  near  Calcasieu  Lake,  Louisiana, 
where  tidal  flux  was  considerably  lower  (mean  tide  range 
<0.6  m)  and  the  system  (marsh  impoundments)  was  less 
open  than  that  in  the  present  study.  Knudsen  et  al.  (1996) 
marked  batches  of  juvenile  white  shrimp  (45-69  mm  TL) 
by  injection  of  colored  pigments  and  released  them  into  a 
pair  of  35-ha.  impoundments,  each  connected  to  the  open 
estuary  through  a  narrow  channel  that  was  fitted  with 
screen  deflectors  and  traps  designed  to  collect  all  emigrat- 
ing nekton.  The  mean  time  from  release  to  emigration  of 
juvenile  white  shrimp  ranged  from  30.2  to  59.9  days.  Our 
estimates  of  tidal  creek  residence  time  for  juvenile  shrimp 
in  Georgia  tidal  creeks  was  about  half  that  reported  for 
impoundments  in  Louisiana  and  may  be  explained  by  the 
differences  in  tidal  flux  and  openness  between  the  two 
systems.  However,  the  values  we  observed  were  likely 
underestimates  of  the  actual  residence  period  of  survivors 
within  the  creeks  because  they  included  losses  due  to  mor- 
tality as  well  as  emigration. 

It  seems  clear  from  the  studies  conducted  thus  far  that 
juvenile  penaeids,  once  having  entered  the  estuarine 
nursery,  tend  to  remain  within  a  limited  spatial  range 
where  they  are  exposed  to  local  conditions  for  several 
weeks.  Our  findings  also  provide  evidence  of  spatial 
variation  for  both  residence  time  and  growth  rate  of  ju- 
venile white  shrimp  that  is  possibly  attributable  to  struc- 
tural differences  in  tidal  creek  subsystems.  We  suggest 
there  may  be  an  optimal  value  for  the  ratio  of  subtidal  to 
intertidal  drainage  area  within  marsh  creek  systems  that 
can  achieve  a  favorable  balance  between  suitable  habitat 
(space)  at  low  tide,  which  tends  to  enhance  residence  time 
and  density  of  juvenile  shrimp,  while  providing  sufficient 
intertidal  foraging  habitat  and  predator  refugia  at  high 
tide  to  promote  high  rates  of  juvenile  shrimp  growth  and 
survival.  Spatially  explicit  information  on  growth  rates 
and  the  extent  to  which  individual  shrimp  move  within 
their  estuarine  nurseries  are  necessary  initial  steps  to- 
ward meeting  the  challenge  of  maintaining  quality  nurs- 
ery habitat  for  a  sustainable  shrimp  fishery  and  satisfying 
other  demands  associated  with  human  development  in 
and  around  estuarine  watersheds. 


Acknowledgments 

Several  individuals  provided  field  and  laboratory  assis- 
tance for  this  project,  but  we  especially  thank  K.  Feeley, 
J.  Kneib,  and  J.  Spicer  for  helping  on  a  regular  basis. 
The  primary  source  of  funding  was  a  National  Estuarine 
Research  Reserve  System  Graduate  Research  Fellowship 
to  S.  Webb  (NA870R0284)  (Estuarine  Reserves  Division, 
Office  of  Ocean  and  Coastal  Resource  Management,  NOS, 
NOAA),  and  matching  funds  provided  by  the  University 
of  Georgia  Marine  Institute.  The  Georgia  Sea  Grant  Col- 


lege Program  contributed  funds  for  the  purchase  of  an 
additional  tag  injector  unit,  which  substantially  improved 
the  effectiveness  of  the  mark-recapture  program.  The  con- 
ceptual basis  for  this  project  was  derived  from  research 
conducted  under  a  grant  from  the  National  Science  Foun- 
dation ( DEB-9629621 ),  which  also  contributed  supplemen- 
tal student  support. 


Literature  cited 

Dall,  W.,  B.  J.  Hill.  P.  C.  Rothlisberg,  and  D.  J.  Staples. 

1990.  The  biology  of  the  Penaeidae.  In  Advances  in  marine 
biology,  vol.  27  (J.  H.  S.  Blaxter  and  A.  J.  Southward,  eds.), 
489  p.     Academic  Press,  London. 

Fitz,  H.  C,  and  R.  G.  Weigert. 

1991.  Tagging  juvenile  blue  crabs,  Callinectes  sapidus,  with 
microwire  tags:  retention,  survival,  and  growth  through 
multiple  molts.     J.  Crust.  Biol.  11(21:229-235. 

Fry,  B.,  D.  M.  Baltz,  M.  C.  Benfield,  J.  W.  Fleeger,  A.  Gace, 
H.  L.  Haas,  and  Z.  J.Quinones-Rivera. 

2003.     Stable  isotope  indicators  of  movement  and  residency 
for  brown  shrimp  (Farfantepenaeus  aztecus)  in  coastal 
Louisiana  marshscapes.     Estuaries  26:82-97. 
Fry,  B.,  P.  L.  Mumford,  and  M.  B.  Robblee. 

1999.     Stable  isotope  studies  of  pink  shrimp  (Farfante- 
penaeus  duorarum  Burkenroad)  migrations  on  the  south- 
western Florida  shelf.     Bull.  Mar.  Sci.  65(2):419-430. 
Garcia,  S.,  and  L.  Le  Reste. 

1981.     Life  cycles,  dynamics,  exploitation  and  management 
of  coastal  penaeid  shrimp  stocks.     FAO  Fish.  Tech.  Paper 
203,  215  p.     FAO,  Rome. 
Gunter,  G. 

1950.     Seasonal  population  changes  and  distributions  as 
related  to  salinity,  of  certain  invertebrates  of  the  Texas 
coast,  including  the  commercial  shrimp.     Publ.  Inst.  Mar. 
Sci.,  Univ.  Texas  1:7-31. 
Hackney,  C.  T,  and  W.  D.  Burbanck. 

1976.     Some  observations  on  the  movement  and  location  of 
juvenile  shrimp  in  coastal  waters  of  Georgia.     Bull.  GA 
Acad.  Sci.  34:129-136. 
Harris,  C.  D. 

1974.     Observations  on  the  white  shrimp  {Penaeus  setiferus) 
in  Georgia.     Georgia  Dep.  Nat.  Res.  Contr.  Ser.  27:1-47. 
Haywood,  M.  D.  E.,  and  D.  J.  Staples. 

1993.     Field  estimates  of  growth  and  mortality  of  juvenile 
banana  prawn  (Penaeus  merguiensis).     Mar.  Biol.  116: 
407-416. 
Isely,  J  .J.,  and  A.  G.  Eversole. 

1998.     Tag  retention,  growth,  and  survival  of  red  swamp 
crayfish  Procambarus  clarkii  marked  with  coded  wire 
tags.    Trans.  Am.  Fish.  Soc.  127:658-660. 
Jefferts,  K.  B..  P.  K.  Bergmann,  and  H.  F.  Fiscusest. 

1963.     A  coded  wire  identification  system  for  macro-organ- 
isms.    Nature  198:460-462. 
Johnson,  M.  C,  and  J.  R.  Fielding. 

1956.     Propagation  of  the  white  shrimp,  Penaeus  setiferus 
I  Linn.  I,  in  captivity.     Tul.  Stud.  Zool.  2:49-56. 
Klima,  E.  F. 

1974.     A  white  shrimp  mark-recapture  study.     Trans.  Am. 
Fish.  Soc.  103(11:107-113. 
Kneib,  R.  T. 

1991.  Flume  weir  for  quantitative  collection  of  nekton 
from  vegetated  intertidal  habitats.  Mar.  Ecol.  Prog.  Ser. 
75:29-38. 


Webb  and  Kneib:  Individual  growth  rates  and  movement  of  Litopenaeus  setiferus  in  a  tidal  marsh  nursery 


387 


1995.  Behaviour  separates  potential  and  realized  effects  of 
decapod  crustaceans  in  salt  marsh  communities.  J.  Exp. 
Mar.  Biol.  Ecol.  193:239-256. 

1997.  The  role  of  tidal  marshes  in  the  ecology  of  estuarine 
nekton.     Oceanogr.  Mar.  Biol.  Ann.  Rev.  35:163-220. 

2000.  Salt  marsh  ecoscapes  and  production  transfers  by 
estuarine  nekton  in  the  southeastern  United  States.  In 
Concepts  and  controversies  in  tidal  marsh  ecology  (M.  P. 
Weinstein  and  D.  A.  Kreeger.  eds.),  p.  267-291.  Kluwer 
Academic  Publishers,  Dordrecht,  The  Netherlands. 

Kneib,  R.  T..  and  M.  C.  Huggler. 

2001.  Tag  placement,  mark  retention,  survival  and  growth 
of  juvenile  white  shrimp  {Litopenaeus  setiferus  Perez-Far- 
fante,  1969)  injected  with  coded  wire  tags.  J.  Exp.  Mar. 
Biol.  Ecol.  266:109-120. 

Kneib,  R.  T,  and  S.  L.  Wagner. 

1994.     Nekton  use  of  vegetated  marsh  habitats  at  differ- 
ent stages  of  tidal  inundation.     Mar.  Ecol.  Prog.  Ser. 
106:227-238. 
Knudsen,  E.  E.,  B.  D.  Rogers,  R.  F.  Paille,  W.  H.  Herke,  and 
J.  P.  Geaghan. 

1996.  Juvenile  white  shrimp  growth,  mortality,  and  emi- 
gration in  weired  and  unweired  Lousiana  marsh  ponds. 
North  Am.  J.  Fish.  Manag.  16:640-652. 

Krouse,  J.  S.,  and  G.  E.  Nutting. 

1990.     Evaluation  of  coded  microwire  tags  inserted  in  legs 
of  small  juvenile  American  lobsters.     //;  Fish-marking 
techniques  (N.  C.  Parker,  et  al..  eds.),  p.  304-310.     Am. 
Fish.  Soc.  Symp.  7,  Bethesda,  MD. 
Kutkuhn,  J.  H. 

1966.     The  role  of  estuaries  in  the  development  and  per- 
petuation of  commercial  shrimp  resources.     Am.  Fish. 
Soc.  Spec.  Publ.  3:16-36. 
Lindner,  M.  J.,  and  W.  W.  Anderson. 

1956.     Growth,  migrations,  spawning  and  size  distribution 
of  shrimp,  Penaeus  setiferus.     Fish.  Bull.  56:555-645. 
Loesch,  H. 

1965.     Distribution  and  growth  of  penaeid  shrimp  in 
Mobile  Bay,  Alabama.     Publ.  Inst.  Mar.  Sci.,  Univ.  Texas 
10:41-58." 
Mayer,  M.  A. 

1985.     Ecology  of  juvenile  white  shrimp,  Penaeus  setiferus 
Linnaeus,  in  the  Salt  Marsh  Habitat.     M.S.  thesis,  62 
p.     Georgia  Inst.  Technology,  Atlanta,  GA. 
McTigue,  T  A.,  and  R.  J.  Zimmerman. 

1998.  The  use  of  infauna  by  juvenile  Penaeus  aztecus  Ives 
and  Penaeus  setiferus  (Linnaeus).  Estuaries  21:160- 
175. 

Minello,  T.  J,  and  R.  J.  Zimmerman. 

1985.     Differential  selection  for  vegetative  structure  between 
juvenile  brown  shrimp  {Penaeus  azteeus)  and  white  shrimp 
{Penaeus  setiferus),  and  implications  in  predator-prey 
relationships.     Estuar.  Coastal  Shelf  Sci.  20:707-716. 
Muncy,  R.  J. 

1984.  Species  profiles:  life  histories  and  environmental 
requirements  of  coastal  fishes  and  invertebrates  (South 
Atlantic) — white  shrimp.  U.S.  Fish  and  Wildlife  Service 
FWS/OBS-82/11.27,  19  p.  U.S.  Army  Corps  of  Engineers 
TR  EL-82-4,  19  p. 
Pearson,  J.  C. 

1939.     The  early  life  history  of  some  American  Penaei- 
dae,  chiefly  the  commercial  shrimp  Penaeus  setiferus 
(Linn.).     Fish.  Bull.  30:1-73. 
Perez-Farfante,  I. 

1969.  WesternAtlanticshrimpsofthegenusfc/iaci/s.  Fish. 
Bull.  67:461-591. 


Perez-Farfante,  I.,  and  B.  F.  Kensley. 

1997.  Penaeoid  and  sergestoid  shrimps  and  prawns  of  the 
world:  keys  and  diagnoses  for  the  families  and  genera. 
Memoires  du  Museum  National  d'Historie  Naturelle  175: 
1-233. 

Prentice,  E.  F.,  and  J.  E.  Rensel. 

1977.     Tag  retention  of  the  spot  prawn,  Pandalus  platyceros, 
injected  with  coded  wire  tags.     J.  Fish.  Res.  Board  Can. 
34:2199-2203. 
Ragotzkie,  R.  A.,  and  R.  A.  Bryson. 

1955.     Hydrography  of  the  Duplin  River,  Sapelo  Island. 
Georgia.     Bull.  Mar.  Sci.  Gulf  Caribb.  5:297-314. 
Shipman,  S. 

1983.  Mark-recapture  studies  of  penaeid  shrimp  in  Geor- 
gia, 1978-1981.  In  Studies  and  assessment  of  Georgia's 
marine  fisheries  resources,  1977-1981  (S.  Shipman.  V. 
Baisden,  and  H.  Ashley,  eds. ),  p.  1-176.  Georgia  Dep.  Nat. 
Res.  Contr.  Ser.  35. 
Sokal,  R.  R.,  and  F.  J.  Rohlf. 

1995.     Biometry:  the  principles  and  practice  of  statistics 
in  biological  research,  3rd  ed.,  887  p.     W.H.  Freeman  and 
Co.,  New  York,  NY. 
Turner,  R.  E. 

1977.  Intertidal  vegetation  and  commercial  yields  of  pen- 
aeid shrimp.  Trans.  Am.  Fish.  Soc.  106:411-416. 
1992.  Coastal  wetlands  and  penaeid  shrimp  habitat.  In 
Stemming  the  tide  of  coastal  fish  habitat  loss  ( R.  H.  Stroud, 
ed.),  p.  97-104.  National  Coalition  for  Marine  Conserva- 
tion, Savannah,  GA. 
Uglem,  I.,  and  S.  Grimsen. 

1995.     Tag  retention  and  survival  of  juvenile  lobsters,  Hom- 
marus  gammarus,  marked  with  coded  wire  tags.    Aquacult. 
Res.  26:837-841. 
van  Montfrans,  J.,  J.  Capelli.  R.  J.  Orth,  and  C.  H.  Ryer. 

1986.     Use  of  microwire  tags  for  tagging  juvenile  blue  crabs 
{Callinectes  sapidus).     J.  Crust.  Biol.  6:370-376. 
Vetter,  E.  A. 

1983.  The  ecology  of  Penaeus  setiferus:  habitat  selec- 
tion, carbon  and  nitrogen  metabolism,  and  simulation 
modeling.  Ph.D.  diss..  151  p.  Univ.  Georgia.  Athens. 
GA. 

Wadsworth,  J.  R. 

1980.     Geomorphic  characteristics  of  tidal  drainage  net- 
works in  the  Duplin  River  system,  Sapelo  Island,  Georgia. 
Ph.D.  diss.,  247  p.    Univ.  Georgia,  Athens,  GA. 
Webb,  S.  R.,  and  R.  T  Kneib. 

2002.     Abundance  and  distribution  of  juvenile  white 
shrimp  Litopenaeus  setiferus  within  a  tidal  marsh  land- 
scape.    Mar.  Ecol.  Prog.  Ser.  232:213-223. 
Weinstein,  M.  P. 

1979.     Shallow  marsh  habitats  as  primary  nurseries  for 
fishes  and  shellfish.  Cape  Fear  River,  North  Carolina. 
Fish.  Bull.  77:339-357. 
Wenner.  E.  L„  D.  Knott,  J.  O.  Blanton,  C.  Barans,  and  J.  Amft. 

1998.  Roles  of  tidal  and  wind-generated  currents  in  trans- 
porting white  shrimp  {Penaeus  setiferus)  postlarvae 
through  a  South  Carolina  (USA)  inlet.  J.  Plank.  Res. 
20(121:2333-2356. 

Williams,  A.  B. 

1955.  Contribution  to  the  life  histories  of  commercial 
shrimp  (Penaeidae)  in  North  Carolina.  Bull.  Mar.  Sci. 
Gulf  Caribb.  5:116-146. 

1984.  Shrimps,  lobsters,  and  crabs  of  the  Atlantic  coast 
of  the  eastern  United  States,  Maine  to  Florida,  550  p. 
Smithsonian  Institution  Press,  Washington,  D.C. 


388  Fishery  Bulletin  102(2) 


Zein-Eldin,  Z.  P.,  and  G.  W.  Griffith.  Zein-Eldin,  Z.  P.,  and  M.  L.  Renaud. 

1969.     An  appraisal  of  the  effects  of  salinity  and  tempera-  1986.     Inshore  environmental  effects  on  brown  shrimp, 

ture  on  growth  and  survival  of  postlarval  penaeids.     FAO  Penaeus  aztecus,  and  white  shrimp.  P.  setiferus,  popula- 

Fish.  Rep.  57(3):1015-1026.  tions  in  coastal  waters,  particularly  of  Texas.     Mar.  Fish. 

Rev.  48(3):9-19. 


389 


Does  the  California  market  squid 
(Loligo  opalescens)  spawn  naturally  during 
the  day  or  at  night?  A  note  on  the  successful  use 
of  ROVs  to  obtain  basic  fisheries  biology  data 


John  Forsythe 

National  Resource  Center  for  Cephalopods 
University  of  Texas  Medical  Branch  at  Galveston 
301  University  Blvd 
Galveston,  Texas  77555-1163 
E-mail  address:  |ohn  forsythetg' utmb.edu 

Nuutti  Kangas 

Roger  T.  Hanlon 

Marine  Resources  Center 
Marine  Biological  Laboratory 
Woods  Hole,  Massachusetts  02543 


The  California  market  squid  iLoligo 
opalescens  Berry),  also  known  as  the 
opalescent  inshore  squid  ( FAO ),  plays  a 
central  role  in  the  nearshore  ecological 
communities  of  the  west  coast  of  the 
United  States  (Morejohn  et  al.,  1978; 
Hixon,  1983)  and  it  is  also  a  prime 
focus  of  California  fisheries,  ranking 
first  in  dollar  value  and  tons  landed 
in  recent  years  (Vojkovich,  1998).  The 
life  span  of  this  species  is  only  7-10 
months  after  hatching,  as  ascertained 
by  aging  statoliths  ( Butler  et  al.,  1999; 
Jackson,  1994;  Jackson  and  Domier, 
2003)  and  mariculture  trials  (Yang,  et 
al.,  1986).  Thus,  annual  recruitment  is 
required  to  sustain  the  population.  The 
spawning  season  ranges  from  April  to 
November  and  spawning  peaks  from 
May  to  June.  In  some  years  there  can 
be  a  smaller  second  peak  in  November. 
In  Monterey  Bay,  the  squids  are  fished 
directly  on  the  egg  beds,  and  the  con- 
sequences of  this  practice  for  conser- 
vation and  fisheries  management  are 
unknown  but  of  some  concern  (Hanlon, 
1998).  Beginning  in  April  2000,  we 
began  a  study  of  the  in  situ  spawning 
behavior  of  L.  opalescens  in  the  south- 
ern Monterey  Bay  fishing  area. 

The  prevailing  thought  is  that  the 
majority  of  spawning  activity  takes 
place  at  night  because  fishermen  have 
observed  these  squids  mating  under 
their  bright  lights  (which  are  used  to 


attract  and  capture  squids)  and  be- 
cause television  documentaries  have 
revealed  mating  and  spawning  activ- 
ity in  large  aggregations  at  night.  The 
scientific  literature  on  reproductive 
behavior  is  sparse.  There  are  some  cur- 
sory observations  of  actively  spawning 
L.  opalescens  during  diver  surveys 
of  egg  beds  (McGowan,  1954;  Fields, 
1965;  Hobson,  1965;  Hurley,  1977). 
Some  daytime  spawning  has  been 
seen  both  in  southern  and  northern 
California  but  Fields  ( 1965 )  and  Hixon 
(1983)  suggested  indirectly  that  most 
spawning  occurs  at  night.  Shimek  et 
al.  (1984)  also  suggested  night  spawn- 
ing by  L.  opalescens  in  Canada.  Other 
loliginid  squids  whose  natural  behav- 
ior has  been  studied  in  the  field  were 
found  to  be  daytime  spawners  (e.g.,  L. 
pealeii,  L.  vulgaris  reynaudii,  Sepioteu- 
this  sepioidea;  summarized  in  Hanlon 
and  Messenger,  1996). 

To  help  resolve  this  issue,  we  con- 
ducted three  field  expeditions  (28 
April-8  May  2000,  10-17  September 
2000,  and  16-21  August  2001)  using 
remotely  operated  vehicles  (ROVs)  de- 
ployed either  from  the  RV  John  Martin 
(Moss  Landing  Marine  Laboratory)  or 
the  commercial  squid  FV  Lady  J.  The 
ROVs  were  tethered  vehicles  with  on- 
board video  cameras  and  lights.  Live 
video  signals  were  transmitted  by 
the  tether  to  shipboard  VCRs  where 


observational  data  were  viewed  and 
recorded.  For  the  first  field  trip,  a 
large  S4  Phantom  ROV  was  used;  it 
was  outfitted  with  a  video  camera  and 
zoom  lens  with  tilt  capability,  and  the 
video  was  recorded  on  Hi8  format  video 
decks.  For  the  second  and  third  trips, 
a  smaller  S2  inspection-class  Phantom 
ROV  on  loan  from  the  NOAA  Sustain- 
able Seas  Expeditions  was  used;  this 
ROV  had  a  customized  fiber-optic  teth- 
er and  the  video  data  were  recorded  on 
mini-digital  video  cassettes.  Our  goal 
was  to  make  ROV  dives  each  day  from 
approximately  dawn  to  dusk  and  to 
make  a  few  comparable  all-night  sur- 
veys. A  combination  of  adverse  weather 
conditions  and  technical  problems  with 
the  ROVs  rarely  allowed  continuous 
video  observations.  During  dives,  if 
squids  were  encountered,  we  used 
video  to  conduct  focal  animal  samples 
on  females  (which  were  paired)  for  as 
long  as  squids  were  present,  or  as  long 
as  we  could  keep  track  of  the  same 
individuals.  Unless  absolutely  neces- 
sary to  see  the  squids  (for  instance  at 
night  or  at  depths  greater  than  30  m 
in  turbid  daytime  conditions),  lights 
were  not  used  for  video  taping  in  an 
effort  to  minimize  their  impact  on 
the  mating  squids.  Squids  acclimated 
within  minutes  to  the  ROVs.  After  the 
expeditions,  the  videotapes  were  stud- 
ied and  the  behavioral  and  biological 
data  were  quantified  on  a  multimotion 
playback  VCR. 

By  "mating"  we  refer  to  the  peculiar 
mating  behavior  of  this  species  that 
is  unique  among  loliginid  squids.  The 
male  firmly  grasps  the  female  from  her 
ventral  side  and  holds  her  for  minutes 
or  hours  in  a  "copulatory  embrace"  in 
a  nearly  vertical  position.  Both  copula- 
tion (i.e.,  transfer  of  spermatophores) 
and  deposition  of  egg  capsules  occur 
in  this  posture.  For  example,  as  the 
female  exudes  a  new  egg  capsule,  the 
male  and  female  lower  themselves  in 
unison  to  the  egg  bed  where  the  female 
deposits  the  egg  capsule  in  the  sand. 
We  have  reported  elsewhere  on  egg- 


Manuscript  approved  for  publication 
20  January  2004  by  Scientific  Editor. 

Manuscript  received  25  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:389-392  (2004). 


390 


Fishery  Bulletin  102(2) 


No  Maling  Squids 


Mating  Squids  Present 


29Apr00-, 
30Apr00 

n  i  May Uli 
04May00 
OSMayOO 
OtSMayOO 
07May00 


•Wk 


^mmm-M 


Sunset 


b       mw?. 


wmmmmr* 


lOSepOO 
12Sep00 
USepOO 
16SepOO 


0000      0200       0400       0600       0800       1000       1200       1400 

Time  of  day  (h) 


1600       1800       2000       2200      2400 


Figure  1 

A  summary  of  154  spawning  groups  of  Lot igo  opalescens  showing  the  daily  presence  or 
absence  of  mating  squids  on  egg  beds  in  Monterey  Bay,  California.  The  horizontal  bars  repre- 
sent the  time  periods  when  a  ROV  was  on  the  bottom  searching  for  mating  squids  during  three 
different  expeditions.  Note  three  occasions  in  which  operations  were  conducted  continuously 
through  the  day  and  night,  and  also  that  spawning  ceased  at  dusk. 


laying  frequency  (Hanlon  et  al.,  in  press).  Only  very  rarely 
did  we  observe  females  laying  eggs  while  unattended  by  a 
male;  in  these  cases  the  female  was  moribund  and  laying 
her  last  few  egg  strands. 


Results  and  discussion 

We  examined  28  hours  of  videotape  recorded  during  50 
ROV  dives  over  18  days  (divided  into  three  expeditions). 
Figure  1  illustrates  the  relative  presence  or  absence  of 
mating Lo//go  opalescens  throughout  24-hour  periods.  The 
large  gaps  in  the  daytime  observation  record  were  due  to 
ROV  problems.  Although  observation  times  varied  daily, 
it  is  clear  that  normal  mating  and  egg-laying  behaviors 
were  exclusively  observed  during  daylight  hours  (ca. 
0800-1800  hours  but  with  some  seasonal  variation)  and 
concluded  near  dusk.  In  all  instances  in  Figure  1  where 
egg-laying  extended  into  the  early  evening,  these  mating 
assemblages  had  formed  during  daylight  hours  and  per- 
sisted slightly  past  sunset  and  the  number  of  participating 
squids  constantly  decreased  as  sunset  approached  and 
passed.  Observations  were  made  throughout  the  night 
on  three  nights.  Not  only  were  no  mating  squids  ever 
encountered  around  the  egg  beds  at  night,  but  generally 
no  squids  were  encountered  at  all  near  the  seabed,  despite 
large  aggregations  that  were  present  higher  in  the  water 
column.  Thus  the  200-400  watt  lights  on  the  ROVs  never 


induced  any  artificial  spawning  behavior  because  there 
were  no  squids  present. 

Figure  2  provides  some  quantification  of  Figure  1.  This 
graph  is  based  on  154  spawning  groups  that  were  vid- 
eotaped and  includes  all  three  trips  as  well  as  the  three 
"all  night  observations"  illustrated  in  Figure  1.  We  were 
studying  discrete  groups  of  squids  to  examine  mating  dy- 
namics and  thus  were  sometimes  biased  to  smaller  groups 
of  squids  that  could  be  kept  in  view.  Overall,  we  observed 
that  squids  were  present  in  greatest  numbers  in  mid  to 
late  afternoon  and  absent  during  the  night. 

Our  findings  strongly  indicate  that  the  extensive  egg 
beds  produced  at  depths  of  20-60  m  in  southern  Monterey 
Bay  (just  beyond  the  kelp  beds)  are  the  result  of  daytime 
aggregations  of  mating  Loligo  opalescens.  These  benthic 
aggregations  begin  forming  in  the  early  morning  hours 
and  tend  to  be  larger  in  the  afternoon.  Reproductive  ac- 
tivity begins  to  wane  toward  sunset  and  comes  to  a  near 
halt  at  sunset.  We  could  find  no  evidence  that  egg  laying 
occurs  naturally  during  the  night.  All  observations  that 
we  are  aware  of  (mainly  television  documentaries)  have 
occurred  in  the  presence  of  artificial  light  sources  near  the 
surface  provided  either  by  fishermen  or  cinematographers. 
In  the  absence  of  artificial  lighting,  L.  opalescens  in  South 
Monterey  Bay  does  not  aggregate  into  mating  and  spawn- 
ing groups  at  night.  Thus,  we  conclude  that  all  significant 
egg  deposition  in  the  Monterey  Bay  fishery  is  the  result  of 
daytime  aggregations  of  squids. 


NOTE     Forsythe  et  al.:  Spawning  patterns  of  Loligo  opa/escens 


391 


Two  other  ascribed  characteristics  of  L. 
opalescens  spawning  are  mass  aggregations 
at  the  sea  floor  and  subsequent  die-offs  after 
squids  have  spawned.  Mass  aggregations  can 
be  detected  by  standard  fathometers  used  by 
commercial  fishermen,  who  report  that  mass 
aggregations  on  the  sea  floor  are  rare  in  Mon- 
terey Bay.  During  our  ROV  operations  we  en- 
countered only  one  large  aggregation,  which 
occurred  on  21  August  2001.  We  estimated 
from  our  video  recordings  that  there  were  ap- 
proximately 3000-4000  squid  in  a  50-m2  area 
on  the  sea  floor  and  that  intermittent  egg  lay- 
ing was  occurring  over  an  area  of  ca.  2000  m2 
during  a  period  of  about  3  hours.  Collectively, 
then,  we  recorded  154  very  small  spawning 
groups  and  one  large  spawning  group.  There 
was  no  mass  die-off  during  or  after  this  large 
spawning  aggregation.  Instead,  we  consis- 
tently observed  in  all  spawning  groups  that 
females  actively  broke  the  embrace  of  the 
paired  male  and  jetted  strongly  upwards  away 
from  the  spawning  groups  and  rejoined  large 
schools  in  the  upper  water  column.  Thus, 
squids  that  dispersed  from  the  egg  beds  were 
consistently  in  excellent  condition  —  certainly 
not  senescent  or  moribund.  These  observa- 
tions corroborate  the  results  of  other  studies  on  loliginid 
squids  that  spawn  intermittently  (Moltschaniwskyj, 
1995;  Maxwell  and  Hanlon,  2000 ).  Twice  we  encountered 
large  numbers  of  dead  squids  on  the  sea  floor  in  the  early 
morning,  but  in  both  instances  the  squid  fishing  fleet 
had  been  working  in  the  same  area  the  night  before  and 
it  appeared  as  though  these  mortalities  were  associated 
with  the  purse-seine  fishery;  there  were  few  eggs  in  those 
localities.  McGowan  ( 1954 ),  Hobson  ( 1965 ),  and  Cousteau 
and  Diole  (1973)  reported  that  squids  died  after  spawn- 
ing in  S.  California.  Various  Loligo  spp.  are  noted  for 
flexible  reproductive  strategies  (cf.  Hanlon  and  Messen- 
ger, 1996)  so  it  should  not  be  surprising  if  L.  opalescens 
occasionally  engaged  in  large  reproductive  events.  Our 
data  suggest  that  small  groups  of  squids  (20-200  indi- 
viduals) generally  descend  during  the  day  and  lay  eggs 
for  several  hours  before  rejoining  squids  in  the  water 
column.  We  encourage  other  researchers  to  use  ROVs  or 
SCUBA  without  lights  and  with  stealthy  approaches  to 
determine  the  natural  diurnal  spawning  of  L.  opalescens 
throughout  its  range.  Given  our  findings  that  active  sex- 
ual selection  processes  are  occurring  during  the  day  and 
that  there  is  vertical  migration  between  the  large  schools 
of  squid  in  the  water  column  and  the  small  spawning 
groups  at  the  substrate,  it  would  be  prudent,  at  the  very 
least,  to  restrict  daytime  fishing  directly  over  egg  beds 
or  to  create  protected  spawning  areas  in  southern  Mon- 
terey Bay.  This  strategy  would  allow  the  complex  mating 
system  of  L.  opalescens  to  be  played  out  without  direct 
disruption  by  fishing  activity.  In  such  a  short-lived  spe- 
cies, annual  recruitment  to  the  population  is  necessary; 
thus  sufficient  eggs  must  be  laid  for  each  new  generation 
to  ensure  a  viable  living  resource. 


30    - 

n  =  1 54  groups 

25 

- 

CO 

Q. 

=J 

o 

oi    20 

- 

en 

c 

"c 

I      15 

■ 

■ 

Q. 

(fi 

O 

_■ 

5      10 

ll 

E 

3 

■ 

z 

- 

5 

1 

III 

III! 

o 

"' 

III 

1 

Illl 

0000    02O0    0400    0600    0800     1000    1200     1400     1600    1800    2000    2200    2400 

Time  of  day  (h) 

Figure  2 

The  number  of  squid  spawning  groups  at  the  egg  beds  at  different  times 

of  day  (data  pooled  from  three  expeditions  over  2  years;  n= 154  groups). 

A  group  is  denned  as  a  group  of  two  or  more  mating  pairs.  These  data 

correspond  hour  by  hour  with  data  in  Figure  1. 

Acknowledgments 

We  are  most  grateful  for  funding  on  NOAA  grant  UAF  98 
0037  from  the  National  Undersea  Research  Center  (West 
Coast).  Additional  funding  was  provided  by  the  David  and 
Lucile  Packard  Foundation  and  the  Sholley  Foundation. 
J.  Forsythe  gratefully  acknowledges  financial  support  for 
travel  from  the  National  Institutes  of  Health,  National 
Center  for  Research  Resources  (grant  P40  RR0102423- 
23),  and  the  Marine  Medicine  General  Budget  account  of 
the  Marine  Biomedical  Institute.  N.  Kangas  gratefully 
acknowledges  financial  support  from  the  Academy  of 
Finland.  We  thank  Sylvia  Earle  for  loan  of  the  Sustain- 
able Seas  ROV  and  we  appreciate  the  professional  efforts 
of  Deep  Ocean  Exploration  and  Research  (DOER)  who 
supported  the  ROV  operations.  We  especially  thank  John 
Rummel  who  helped  begin  this  project,  and  Brett  Hobson 
who  kept  it  going  at  a  critical  juncture.  We  are  thankful  for 
expert  shipboard  assistance  from  the  captains  and  crew  of 
the  KVJohn  Martin  and  the  FV Lady  J  (especially  Captain 
Tom  Noto).  We  benefitted  from  discussions  with  Bob  Leos, 
Bill  Gilly,  Annette  Henry,  John  Butler,  Teirney  Thies,  and 
Sue  Houghton. 


Literature  cited 

Butler,  J.,  D.  Fuller,  and  M.  Yaremko. 

1999.     Age  and  growth  of  market  squid  iLoligo  opalescens) 
off  California  duringl998.     CalCOFI  Rep.  40:191-195. 
Cousteau  J.-Y..  and  P.  Diole. 

1973.     Octopus  and  squid:  the  soft  intelligence,  304  p.     Cas- 
sell,  London. 


392 


Fishery  Bulletin  102(2) 


Fields,  W.  G. 

1965.     The  structure,  development,  food  relations,  repro- 
duction, and  life  history  of  the  squid  Loiigo  opalescens 
Berry.     Fish.  Bull.  131:1-108. 
Hanlon,  R.  T. 

1998.     Mating  systems  and  sexual  selection  in  the  squid 
Loiigo:  How  might  commercial  fishing  on  spawning  squids 
affect  them?     CalCOFI  Rep.  39:92-100. 
Hanlon,  R.T.,  N.  Kangas,  and  J.  W.  Forsythe. 

In  press.     Rate  of  egg  capsule  deposition,  changing  OSR 
and  associated  reproductive  behavior  of  the  squid  Loiigo 
opalescens  on  spawning  grounds.     Mar.  Biol. 
Hanlon,  R.  T.,  and  J.  B.  Messenger. 

1996.     Cephalopod  behaviour,  232  p.     Cambridge  Univ. 
Press,  Cambridge. 
Hixon.  R.  F. 

1983.     Loiigo  opalescens.     In  Cephalopod  life  cycles,  vol.  I, 
species  accounts,  475  p.     Academic  Press,  London. 
Hobson,  E.  S. 

1965.     Spawning  in  the  Pacific  Coast  squid.  Loiigo  opal- 
escens.   Underwater  Naturalist  3:20-21. 
Hurley,  A.  C. 

1977.     Mating  behavior  of  the  squid  Loiigo  opalescens. 
Mar.  Behav.  Physiol.  4:195-203. 
Jackson,  G.  D. 

1994.     Statolith  age  estimates  of  the  loliginid  squid  Loiigo 
opalescens:  corroboration  with  culture  data.     Bull.  Mar. 
Sci.  54:554-557. 
Jackson,  G.  D..  and  M.  L.  Domeier. 

2003.  The  effects  of  an  extraordinary  El  Nino/La  Nina 
event  on  the  size  and  growth  of  the  squid  Loiigo  opalescens 
off  Southern  California.     Mar.  Biol.  142:925-935. 


Maxwell,  M.  R.,  and  R.  T.  Hanlon. 

2000.     Female  reproductive  output  in  the  squid  Loiigo  pea- 
leii:  multiple  egg  clutches  and  implications  for  a  spawning 
strategy.     Mar.  Ecol.  Prog.  Ser.  199:159-170. 
McGowan,  J.  A. 

1954.     Observations  on  the  sexual  behavior  and  spawning 
of  the  squid,  Loiigo  opalescens,  at  LaJolla,  CA.     Cal.  Fish 
Game  40:47-54. 
Moltschaniwskyj,  N.  A. 

1995.     Multiple  spawning  in  the  tropical  squid  Photolo- 
ligo  sp.:  what  is  the  cost  in  somatic  growth?     Mar.  Biol. 
124:127-135. 
Morejohn,  G.  V.,  J.  T.  Harvey,  and  L.  T.Krasnow. 

1978.  The  importance  of  Loiigo  opalescens  in  the  food  web 
of  marine  vertebrates  in  Monterey  Bay,  California.  In 
Biological,  oceanographic,  and  acoustic  aspects  of  the 
market  squid,  Loiigo  opalescens  Berry  (C.  W.  Recksiek 
and  H.  W  .  Frey,  eds.),  p.  67-98.  Calif.  Dep.  Fish  Game 
Fish  Bull.  169. 
Shimek,  R.  L.,  D.  Fyfe.  L.  Ramsey,  A.  Bergey,  J.  Elliott, 
and  S.  Guy. 

1984.     A  note  on  spawning  of  the  Pacific  market  squid.  Loiigo 
opalescens  (Berry,  1911),  in  the  Barkley  sound  region,  Van- 
couver Island,  Canada.     Fish.  Bull.  82:445-446. 
Vqjkovich,  M. 

1998.     The  California  fishery  for  market  squid  (Loiigo 
opalescens).     CalCOFI  Rep.  39:55-60. 
Yang,  W.  T,  R.  F.  Hixon,  P.  E.  Turk.  M.  E.  Krejci,  W.  H.  Hulet, 
and  R.  T.  Hanlon. 

1986.  Growth,  behavior,  and  sexual  maturation  of  the 
market  squid,  Loiigo  opalescens,  cultured  through  the  life 
cycle.     Fish.  Bull.  84:771-798. 


393 


Incidental  capture  of  loggerhead  (Caretta  caretta) 
and  leatherback  (Dermochelys  coriacea) 
sea  turtles  by  the  pelagic  longline  fishery 
off  southern  Brazil 


Jorge  E.  Kotas 

IBAMA/Acordo  Projeto  TAMAR- 

Instituto  de  Pesca/CPPM 
Programa  REVIZEE-SCORE  SUL 
Rodovia  Osvaldo  Reis  345  apt.  22  C 
Itajaf-SC  88306-001.  Brazil 

Silvio  dos  Santos 

DTI-CNPq 

Programa  REVIZEE-SCORE  SUL 
Rua  Ezio  Testlni  320 
Santos-SP  11089-210,  Brazil 


Berenice  M.  G.  Gallo 

Fundacao  Pro-TAMAR 
Rua  Antonio  Athanasio  273 
Ubatuba-SP  11680-000,  Brazil 


Paulo  C.  R.  Barata 

Fundacao  Oswaldo  Cruz 

Rua  Leopoldo  Bulhoes  1480  -  8A 

Rio  de  Janeiro  -  RJ  21041-210,  Brazil 

E-mail  address  (for  P.  C.  R.  Barata,  contact  author): 

pbarataigialternex.com.br 


back  sea  turtles  by  the  surface  longline 
fishery  operating  off  the  southern  coast 
of  Brazil,  within  Brazil's  200  mile 
exclusive  economic  zone  (EEZ)  and 
in  international  waters,  and  present 
catch-per-unit-of-effort  (CPUE)  data 
and  estimates  of  average  probability 
of  death  at  capture  for  these  species. 
Preliminary  results  of  incidental  cap- 
tures of  sea  turtles  by  longliners  dur- 
ing one  longline  trip  in  this  area  were 
presented  by  Barata  et  al.2  In  the 
present  study  we  provide  more  detailed 
data  from  additional  trips,  including 
information  concerning  leatherback 
sea  turtles,  as  well  as  analyses  of  these 
data.  To  our  knowledge,  this  is  the  first 
detailed  report  about  the  incidental 
capture  of  sea  turtles  by  the  Brazilian 
commercial  longline  fleet. 


Venancio  G.  de  Azevedo 

DTI-CNPq 

Programa  REVIZEE-SCORE  SUL 

Av.  Pavao  1 64 

Caraguatatuba-SP  11676-520,  Brazil 


Incidental  capture  in  fishing  gear  is 
one  of  the  main  sources  of  injury  and 
mortality  of  juvenile  and  adult  sea 
turtles  (NRC,  1990;  Lutcavage  et  al., 
1997;  Oravetz,  1999).  Six  out  of  the 
seven  extant  species  of  sea  turtles — the 
leatherback  (Dermochelys  coriacea), 
the  green  turtle  (Chelonia  mydas), 
the  loggerhead  (Caretta  caretta),  the 
hawksbill  (Eretmochelys  imbricata), 
the  olive  ridley  (Lepidochelys  olivacea). 
and  the  Kemp's  ridley  (Lepidochelys 
kempii)  —  are  currently  classified  as 
endangered  or  critically  endangered  by 
the  World  Conservation  Union  (IUCN, 
formerly  the  International  Union  for 
Conservation  of  Nature  and  Natural 
Resources),  which  makes  the  assess- 
ment and  reduction  of  incidental  cap- 
ture and  mortality  of  these  species  in 
fisheries  priority  conservation  issues 
(IUCN/Species  Survival  Commission, 
1995). 

Several  studies  have  examined  sea 
turtle  bycatch  by  pelagic  longline  fish- 
eries, especially  in  the  North  Atlantic 
and  Pacific  oceans  (NRC,  1990;  Nish- 
emura  and  Nakahigashi,  1990;  Tobias, 
1991;  Bolten  et  al.,  1996;  Williams  et 


al.,  1996;  Lutcavage  et  al.,  1997),  but 
little  is  known  about  sea  turtle  bycatch 
in  the  South  Atlantic.  One  of  the  most 
detailed  reports  on  longline  incidental 
captures  in  that  area  is  that  by  Acha- 
val  et  al.  (2000),  which  documents  the 
incidental  capture  of  loggerhead  and 
leatherback  sea  turtles  in  the  south- 
western Atlantic  by  longliners  target- 
ing swordfish  (Xiphias  gladius),  tuna 
(Thunnus  obesus),  and  other  related 
species.  Additional  references,  some- 
times with  scant  detail,  can  be  found  in 
Weidner  and  Arocha  ( 1999 ),  Fallabrino 
et  al.  (2000),  and  Domingo  et  al.1 

In  this  study,  we  report  the  inciden- 
tal capture  of  loggerhead  and  leather- 


1  Domingo,  A..  A.  Fallabrino,  R.  Forselledo, 
and  V.  Quirici.  2002.  Incidental  cap- 
ture of  loggerhead  (Caretta  caretta)  and 
leatherback  (Dermochelys  coriacea)  sea 
turtles  in  the  Uruguayan  long-line  fish- 
ery in  Southwest  Atlantic.  Presented 
at  the  22nd  Annual  Symposium  on  Sea 
Turtle  Biologv  and  Conservation,  Miami, 
USA,  4-7  April  2002.  [Available  from  A. 
Domingo:  Direccidn  Nacional  de  Recur- 
sos  Acuaticos,  Constituyente  1497,  C.R 
11.200,  Montevideo,  Uruguay.] 


Materials  and  methods 

Observations  were  carried  out  by  three 
of  the  authors  (JEK.  SS,  and  VGA) 
during  three  trips  aboard  Brazil- 
flagged  commercial  longline  vessels 
based  in  Itajai,  State  of  Santa  Cata- 
rina,  southern  Brazil  (Fig.  1).  The 
trips  occurred  in  1998,  the  first  (10 
sets)  between  13  March  and  12  April 
(summer-fall),  the  second  (13  sets) 
between  15  June  and  5  July  (fall- 
winter  I,  and  the  third  (11  sets  (between 
28  September  and  13  October  (spring), 
and  took  place  between  latitudes 
27°30'S  and  34°30'S  and  longitudes 
36°00'W  and  52°00'W  (Fig.  1).  The 


Barata,  P.  C.  R.,  B.  M.  G.  Gallo,  S.  dos 
Santos,  V.  G.  Azevedo,  and  J.  E.  Kotas. 
1998.  Captura  acidental  da  tartaruga 
marinha  Caretta  caretta  (Linnaeus,  1758) 
na  pesca  de  espinhel  de  superficie  na  ZEE 
brasileira  e  em  aguas  internacionais.  In 
Resumos  Expandidos  da  XI  Semana 
Nacional  de  Oceanografia,  Rio  Grande, 
RS,  outubro  de  1998,  p.  579-581.  Edi- 
tora  Universitaria-UFPel,  Pelotas,  RS, 
Brazil.  [Available  from  FURG,  Oceano- 
logia.  Av.  Italia,  km  8,  Campus  Carreiros, 
C.P  474,  Rio  Grande,  RS  96201-900, 
Brazil.! 


Manuscript  approved  for  publication 
22  December  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:393-399(2004). 


394 


Fishery  Bulletin  102(2) 


t — i — i — i — i — i — vr—r 
Atlantic  Ocean 


J. 


10  N 


40    - 


50  S  - 


Ascension 


St.  Helena 


Tristan  da  Cunha 


South  Georgia 


N 

t 


j i i i i i i    i     i 


80  W 


20 


10 


10 


20  E 


Figure  1 

Fishing  locations.  Numbers  1,  2,  and  3  indicate  locations  of  the  first,  second,  and  third 
longline  trips  respectively;  for  each  location,  one  or  more  sets  were  performed.  Circled 
numbers  indicate  international  waters  outside  the  200-mile  Brazilian  exclusive  economic 
zone.  The  rectangular  ocean  area  is  limited  by  latitudes  25°S  and  35°S  and  longitude  35°W. 
The  fishing  location  farthest  to  the  east  is  about  1320  km  (713  nautical  miles)  from  Itajai, 
State  of  Santa  Catarina,  Brazil,  the  home  port  of  the  fishing  vessels. 


seabed  in  this  area  ranged  from  the  continental  shelf 
border  to  abyssal  plains,  including  submarine  elevations 
(e.g.,  Rio  Grande).  Operation  depths,  ranging  from  170  to 
4000  m,  were  obtained  from  nautical  charts. 

The  first  and  second  trips  were  aboard  the  Yamaya  III, 
a  20.7-m,  325-hp  engine,  30-t  hold  capacity,  10-crew  long- 
liner,  and  the  third  trip  was  aboard  the  Basco,  a  24.4-m, 
330-hp  engine,  70-t  hold  capacity,  11-crew  longliner. 
The  vessels  targeted  swordfishes,  sharks  (mainly  blue 
sharks,  Prionace  glauca)  and  tunas  (Thunnus  albacares, 
T.  alalunga  and  T.  obesus).  Their  fishing  gear  was  the 
U.S. -style  monofilament  nylon  longline,  with  200- 
300  m  sections  between  buoys,  and  each  section  contained 
four  to  five  gangions  set  40-60  m  apart.  Buoy  dropper 
length  ranged  between  10  and  20  m,  and  gangion  length 
ranged  between  13  and  20  m.  Each  non-offset  "J"  hook 
(Swordfish  9/0)  was  baited  with  Argentine  shortfin  squid 
(Illex  argentinus)  and  had  a  yellow  chemical  light  stick 
hung  over  it.  The  average  number  of  hooks  per  set  was 
1030,  992,  and  950  on  the  first,  second,  and  third  trips, 
respectively. 

On  the  first  and  second  trips,  the  mainline  was  set  off  the 
stern  by  means  of  a  line  shooter  so  that  a  marked  catenary 
was  formed  between  buoys,  allowing  the  hooks  to  operate 
at  a  greater  depth.  In  this  case,  the  maximum  hook  depth 
may  have  reached  more  than  40  m.  On  the  third  trip,  the 
vessel  Basco  did  not  use  a  line  shooter,  and  thus  the  hook 


depth  for  that  trip  may  have  been  shallower.  The  longline 
gear  was  set  around  5:30  PM,  and  was  retrieved  early  in 
the  morning.  The  average  soak  time  was  7  h  30  min.  For 
each  set,  the  date,  time,  geographical  position,  number  of 
hooks,  and  sea  surface  temperature  were  recorded.  The 
species  and  condition  (i.e.,  if  the  animal  was  alive  or  dead) 
of  captured  turtles  were  recorded;  specimens  with  no  ap- 
parent movement  were  considered  dead. 

Incidentally  captured  loggerhead  turtles  were  taken 
aboard  and  hooks  and  lines  were  then  removed.  Whenever 
possible  curved  carapace  length  (CCL)  and  width  were 
measured,  and  the  turtles  were  double  tagged  (inconel 
tags  style  681,  National  Band  and  Tag  Co.,  Newport, 
KY),  according  to  Projeto  TAMAR's  (Projeto  Tartaruga 
Marinha,  the  Brazilian  sea  turtle  conservation  program) 
standard  methods  (Marcovaldi  and  Laurent,  1996).  In 
some  cases,  it  was  not  possible  to  bring  loggerhead  sea 
turtles  on  board  the  fishing  vessel  and,  because  of  their 
great  size,  no  leatherback  sea  turtles  were  brought  on 
board.  On  these  occasions,  the  turtles  were  pulled  close  to 
the  boat  and  the  gangions  were  then  cut  to  free  the  turtles 
with  the  hooks  still  attached  to  them;  however  the  length 
of  the  line  remaining  on  the  turtle  was  not  recorded.  None 
of  these  turtles  was  measured  or  tagged,  although  some  of 
the  leatherback  sea  turtles  were  filmed  on  video.  No  addi- 
tional data  and  measurements,  other  than  those  presented 
in  this  study,  were  obtained. 


NOTE     Kotas  et  al.:  Incidental  capture  of  Caretta  caretta  and  Dermochelys  coriacea  by  the  pelagic  longline  fishery 


395 


Table  1 

Data  referring  to  fishing  practices,  sea  surface  temperature  CO,  and  capture  of  loggerhead  and  leatherback  sea  turtles. 
CPUE  =  catch-per-unit-of-effort  (number  of  captured  turtles/1000  hooks). 

by  trip. 

Trip 

Date 

No.  of 
sets 

Average 
hooks/set 

Average 
sea  surface 
temperature 

Loggerheads 

Leatherbacks 

Alive 
(tagged) 

Dead 

Condition 
not  recorded 

CPUE 

Condition 
Alive     Dead      not  recorded 

CPUE 

1 

13  Mar  98- 
12  Apr  98 

10 

1030 

13.6 

84(17) 

15 

9 

10.49 

1                             — 

0.10 

2 

15  Jun  98- 
5  Jul  98 

13 

992 

21.4 

28(12) 

4 



2.48 

13           1 

1.09 

3 

28  Sep  98- 
13  Oct  98 

11 

950 

18.9 

5(5) 

_ 

_ 

0.48 

5 

0.48 

Total 

34 

990 

117(34) 

19 

9 

4.31 

19           1                 — 

0.59 

CPUE  (number  of  captured  turtles/1000  hooks)  was 
calculated  separately  for  each  species.  Straight  carapace 
lengths  in  published  data  were  converted  to  CCL  by  us- 
ing the  formula  in  Teas  (1993)  to  compare  the  CCL  of 
captured  loggerhead  sea  turtles  to  carapace  length  data 
found  in  the  literature.  To  assess  the  significance  of  the 
difference  in  the  proportion  of  dead  loggerhead  or  leath- 
erback sea  turtles  among  trips,  exact  tests  were  applied, 
because  ordinary  chi-square  tests  are  not  reliable  when 
expected  cell  frequencies  are  too  small.  The  test  statistics 
were  x2  =  ^[(Observed  -  Expected )'2IExpected] ,  and  exact 
probabilities  were  computed  for  all  tables  with  marginal 
frequencies  fixed  at  the  observed  values  (Lindgren,  1993, 
p.  376).  These  probability  calculations  were  performed  by 
a  Turbo  Pascal  vers.  7  program  (Borland  International. 
Scotts  Valley,  CA).  The  confidence  interval  for  overall  prob- 
ability of  death  at  capture  was  calculated  by  the  method  in 
Zar  ( 1996,  p.  524 ).  Ordinary  chi-square  tests  and  analysis 
of  variance  (ANOVA)  tests  followed  Zar  (1996)  and  were 
carried  out  with  the  software  Systat  vers.  9  (SPSS  Inc., 
Chicago,  IL).  In  the  statistical  tests,  type-I  error  a  was 
equal  to  0.05.  In  the  construction  of  Figure  2,  to  avoid 
overlapping  of  data  points,  the  temperatures  (but  not  the 
CPUEs)  were  jittered,  that  is,  a  small  amount  of  uniform 
random  noise  was  added  to  the  temperature  measure- 
ments (Cleveland,  1993). 


Results 

From  a  total  of  34  sets  and  33,650  hooks,  145  logger- 
head (CPUE  =  4. 31/1000  hooks)  and  20  leatherback 
(CPUE  =  0. 59/1000  hooks)  sea  turtles  were  captured. 
There  was  a  significant  difference  in  loggerhead  CPUE 
among  the  trips  (chi-square  test,  x2=137.3,  P<0.001),  but 
the  proportion  of  dead  loggerhead  sea  turtles  was  not  sig- 
nificantly different  among  the  trips  (exact  test,  P= 0.656). 
The  average  probability  of  death  at  capture  for  loggerhead 
sea  turtles  for  the  three  trips  was  0.140  (95%  confidence 
interval=  [0.086,  0.210]).  For  leatherback  sea  turtles,  the 


Table  2 

Curved  carapace  length  (CCL,  cm)  for  loggerhead  sea 

turtles,  by  trip. 

Sample 

Average 

Standard 

Trip     size 

CCL 

deviation 

Minimum 

Maximum 

1            19 

56.9 

7.3 

46.0 

70.0 

2            30 

57.2 

7.5 

46.0 

68.0 

3             5 

67.0 

5.9 

58.0 

73.0 

Total    54 

58.0 

7.7 

46.0 

73.0 

difference  in  CPUE  among  the  trips  was  significant  (chi- 
square  test,  x2=9.76,  P<0.01),  and  the  proportion  of  dead 
leatherback  sea  turtles  was  not  significantly  different 
among  the  trips  (exact  test,  P=1.000).  The  average  prob- 
ability of  death  at  capture  for  leatherback  sea  turtles  for 
the  three  trips  was  0.050  ( 95r>i  confidence  interval=  [0.001, 
0.249]). 

The  average  sea  surface  temperature  (Table  1)  was 
significantly  different  among  the  trips  (ANOVA,  ;?  =  34. 
F=55.37,  P<0.001).  The  average  temperature  on  the  first 
trip  was  significantly  lower  than  those  on  the  second  and 
third  trips,  and  the  average  temperature  on  the  second 
trip  was  significantly  higher  than  that  on  the  third  trip 
(Tukey's post  hoc  test).  For  loggerhead  sea  turtles,  CPUEs 
were  generally  higher  on  the  first  trip,  which  had  the 
lowest  average  temperature  (Fig.  2).  For  leatherback  sea 
turtles,  on  the  contrary,  the  lowest  CPUEs  were  found  on 
the  first  trip,  on  which  only  one  leatherback  sea  turtle  was 
captured  (Table  1). 

CCLs  of  captured  loggerheads  were  in  the  range  of 
46-73  cm.  Detailed  loggerhead  CCL  data  are  presented  in 
Table  2.  There  was  a  significant  difference  in  average  log- 
gerhead CCL  among  the  trips  (Table  2);  the  average  CCL 
on  the  third  trip  was  greater  than  those  on  the  first  and 
second  trips  (ANOVA,  n  =  54,  F=4.209,  P=0.020,  Tukey's 


396 


Fishery  Bulletin  102(2) 


D 

a. 
O 


25 
20 


5  - 


Caretta 


O 
O 


o 

— r- 


— l— 
15 


— r~ 
25 


Temperature  (:C) 


Figure  2 

CPUE  (number  of  captured  turtles/1000  hooks  I  by  sea  surface  temperature  (°C)  in 
each  set,  by  species.  Circles  =  first  trip,  triangles  =  second  trip,  stars  =  third  trip.  In 
each  graph,  the  dashed  vertical  line,  arbitrarily  placed  at  16.7  °C,  marks  a  separation 
between  the  temperatures  in  the  first  trip  and  those  in  the  second  and  third  trips 
(except  for  one  set  in  the  first  trip).  Note  that  the  two  graphs  have  different  vertical 
scales,  and  that,  in  the  construction  of  this  figure,  temperature  measurements  (but 
not  the  CPUEsl  have  been  jittered  (see  "Materials  and  methods"  section i. 


post  hoc  test).  Although  leatherback  sea  turtles  could  not 
be  hauled  aboard  for  measurements,  on  board  observa- 
tions and  video  recordings  indicated  that  they  were  sub- 
adult  or  adult  animals. 

Most  of  the  loggerhead  turtles  were  hooked  through 
their  mouths  or  esophagus,  but  a  small  number  were 
hooked  through  their  flippers  or  were  found  to  be  simply 
entangled  in  the  lines.  Loggerhead  sea  turtles  taken 
aboard  had  their  hooks  removed,  sometimes  in  a  care- 
less way  that  caused  severe  injury,  and  they  were  then 
returned  to  the  sea.  Leatherback  sea  turtles  were  found 
entangled  in  the  lines  or  hooked  either  through  the  flippers 
or  carapace  or  through  the  mouth.  Because  no  leatherback 
sea  turtle  was  hauled  aboard,  we  could  not  tell  if  any  were 
hooked  in  the  esophagus. 


Discussion 

Achaval  et  al.  (2000 )  reported  data  obtained  from  nine  trips 
aboard  two  different  longline  vessels  operating  within  the 
Uruguayan  EEZ  and  in  international  waters  in  the  South 
Atlantic  in  different  seasons  of  the  year,  and  employing 
different  longline  methods.  Those  authors  reported  that 
28  loggerhead  and  28  leatherback  sea  turtles  were  cap- 
tured in  86  sets  with  75.033  hooks  in  zones  I  and  II,  that 
correspond  approximately  to  the  fishing  area  covered  in 
this  study,  yielding  a  CPUE  of  0.37/1000  hooks  for  both 
loggerhead  and  leatherback  sea  turtles.  For  loggerhead 
sea  turtles,  there  was  a  significant  difference  between  our 
CPUE  (Table  1)  and  that  of  Achaval  et  al.  (chi-square  test, 
X2=226.4,  P<0.001 );  whereas  for  leatherback  sea  turtles  no 
significant  difference  was  found  i  chi-square  test,  x2=1.97, 
P=0.161). 

Although  the  variations  in  CPUE  observed  in  our  study 
could  be  explained  by  differences  in  temperatures  (Fig.  2), 
other  physical,  spatial,  or  temporal  factors  (or  a  combina- 


tion of  these  factors)  could  be  involved.  The  trips  were  car- 
ried out  at  different  times  of  the  year  (Table  1);  the  third 
trip  was  more  to  the  south  and  closer  to  the  coast,  and  the 
first  trip  had  sets  more  to  the  east  (Fig.  1). 

Our  estimates  of  sea  turtle  mortality  at  capture  may 
be  lower  than  the  actual  mortality  rates  from  longlines 
because  our  estimates  do  not  consider  postrelease  deaths 
derived  from  1)  wounds  caused  by  hooks  removed  from 
turtles  on  board,  2)  embedded  hooks  and  lines,  and  3) 
stress  caused  by  capture  itself.  Other  researchers  have 
also  recognized  that,  because  of  factors  such  as  these, 
there  is  great  uncertainty  in  the  estimates  of  mortality 
levels  for  sea  turtles  captured  in  longline  gear  (Balazs  and 
Pooley,  1994;  Eckert,  1994). 

Captured  loggerhead  sea  turtles  were  smaller  (Table  2 1 
than  loggerhead  sea  turtles  nesting  in  Brazil  (minimum 
CCL  =  83.0  cm,  average  CCL=  103.0  cm,  nesting  season 
1982-83  through  nesting  season  1999-2000;  Projeto 
TAMAR3)  and  in  several  places  in  the  North  Atlantic  and 
the  Caribbean  (minimum  CCL=75.4  cm,  average  CCL  in 
the  range  of  94.0-105.1  cm;  Dodd,  1988).  However,  logger- 
head sea  turtles  nesting  in  Cape  Verde,  in  the  northeast- 
ern Atlantic,  are  smaller  than  those  nesting  in  those  other 
places:  minimum  CCL  =  68.0  cm,  average  CCL=82.9  cm. 
data  from  1998  (Cejudo  et  al.,  20001.  There  is  an  overlap 
between  the  observed  CCL  range  and  that  of  adult  Cape 
Verde  loggerhead  sea  turtles  (seven  loggerhead  turtles 
out  of  54  observed,  or  13.0'7f ,  had  a  CCL  equal  or  greater 
than  the  minimum  Cape  Verde  CCL),  but  the  average 
CCL  of  the  captured  loggerhead  sea  turtles  (Table  2  i  was 
well  below  that  of  loggerhead  sea  turtles  nesting  in  Cape 
Verde.  We  estimate  that  the  captured  loggerhead  sea 
turtles  were  generally  juveniles,  although  a  small  number 
of  them  could  have  been  adult  turtles.  However,  size  is 


3  Projeto  TAMAR.     2000.     Unpubl.  data 

Salvador,  BA  40210-970,  Brazil. 


Caixa  Postal  2219. 


NOTE     Kotas  et  al.:  Incidental  capture  of  Caretta  caretta  and  Dermochelys  conacea  by  the  pelagic  longline  fishery 


397 


not  a  reliable  indicator  of  maturity  or  breeding  status  for 
sea  turtles  (Miller,  1997). 

Along  the  southern  coast  of  Brazil  (between  latitudes 
23°S  and  33°S),  loggerhead  sea  turtles  stranded  or  in- 
cidentally captured  in  fishing  gear  with  CCLs  as  small 
as  32.5  cm  have  been  observed  (Projeto  TAMAR4),  but 
usually  loggerhead  sea  turtles  found  in  that  region  have 
CCLs  greater  than  50  cm,  most  commonly  in  the  range  of 
60-90  cm  (Pinedo  et  al.,  1998;  Bugoni  et  al.,  2001;  Projeto 
TAMAR4 ).  Loggerhead  sea  turtles  have  also  been  found  in 
Uruguay  and  Argentina  (Frazier,  1984;  Fallabrino  et  al., 
2000).  Their  CCLs  in  those  countries  have  been  reported 
to  be  in  the  approximate  range  of  50-115  cm  (Frazier, 
1984).  The  loggerhead  sea  turtles  reported  here  have 
an  average  CCL  smaller  than  that  usually  observed  for 
loggerhead  sea  turtles  stranded  or  captured  in  southern 
Brazil,  Uruguay,  and  Argentina,  although  most  of  the 
turtles  (45  out  of  54,  or  83%)  had  CCLs  equal  to  or  greater 
than  50  cm,  that  is,  they  were  within  the  size  range  for 
that  region. 

Cumulative  evidence  obtained  from  genetic  and  size- 
distribution  data  around  oceanic  basins,  as  well  as  tag 
returns,  shows  that  the  ontogenetic  development  of  log- 
gerhead sea  turtles  involves  a  pelagic  juvenile  stage  (Carr, 
1987;  Musick  and  Limpus,  1997;  Bolten  et  al.,  1998).  Trans- 
oceanic developmental  migrations  establishing  a  link  be- 
tween juveniles  in  feeding  grounds  and  hatchlings  from 
nesting  beaches  on  opposite  sides  of  the  ocean  basin  have 
been  demonstrated  through  genetic  analysis  for  the  North 
Atlantic  and  North  Pacific  (Bowenet  al.,  1995;  Bolten  et  al., 
1998).  It  has  been  suggested  that  a  similar  pattern  may  be 
expected  for  the  South  Atlantic  (Bolten  et  al.,  1998),  where 
loggerhead  sea  turtles  nest  in  Brazil  and  possibly  in  Africa 
(Marcovaldi  and  Laurent,  1996;  Fretey,  2001).  The  inciden- 
tal captures  reported  in  our  study,  indicating  the  use  of  the 
pelagic  environment  by  juvenile  loggerhead  sea  turtles  in 
the  South  Atlantic,  support  the  hypothesis  of  transoceanic 
developmental  migrations  for  those  turtles  in  that  ocean. 
Future  genetic  analysis  of  turtles  incidentally  captured  in 
the  South  Atlantic  would  help  to  clarify  their  natal  origin. 

For  leatherback  sea  turtles,  there  are  important  nesting 
grounds  in  the  Atlantic,  mainly  in  French  Guiana  and  Su- 
riname  in  South  America,  and  Gabon  and  Congo  in  Africa 
(Spotila  et  al.,  1996;  Fretey,  2001).  Leatherback  sea  turtles 
are  known  to  travel  long  distances  from  their  nesting 
beaches  into  pelagic  waters  (Goff  et  al.,  1994;  Morreale  et 
al.,  1996;  Eckert  and  Sarti,  1997;  Eckert,  1998).  Satellite 
telemetry  data  indicate  that  leatherback  sea  turtles  nest- 
ing in  eastern  South  Africa  can  enter  the  South  Atlantic 
(Hughes  et  al.,  1998;  Hughes5).  In  the  southwestern  Atlan- 
tic, leatherback  sea  turtles  have  been  observed  or  captured 
in  Brazil,  Uruguay,  and  Argentina  (Frazier,  1984;  Pinedo 
et  al.,  1998;  Achaval  et  al.,  2000;  Fallabrino  et  al.,  2000; 
Bugoni  etal.,  2001). 


Some  measure  of  the  significance  of  the  three  trips  re- 
ported in  the  present  study  in  terms  of  the  potential  for 
turtle  capture  and  mortality  in  the  South  Atlantic  longline 
fishery  can  be  obtained  by  looking  at  information  concern- 
ing the  total  fishing  effort  in  the  study  area.  In  1999,  the 
Brazilian  longline  fleet  consisted  of  70  longliners  (42  Bra- 
zilian and  28  leased  foreign  vessels);  among  them,  33  ves- 
sels were  operating  out  of  ports  in  southern  Brazil,  in  the 
states  of  Sao  Paulo,  Santa  Catarina,  and  Rio  Grande  do 
Sul.  In  that  year,  the  total  number  of  hooks  of  that  long- 
line  fleet  (both  Brazilian  and  leased  vessels)  amounted  to 
13,598,260  hooks  (ICCAT6).  However,  the  southwestern 
Atlantic  is  fished  not  only  by  Brazil-based  longliners,  but 
also  by  longliners  from  Uruguay,  Chile,  Japan,  Taiwan,  and 
Spain  (Folsom,  1997;  Weidner  and  Arocha,  1999;  Weidner 
et  al.,  1999).  According  to  ICCAT's  (International  Commis- 
sion for  the  Conservation  of  Atlantic  Tunas)  CATDIS  data 
set  (ICCAT)7  longliners  operating  during  1995-97  in  the 
area  delineated  by  the  present  study  (latitudes  25°S  and 
35°S  and  longitude  35°W,  or  eight  ICCAT  5x5°  statistical 
blocks.  Fig.  1)  had  an  average  annual  catch  of  tunas  and 
swordfishes  of  6885  metric  tons  (t)  (the  total  hold  capacity 
of  the  vessels  on  the  three  trips  reported  in  this  study  was 
130  t).  However,  due  to  unreported  landings  by  vessels 
flying  flags  of  convenience  (FAO,  2001;  FAO8)  and  other 
sources,  the  estimate  obtained  from  ICCAT  data  should  be 
considered  a  minimum  estimate  of  the  total  annual  tuna 
and  swordfish  catch  ( ICCAT9 ).  Furthermore,  because  North 
Atlantic  stocks  of  swordfishes  and  some  species  of  tuna  are 
considered  overfished  (NMFS10),  quota  or  closure  regula- 
tions (or  both)  in  the  North  Atlantic  may  be  driving  longline 
fleets  to  the  South  Atlantic,  increasing  the  risk  of  incidental 
capture  of  sea  turtles  there. 

In  Brazil,  sea  turtle  capture  is  prohibited  by  federal 
legislation  (Marcovaldi  and  Marcovaldi,  1999),  and  mea- 
sures have  been  taken  to  address  the  problem  of  inci- 
dental capture  by  longlines  and  other  kinds  of  fishing 


4  Projeto  TAMAR.     2000.     Unpubl.   data.     Rua  Antonio 
Athanasio  273,  Ubatuba,  SP  11680-000,  Brazil. 

5  Hughes,  G.  R.     2002.     Personal  commun.     Ezemvelo  KZN 
Wildlife,  P  O  Box  13053,  Cascades  3202,  South  Africa. 


6  ICCAT  (International  Commission  for  the  Conservation  of 
Atlantic  Tunas).  2001.  National  report  of  Brazil.  Report 
for  biennial  period,  2000-2001,  part  I  (2000),  vol.  1,  English 
version,  p.  312-315.  Calle  Corazon  de  Maria,  8,  28002 
Madrid,  Spain. 

7  ICCAT  (International  Commission  for  the  Conservation  of 
Atlantic  Tunas).  2002.  CATDIS  dataset.  Calle  Corazon  de 
Maria,  8,  28002  Madrid,  Spain.  (Available  from  http://www. 
iccat.org.] 

s  FAO  (Food  and  Agriculture  Organization  of  the  United 
Nations).  2001.  International  plan  of  action  to  prevent, 
deter  and  eliminate  illegal,  unreported  and  unregulated  fish- 
ing, 24  p.  FAO,  Rome.  (Available  from  http://www.fao.org/ 
docrep/003/yl224e/yl224e00.htm.] 

9  ICCAT  (International  Commission  for  the  Conservation  of 
Atlantic  Tunas).  1999.  Detailed  report  for  swordfish, 
ICCAT  SCRS  swordfish  stock  assessment  session  (Madrid, 
Spain,  September  27  to  October  4,  1999),  176  p.  Calle 
Corazon  de  Maria,  8,  28002  Madrid.  Spain. 

10  NMFS  (National  Marine  Fisheries  Service).  2000.  2000 
stock  assessment  and  fishery  evaluation  for  Atlantic  highly 
migratory  species,  150  p.  U.S.  Dep.  Commer.,  NOAA,  NMFS. 
Highly  Migratory  Species  Management  Division,  1315  East- 
West  Highway,  Silver  Spring,  MD  20910. 


398 


Fishery  Bulletin  102(2) 


gear.  Since  2001,  Projeto  TAMAR  has  been  developing  and 
implementing  (through  partnerships  with  other  institu- 
tions) an  action  plan  whose  main  objective  is  to  reduce 
incidental  sea  turtle  capture,  including  captures  occurring 
in  the  open  sea  (Marcovaldi  et  al.,  2002).  The  action  plan 
includes,  among  other  things,  an  assessment  of  fishery- 
related  sea  turtle  mortality,  the  development  of  mitigation 
methods,  and  a  proposal  of  adequate  conservation  and 
enforcement  policies  (Marcovaldi  et  al.,  2002).  However, 
because  the  longline  fleet  is  composed  of  vessels  from  many 
nations,  the  reduction  of  incidental  capture  in  the  open  sea 
calls  for  international  cooperation  ( Eckert  and  Sarti,  1997; 
Trono  and  Salm,  1999;  Crowder,  2000). 

The  observations  reported  in  this  study  and  the  pres- 
ence of  a  sizable  longline  fleet  operating  in  the  southwest- 
ern Atlantic  indicate  1)  the  need  for  research  to  clarify 
habitat  use  by  sea  turtles  in  that  part  of  the  ocean  (Eckert 
and  Sarti,  1997;  Bolten  et  al.,  1998),  2)  the  need  for  contin- 
ued research  to  quantify  the  impact  of  longline  fishing  on 
sea  turtles  in  the  pelagic  realm  of  that  ocean  (Balazs  and 
Pooley,  1994;  Eckert,  1994),  and  3)  the  implementation  of 
conservation  measures  for  sea  turtles  in  that  environment. 
We  suggest  the  implementation  of  an  International  Ob- 
servers Program  on  board  longliners  operating  throughout 
the  South  Atlantic  ocean. 


Acknowledgments 

This  note  is  the  result  of  observations  made  possible 
through  an  agreement  between  the  REVIZEE  Program 
(National  Program  for  the  Assessment  of  the  Sustain- 
able Fishing  Potential  of  the  Exclusive  Economic  Zone 
Live  Resources,  a  Brazilian  Government  program)  and 
Projeto  TAMAR's  station  at  Ubatuba,  State  of  Sao  Paulo. 
We  would  like  to  thank  Jose  Kowalsky  of  the  Kowalsky 
fishing  company  and  Marcelino  Talavera  (Itajai,  State  of 
Santa  Catarina),  owners  of  the  vessels  Yamaya  III  and 
Basco,  respectively,  for  kindly  allowing  access  to  the  fish- 
ing vessels,  and  the  crew  of  the  two  longliners,  and  also 
the  fishing  research  center  Centro  de  Pesquisa  e  Extensao 
Pesqueira  do  Sudeste-Sul-CEPSUL/IBAMA  ( Itajai,  State 
of  Santa  Catarina),  and  particularly  Jorge  Almeida  de 
Albuquerque,  for  making  this  research  possible.  We  also 
thank  Larisa  Avens  and  Matthew  Godfrey  for  their  gener- 
ous reviews  of  the  paper,  and  the  two  anonymous  referees, 
whose  suggestions  helped  to  improve  our  work.  Projeto 
TAMAR  is  affiliated  with  IBAMA  (the  Brazilian  Institute 
for  the  Environment  and  Renewable  Natural  Resources), 
is  co-managed  by  Fundacao  Pro-TAMAR,  and  officially 
sponsored  by  Petrobras.  In  Ubatuba.  TAMAR  is  supported 
by  Ubatuba's  municipal  government  ( Prefeitura  Municipal 
de  Ubatuba).  S.S.  and  V.G.A.  were  supported  by  CNPq 
(Brazilian  National  Research  Council). 


Literature  cited 

Achaval,  F.,  Y.  H.  Marin,  and  L.  C.  Barea. 

2000.     Captura  incidental  de  tortugas  con  palangre  pela- 


gico  oceanico  en  el  Atlantico  sudoccidental.  In  Captura  de 
grandes  peces  pelagicos  (pez  espada  y  atunesi  en  el  Atlan- 
tico Sudoccidental.  y  su  interaccion  con  otras  poblaciones 
(G.  Arena  and  M  Rey,  eds.),  p.  83-88.  Instituto  Nacional 
de  Pesca,  Programa  de  las  Naciones  Unidas  para  el  Desar- 
rollo,  Montevideo.  Uruguay. 
Balazs,  G.  H.,  and  S.  G  Pooley  (comps.). 

1994.  Research  plan  to  assess  marine  turtle  hooking  mor- 
tality: results  of  an  expert  workshop  held  in  Honolulu. 
Hawaii,  November  16-18,  1993.  NOAA  Tech.  Memo. 
NOAA-TM-NMFS-SWFSC-201,  166  p. 

Bolten,  A.  B.,  K.  A.  Bjorndal.  H.  R.  Martins,  T.  Dellinger, 
M.  J.  Biscoito,  S.  E.  Encalada,  and  B.  W.  Bowen. 

1998.     Transatlantic  developmental  migrations  of  log- 
gerhead sea  turtles  demonstrated  by  mtDNA  sequence 
analysis.     Ecol.  Appl.  8:1-7. 
Bolten,  A.  B.,  J.  A.  Wetherall,  G.  H.  Balazs.  and 
S.  G.  Pooley  (comps.). 

1996.  Status  of  marine  turtles  in  the  Pacific  Ocean  relevant 
to  incidental  take  in  the  Hawaii-based  pelagic  longline 
fishery.  NOAA  Tech.  Memo.  NOAA-TM-NMFS-SWFSC- 
230,  167  p. 

Bowen,  B.  W..  F.  A.  Abreu-Grobois,  G.  H.  Balazs,  N.  Kamezaki. 
C.  J.  Limpus,  and  R.  J.  Ferl. 

1995.  Trans-Pacific  migrations  of  the  loggerhead  turtle 
[Caretta  caretta)  demonstrated  with  mitochondrial  DNA 
markers.     Proc.  Natl.  Acad.  Sci.  USA  92:  3731-3734. 

Bugoni,  L.,  L.  Krause,  and  M.  V.  Petrv. 

2001.     Marine  debris  and  human  impacts  on  sea  turtles  in 
southern  Brazil.     Mar.  Pollut.  Bull.  42:1330-1334. 
Carr,  A. 

1987.  New  perspectives  on  the  pelagic  stage  of  sea  turtle 
development.     Conserv.  Biol.  1:  103-121. 

Cejudo,  D.,  I.  Cabrera,  L.  F  Lopez-Jurado.  C.  Evora, 
and  P.  Alfama. 

2000.     The  reproductive  biology  of  Caretta  caretta  on  the 
Island  of  Boavista  ( Republic  of  Cabo  Verde.  Western  Afri- 
ca).    NOAA  Tech.  Memo.  NMFS-SEFSC-443:244-245. 
Cleveland,  W.  S. 

1993.  Visualizing  data,  360  p.  Hobart  Press,  Summit, 
NJ. 

Crowder.  L. 

2000.     Leatherback's  survival  will  depend  on  an  interna- 
tional effort.     Nature  405:881. 
Dodd,  C.  K.,  Jr. 

1988.  Synopsis  of  the  biological  data  on  the  loggerhead  sea 
turtle  Caretta  caretta  (Linnaeus  1758).  U.S.  Fish  Wildl. 
Serv.  Biol.  Rep.  88.  110  p. 

Eckert,  S.  A. 

1994.  Evaluating  the  post-release  mortality  of  sea  turtles 
incidentally  caught  in  pelagic  longline  fisheries.  NOAA 
Tech.  Memo.  NMFS-SWFSC-201:106-110. 

1998.  Perspectives  on  the  use  of  satellite  telemetry  and 
other  electronic  technologies  for  the  study  of  marine 
turtles,  with  reference  to  the  first  year  long  tracking  of 
leatherback  sea  turtles.  NOAA  Tech.  Memo.  NMFS- 
SEFSC-415:44-46. 
Eckert,  S.  A.,  and  M.  L.  Sarti. 

1997.  Distant  fisheries  implicated  in  the  loss  of  the  world's 
largest  leatherback  nesting  population.  Mar.  Turtle 
Newsl.  78:2-7. 

Fallabrino.  A.,  A.  Bager.  A.  Estrades,  and  F  Achaval. 

2000.  Current  status  of  marine  turtles  in  Uruguay.  Mar. 
Turtle  Newsl.  87:4-5. 

FAO  (Food  and  Agriculture  Organization  of  the  United  Nations). 

2001.  Report  of  the  second  technical  consultation  on  illegal, 


NOTE     Kotas  et  al.:  Incidental  capture  of  Caretta  caretta  and  Dermochelys  conacea  by  the  pelagic  longlme  fishery 


399 


unreported  and  unregulated  fishing.  Rome,  22-23  Febru- 
ary 2001.     FAO  Fisheries  Rep.  646,  38  p.     FAO,  Rome. 
Folsom,  W.  B. 

1997.  World  swordfish  fisheries:  an  analysis  of  swordfish 
fisheries,  market  trends  and  trade  patterns,  vol.  VI: 
Western  Europe.  NOAA  Tech.  Memo.  NMFS-F/SPO-29, 
324  p. 

Frazier,  J. 

1984.     Las  tortugas  marinas  en  el  Atlantico  sur  occidental. 

In  Serie  divulgacidn  no.  2,  p.  2-21.     Associacion  Herpeto- 

logica  Argentina,  La  Plata,  Argentina. 
Fretey,  J. 

2001.  Biogeography  and  conservation  of  marine  turtles  of 
the  Atlantic  coast  of  Africa.  CMS  Technical  Series  Publ. 
no.  6,  429  p.     UNEP/CMS  Secretariat.  Bonn.  Germany. 

Goff,  G.  P.,  J.  Lien,  G.  B.  Stenson.  and  J.  Fretey. 

1994.  The  migration  of  a  tagged  leatherback  turtle,  Der- 
mochelys coriacea,  from  French  Guiana,  South  America, 
to  Newfoundland,  Canada,  in  128  days.  Can.  Field  Nat. 
108:72-73. 

Hughes,  G.  R.,  P.  Luschi,  R.  Mencacci,  and  F.  Papi. 

1998.  The  7000  km  oceanic  journey  of  a  leatherback 
turtle  tracked  by  satellite.  J.  Exp.  Mar.  Biol.  Ecol.  229: 
209-217. 

IUCN  (International  Union  for  Conservation  of  Nature  and 
Natural  Resources!  Species  Survival  Commission. 

1995.  A  global  strategy  for  the  conservation  of  marine 
turtles,  25  p.     IUCN,  Gland,  Switzerland. 

Lindgren,  B.  W. 

1993.     Statistical  theory,  4th  ed.,  633  p.     Chapman  &  Hall, 
New  York,  NY. 
Lutcavage,  M.  E.,  P.  Plotkin,  B.  Witherington,  and  P.  L.  Lutz. 
1997.     Human  impacts  on  sea  turtle  survival.     In  The  biol- 
ogy of  sea  turtles  (P.  L.  Lutz  and  J.  A.  Musick,  eds.),  p. 
387-409.     CRC  Press,  Boca  Raton,  FL. 
Marcovaldi,  M.  A.,  and  A.  Laurent. 

1996.  A  six  season  study  of  marine  turtle  nesting  at  Praia 
do  Forte,  Bahia,  Brazil,  with  implications  for  conservation 
and  management.     Chelonian  Conserv.  Biol.  2:55-59. 

Marcovaldi,  M.  A.,  and  G.  G.  dei  Marcovaldi. 

1999.  Marine  turtles  of  Brazil:  the  history  and  structure  of 
Projeto  TAMAR-IBAMA.     Biol.  Conserv.  91:35-41. 

Marcovaldi,  M.  A.,  J.  C.  Thome,  G.  Sales,  A.  C.  Coelho.  B.  Gallo. 
and  C.  Bellini. 

2002.  Brazilian  plan  for  reduction  of  incidental  sea  turtle 
capture  in  fisheries.     Mar.  Turtle  Newsl.  96:24-25. 

Miller.  J.  D. 

1997.  Reproduction  in  sea  turtles.  In  The  biology  of  sea 
turtles  (P.  L.  Lutz  and  J.  A.  Musick,  eds.),  p.  51-81.  CRC 
Press,  Boca  Raton,  FL. 

Morreale,  S.  J.,  E.  A.  Standora,  J.  R.  Spotila,  and  F.  V.  Paladino. 

1996.  Migration  corridor  for  sea  turtles.  Nature  384: 
319-320. 

Musick,  J.  A.,  and  C.  J.  Limpus. 

1997.  Habitat  utilization  and  migration  in  juvenile  sea 
turtles.  In  The  biology  of  sea  turtles  (P.  L.  Lutz  and  J.  A. 
Musick,  eds.),  p.  137-163.     CRC  Press,  Boca  Raton,  FL. 


NRC  (National  Research  Council  I. 

1990.     Decline  of  the  sea  turtles:  causes  and  prevention, 
259  p.     National  Academy  Press,  Washington,  D.C. 
Nishemura,  W„  and  S.  Nakahigashi. 

1990.  Incidental  capture  of  sea  turtles  by  Japanese  research 
and  training  vessels:  results  of  a  questionaire.  Mar. 
Turtle  Newsl.  51:1-4. 

Oravetz.  C.  A. 

1999.     Reducing  incidental  catch  in  fisheries.     In  Research 

and  management  techniques  for  the  conservation  of  sea 

turtles  (K.  L.  Eckert,  K.  A.  Bjorndal,  F.  A.  Abreu-Grobois. 

and  M.  Donnelly,  eds.).  p.  189-193.     IUCN  SSC  Marine 

Turtle  Specialist  Group  publication  no.  4. 
Pinedo,  M.  C,  R.  Capitoli,  A.  S.  Barreto,  and  A.  L.  V.  Andrade. 

1998.  Occurrence  and  feeding  of  sea  turtles  in  southern 
Brazil.  NOAA  Tech.  Memo.  NMFS-SEFSC-412:117- 
118. 

Spotila,  J.  R..  A.  E.  Dunham,  A.  J.  Leslie.  A.  C.  Steyermark, 
P.  T.  Plotkin,  and  F.V.  Paladino. 

1996.     Worldwide  decline  of  Dermochelys  coriacea:  are 
leatherback  turtles  going  extinct?     Chelonian  Conserv. 
Biol.  2:209-222. 
Teas,  W.  G. 

1993.     Species  composition  and  size  class  distribution 
of  marine  turtle  strandings  on  the  Gulf  of  Mexico  and 
southeast  United  States  coasts,  1985-1991.    NOAA  Tech. 
Memo.  NMFS-SEFSC-315,  43  p. 
Tobias,  W. 

1991.  Turtles  caught  in  Caribbean  swordfish  net  fishery. 
Mar.  Turtle  Newsl.  53:10-12. 

Trono,  R.  B.,  and  R.  V.  Salm. 

1999.  Regional  collaboration.  In  Research  and  manage- 
ment techniques  for  the  conservation  of  sea  turtles  (K. 
L.  Eckert,  K.  A.  Bjorndal,  F.  A.  Abreu-Grobois,  and  M. 
Donnelly,  eds.),  p.  224-227.  IUCN  SSC  Marine  Turtle 
Specialist  Group  publication  no.  4. 

Weidner,  D.  M„  and  F.  Arocha. 

1999.     World  swordfish  fisheries:  an  analysis  of  swordfish 
fisheries,  market  trends  and  trade  patterns.     Vol.  IV, 
Part  A2b:  Brazil.     NOAA  Tech.  Memo.  NMFS-F/SPO- 
35:237-628. 
Weidner,  D.  M..  F.  J.  Fontes,  and  J.  Serrano. 

1999.     World  swordfish  fisheries:  an  analysis  of  swordfish 
fisheries,  market  trends  and  trade  patterns.  Vol.  IV  part 
A2c:  Uruguay,  Paraguay  and  Argentina.     NOAA  Tech. 
Memo.  NMFS-F/SPO-36:631-916. 
Williams,  P.,  P.  J.  Anninos,  P.  T.  Plotkin,  and  K.  L.  Salvini 
(comps.l. 

1996.  Pelagic  longline  fishery — sea  turtle  interactions. 
Proceedings  of  an  industry,  academic  and  government 
experts,  and  stakeholders  workshop  held  in  Silver  Spring, 
Maryland.  24-25  May  1994.  NOAA  Tech.  Memo.  NMFS- 
OPR-7,  77  p. 
Zar,  J.  H. 

1996.  Biostatistical  analysis,  3rd  ed.,  662  p.  Prentice  Hall, 
LTpper  Saddle  River,  NJ. 


400 


Diet  changes  of  Pacific  cod 

(Gadus  macrocephalus)  in  Pavlof  Bay 

associated  with  climate  changes 

in  the  Gulf  of  Alaska  between  1980  and  1995 


Mei-Sun  Yang 

Alaska  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

7600  Sand  Point  Way  NE 

Seattle.  Washington  98115 

E  mail  address  mei-sunyangffinoaa  gov 


The  diet  of  Pacific  cod  (Gadus  mac- 
rocephalus) in  the  area  of  Pavlof  Bay, 
Alaska,  was  studied  in  the  early  1980s 
by  Albers  and  Anderson  (1985).  They 
found  that  the  dominant  prey  spe- 
cies were  forage  species  like  pandalid 
shrimp,  capelin  iMallotus  villosus), 
and  walleye  pollock  [Theragra  chal- 
cogramma).  The  shrimp  fishery  in 
Pavlof  Bay  began  in  1968  and  closed 
in  1980  because  of  low  shrimp  abun- 
dance (Ruccio  and  Worton1).  Survey 
data  indicate  that,  during  the  period 
between  1972  and  1997,  the  abun- 
dance of  forage  species  such  as  pan- 
dalid shrimp  and  capelin  declined 
and  higher  trophic-level  groundfish 
such  as  Pacific  cod  increased.  There 
is  a  general  recognition  that  a  long- 
term  ocean  climate  shift  in  the  Gulf  of 
Alaska  has  been  partially  responsible 
for  the  observed  reorganization  of  the 
community  structure  (Anderson  and 
Piatt,  1999). 

Because  there  has  been  an  appar- 
ent shift  in  the  abundance  of  both 
predators  and  prey  in  Pavlof  Bay,  it  is 
important  to  understand  how  trophic 
relationships  may  also  have  changed. 


1  Ruccio,  M.  P..  and  C.  L.  Worton.  1999. 
Annual  management  report  for  the  shell- 
fish fisheries  of  the  Alaska  peninsula 
area,  1998.  In  Annual  management 
report  for  the  shellfish  fisheries  of  the 
westward  region.  1998.  Regional  Infor- 
mation Report  4K99-49,  312  p.  Alaska 
Department  of  Fish  and  Game,  Division  of 
Commercial  Fisheries,  211  Mission  Road, 
Kodiak,  Alaska  99615. 


In  order  to  partially  address  this  ques- 
tion, stomach  samples  of  Pacific  cod 
and  other  groundfishes  were  taken  in 
1995.  By  performing  a  comparison  of 
the  diet  of  Pacific  cod  right  after  the 
climate  shift  with  Pacific  cod  and  other 
groundfishes  well  after  the  shift,  this 
analysis  may  demonstrate  how  the 
relative  abundance  of  prey  in  the  Gulf 
of  Alaska  may  have  changed. 


Methods 

Stomachs  of  Pacific  cod,  walleye  pol- 
lock, and  arrowtooth  flounder  (Atheres- 
thes  stomias)  were  collected  by  National 
Marine  Fisheries  Service  (NMFS)  sci- 
entists on  board  the  chartered  vessel 
FV  Arcturus  conducting  a  trawl  survey 
in  Pavlof  Bay,  Alaska.  (Fig.  1)  from  5 
August  to  7  August  1995.  The  survey 
targeted  shrimp  and  used  a  high- 
opening  net  with  small  mesh  (32-mm 
stretched  mesh).  Each  tow  was  about 
1.2  km  in  length.  The  average  depth  of 
the  13  hauls  where  stomachs  were  col- 
lected was  108.9  (±9.5)  m  with  a  range 
from  90  to  123  m.  When  a  sampled 
stomach  was  retained,  it  was  put  in  a 
cloth  stomach  bag.  A  field  tag  with  the 
species  name,  fork  length  (FL  in  cm)  of 
the  fish,  and  haul  data  (vessel,  cruise, 
haul  number,  specimen  number)  was 
also  put  in  the  bag.  All  the  samples  col- 
lected were  then  preserved  in  buckets 
containing  a  10%  formalin  solution. 
When  the  samples  arrived  at  the  labo- 
ratory, they  were  transferred  into  70% 
ethanol  before  the  stomach  contents 


were  analyzed.  In  the  laboratory,  the 
stomach  was  cut  open,  the  contents 
were  removed  and  blotted  with  a  paper 
towel.  Wet  weight  was  then  recorded  to 
the  nearest  0.1  g.  After  obtaining  the 
total  weight  for  a  stomach's  contents, 
the  contents  were  placed  in  a  Petri  dish 
and  examined  under  a  microscope. 
Each  prey  item  was  classified  to  the 
lowest  practical  taxonomic  level.  Prey 
weights  and  numbers  of  commercially 
important  fish  were  recorded.  Stan- 
dard lengths  of  prey  fish  and  carapace 
width  of  crabs  were  also  recorded.  The 
diet  of  Pacific  cod  was  summarized  to 
show  the  percent  frequency  of  occur- 
rence, the  percentage  by  number,  and 
the  percentage  of  the  total  weight  of 
each  prey  item  found  in  the  stomachs. 
Stomach  contents  of  walleye  pollock 
and  arrowtooth  flounder  were  ana- 
lyzed for  comparisons. 


Results 

Of  130  Pacific  cod  stomachs  analyzed. 
129  contained  food.  Pacific  cod  sizes 
ranged  from  40  to  80  cm  FL  (fork 
length);  a  mean  size  was  55.4  (SD 
±7.2)  cm. 

Polychaetes,  crangonid  shrimp,  pea 
crab,  and  clams  were  the  most  fre- 
quently found  prey  items  in  Pacific  cod 
stomachs  (Table  1).  However,  in  terms 
of  weight,  eelpouts  (zoarcids).  Tanner 
crab  (Chionoecetes  bairdi),  crangonid 
shrimp,  hermit  crab,  and  polychaetes 
were  the  most  important  prey  of  Pa- 
cific cod.  Pandalid  shrimp,  spinyhead 
sculpin  (Dasycottus  setiger),  prickle- 
backs  (stichaeid).  Pacific  sandlance 
(Ammodytes  hexapterus),  arrowtooth 
flounder  lAtheresthes  stomias),  and 
flathead  sole  (Hippoglossoides  elasso- 
don )  were  minor  prey. 

Invertebrates  (mainly  crangonid 
shrimp,  polychaetes,  and  crabs)  were 
the  principal  prey  of  Pacific  cod  smaller 
than  60  cm  (Fig.  2).  There  were  nine 
prey  categories  as  shown  in  Figure  2. 
The  miscellaneous  prey  included  Si- 
puncula,  Echiura.  fish  offal  (processed 


Manuscript  approved  for  publication 
2  I  I  Vcember  2003  by  Scientific  Editor. 

Manuscript  received  20  January  2004 
at  NMFS  Scientific  Publications  Office. 

Fish.  Bull.  102:400-405(2004). 


NOTE     Yang:  Diet  changes  of  Gadus  macrocephalus  associated  with  climate  changes  in  Pavlof  Bay 


401 


Figure  1 

Location  of  study  area  in  1980,  1981,  and  1995. 


100 


fish  parts  like  head,  tail,  pyloric  caeca, 
etc.),  and  all  other  prey  organisms  not 
included  in  the  other  eight  prey  catego- 
ries. The  importance  of  fish  in  the  diet 
of  Pacific  cod  increased  after  60  cm  FL. 
Walleye  pollock  were  consumed  only  by 
Pacific  cod  >60  cm  FL. 

In  general,  Pacific  cod  ate  prey  of  small 
individual  size  (Table  2).  Tanner  crabs 
iChionoecetes  bairdi)  ranged  from  4.5 
to  42.3  mm  carapace  width.  Eelpouts 
ranged  in  length  from  36.2  to  256.6  mm 
standard  length.  Other  fish  prey  ranged 
in  length  from  32.7  to  81.5  mm.  Walleye 
pollock  were  consumed  by  Pacific  cod  but 
were  not  measurable. 

In  1995,  when  Pacific  cod  stomachs 
were  collected  in  Pavlof  Bay,  218  wall- 
eye pollock  and  80  arrowtooth  flounder 
stomachs  were  also  collected.  Similar 
to  the  results  for  Pacific  cod,  pandalid 
shrimp  and  capelin  were  not  important 
food  of  walleye  pollock  and  arrowtooth 
flounder  either  (Fig.  3).  These  prey  each 
comprised  less  than  39c  of  the  total 
stomach  content  weight  of  walleye  pol- 
lock and  arrowtooth  flounder.  Instead, 
eelpouts,  pricklebacks,  euphausiids,  and  walleye  pollock 
were  important  food  of  arrowtooth  flounder,  and  euph- 
ausiids (83%  by  weight)  were  the  main  food  of  walleye 
pollock. 


N=26 


40-49 


□  Polychaete 

□  Pollock 

□  Misc.  fish 

□  Tanner  crab 
S  Pagurid 

□  Other  crab 
El  Pandalid 
O  Crangonid 

□  Misc.  prey 


50-59 
Predator  fork  length  (cm) 


60-80 


Figure  2 

Variations  in  the  main  food  items  of  Pacific  cod,  by  predator  size,  in  Pavlof 
Bay  in  1995.  /i=sample  size. 


Discussion 

This  study  shows  that  eelpouts 
hermit  crabs,  polychaetes,  and 


,  Tanner  crabs,  crangonids, 
echiuroids  were  the  princi- 


402 


Fishery  Bulletin  102(2) 


Table  1 

Percent  frequency  of  occurrence  (%F),  percentage  by 

number  (7cN),  and  percentage  by 

weight  C7rW)  of  prey 

items  of  Pacific  cod 

collected  in  Pavlof  Bay,  Alaska,  1995. 

Prey  name 

9c  F 

%N 

%W 

Polychaeta  (worm) 

79.8 

11.4 

9.2 

Gastropoda  (snail) 

14.0 

0.8 

0.4 

Bivalvia  (clam) 

55.0 

6.1 

2.1 

Cephalopoda  (squid  and  octopus) 

10.1 

0.5 

2.1 

Copepoda 

0.8 

0.0 

0.0 

Peracarida  Mysidacea  (mysid) 

31.8 

11.5 

0.2 

Cumacea  (cumaceanl 

13.2 

0.9 

0.0 

Amphipoda  (amphipod) 

17.1 

1.2 

0.0 

Euphausiacea  leuphausiid) 

15.5 

10.0 

0.7 

Natantia  (unidentified  shrimp) 

12.4 

0.7 

0.1 

Caridea  (shrimp) 

12.4 

1.2 

1.2 

Hippolytidae  (shrimp) 

17.8 

1.2 

0.2 

Pandalidae  (shrimp) 

41.1 

5.7 

2.3 

Crangonidae  (shrimp) 

76.0 

18.9 

13.3 

Reptantia  (unidentified  crab) 

11.6 

0.5 

1.9 

Paguridae  (hermit  crab) 

22.5 

1.3 

9.5 

Decapoda  Brachyura  (crab) 

0.8 

0.0 

0.1 

Hyas  sp.  (lyre  crab) 

0.8 

0.0 

0.9 

Hyas  lyratus  (lyre  crab) 

1.6 

0.1 

0.6 

Chionoecetes  sp.  (snow  and  Tanner  crab) 

40.3 

3.5 

13.9 

Pinnotheridae  (pea  crab) 

1.6 

0.1 

0.1 

Pinnixa  sp.  (pea  crab) 

68.2 

8.4 

3.2 

Sipuncula  I  marine  worm) 

0.8 

0.0 

0.6 

Echiura  (marine  worm) 

24.0 

1.4 

6.6 

Ophiuroidea  (basket  and  brittle  star) 

9.3 

0.7 

0.1 

Chaetognatha  (arrow  worm) 

1.6 

0.2 

0.0 

Rajidae  (skate) 

2.3 

0.1 

0.4 

Osteichthyes  Teleostei  (fish) 

12.4 

1.1 

0.6 

Nongadoid  fish  remains 

47.3 

6.5 

2.3 

Gadidae  (unidentified) 

1.6 

0.1 

0.4 

Theragra  chalcogramma  (walleye  pollock) 

2.3 

0.1 

1.4 

Zoarcidae  (eelpout) 

16.3 

0.9 

14.0 

Cottoidei  (Sculpim 

2.3 

0.1 

0.2 

Dasycottus  setiger  (spinyhead  sculpin) 

0.8 

0.0 

0.2 

Stichaeidae  (prickleback) 

8.5 

1.5 

0.6 

Lumpenus  sp.  (prickleback) 

0.8 

0.0 

0.0 

Ammodytes  hexapterus  (Pacific  sand  lance) 

0.8 

0.0 

0.0 

Pleuronectidae  (flatfish) 

2.3 

0.2 

0.2 

Atheresthes  stomias  (arrowtooth  flounder) 

1.6 

0.1 

0.0 

Hippoglossoides  elassodon  (flathead  sole) 

3.9 

0.2 

0.5 

Unidentified  organic  material 

10.1 

0.5 

0.6 

Unidentified  worm-like  organism 

5.4 

0.3 

0.5 

Fish  offal  (processed  fish  parts,  e.g.,  head,  tail) 

0.8 

0.0 

8.1 

Total  prey  weight 

2715  g 

Total  stomachs 

1.30 

Total  empty  stomachs 

1 

NOTE     Yang:  Diet  changes  of  Gadus  macrocephalus  associated  with  climate  changes  in  Pavlof  Bay 


403 


D  Pacific  cod 
G  Walleye  pollock 
■  Arrowtooth  flounder 


Main  prey  items 

Figure  3 

Percentage  by  weight  of  the  main  prey  in  the  diet  of  Pacific  cod  (n=129),  walleye  pollock  (n= 
and  arrowtooth  flounder  <h  =  43)  collected  in  Pavlof  Bay  in  1995. 


216) 


pal  prey  of  Pacific  cod  collected  in  Pavlof  Bay  in  1995.  This 
is  a  large  change  in  diet  composition  compared  with  that 
observed  15  years  earlier  (Albers  and  Anderson,  1985). 
In  Albers  and  Anderson's  (1985)  study,  pandalid  shrimp, 
capelin.  and  walleye  pollock  were  the  main  prey  of  Pacific 
cod  (Fig.  4).  The  change  in  main  prey  from  pelagic  prey  in 
the  1980s  to  benthic  prey  in  1995  corresponds  to  changes 
in  species  abundance  trends  in  nearshore  small-mesh 
trawl  surveys  observed  by  Anderson  and  Piatt  (1999).  In 
that  study,  they  described  that  the  community  reorgani- 
zation in  the  Gulf  of  Alaska  was  triggered  by  a  shift  in 
ocean  climate  during  the  late  1970s.  They  showed  that  the 
abundance  of  species  such  as  pandalid  shrimp  and  capelin 
declined  while  the  abundance  of  predators  such  as  Pacific 
cod,  walleye  pollock,  and  flatfish  increased  between  1972 
and  1997. 

The  mean  weight  of  pandalid  shrimp  consumed  by  Pa- 
cific cod  in  1995  was  only  0.5  g/cod.  In  contrast,  the  mean 
weights  of  undigested  pink  shrimp  in  Pacific  cod  stomachs 
ranged  between  4.5  g/cod  and  24.4  g/cod  during  1980  and 
1981.  This  finding  corroborates  those  of  Anderson  (2000) 
and  show  that  pandalid  shrimp  abundance  continued  to 
decrease  in  the  late  1990s  and  Pacific  cod  abundance  con- 
tinued to  increase  during  that  same  period. 

The  diet  of  Pacific  cod  in  the  present  study  was  also 
compared  with  the  diet  of  Pacific  cod  in  the  broader  Gulf 
of  Alaska  shelf  area  (Fig.  5)  (Yang  and  Nelson,  2000).  The 
values  of  the  percentage  by  weight  of  capelin  in  Pacific  cod 
stomachs  in  the  Gulf  of  Alaska  in  1990,  1993,  and  1996 
were  similar  (all  were  less  than  3%)  to  that  in  Pavlof  Bay 
in  1995.  However,  pandalid  shrimp  were  an  important 


Table  2 

Mean  standard  length 

or  carapace  width),  standard  devi- 

ation, and  the 

size  ran 

ce  of  prey 

consumed  by  Pacific  cod 

in  Pavlof  Bay  1995. 

Mean 

SD 

Range 

No.  of 

Prey  name 

i  m  m  I 

(mm) 

(mm) 

individuals 

Tanner  crab 

22.1 

10.5 

4.5-42.3 

70 

Zoarcid 

86.4 

73.5 

36.2-256.6 

15 

Cottid 

48.2 

9.4 

41.8-59 

3 

Stichaeid 

46.9 

19.6 

32.7-81.5 

5 

Pacific  sand 

44.8 

0.0 

44.8-44.8 

1 

lance 

Arrowtooth 

39.1 

7.4 

33.8-44.3 

2 

flounder 

Flathead  sole 

58.2 

14.4 

47.5-80.4 

4 

food  item  of  Pacific  cod  throughout  the  Gulf  of  Alaska, 
comprising  from  11%  to  15%  by  weight  of  the  total  stom- 
ach contents  of  Pacific  cod  in  the  Gulf  of  Alaska  in  3  years 
(1990,  1993,  and  1996)  (Yang  and  Nelson,  2000).  These 
values  are  higher  than  that  in  Pavlof  Bay  (2%  by  weight). 
By  comparing  the  depths  of  the  sampling  locations  of  the 
Pacific  cod,  high  percentages  of  pandalid  shrimp  were 
found  in  the  cod  diet  in  deeper  offshore  areas  of  the  Gulf 
of  Alaska  in  1990,  1993,  and  1996,  whereas  low  percent- 
ages of  pandalid  shrimp  were  found  in  cod  diet  in  much 


404 


Fishery  Bulletin  102(2) 


shallower  areas  in  the  Pavlof  Bay  area  <Fig.  6).  From  the 
shrimp  survey  data,  Anderson  (2000)  showed  that  pan- 
dalid  shrimp  occupying  inshore  and  shallower  water  (e.g., 
Pavlof  Bay  area)  declined  to  near  extinction  (<0.1  kg/km) 
from  1978  to  1982,  while  offshore  and  deepwater  pandalid 


DAug-80 

■  May-81 
QSep-81 

■  Aug-95 


Main  prey  items 

Figure  4 

Percentage  by  volume  (for  the  values  in  1980s)  and  the  percentage  by  weight 
(for  the  values  in  1995)  of  the  main  prey  items  of  Pacific  cod  collected  in  Pavlof 
Bay.  Alaska. 


a 


Main  prey  items 

Figure  5 

Percentage  by  weight  of  the  main  prey  items  in  the  diet  of  Pacific  cod  collected 
in  1990  (GOA90),  1993  (GOA93),  and  1996  (GOA96)  in  the  Gulf  of  Alaska  and 
in  1995  (PAV95)  in  Pavlof  Bav. 


shrimp  species  maintained  low  population  levels  (>0.1 

kg/km ).  The  data  from  this  study  corroborates  Anderson's 

(2000)  results. 
Anderson  (2000)  also  reported  that  during  the  period 

of  the  decline  of  pandalid  shrimp  in  inshore  waters  of  the 
Gulf  of  Alaska,  the  abundance  of  some 
pleuronectids,  Pacific  cod,  and  walleye 
pollock  increased.  These  species  are 
predators  of  pandalid  shrimp  (Yang 
and  Nelson,  2000).  One  hypothesis  is 
that  predators  keep  pandalid  shrimp 
populations  low.  Albers  and  Anderson 
(1985)  suggested  that  cod  predation 
was  one  reason  for  the  failure  of  the 
pink  shrimp  stock  to  rebuild  in  Pavlof 
Bay.  In  the  Northwest  Atlantic.  Lilly 
et  al.  (2000)  showed  that  the  large  in- 
crease in  shrimp  biomass  seen  in  the 
1990s  was  related  to  the  collapse  of  cod 
\Gctdus  morhua)  populations  during  the 
late  1980s  and  1990s  in  the  northeast 
Newfoundland  shelf.  The  impact  of  cod 
on  Barents  Sea  shrimp  (P.  borealis) 
was  also  reported  by  Berenboim  et 
al.  (2000).  They  found  that  when  cod 
biomass  is  high,  the  shrimp  frequency 
of  occurrence  in  cod  stomachs  declines; 
there  is  a  significant  inverse  correla- 
tion between  the  abundance  of  cod  and 
shrimp. 

Tanner  crabs  consumed  by  Pacific  cod 
in  this  study  ranged  from  5  to  42  mm 
carapace  width  (CW).  In  general,  the 
size  of  Tanner  crabs  consumed  in- 
creases as  Pacific  cod  size  increases. 
The  size  range  of  Tanner  crabs  con- 
sumed by  Pacific  cod  in  this  study  is 
similar  to  that  (5-45  mm)  found  in 
Pacific  cod  stomachs  in  Albers  and  An- 
derson's ( 1985 )  study  and  is  also  similar 
to  that  (1-40  mm)  found  in  Hunter's 
(1979)  study  near  Kodiak  Island. 

Jewett's  ( 1978 )  Pacific  cod  diet  study 
around  Kodiak  Island  from  1973  to 
1976  showed  that  Tanner  crabs  were 
the  most  frequent  (37%)  prey  of  Pa- 
cific cod;  pandalid  shrimp  occurred  in 
8-10%  of  the  stomachs  examined  from 
1973  to  1975;  and  walleye  pollock  were 
found  in  49r  of  the  stomachs  examined. 
The  importance  of  Tanner  crabs  as  food 
of  Pacific  cod  in  Jewett's  ( 1978 1  study  is 
coincident  with  our  study. 

This  study  suggests  that  there  were 
substantial  differences  between  the 
diets  of  Pacific  cod  in  Pavlof  Bay  be- 
tween the  early  1980s  and  1995.  In 
the  1980s,  pandalid  shrimp  and  cap- 
elin  were  the  main  food  of  Pacific  cod, 
whereas  benthic  species  (polychaetes, 


NOTE     Yang:  Diet  changes  of  Gadus  macrocephalus  associated  with  climate  changes  in  Pavlof  Bay 


405 


80 


Figure  6 

Pandalid  shrimp  consumed  by  Pacific  cod  sampled  at  different  bottom  depths  (m)  in 
Pavlof  Bay  in  1995,  and  in  the  Gulf  of  Alaska  in  1990,  1993,  and  1996. 


hermit  crabs,  Tanner  crabs,  and  eelpouts)  were  the  domi- 
nant food  in  1995.  This  change  was  probably  due  to  the 
climate  shift  from  cold  to  warm  in  the  Gulf  of  Alaska. 


Acknowledgments 

I  would  like  to  thank  Paul  Anderson,  Troy  Buckley,  and 
Patricia  Livingston  for  reviewing  the  manuscript  and  for 
their  very  helpful  suggestions.  I  also  want  to  thank  the  two 
anonymous  reviewers  for  their  comments  and  suggestions. 


Literature  cited 

Albers  W.  D.,  and  P.  J.  Anderson. 

1985.     Diet  of  the  Pacific  cod.  Gadus  macrocephalus.  and 
predation  on  the  northern  pink  shrimp,  Pandalus  borealis, 
in  Pavlof  Bay,  Alaska.     Fish.  Bull.  83:601-610. 
Anderson,  P.  J  . 

2000.     Pandalid  shrimp  as  indicators  of  ecosystem  regime 
shift.     J.  Northw.  Atl.  Fish.  Sci.  27:1-10. 


Anderson,  P.  J„  and  J.  F.  Piatt. 

1999.  Community  reorganization  in  the  Gulf  of  Alaska  fol- 
lowing ocean  climate  regime  shift.  Mar.  Ecol.  Prog.  Ser. 
189:117-123. 

Berenboim,  B.  I.,  A.  V.  Dolgov,  V.  A.  Korzhev,  and  N.  A.  Yaragina. 

2000.  The  impact  of  cod  on  the  dynamics  of  Barents  Sea 
shrimp  (Pandalus  borealis)  as  determined  by  multispecies 
models.     J.  Northw.  Atl.  Fish.  Sci.  27:69-75. 

Hunter,  M.  A. 

1979.     Food  resource  partitioning  among  demersal  fishes 
in  the  vicinity  of  Kodiak  Island,  Alaska.  M.S.  thesis,  120 
p.     Univ.  Washington,  Seattle,  WA. 
Jewett,  S.  C. 

1978.     Summer  food  of  the  Pacific  cod,  Gadus  macrocepha- 
lus, near  Kodiak  Island,  Alaska.     Fish.  Bull.  76:700-706. 
Lilly,  G.  R.,  D.  G.  Parsons,  and  D.  W.  Kulka. 

2000.     Was  the  increase  in  shrimp  biomass  on  the  North- 
east Newfoundland  shelf  a  consequence  of  a  release  in 
predation  pressure  from  cod?    J.  Northw.  Atl.  Fish.  Sci. 
27:45-61. 
Yang,  M-S.,  and  M.  W.  Nelson. 

2000.  Food  habits  of  the  commercially  important  ground- 
fishes  in  the  Gulf  of  Alaska  in  1990,  1993,  and  1996. 
NOAA  Tech.  Memo.  NMFS-AFSC-112,  174  p. 


406 


Fishery  Bulletin  102(2) 


Superintendent  of  Documents  Publications  Order  Form 
*5178 

I |  YES,  please  send  me  the  following  publications: 


Subscriptions  to  Fishery  Bulletin 
for  $55.00  per  year  ($68.75  foreign) 


The  total  cost  of  my  order  is  $ 


Prices  include  regular  domestic 


postage  and  handling  and  are  subject  to  change. 


(Company  or  Personal  Name) 


(Please  type  or  print! 


(Additional  address/attention  line) 


( Street  address  I 


(City,  State,  ZIP  Code) 


(Daytime  phone  including  area  code) 


( Purchase  Order  No. ) 


Charge 
your 
order. 


MasterCard- 


Please  Choose  Method  of  Payment: 

]  Check  Payable  to  the  Superintendent  of  Documents 
]  GPO  Deposit  Account   [ 


]  VISA  or  MasterCard  Account 


(Credit  card  expiration  date) 


To  fax 

your  orders 

(202)  512-2250 


(Authorizing  Signature) 

Mail  To:    Superintendent  of  Documents 

P.O.  Box  371954,  Pittsburgh,  PA  15250-7954 


Thank  you  for 
your  order! 


.■■■Ill 


Fishery  Bulletin 

Guidelines  for  contributors 


Content  of  papers 

Articles 

Articles  are  reports  of  10  to  30  pages  (double 
spaced)  that  describe  original  research  in  one  or 
a  combination  of  the  following  fields  of  marine 
science:  taxonomy,  biology,  genetics,  mathematics 
(including  modeling!,  statistics,  engineering,  eco- 
nomics, and  ecology. 

Notes 

Notes  are  reports  of  5  to  10  pages  without  an 
abstract  that  describe  methods  and  results  not 
supported  by  a  large  body  of  data.  Although  all 
contributions  are  subject  to  peer  review,  responsi- 
bility for  the  contents  of  articles  and  notes  rests 
upon  the  authors  and  not  upon  the  editor  or  the 
publisher.  It  is  therefore  important  that  authors 
consider  the  contents  of  their  manuscripts  care- 
fully. Submission  of  an  article  is  un-derstood  to 
imply  that  the  article  is  original  and  is  not  being 
considered  for  publication  elsewhere.  Manuscripts 
must  be  written  in  English.  Authors  whose  native 
language  is  not  English  are  strongly  advised  to 
have  their  manuscripts  checked  for  fluency  by 
English-speaking  colleagues  prior  to  submission. 

Preparation  of  papers 
Text 

Title  page  should  include  authors'  full  names  and 
mailing  addresses  I  street  address  required)  and 
the  senior  authors  telephone,  fax  number,  e-mail 
address,  as  well  as  a  list  of  key  words  to  describe  the 
contents  of  the  manuscript.  Abstract  must  be  less 
than  one  typed  page  (double  spaced)  and  must  not 
contain  any  citations.  It  should  state  the  main  scope 
of  the  research  but  emphasize  the  authors  con- 
clusions and  relevant  findings.  Because  abstracts 
are  circulated  by  abstracting  agencies,  it  is  impor- 
tant that  they  represent  the  research  clearly  and 
concisely  General  text  must  be  typed  in  double- 
spaced  format.  A  brief  introduction  should  state  the 
broad  significance  of  the  paper;  the  remainder  of 
the  paper  should  be  divided  into  the  following  sec- 
tions: Materials  and  methods.  Results,  Discussion 
(or  Conclusions),  and  Acknowledgments.  Headings 
within  each  section  must  be  short,  reflect  a  logical 
sequence,  and  follow  the  rules  of  multiple  subdi- 
vision (i.e.  there  can  be  no  subdivision  without  at 
least  two  subheadings).  The  entire  text  should  be 
intelligible  to  interdisciplinary  readers;  therefore, 
all  acronyms  and  abbreviations  should  be  written 
out  and  all  lesser  known  technical  terms  should  be 
defined  the  first  time  they  are  mentioned.  The 
scientific  names  of  species  must  be  written  out  the 
first  time  they  are  mentioned;  subsequent  mention 
of  scientific  names  may  be  abbreviated.  Follow  Sci- 
entific style  and  format:  CBE  manual  for  authors, 
editors,  and  publishers  (6th  ed.i  for  editorial  style 
and  the  most  current  issue  of  the  American  Fish- 
eries Society's  common  and  scientific  frames  of 
from  the  United  States  and  Canada  for 
fish  nomenclature.  Dates  should  be  written  as  fol- 
lows: 11  November  1991.  Measurements  should  be 
expressed  in  metric  units,  e.g.  metric  tons  (t).  The 
numeral  one  1 1 )  should  be  typed  as  a  one,  not  as  a 
lower-case  el  (1). 

Footnotes 

Use  footnotes  to  add  editorial  comments  regarding 
claims  made  in  the  text  and  to  document  unpub- 


lished works  or  works  with  local  circulation.  Foot- 
notes should  be  numbered  with  Arabic  numerals 
and  inserted  in  10-point  font  at  the  bottom  of  the 
first  page  on  which  they  are  cited.  Footnotes  should 
be  formatted  in  the  same  manner  as  citations. 
If  a  manuscript  is  unpublished,  in  the  process 
of  review,  or  if  the  information  provided  in  the 
footnote  has  been  conveyed  verbally,  please  state 
this  information  as  "unpubl.  data,"  "'manuscript 
in  review,"  and  "personal  commun,"  respectively. 
Authors  are  advised  wherever  possible  to  avoid  ref- 
erences to  nonstandard  literature  (unpublished  lit- 
erature that  is  difficult  to  obtain,  such  as  internal 
reports,  processed  reports,  administrative  reports, 
ICES  council  minutes,  IWC  minutes  or  working 
papers,  any  "research"  or  "working"  documents, 
laboratory  reports,  contract  reports,  and  manu- 
scripts in  review).  If  these  references  are  used, 
please  indicate  whether  they  are  available  from 
NTIS  (National  Technical  Information  Sen 
from  some  other  public  depository.  Footnote  format: 
author  (last  name,  followed  by  first-name  initials); 
year;  title  of  report  or  manuscript;  type  of  report 
and  its  administrative  or  serial  number;  name  and 
address  of  agency  or  institution  where  the  report  is 
filed. 

Literature  cited 

The  literature  cited  section  comprises  works  that 
have  been  published  and  those  accepted  for  pub- 
lication 'works  in  press  I  in  peer-reviewed  jour- 
nals and  books.  Follow  the  name  and  year  system 
For  citation  format.  In  the  text,  write  "Smith  and 
Jones  (1977)  reported"  but  if  the  citation  takes 
the  form  of  parenthetical  matter,  write 
and  Jones,  19771."  In  the  literature  cited  section. 
ations  alphabetically  by  last  name  of  senior 
author:  For  example.  Alston.  1952;  Mannly,  1988; 
Smith,  1932;  Smith.  1947:  Stalinsk 

Abbreviations  of  journals  should  conform 
to  the  abbreviations  given  in  the  Serial 
for  the  BIOSIS  previews  database.  Authoi 
responsible  for  the  accuracy  and  completeness  of 
all  citations.  Literature  citation  format:  author 
(last  name,  followed  by  first-name  initials 
title  of  report  or  article;  abbreviated  title  of  the 
journal  in  which  the  article  was  published,  volume 
number,  page  numbers.  For  books,  please  provide 
publisher,  city,  and  state. 

Tables 

Tables  should  not  be  exce  and  must  be 

cited  in  numerical  order  in  the  text.  Headings  in 
tables  should  be  short  but  ample  enough  to  allow 
the  table  to  be  intelligible  on  its  own.  All  unusual 
symbols  must  be  explained  in  the  table  legend. 
Other  incidental  comments  may  be  footnoted  (use 
italic  arabic  numerals  for  footnote  markers).  Use 
asterisks  only  to  indicate  probability  in  statistical 
data.  Place  table  legends  on  the  same  page  as  the 
table  data.  We  accept  tables  saved  in  most  spread- 
sheet software  programs  (e.g.  Microsoft  Excel). 
Please  note  the  following: 

•  Use  a  comma  in  numbers  of  five  digits  or  more 
le.g.  13.000  but  3000). 

•  Use  zeros  before  all  decimal  points  for 
less  than  one  (e.g.  0.31 1. 

Figures 

Figures  include  line  illustrations,  computer-gener- 
ated line  graphs,  and  photographs  (or  slides  >.  They 


must  be  cited  in  numerical  order  in  the  text  Line 
illustrations  are  best  submitted  as  original  draw- 
ings. Computer-generated  line  graphs  should  be 
printed  on  laser-quality  paper.  Photographs  should 
be  submitted  on  glossy  paper  with  good  contrast. 
All  figures  are  to  be  labeled  with  senior  authors 
name  and  the  number  of  the  figure  (e.g.  Smith, 
Fig.  4).  Use  Helvetica  or  Arial  font  to  label  ana- 
tomical parts  (line  drawings)  or  variables  (graphs) 
within  figures;  use  Times  Roman  bold  font  to  label 
the  different  sections  of  a  figure  (e.g.  A,  B,  C». 
Figure  legends  should  explain  all  symbols  and 
abbreviations  seen  within  the  figure  and  should  be 
typed  in  double-spaced  format  on  a  separate  page 
at  the  end  of  the  manuscript.  We  advise  authors  to 
peruse  a  recent  issue  of  Fishery  Bulletin  for  stan- 
dard formats.  Please  note  the  following: 

•  Capitalize  the  first  letter  of  the  first  word  of 
axis  labels. 

•  Do  not  use  overly  large  font  sizes  to  label  axes 
or  parts  within  figures. 

•Do  not  use  boldface  fonts  within  figures. 
•Do  not  create  outline  rules  around  graphs. 
•Do  not  use  horizontal  lines  through  graphs. 
•Do  not  use  large  font  sizes  to  label  degrees  of 
longitude  and  latitude  on  maps. 

•  Indicate  direction  of  degrees  longitude  and 
latitude  on  maps 

•Avoid  placing  labels  on  a  vertical  plane 

(except  on  ) 
•Avoid  odd  (nonstandard)  patterns  to  mark 

sections  of  bar  graphs  and  pie  charts. 

Copyright  law 

Fishery  Bulletin,  a  U.S.  government  publication,  is 
not  subject  to  copyright  law.  If  an  author  wishes  to 
<>t' Fishery  Bulletin  in  his  or  her 
work,  he  or  she  is  obliged,  however,  to  acknowledge 
the  source  of  the  extracted  literature. 

Submission  of  papers 

Send  four  printed  copies  (one  original  plus  three 
copies) — clipped,  not  stapled — to  the  Scientific  Edi- 
tor, at  the  address  shown  below.  Send  photocopies 
of  figures  with  initial  submission  of  manuscript. 
Original  figures  will  be  requested  later  when  the 
manuscript  has  been  accepted  for  publication. 
Do  not  send  your  manuscript  on  diskette  until 
requested  to  do  so. 

Dr.  Norman  Bartoo 

National  Marine  Fisheries  Service,  NOAA 

8604  La  Jolla  Shores  Drive 

LaJolla,  CA  92037 

Once  the  manuscript  has  been  accepted  for  pub- 
lication, you  will  be  asked  to  submit  a  software 
copy  of  your  manuscript.  The  software  copy  should 
be  submitted  in  WordPerfect  or  Word  format  (in 
Word,  save  as  Rich  Text  Format).  Please  note  that 
we  do  not  accept  ASCII  text  files 

Reprints 

Copies  of  published  articles  and  notes  are  avail- 
able free  of  charge  to  the  senior  author  (50  copies) 
and  to  his  or  her  laboratory  <  50  copies ).  Additional 
copies  may  be  purchased  in  lots  of  100  when  the 
author  receives  page  proofs. 


U.S.  Department 
of  Commerce 

Volume  102 
Number  3 
July  2004 


AUG  0  2  2004 


U.S.  Department 
of  Commerce 

Donald  L.  Evans 
Secretary 


National  Oceanic 
and  Atmospheric 
Administration 

Vice  Admiral 

Conrad  C.  Lautenbacher  Jr., 

USN  (ret.) 

Under  Secretary  for 
Oceans  and  Atmosphere 


National  Marine 
Fisheries  Service 

William  T.  Hogarth 

Assistant  Administrator 
for  Fisheries 


^tofc 


**rEs  of  * 


The  Fishery  Bulletin  (ISSN 
is  published  qua>  I 
Publications  Office.  National  M 

and  at  additional  mailing  ofl 

scriptions   to  Fishery   Built  i 
tendent  of  Docum 
inch.  Mail 
ton.  DC  20402-:' 

Although  the  contents  of  this  publica- 
tion ha  ii  copyrighted  and  may 
he  reprinted 
is  apprecial 

The  S 
mined  that  the  publication  of  tin 
odical  is  nei 

nsaction  of  public  business  of  this 
of  funds  for  pnn 
this  periodica]  has  been  approved  by  the 
or  of  the  Office  of  Management  and 
Budget 

For  sale 
Documents.  U.S.  Government  Printing 
Office.  Washington.  DC  20402.  Subscrip- 
tion price  pi  i  i  Lc  and 
$68.75  foreign.  Cost  per  single  issue: 
$28.00  domestic  and  $35.00  foreign.  See 
back  for  order  form. 


Scientific  Editor 

Norman  Bartoo,  PhD 

Associate  Editor 

Sarah  Shoffler 

National  Marine  Fisheries  Service,  NOAA 
8604  La  Jolla  Shores  Drive 
La  Jolla,  California  92037 

Managing  Editor 

Sharyn  Matriotti 

National  Marine  Fisheries  Service 
Scientific  Publications  Office 
7600  Sand  Point  Way  NE,  BIN  C15700 
Seattle,  Washington  98115-0070 


Editorial  Committee 

Harlyn  O.  Halvorson,  PhD 
Ronald  W.  Hardy,  PhD 
Richard  D.  Methot,  PhD 
Theodore  W.  Pietsch,  PhD 
Joseph  E.  Powers,  PhD 
Harald  Rosenthal,  PhD 
Fredric  M.  Serchuk,  PhD 
George  Watters,  PhD 


University  of  Massachusetts,  Boston 
University  of  Idaho,  Hagerman 
National  Marine  Fisheries  Service 
University  of  Washington,  Seattle 
National  Marine  Fisheries  Service 
Universitat  Kiel,  Germany 
National  Marine  Fisheries  Service 
National  Marine  Fisheries  Service 


Fishery  Bulletin  web  site:  www.fishbull.noaa.gov 


The  /•'  iid  technical  not  ions  in 

[I  began  as  the  Bulletin  of  the  United  Stai 
Commission  in  1881;  it  became  the  Bulletin  of  the  Bureau  of  Fisheries  m  1904  and  the  I 
Bulletin  of  the  Fish  and  Wildlife  S  941.  Separates  were  issued  as  documents  tl 

volume   Hi.  the  last  document  was  No.  1103.  Beginning  with  volume  47  in  1931  and  continuing 

Ii   volume  82  in  IS  a  numbered  bulletin.  A  new    system 

i   which  papers  are  bound  together  in  a   single  issue  of  the 
bulletii  f0,  number  1.  January  1972.  the  Fishery  Bulletin  bee: 

periodical,   -  terly.  In  this  form,  it  is  available  by  subscription  from  the  Superintendent 

uments,  U.S.  Gove nt  Printing  Office,  Washington.  DC  20402.  Ii  i-  also  available  free  in 

limited  numbers  to  hi"  arch  institutions,  Slate  and  Federal  agencies,  and  in  exchange 

lor  other  scientific  publications 


U.S.  Department 
of  Commerce 

Seattle,  Washington 

Volume  102 
Number  3 
July  2004 


Fishery 
Bulletin 


Contents 


Articles 


407—417  Abascal,  Francisco  J.,  Cesar  Megina,  and  Antonio  Medina 

Testicular  development  in  migrant  and  spawning  bluefin 
tuna  (Thunnus  thynnus  (U)  from  the  eastern  Atlantic  and 
Mediterranean 


418-429  Bobko,  Stephen  J.,  and  Steven  A.  Berkeley 

Maturity,  ovarian  cycle,  fecundity,  and  age-specific  parturition 
of  black  rockfish  (Sebastes  melanops) 


430-440  Brock,  Daniel  J.,  and  Timothy  M.  Ward 

Maori  octopus  (Octopus  maorum)  bycatch  and  southern  rock 
lobster  (Jasus  edwardsii)  mortality  in  the  South  Australian 
lobster  fishery 


The  conclusions  and  opinions  expressed 
in  Fishery  Bulletin  are  solely  those  of  the 
authors  and  do  not  represent  the  official 
position  of  the  National  Marine  Fisher- 
ies Service  INOAA)  or  any  other  agency 
or  institution. 

The  National  Marine  Fisheries  Service 
(NMFSi  does  not  approve,  recommend,  or 
endorse  any  proprietary  product  or  pro- 
prietary material  mentioned  in  this  pub- 
lication. No  reference  shall  be  made  to 
NMFS.  or  to  this  publication  furnished  by 
NMFS,  in  any  advertising  or  sales  pro- 
motion which  would  indicate  or  imply 
that  NMFS  approves,  recommends,  or 
endorses  any  proprietary  product  or  pro- 
prietary material  mentioned  herein,  or 
which  has  as  its  purpose  an  intent  to 
cause  directly  or  indirectly  the  advertised 
product  to  be  used  or  purchased  because 
of  this  NMFS  publication. 


441—451  Dawson,  Stephen,  Elisabeth  Slooten,  Sam  DuFresne, 

Paul  Wade,  and  Deanna  Clement 

Small-boat  surveys  for  coastal  dolphins;  line-transect  surveys 
for  Hector's  dolphins  (Cephalorhynchus  hector/) 


452—463  Laidig,  Thomas  E.,  Keith  M.  Sakuma,  and  Jason  A.  Stannard 

Description  and  growth  of  larval  and  pelagic  juvenile  pygmy 
rockfish  (Sebastes  wilsoni)  (family  Sebastidae) 


464—472  McGarvey,  Richard 

Estimating  the  emigration  rate  of  fish  stocks  from  marine 
sanctuaries  using  tag-recovery  data 


473-487  Roumillat,  William  A.,  and  Myra  C.  Brouwer 

Reproductive  dynamics  of  female  spotted  seatrout 
(Cynoscion  nebutosus)  in  South  Carolina 


Fishery  Bulletin  102(3) 


488-497  Taggart,  S.  James,  Charles  E.  O'Clair,  Thomas  C.  Shirley,  and  Jennifer  Mondragon 

Estimating  Dungeness  crab  (Cancer  magister)  abundance:  crab  pots  and  dive  transects  compared 

Companion  papers 

498-508  Tollit,  Dominic  J.,  Susan  G.  Heaslip,  Tonya  K.  Zeppelin,  Ruth  Joy,  Katherine  A.  Call, 

and  Andrew  W.  Trites 

A  method  to  improve  size  estimates  of  walleye  pollock  (Theragra  chalcogramma)  and 
Atka  mackerel  (Pleurogrommus  monopterygius)  consumed  by  pinnipeds: 
digestion  correction  factors  applied  to  bones  and  otoliths  recovered  in  scats 

509-521  Zeppelin,  Tonya  K.,  Dominic  J.  Tollit,  Katherine  A.  Call,  Trevor  J.  Orchard, 

and  Carolyn  J.  Gudmundson 

Sizes  of  walleye  pollock  (Theragra  chalcogramma)  and  Atka  mackerel 
(Pleurogrommus  monopterygius)  consumed  by  the  western  stock  of  Steller  sea  lions 
(Eumetopias  /ubatus)  in  Alaska  from  1999  to  2000 

522-532  Tollit,  Dominic  J.,  Susan  G.  Heaslip,  and  Andrew  W.  Trites 

Sizes  of  walleye  pollock  (Theragra  chalcogramma)  consumed  by  the  eastern  stock  of 
Steller  sea  lions  (Eumetopias  /ubatus)  in  Southeast  Alaska  from  1994  to  1999 

533-544  Tremain,  Derek  M.,  Christopher  W.  Harnden,  and  Douglas  H.  Adams 

Multidirectional  movements  of  sportfish  species  between  an  estuanne  no-take  zone  and 
surrounding  waters  of  the  Indian  River  Lagoon,  Florida 

545-554  Wells,  R.  J.  David,  and  Jay  R.  Rooker 

Distribution,  age,  and  growth  of  young-of-the  year  greater  amberjack  (Seriola  dumerili)  associated  with 
pelagic  Sargassum 


Note 

555-560  Hiroishi,  Shingo,  Yasutaka  Yuki,  Eriko  Yuruzume,  Yosuke  Onishi,  Tomoji  Ikeda,  Hironobu  Komaki, 

and  Muneo  Okiyama 

Identification  of  formalin-preserved  eggs  of  red  sea  bream  (Pagrus  ma/or)  (Pisces:  Spandae)  using 
monoclonal  antibodies 

561  Subscription  form 


407 


Abstract— Testis  histological  structure 
was  studied  in  bluefin  tuna  {.Thunnus 
thynnus  I  from  the  eastern  Atlantic  and 
Mediterranean  during  the  reproductive 
season  (from  late  April  to  early  Junei. 
Testicular  maturation  was  investi- 
gated by  comparing  samples  from 
bluefin  tuna  caught  on  their  eastward 
reproductive  migration  off  Barbate 
i  Strait  of  Gibraltar  area)  with  samples 
of  bluefin  tuna  fished  in  spawning 
grounds  around  the  Balearic  Islands. 
Histological  evaluations  of  cross  sec- 
tions showed  that  the  testis  consists  of 
two  structurally  different  regions,  an 
outer  proliferative  region  where  germ 
cells  develop  synchronously  in  cysts, 
and  a  central  region  made  up  of  a  well- 
developed  system  of  ducts  that  convey 
the  spermatozoa  produced  in  the  prolif- 
erative region  to  the  main  sperm  duct. 
Ultrastructural  features  of  the  differ- 
ent stages  of  the  male  germ  cell  line  are 
very  similar  to  those  described  in  other 
teleost  species.  The  bluefin  tuna  testis 
is  of  the  unrestricted  spermatogonial 
testicular  type,  where  primary  sper- 
matogonia are  present  all  along  the 
germinative  portion  of  the  lobules.  All 
stages  of  spermatogenesis  were  pres- 
ent in  the  gonad  tissue  of  migrant  and 
spawning  bluefin  tuna,  although  sper- 
matids were  more  abundant  in  spawn- 
ing fish.  The  testis  size  was  found  to 
increase  by  a  factor  of  four  ( on  average ) 
during  migration  to  the  Mediterranean 
spawning  grounds,  whereas  the  fat 
bodies  (mesenteric  lipid  stores  associ- 
ated with  the  gonads)  became  reduced 
to  half  their  weight,  and  the  liver  mass 
did  not  change  significantly  with  sexual 
maturation.  Linear  regression  analy- 
sis of  the  pooled  data  of  migrant  and 
spawning  bluefin  tuna  revealed  a  sig- 
nificant negative  correlation  between 
the  gonad  index  (IG)  and  the  fat  tissue 
index  (IF),  and  a  weaker  positive  cor- 
relation between  the  gonad  index  (IG) 
and  the  liver  index  (IL).  Our  analyses 
indicate  that  the  liver  does  not  play  a 
significant  role  in  the  storage  of  lipids 
and  that  mesenteric  lipid  reserves  con- 
stitute an  important  energy  source  for 
gametogenesis  in  bluefin  tuna. 


Testicular  development  In  migrant  and  spawning 
bluefin  tuna  {Thunnus  thynnus  (L.))  from 
the  eastern  Atlantic  and  Mediterranean 


Francisco  J.  Abascal 

Cesar  Megina 

Antonio  Medina 

Departamento  de  Biologia 

Facultad  de  Ciencias  del  Mar  y  Ambientales 

Universidad  de  Cadiz 

Av  Republica  Saharaui 

11510  Puerto  Real 

Cadiz,  Spain 

E-mail  address  (for  A  Medina,  contact  author):  antonio.medina@uca.es 


Manuscript  submitted  27  April  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
25  March  2004  by  the  Scientific  Editor. 
Fish.  Bull.  102:407-417  (2004). 


The  Atlantic  northern  bluefin  tuna 
(Thunnus  thynnus  thynnus  (L.)),  is 
one  of  the  most  commercially  valu- 
able wild  animals  in  the  world.  In  the 
last  two  decades  this  species  has  been 
subject  to  intense  over-fishing,  which 
has  caused  a  decline  in  both  the  east- 
ern and  western  populations  because 
of  lowered  recruitment  (Mather  et 
al.,  1995;  Sissenwine  et  al.,  1998). 
The  bluefin  tunas  (7!  thynnus  and  T. 
maccoyii)  are  unique  among  tuna  spe- 
cies in  that  they  live  mainly  in  cold 
waters  and  move  into  warmer  waters 
to  spawn  (Olson,  1980;  Lee,  1998; 
Schaefer,  2001);  therefore  the  migra- 
tory pattern  of  these  species  depends 
substantially  on  reproduction.  The 
eastern  stock  of  Atlantic  bluefin  tuna 
spawns  from  June  through  August  in 
the  Mediterranean  Sea.  where  natural 
conditions  are  apparently  optimal  for 
the  survival  of  offspring.  From  late 
April  to  mid  June,  bluefin  tuna  breed- 
ing stocks  migrate  from  the  North 
Atlantic  to  spawning  grounds  in  the 
Mediterranean  (Mather  et  al.,  1995; 
Ravier  and  Fromentin,  2001).  A  good 
understanding  of  the  reproductive 
parameters  (especially  sexual  matu- 
ration, fecundity,  and  spawning)  of 
tunas  is  of  paramount  importance  for 
population  dynamics  studies  and  the 
management  of  fisheries  that  target 
tunas.  Nevertheless,  "a  very  limited 
amount  of  scientifically  useful  infor- 


mation is  available  on  the  reproduc- 
tive biology  for  most  tunas"  ( Schaefer, 
2001).  Recent  work  has  increased  our 
knowledge  on  the  reproductive  biol- 
ogy of  female  Thunnus  thynnus  in 
the  eastern  Atlantic  and  the  Medi- 
terranean (Susca  et  al„  2000,  2001a, 
2001b;  Hattour  and  Macias,  2002; 
Medina  et  al.,  2002;  Mourente  et  al., 
2002),  but  many  questions  remain 
still  to  be  answered  regarding  male 
reproductive  activity  in  this  and  other 
tuna  species. 

Histological  examination  of  gonads 
is  a  useful  tool  for  assessing  the  ma- 
turity state  offish.  However,  very  few 
light-microscopy  studies  have  been 
published  on  bluefin  tuna  and  no  ul- 
trastructural studies  of  reproductive 
organs  are  yet  available.  The  male 
reproductive  cycle  of  T  thynnus  has 
been  characterized  histologically  by 
Santamaria  et  al.  (2003),  and  Ratty 
et  al.  ( 1990 )  and  Schaefer  ( 1996,  1998 ) 
have  reported  valuable  histological 
descriptions  on  male  and  female  go- 
nads of  the  Pacific  albacore  (Thun- 
nus alalunga)  and  the  yellowfin  tuna 
(Thunnus  albacares),  respectively.  In 
this  article  we  report  biometric  and 
histological  data  on  male  T.  thynnus 
caught  during  their  reproductive  mi- 
gration and  spawning  period  in  order 
to  provide  further  information  on  the 
biological  aspects  of  reproduction  for 
this  species. 


408 


Fishery  Bulletin  102(3) 


Materials  and  methods 


Statistical  analysis 


Samples  and  condition  indices 

During  the  eastward  migration,  62  adult  male  bluefin 
tuna  weighing  between  71  and  273  kg  (mean  195.17  kg) 
were  obtained  from  the  trap  fishery  in  the  area  of  the 
Strait  of  Gibraltar  (Barbate.  Cadiz,  southwestern  Spain) 
from  late  April  to  early  June  1999.  2000,  and  2001. 
Thirty-four  mature  males,  weighing  between  19  and 
349  kg  (mean  115.11  kg),  were  sampled  in  June-July 
1999-2001  from  the  purse-seine  fleet  operating  in  the 
Mediterranean  spawning  grounds  of  bluefin  tuna  off 
the  Balearic  Islands.  Whenever  possible,  the  total  body 
weight  (W)  was  recorded  to  the  nearest  kg.  When  indi- 
vidual body  weights  were  not  available,  W  was  estimated 
from  the  fork  length  (LF)  measurements  (recorded  to  the 
nearest  cm),  according  to  the  formula:  W  =  0.000019  x 
LF3  (Table  VIII  in  Rodriguez-Roda,  1964).  Following 
dissection,  the  liver,  testes,  and  the  fat  bodies  associ- 
ated with  the  gonads  were  removed  and  weighed  to  the 
nearest  g.  The  condition  of  the  fish  was  assessed  by  three 
different  indices.  The  gonad  index  (gonadosomatic  index) 
(IG)  is  indicative  of  the  maturation  state  and  was  calcu- 
lated as:  IG  =  (WG  I  W)  x  100,  where  WG  =  gonad  weight. 
The  liver  index  (hepatosomatic  index)  (IL)  and  fat-body 
index  (IF)  were  calculated  as  IL  =  (WL  I  W)  x  100,  and 
Ip-  =  (WF  I  W)  x  100  (where  WL  and  WF  represent  liver 
and  fat-body  weights),  respectively,  and  are  considered 
as  good  indicators  of  the  metabolic  condition  and  energy 
reserves  of  the  fish.  All  measurements  are  expressed  as 
means  +SD. 


The  bluefin  tuna  specimens  used  in  this  study  showed 
considerable  variability  in  size,  especially  those  caught 
by  purse  seine  in  Balearic  waters,  where  weight  ranged 
between  12  and  349  kg.  The  purse-seine  fishery  is,  in 
fact,  much  less  size-selective  than  are  traps,  which 
seldom  catch  small  bluefin  tuna  (Rodriguez-Roda,  1964: 
Mather  et  al.,  1995).  Analysis  of  covariance  (ANCOVA), 
with  body  weight  as  covariate,  was  used  as  the  most 
suitable  method  (Garcia-Berthou,  2001)  to  test  interan- 
nual  differences  in  the  weight  of  the  organs  within  the 
two  sampling  sites.  ANCOVA  was  likewise  applied  to 
compare  the  weights  of  the  three  organs  between  both 
areas.  All  data  were  previously  log-transformed  to  meet 
the  prerequisites  of  normality  and  homoscedasticity 
(Zar,  1996).  Linear  least-squares  regression  analyses 
were  performed  to  test  possible  correlations  between 
IG  and  the  two  other  indices  (IL  and  IF)  by  using  the 
pooled  data  of  Barbate  and  the  Balearic  Islands.  In 
the  regression  between  IG  and  IF,  the  Balearic  samples 
corresponding  to  year  2001  were  excluded  because  the 
reduced  fat-body  size  (adipose  tissue  was  almost  non- 
existent in  the  mesentery)  of  these  small  bluefin  tuna 
did  not  permit  an  accurate  weight  measurements  on 
board.  The  values  of  the  indices  were  arcsine-trans- 
formed  prior  to  the  statistical  analysis  (Zar,  1996).  A 
P-value  <0.05  was  considered  statistically  significant 
for  all  tests. 


Results 


Histology 

For  light  microscopy,  tissue  samples  from  the  central 
part  of  the  testes  were  fixed  for  48-96  hours  in  10% 
formalin  in  phosphate  buffer,  0.1  M,  pH  7.2.  After  dehy- 
dration in  ascending  concentrations  of  ethanol,  a  part 
of  each  sample  was  embedded  in  paraffin  wax  and  the 
remainder  was  embedded  in  plastic  medium  (2-hydroxy- 
ethyl-methacrylate).  Paraffin  sections  (6  /im  thick)  were 
stained  with  haematoxylin-eosin,  and  plastic  sections  (3 
/jm  thick)  were  stained  with  toluidine  blue.  These  were 
examined  and  photographed  on  a  Leitz  DMR  BE  light 
microscope. 

For  electron  microscopy,  small  fragments  of  testis  were 
fixed  for  3-4  hours  in  2.5%  glutaraldehyde  buffered  with 
0.1  M  sodium  cacodylate  buffer  (pH  7.2).  Following  two 
30-min  washes  in  cacodylate  buffer,  they  were  postfixed 
for  1  hour  at  4C  in  cacodylate-buffered  1%  osmium 
tetroxide,  rinsed  several  times  in  buffer,  dehydrated  in 
ascending  concentrations  of  acetone,  and  embedded  in 
epoxy  resin  (either  Epon  812  or  Spurr).  Thin  sections 
(-80  nm  thick)  were  picked  up  on  copper  grids,  stained 
with  uranyl  acetate  and  lead  citrate,  and  examined  in 
a  Jeol  1200  EX  transmission  electron  microscope.  Ap- 
proximate dimensions  provided  for  germ  cells  are  mea- 
surements (means  ±SD)  of  the  largest  cell  diameters  on 
electron  micrographs. 


Condition  indices 

ANCOVA  did  not  reveal  significant  interannual  differ- 
ences in  gonad,  liver,  and  fat-body  weight  in  the  samples 
of  Barbate  as  well  as  in  those  of  the  Balearic  Islands.  In 
contrast,  a  strongly  significant  difference  in  testicular 
size  (P<0.0001)  was  found  in  comparing  data  of  matur- 
ing bluefin  tuna  from  Barbate  (migrant  tuna)  with  fully 
mature  fish  from  the  Balearic  Islands  (spawning  fish). 
In  fact,  as  shown  in  Figure  1,  the  average  I0  was  more 
than  fourfold  higher  in  the  Balearic  Islands  than  it  was 
in  Barbate  (4.81  ±1.77  vs.  1.12  ±0.57).  This  finding  may 
indicate  a  noticeable  increase  in  sperm  production  during 
reproductive  migration  to  the  Mediterranean  spawning 
grounds.  Significant  differences  between  maturing  and 
spawning  tuna  were  also  found  in  fat-body  weight,  the 
volume  of  which  dropped  to  about  half  by  spawning 
time.  Thus,  IF  fell  from  0.36  ±0.24  in  migrating  fish  to 
0.16  ±0.12  in  spawning  fish  isee  Fig.  1).  The  liver  mass, 
however,  did  not  differ  significantly  (P=0.31>  between 
the  two  samples. 

Figure  2  illustrates  linear  regression  analysis  be- 
tween IG  and  I,  ,  and  between  IG  and  IF.  A  significant 
negative  correlation  (;-'-  =  0.34;  P<0.0001)  was  found  be- 
tween Iq  and  Ip,  indicating  that  the  amount  of  mesenteric 
fat  tissue  decreases  as  the  gonad  matures.  In  contrast, 
there  was  a  positive,  though  somewhat  weak,  correlation 


Abascal  et  al.:  Testicular  development  in  Thunnus  thynnus 


409 


V? 


™i  'G 
V////A  l|_ 

c=i  If 


j&A 

Barbate 


Balearic  Islands 


Figure    1 

Differences  in  gonad  index  (IG) 


liver  index  1 1,  ), 


bluefin  tuna  [Thunnus  thynnus)  from  Barbate  and 
male  tuna  from  the  Balearic  Islands. 


(r2=0.21;  P<0.0001)  between  IG  and  IL,  which  suggests  a 
slight  growth  of  the  liver  with  sexual  maturation. 

Histology 

The  testes  of  Thunnus  thynnus  are  paired,  elongate 
organs  that  appear  attached  to  the  dorsal  body  wall  by 
a  mesentery.  The  fat  body,  which  is  closely  associated 
with  the  gonad,  consists  of  a  variable  amount  of  adipose 
tissue.  The  testis  is  composed  of  a  dense  array  of  lobules 
converging  on  the  main  sperm  duct  (vas  deferens)  and 
terminating  blindly  beneath  the  tunica  albuginea  at  the 
periphery  ( Fig.  3,  A  and  B).  Two  distinct  zones  can  be 
distinguished  in  cross  sections  of  the  testes  (Fig.  3A).  At 
the  outer  region,  the  seminiferous  lobules  have  a  thick 
wall  formed  by  the  germinal  epithelium,  where  germ  cells 
develop  in  association  with  Sertoli  cells;  the  lumina  of 
the  lobules  are  filled  with  spermatozoa  that  have  been 
released  after  completion  of  the  spermiogenetic  process 
(Fig.  3,  B  and  C).  As  a  result  of  the  release  of  mature 
sperm  from  spermatocysts  into  the  lobule  lumina,  the 
germinal  epithelium  becomes  discontinuous  (Fig.  3B). 
The  transition  from  the  outer  to  the  central  region  of  the 
testis  is  marked  by  an  abrupt  change  in  the  configuration 
of  the  testicular  lobules,  which  lose  the  germinal  epithe- 
lium and  become  ducts  where  lobule  function  has  shifted 
from  sperm  production  to  sperm  storage  (Fig.  3C).  Thus, 
the  only  sex  cells  that  are  found  in  the  central  part  of 
the  testis  are  mature  spermatozoa,  which  fill  the  swollen 
lumina  of  the  lobules.  In  this  zone  the  testis  ducts  con- 
stitute an  intricate  network  of  channels  that  convey  the 
spermatozoa  produced  in  the  proliferative  region  to  the 
main  sperm  duct  (Fig.  3,  A  and  D),  which  is  thick  walled 
and  located  in  the  center  of  the  testis  (Fig.  3D). 


0,25 

A 

0,20 

o 

^    0,10 
o 

0,05 

• 

0,00 

0,00        0.05        0.10        0,15        0,20        0.25        0,30        0,35 

arcsin[(lG/100)1'2] 

0,12  • 

B 

• 

0.10 

• 

c\j 

5-    0.08 

o 

o 

jL   0,06  - 

o     0.04  - 

03 

•  > 

•  •  •  .^>^»    • 

•                              ^"~\^^ 

0.02 

•\. 

•                                    ^--^ 

0,00         0,05        0,10        0,15        0,20         0,25         0,30         0,35 

arcsin[(lG/100)"2] 

Figure    2 

Linear  regression  between  gonad  index  (IG)  and  liver 

index  (ILl  I  Al,  and  between  gonad  index  (I,-. )  and  fat  body 

index  (IFI  (Bi  (data  were  pooled  from  the  two  areas).  In 

B,  samples  of  bluefin  tuna  (Thunnus  thynnus)  collected 

off  the  Balearic  Islands  in  2001  were  excluded. 

The  gametes  develop  in  groups  of  isogenic  cells  called 
germinal  cysts  or  spermatocysts,  where  the  process  of 
differentiation  is  synchronous  (Fig.  4).  Primary  sper- 
matogonia are  large,  single  cells  (Fig.  4A)  that  are  dis- 
tributed all  along  the  germinal  epithelium,  as  is  char- 
acteristic of  the  teleost  unrestricted  testicular  type. 
Spermatogonia  B  resulting  from  successive  mitoses  of 
spermatogonia  A  are  found  in  small  groups,  whereas 
spermatocytes  and  spermatids  are  grouped  within  larger 
spermatocysts  (Figs.  3,  B  and  C,  4).  The  cysts  contain- 
ing late  spermatids  and  spermatozoa,  prior  to  spermia- 
tion,  display  a  particular  alveolar  appearance  due  to 
the  orientation  of  the  spermatid  heads  facing  the  lobule 
walls  and  the  bundles  of  fiagella  directed  toward  the 
seminiferous  lobule  lumen  (Fig.  4,  A,  C,  and  D). 

Active  spermatogenesis  was  observed  to  occur  both  in 
migrant  bluefin  tuna  from  the  Strait  of  Gibraltar  (Fig.  4, 
A  and  B)  and  spawning  fish  from  the  Mediterranean 
(Fig.  4,  C  and  D).  In  both  cases,  all  stages  of  the  male 


410 


Fishery  Bulletin  102(3) 


Figure    3 

Light  micrographs  depicting  the  histological  organization  of  the  bluefin  tuna  (Thunnus  thyn- 
nus)  testis.  iAi  Transverse  section  showing  the  outer  proliferative  region  (PR)  of  the  testis  and 
the  inner  region,  which  includes  the  testis  duct  system  (D)  and  the  main  sperm  duct  (MSD). 
(B)  Peripheral  zone  of  the  testis  where  the  distal  ends  of  some  tubules  i  dotted  lines!  terminate 
beneath  the  tunica  albuginea  (TA).  (C)  Transition  (dotted  line)  between  the  outer  region  (PR) 
of  the  testis,  where  the  lobules  contain  developing  germinal  cysts,  and  the  inner  region  (Dl, 
whose  lobule  walls  enclose  only  mature  spermatozoa.  (Dl  Main  sperm  duct  iMSDi  filled  with 
a  compact  mass  of  spermatozoa  that  are  incorporated  (arrowhead)  from  the  testis  ducts  (D). 
Arrows  =  discontinuities  in  the  germinal  epithelium;  sc  =  spermatocytes;  sd  =  spermatids;  sg  = 
spermatogonia;  sz  =  spermatozoa.  All  samples  are  from  Barbate  (A  and  D,  paraffin-embedded 
sections  stained  with  haematoxylin-eosin;  B  and  ('.  toluidine-blue  stained  plastic-embedded 
sections). 


Abascal  et  al.:  Testicular  development  in  Thunnus  thynnus 


411 


L  ,  tf  K  *.T<£W*.  »5r  *  .'50  ufk 


Figure   4 

Spermatocysts  in  the  testis  proliferative  region  of  bluefin  tuna  (Thunnus  thynnus)  from  Barbate 
(A  and  B)  and  the  Balearic  Islands  (C  and  D).  All  stages  of  spermatogenesis  are  present  in  both 
cases,  although  spermatid  cysts  containing  late  spermatids  and  spermatozoa  (asterisks)  are 
somewhat  more  abundant  in  specimens  from  the  Balearic  Islands.  Arrow  =  dividing  sperma- 
tocytes; sc  =  spermatocytes;  sd  =  spermatids;  sg  =  primary  spermatogonion;  sz  =  spermatozoa. 
Plastic-embedded  sections  (A-D)  were  stained  with  toluidine  blue. 


germ  cell  line  were  present  in  the  gonads.  In  addition, 
large  amounts  of  spermatozoa  had  accumulated  in  the 
central  system  of  ducts  and  in  the  main  sperm  duct, 
both  of  which  appear  to  function  as  reservoirs  of  sperm. 
In  specimens  from  Barbate,  spermatocytes  and  sper- 


matids were  abundant  (Fig.  4B),  whereas  in  most  tuna 
collected  in  the  Balearic  area  spermatids  predominated 
over  spermatocytes.  Cysts  containing  late  spermatids 
and  spermatozoa  were  particularly  common  (Fig.  4,  C 
and  D). 


412 


Fishery  Bulletin  102(3) 


sgA 


sd2 


sgA 


B^'-V-   ^ 


4/r 


Se 

(.. 


A 


!."»"  , 


■■■':'.'^' 


m 


/*-■» 


• 


I 


sd3 


mc 


sgA      ^ 


// 


'I 


cf 
Se 


2  urn 


Figure    5 

Electron  micrographs  of  testicular  tissue  from  bluefin  tuna  (Thunnus  thynnus).  i  A  and  Bl  over- 
views displaying  several  spermatocyst  types,  including  single  primary  spermatogonia  (sgA). 
and  clusters  of  spermatogonia  B  IsgB),  primary  spermatocytes  iscl  I,  mid  spermatids  ( sd2 1.  late 
spermatids  (sd.3),  and  spermatozoa  (sz).  The  germ  cells  are  surrounded  by  Sertoli  cells  (Se). 
Arrowheads  =  perinuclear  bodies  ("image");  N  =  nucleus;  n  =  nucleolus;  cf  =  collagen  fibers; 
mc  =  myoid  cell. 


infrastructure 

Primary  spermatogonia  are  large,  ovoid  cells  (8.55  ±1.07 
jim)  whose  nucleus  (>5  Jim  in  its  largest  diameter)  shows 


diffuse  chromatin  and  a  single  central  nucleolus.  The 
cytoplasm  contains  free  ribosomes,  a  few  mitochondria, 
endoplasmic  reticulum  cisternae,  and  several  masses 


Abascal  et  al.:  Testicular  development  in  Thunnus  thynnus 


413 


of  electron-dense  perinuclear  material  ("nuages")  that 
indicate  nucleocytoplasmic  transport  (Fig.  5,  A  and  B). 
Such  chromatoid  bodies  persist  throughout  spermatogen- 
esis until  the  spermatid  stage,  but  their  size  and  number 
is  far  higher  in  primary  spermatogonia.  Spermatogonia 
B  are  grouped  in  clusters  of  a  few  cells.  They  are  per- 
ceptibly smaller  (6.75  ±0.37  um)  than  spermatogonia  A 
and  their  nucleus  contains  patchy  chromatin  (Fig.  5,  A 
and  B). 

Spermatocytes  form  clusters  in  which  the  cells  are 
interconnected  by  cytoplasmic  bridges.  Primary  sperma- 
tocytes (4.84  ±0.45  Jim)  show  a  heterochromatic  nucleus 
(-3.5  um  in  diameter)  that  varies  in  appearance  depend- 
ing on  the  prophase-I  stage.  The  cytoplasm  contains 
free  ribosomes  (mostly  polysomes),  mitochondria,  clear 
vesicles,  and  the  diplosome  (Figs.  5A,  6A).  Synapton- 
emal  complexes  are  clearly  recognizable  at  pachytene 
(Fig.  6A).  Secondary  spermatocytes  are  apparently 
short-lived  cells  because  they  are  rare  in  histological 
samples — a  finding  that  suggests  that  the  second  meiotic 
division  is  triggered  shortly  after  completion  of  the  first 
division.  Spermatocytes  II  are  difficult  to  distinguish 
morphologically  from  early  spermatids,  although  they 
are  slightly  larger  (3.31  ±0.47  /jm).  The  cytoplasm  is 
more  reduced  than  in  spermatocytes  I  and  the  nucleus 
shows  diffuse  chromatin  forming  moderately  electron- 
dense  patches  (Fig.  6B). 

During  spermiogenesis,  the  spermatid  nucleus  changes 
in  shape  and  decreases  in  volume  as  the  chromatin  con- 
denses. In  early  spermatids  (2.39  ±0.28  /Jm)  the  spheri- 
cal nucleus  shows  a  dense  chromatin  with  some  elec- 
tron-lucent areas  (Fig.  6C).  Then  the  chromatin  becomes 
more  homogeneous  in  mid  spermatids  (2.56  ±0.21  fim) 
(Figs.  5A,  6D),  and  eventually  in  late  spermatids  (1.81 
±0.32  /iml  condenses  into  a  coarse  granular  pattern, 
whereas  the  nucleus  assumes  an  ovoid  shape  and  forms 
a  basal  indentation  over  the  proximal  segment  of  the 
axoneme  (Fig.  6E).  Cytoplasmic  changes  involve  elonga- 
tion of  the  flagellum,  reduction  of  the  cytoplasmic  mass, 
and  coalescence  of  the  mitochondria  into  a  few  large 
spherical  units  located  around  the  proximal  portion 
of  the  axoneme.  Rotation  of  the  nucleus  does  not  take 
place  during  spermiogenesis,  therefore  the  flagellum 
axis  remains  parallel  to  the  base  of  the  nucleus  and 
the  spermatozoon  shows  the  typical  ultrastructure  of 
teleostean  type-II  sperm  (Fig.  6F). 


Discussion 

Histologically,  the  bluefin  tuna  testis  is  of  the  unre- 
stricted spermatogonial  testicular  type  found  in  most 
teleosts,  where  spermatogonia  occur  along  the  greater 
part  of  the  testicular  tubules.  In  the  restricted  sper- 
matogonial testicular  type  of  the  atheriniforms,  on  the 
other  hand,  the  spermatogonia  are  confined  to  the  distal 
end  of  the  tubules,  and  spermatogenesis  proceeds  as  the 
germ  cells  approach  the  efferent  ducts  (Grier  et  al.,  1980; 
Grier,  1981).  Efferent  ducts  are  generally  absent  in  unre- 
stricted spermatogonial  testes,  so  that  germinal  cysts 


form  along  the  testicular  tubule  length  (Grier  et  al., 
1980;  Grier,  1981;  Lahnsteiner  et  al.,  1994).  However, 
in  maturing  and  spawning  bluefin  tuna  a  well-developed 
network  of  ducts  collects  the  sperm  produced  by  the  ger- 
minal epithelium  and  voids  them  into  the  main  sperm 
duct.  The  central  ducts  of  the  testis  are  continuous  with 
the  proliferative  segment  of  the  testicular  lobules,  which 
lose  the  germinal  epithelium  in  the  innermost  region 
of  the  testis  and  function  as  sperm  storage  structures. 
This  process  has  been  documented  in  the  common  snook 
iCentropomus  undecimalis)  (Grier  and  Taylor,  1998),  the 
cobia  (Rachycentrum  canadum)  (Brown-Peterson  et  al.. 
2002),  and  the  swamp  eel  (Synbranchus  marmoratus) 
(Lo  Nostro  et  al.,  2003).  Grier  et  al.  (1980)  showed  that 
in  the  atheriniform  Fundulus  grandis  the  efferent  duct 
wall  cells  derive  from  Sertoli  cells.  A  system  of  efferent 
ducts  has  been  described  in  other  species  of  teleosts  pos- 
sessing testes  of  the  unrestricted  spermatogonial  type 
(Rasotto  and  Sadovy,  1995;  Manni  and  Rasotto,  1997). 
As  has  been  shown  in  other  species  of  the  genus  (Ratty 
et  al.,  1990;  Schaefer,  1996;  1998),  the  main  sperm  duct 
of  T.  thynnus  has  a  thick  wall  and  is  located  near  the 
center  of  the  testis,  whereas  in  many  other  teleosts  the 
main  duct  is  dorsal  (Grier  et  al.,  1980). 

Ultrastructural  features  of  bluefin  tuna  spermatogen- 
esis are  comparable  to  those  described  extensively  in 
teleosts  (for  examples  of  recent  literature  see  Gwo  and 
Gwo,  1993;  Stoumboudi  and  Abraham,  1996;  Quagio- 
Grassiotto  et  al.,  2001;  Huang  et  al.,  2002;  Koulish  et 
al.,  2002;  Fishelson,  2003).  The  primary  spermatogonia 
are  the  largest  male  germ  cells  and  exhibit  several 
conspicuous  perinuclear  ("nuage")  bodies.  After  several 
divisions  they  give  rise  to  cysts  of  secondary  spermato- 
gonia that  enter  meiosis  to  produce  successively  primary 
and  secondary  spermatocytes.  Primary  spermatocytes 
are  abundant,  particularly  at  the  pachytene  phase,  and 
are  therefore  thought  to  be  of  long  duration.  In  contrast, 
the  spermatocyte-II  stage  is  thought  to  be  the  shortest 
spermatogenetic  step,  because,  as  occurs  in  teleosts  in 
general,  it  is  the  least  frequent  in  histological  samples. 
Spermiogenesis  develops  without  the  occurrence  of  rotation 
of  the  spermatid  nucleus,  resulting  in  a  teleostean  type-II 
spermatozoon  (Mattei,  1970),  in  which  the  flagellar  axis 
lies  tangential  to  the  nucleus  instead  of  being  inserted 
perpendicular  to  its  base  (Abascal  et  al.,  2002). 

Santamaria  et  al.  (2003)  divided  the  testicular  cycle  of 
T.  thynnus  caught  in  Mediterranean  waters  from  Febru- 
ary to  September  into  five  periods.  Those  developmental 
stages  are  similar  to  stages  2-6  classified  by  Grier  (1981) 
for  a  generalized  teleost  annual  reproductive  cycle.  Most 
probably,  stage  1  (spermatogonial  proliferation)  occurs 
in  Mediterranean  bluefin  tuna  between  October  and 
January.  More  recently,  annual  histological  changes  in 
the  germinal  epithelium  have  been  used  to  identify  five 
distinct  reproductive  classes  in  males  of  several  teleost 
species  (Grier  and  Taylor,  1998;  Taylor  et  al.,  1998: 
Brown-Peterson  et  al.,  2002;  Lo  Nostro  et  al.,  2003).  It 
is  assumed  that  the  most  advanced  maturation  classes  in 
males  are  characterized  by  the  presence  of  a  discontinu- 
ous germinal  epithelium.  According  to  this  criterion,  all 


414 


Fishery  Bulletin  102(3) 


Figure   6 

Electron  micrographs  of  spermatocytes  I  (A),  spermatocytes  II  (B),  early  spermatids  (C),  mid 
spermatids  (D),  late  spermatids  (E),  and  spermatozoon  (F)  from  bluefin  tuna  {Thunnus  thynnus). 
Arrows  =  synaptonemal  complexes;  arrowheads  =  cytoplasmic  bridges  between  spermatids;  ax  = 
axoneme;  c  =  centriole;  cc  =  cytoplasmic  canal;  d  =  diplosome;  dc  =  distal  centriole;  f  =  flagellum; 
Gc  =  Golgi  complex;  m  =  mitochondria;  N  =  nucleus;  pc  =  proximal  centriole. 


of  the  samples  examined  in  the  present  study  correspond 
to  the  mid-  and  late-maturation  stages  proposed  by  Grier 
and  Taylor  (1998),  and  Taylor  et  al.  (1998).  Testes  at 
these  stages  become  storage  organs  that  are  filled  with 
sperm.  The  present  study  encompasses  only  a  short 
period  of  the  reproductive  cycle,  which  comprises  final 
gonad  maturation.  However,  descriptions  of  the  testicu- 
lar histology  throughout  the  annual  cycle  (Santamaria 


et  al.,  2003)  appear  to  indicate  that  different  maturation 
classes  might  be  defined  in  the  bluefin  tuna  based  on 
histological  examination  of  the  germinal  epithelium  (see 
Taylor  et  al,  1998;  Brown-Peterson  et  al.,  2002). 

Final  sexual  maturation  involves  a  considerable  in- 
crease in  testis  size,  but  no  apparent  remarkable  histo- 
logical changes,  with  the  exception  of  a  slightly  higher 
frequency  of  the  most  advanced  stages  of  spermatogen- 


Abascal  et  al.:  Testicular  development  in  Thunnus  thynnus 


415 


esis  in  fully  mature  bluefin  tuna.  The  different  testicular 
development  of  maturing  and  spawning  tuna  is  reflected 
by  their  respective  average  IG,  which  was  fourfold  higher 
in  spawning  fish.  An  equivalent  gonad  growth  was  found 
in  the  females  collected  in  the  same  samplings  (Medina 
et  al.,  2002),  indicating  a  spatiotemporal  parallelism  in 
the  gonad  maturation  cycle  and  a  good  synchronization 
of  the  reproductive  peak  in  the  two  sexes.  The  matura- 
tion schedule  differs  between  the  two  sexes,  however,  in 
that  males  are  capable  of  generating  mature  spermatozoa 
while  still  on  migration,  whereas  females  do  not  appear 
to  develop  fully  mature  oocytes  until  they  have  reached 
the  spawning  grounds  (Medina  et  al.,  2002).  Therefore, 
even  though  mature  spermatozoa  can  be  found  in  tes- 
ticular ducts  during  prolonged  periods  throughout  the 
reproductive  cycle,  it  is  unlikely  that  males  are  actually 
capable  of  spawning  out  of  reproductive  season. 

The  seasonal  IG  profile  of  the  bluefin  tuna  appears 
to  be  similar  to  that  of  the  pelagic,  highly  migratory 
perciform  Rachycentron  canadum  (Brown-Peterson  et 
al.,  2002),  and  the  swamp  eel  (Synbranchus  marmora- 
tus)  (Lo  Nostro  et  al.,  2003),  in  which  peak  Ic;  values 
occur  when  the  reproductive  activity  is  at  a  maximum. 
A  different  situation  has  been  reported  in  the  common 
snook  (Taylor  et  al.,  1998),  where  the  highest  IG  levels 
correspond  with  the  mid  maturation  class  and  decrease 
during  the  latter  part  of  the  reproductive  season.  The 
biological  significance  of  these  different  IG  profiles  in 
terms  of  reproductive  strategies  is  yet  unknown  because 
a  very  limited  number  of  species  have  been  examined 
so  far. 

Because  spermatozoa  are  by  far  the  most  abundant 
cells  in  mature  testes,  the  gonad  weight  becomes  a 
good  indicator  of  the  quantity  of  sperm  produced  by  a 
fish  (Billard,  1986).  Therefore,  the  significant  increase 
in  IG  that  occurred  between  samplings  off  Barbate  and 
the  Balearic  Islands  would  indicate  that,  during  migra- 
tion, bluefin  tuna  can  raise  several  times  the  volume  of 
sperm  accumulated  in  the  testes.  The  apparently  high 
spermatogenetic  activity  observed  in  bluefin  tuna  caught 
on  the  spawning  grounds  suggests  that  bluefin  tuna 
have  the  ability  to  regenerate  testicular  sperm  stores. 
Continuous  sperm  production  could  be  important  be- 
cause external  fertilization  requires  the  release  of  large 
amounts  of  sperm  to  ensure  successful  fertilization  of 
eggs,  especially  when  egg  size  is  small.  In  addition,  it 
should  be  noted  that  tunas  spawn  multiple  times  (June, 
1953;  Yuen,  1955;  Buriag,  1956;  Otsu  and  Uchida,  1959; 
Baglin,  1982;  Stequert  and  Ramcharrun.  1995)  and 
can  spawn  almost  daily  throughout  the  reproductive 
season  (Hunter  et  al..  1986;  McPherson,  1991;  Schaefer, 
1996,  1998,  2001;  Farley  and  Davis,  1998;  Medina  et 
al.,  2002). 

From  histological  examination  of  the  sperm  ducts, 
and  based  on  the  amount  of  sperm  present  and  the 
staining  of  the  epithelium,  Schaefer  (1998)  proposed  a 
spawning  interval  of  1.03  days  for  spawning  male  Thun- 
nus albacares  throughout  the  eastern  Pacific  Ocean. 
The  spawning  rate  estimated  for  reproductively  active 
females  with  the  postovulatory-follicle  method  was  1.19 


days  (Schaefer,  1998),  which  coincides  with  the  spawn- 
ing interval  estimated  for  female  T.  thynnus  around  the 
Balearic  Islands  (Medina  et  al.,  2002).  Unfortunately, 
we  could  not  make  a  reliable  estimation  of  the  male 
spawning  interval  in  our  samples.  Two  possible  reasons 
may  account  for  this  failure.  One  reason  is  that  many 
of  the  samples  of  gonadal  tissue  did  not  include  the 
main  sperm  duct.  On  the  other  hand,  no  clear  evidence 
of  spawning  was  identified  by  histological  examination 
of  those  specimens  processed  that  had  sperm  ducts.  A 
plausible  explanation  for  this  fact  is  that  recent  sperm 
release  can  be  detected  only  within  12  hours  after  the 
spawning  event  (Schaefer,  1996);  hence  for  male  spawn- 
ing to  be  detected  the  fish  would  have  to  be  sampled 
in  a  narrow  range  of  times  following  spawning,  which 
Schaefer  (1996)  established  between  00.01  and  12.00 
hours  after  spawning  for  Thunnus  albacares.  It  would 
be  worth  conducting  further  research  on  bluefin  tuna  at 
their  spawning  grounds,  by  attempting  to  cover  a  broad 
range  of  sampling  times  in  order  to  ensure  collection  of 
specimens  shortly  after  gamete  release.  In  this  way,  use- 
ful information  would  be  obtained  on  such  reproductive 
parameters  as  spawning  schedules,  fecundity,  and  the 
energy  cost  of  spawning,  which  are  essential  for  ecologi- 
cal assessments  of  the  reproductive  potential. 

It  is  noteworthy  that  male  tuna,  as  small  as  20  kg 
in  weight  (-100  cm  LF),  were  caught  on  the  spawning 
grounds  in  our  study.  They  had  gonad  indices  over  5% 
and  histological  features  indicative  of  full  maturity. 
These  observations  indicate  that  the  eastern  stock  of 
Atlantic  northern  bluefin  tuna  can  reach  maturity  at 
age  3  years  and  thus  support  conclusions  of  previous 
studies  (Rodriguez-Roda,  1967;  Hattour  and  Macias, 
2002;  Susca  et  al.,  2001a,  2001b;  Medina  et  al.,  2002); 
western  bluefin  tuna,  on  the  other  hand,  mature  at  an 
older  age,  which  has  been  estimated  at  6  years  (Baglin, 
1982). 

Prior  to  sexual  maturation,  marine  fish  generally 
accumulate  large  lipid  deposits,  primarily  triacylgly- 
cerols,  which  are  subsequently  mobilized  to  support 
gonad  development  and  spawning  migration  (Bell,  1998). 
The  major  lipid  storage  sites  are  the  mesenteric  tissue, 
muscle,  liver,  and  subdermal  fat  layers  (Ackman,  1980). 
In  bluefin  tuna  the  liver  does  not  appear  to  play  an 
important  role  in  lipid  storage  but  is  mainly  involved 
in  processing  fatty  acids  mobilized  from  other  bodily 
sources  (Mourente  et  al„  2002).  This  metabolic  pattern 
is  consistent  with  our  observations  of  weight  modifica- 
tions for  liver  and  fat  body  from  maturation  through  the 
spawning  period.  Although  IL  increases  only  slightly 
with  sexual  maturation,  IF  undergoes  a  marked  decrease 
at  the  time  of  maximum  gonad  development.  Thus,  the 
regression  analysis  of  the  relationship  between  IG  and  IF 
shows  a  significant  negative  correlation,  which  reveals  a 
depletion  of  mesenteric  fat  stores  as  the  testes  grow.  The 
occurrence  of  a  similar  situation  in  females  (Medina  et 
al.,  2002;  Mourente  et  al.,  2002)  and  in  male  and  female 
Thunnus  alalunga  (Ratty  et  al.,  1990)  has  led  to  the  con- 
clusion that  fat-body  lipid  reserves  provide  an  important 
energy  source  for  gametogenesis  in  tunas. 


416 


Fishery  Bulletin  102(3) 


Acknowledgments 

This  study  has  been  funded  by  the  Spanish  government 
and  the  European  Union  (projects  1FD1997-0880-C05- 
04  and  Q5RS-2002-01355).  The  authors  wish  to  thank 
two  anonymous  reviewers  for  helpful  recommendations. 
We  also  thank  Pesquerias  de  Almadraba,  S.  J.  and 
Gines  J.  Mendez  Alcala  (G.  Mendez  Espana,  S.  L.)  for 
co-operation  and  assistance.  The  invaluable  technical 
assistance  of  Agustin  Santos  and  Jose  Luis  Rivero  is 
greatly  appreciated.  G.  Mourente  lent  useful  help  in  our 
discussion  on  lipids. 


Literature  cited 

Abascal,  F.  J..  A.  Medina,  C.  Megina,  and  A.  Calzada. 

2002.     Ultrastructure  of  Thunnus  thynnus  and  Euthynnus 
alletteratus  spermatozoa.     J.  Fish  Biol.  60:147-153. 
Acknian,  R.  G. 

1980.  Fish  lipids,  part  1.  In  Advances  in  fish  science  and 
technology  (J.  J.  Connell,  ed.l.  p.  86-103.  Fishing  News 
Books  Ltd.,  Farnham,  England. 

Baglin,  R.  E.,  Jr. 

1982.     Reproductive  biology  of  western  Atlantic  bluefin 
tuna.    Fish.  Bull.  80:121-134. 
Bell,  J.  C. 

1998.     Current  aspects  of  lipid  nutrition  in  fish  farming.     In 
Biology  of  farmed  fish  (K.  D.  Black  and  A.  D.  Picker- 
ing, eds.l,  p.  114-145.     Sheffield  Academic  Press  Ltd., 
Sheffield,  England. 
Billard,  R. 

1986.     Spermatogenesis  and  spermatology  of  some  teleost 
fish  species.     Reprod.  Nutr.  Develop.  26:877-920. 
Brown-Peterson,  N.  J.,  H.  J.  Grier,  and  R.  M.  Overstreet. 

2002.  Annual  changes  in  germinal  epithelium  determine 
male  reproductive  classes  of  the  cobia.  J.  Fish  Biol. 
60:178-202. 

Bunag,  D.  M. 

1956.     Spawning  habits  of  some  Philippine  tuna  based  on 
diameter  measurements  of  the  ovarian  ova.     Philipp.  J. 
Fish.  26:877-920. 
Farley,  J.  H.,  and  T.  L.  O.  Davis. 

1998.     Reproductive  dynamics  of  southern  bluefin  tuna, 
Thunnus  maccoyii.     Fish.  Bull.  96:223-236. 
Fishelson,  L. 

2003.  Comparison  of  testes  structure,  spermatogenesis, 
and  spermatocytogenesis  in  young,  aging,  and  hybrid  cich- 
lid  fish  (Cichlidae,  Teleostei).     J.  Morphol.  256:285-300. 

Garcia-Berthou,  E. 

2001.  On  the  misuse  of  residuals  in  ecology:  testing  regres- 
sion residuals  vs.  the  analysis  of  covariance.  J.  Anim. 
Ecol.  70:708-711. 

Grier,  H.J. 

1981.  Cellular  organization  of  the  testis  and  spermatogen- 
esis in  fishes.     Am.  Zool.  21:345-357. 

Grier,  H.  J.,  J.  R.  Linton.  J.  F  Leatherland,  and  V.  L.  De  Vlaming. 
1980.     Structural  evidence  for  two  different  testicular  types 
in  teleost  fishes.     Am.  J.  Anat.  159:331-345. 
Grier.  H.  J.,  and  R.  G.  Taylor. 

1998.     Testicular  maturation  and  regression  in  the  common 
snook.    J.  Fish  Biol.  53:521-542. 
Gwo,  J.-C,  and  H.-H.  Gwo. 

1993.     Spermatogenesis  in  the  black  porgy,  Acanthopagrus 


schlegeli  (Teleostei:  Perciformes:  Sparidaei.     Mol.  Reprod. 
Dev.  36:75-83. 
Hattour,  A.,  and  D.  Macias. 

2002.     Bluefin  tuna  maturity  in  Tunisian  waters:  a  prelimi- 
nary approach.     Col.  Vol.  Sci.  Pap.  ICCAT  (International 
Commission  for  the  Conservation  of  Atlantic  Tunas)  54: 
545-553. 
Huang,  J.-D.,  M.-F.  Lee,  and  C.-F.  Chang. 

2002.     The  morphology  of  gonadal  tissue  and  male  germ 
cells  in  the  protandrous  black  porgy,  Acanthopagrus 
schlegeli.     Zool.  Stud.  41:216-227. 
Hunter,  J.  R.,  B.  J.  Macewicz,  and  J.  R.  Sibert. 

1986.     The  spawning  frequency  of  skipjack  tuna,  Kat- 
suwonus  pelamis,  from  the  South  Pacific.     Fish.  Bull. 
84:895-503. 
June,  F.  C. 

1953.     Spawning  of  yellowfin  tuna  in  Hawaiian  waters. 
U.  S.  Fish  and  Wildlife  Service,  Fish.  Bull.  54:47-64. 
Koulish,  S.,  C.  R.  Kramer,  and  H.  J.  Grier. 

2002.  Organization  of  the  male  gonad  in  a  protogynous  fish 
(Teleostei:  Labridael.     J.  Morphol.  254:292-311. 

Lahnsteiner,  F.,  R.  A.  Patzner,  and  T  Weismann. 

1994.  Testicular  main  ducts  and  spermatic  ducts  in 
some  cyprinid  fishes  I.  Morphology,  fine  structure  and 
histochemistry.     J.  Fish  Biol.  44:937-951. 

Lee,  C.  L. 

1998.  A  study  on  the  feasibility  of  the  aquaculture  of  the 
southern  bluefin  tuna,  Thunnus  maccoyii,  in  Australia, 
92  p.  Agriculture  Fisheries  and  Forestry-Australia 
(AFFA),  Canberra,  Western  Australia. 

Lo  Nostro,  F.,  H.  Grier,  L.  Andreone,  and  G.  A.  Guerrero. 

2003.  Involvement  of  the  gonadal  germinal  epithelium 
during  sex  reversal  and  seasonal  testicular  cycling  in 
the  protogynous  swamp  eel,  Synbranchus  marmoratus 
Bloch  1795  (Teleostei,  Synbranchidae).  J.  Morphol. 
257:107-126. 

Manni,  L.,  and  M.  B.  Rasotto. 

1997.  Ultrastructure  and  histochemistry  of  the  testicular 
efferent  duct  system  and  spermiogenesis  in  Opistogna- 
thus  whitehurstii  (Teleostei,  Trachinoidei).  Zoomorphol. 
117:93-102. 

Mather,  F.  J.,  Ill,  J.  M.  Mason,  and  A.  C.  Jones. 

1995.  Historical  document:  life  history  and  fisheries  of 
Atlantic  bluefin  tuna.  NOAATech.  Memo.  NMFS-SEFSC 
370,  165  p. 

Mattei,  X. 

1970.     Spermiogenesecompareedespoissons.     In  Compara- 
tive spermatology  (B.  Baccetti,  ed.),  p.  57-69.     Academic 
Press,  New  York,  NY. 
McPherson,  G.  R. 

1991.     Reproductive  biology  of  yellowfin  tuna  in  the  east- 
ern Australian  fishing  zone,  with  special  reference  to  the 
north-western  Coral  Sea.     Austr.  J.  Mar.  Freshw.  Res. 
42:465-477. 
Medina,  A..  F.  J.  Abascal,  C.  Megina.  and  A.  Garcia. 

2002.  Stereological  assessment  of  the  reproductive  status 
of  female  Atlantic  northern  bluefin  tuna.  Thunnus  thyn- 
nus i  L.i,  during  migration  to  Mediterranean  spawning 
grounds  through  the  Strait  of  Gibraltar.  J.  Fish  Biol. 
60:203-217. 
Mourente,  G.,  C.  Megina.  and  E.  Diaz-Salvago. 

2002.     Lipids  in  female  northern  bluefin  tuna  (Thunnus 
thynnus  thynnus  L.I  during  sexual  maturation.     Fish 
Physiol.  Biochem.  24:351-363. 
Olson,  R.  J. 

1980.     Synopsis  of  biological  data  on  the  southern  hluefin 


Abascal  et  al.:  Testicular  development  in  Thunnus  thynnus 


417 


tuna,  Thunnus  maccoyii  (Castelnau,  1872)      Inter-Am. 
Trop.  Tuna  Comm.,  Spec.  Rep.  2:151-212. 
Otsu.  T.,  and  R.  N.  Uchida. 

1959.     Sexual  maturity  and  spawning  of  albacore  in  the 
Pacific  Ocean.     U.  S.  Fish  and  Wildlife  Service,  Fish.  Bull. 
59:287-305. 
Quagio-Grassiotto,  I.,  J.  N.  C.  Negrao,  E.  D.  Carvalho.  and 
F.  Foresti. 

2001.     infrastructure  of  spermatogenic  cells  and  sperma- 
tozoa in  Hoplias  malabaricus  (Teleostei,  Characiformes, 
Erythrinidae).     J.  Fish  Biol.  59:1494-1502. 
Rasotto.  M.  B..  and  Y.  Sadovy. 

1995.  Peculiarities  of  the  male  urogenital  apparatus  of 
two  grunt  species  (Teleostei,  Haemulidae).  J.  Fish  Biol. 
46:936-948. 

Ratty,  F.  J.,  R.  M.  Laurs,  and  R.  M.  Kelly. 

1990.     Gonad  morphology,  histology,  and  spermatogen- 
esis in  south  Pacific  albacore  tuna  Thunnus  alalunga 
(Scombridae).     Fish.  Bull.  88:207-216. 
Ravier.  C,  and  J.-M.  Fromentin. 

2001.     Long-term  fluctuations  in  the  eastern  Atlantic  and 
Mediterranean  bluefin  tuna  population.     ICES  (Interna- 
tional Council  for  the  Exploration  of  the  Sea).     J.  Mar.  Sci. 
58:1299-1317. 
Rodriguez-Roda.  J. 

1964.     Biologia  del  atiin,  Thunnus  thynnus  (L.),  de  la  costa 
sudatlantica  de  Espana.     Inv.  Pesq.  25:33-146. 
Rodriguez-Roda,  J. 

1967.     Fecundidad  del  atiin,  Thunnus  thynnus  iL.),  de  la 
costa  sudatlantica  de  Espana.     Inv.  Pesq.  31:33-52. 
Santamaria,  N.,  A.  Corriero,  S.  Desantis,  D.  Zubani,  R.  Gentile, 
V.  Sciscioli,  M.  de  la  Serna,  and  C.  R.  Bridges. 

2003.     Testicular  cycle  of  the  Mediterranean  bluefin  tuna 
[Thunnus  thynnus  L.).     Cah.  Opt.  Med.  60:183-185. 
Schaefer,  K.  M. 

1996.  Spawning  time,  frequency,  and  batch  fecundity  of 
yellowfin  tuna,  Thunnus  albacares,  near  Clipperton  Atoll 
in  the  eastern  Pacific  Ocean.     Fish.  Bull.  94:98-112. 

1998.  Reproductive  biology  of  yellowfin  tuna  (Thunnus 
albacares)  in  the  eastern  Pacific  Ocean.  Inter-Am.  Trop. 
Tuna  Comm.,  Bull.  21:201-272. 


2001.     Reproductive  biology  of  tunas,     in  Tuna:  physiology, 
ecology  and  evolution  (B.  A.  Block  and  E.  D.  Stevens,  eds.i, 
p.  225-270.     Academic  Press,  London. 
Sissenwine,  M.  P.,  P.  M.  Mace.  J.  E.  Powers,  and  G.  P.  Scott. 

1998.     A  commentary  on  western  Atlantic  bluefin  tuna 
assessments.     Trans.  Am.  Fish.  Soc.  127:838-855. 
Stequert,  B.,  and  B.  Ramcharrun. 

1995.  Lafecondite  dulistao iKatsuwonuspelamis)de  l'ouest 
de  I'ocean  Indien.     Aquat.  Living  Resour.  8:79-89. 

Stoumboudi,  M.  Th..  and  M.  Abraham. 

1996.  The  spermatogenetic  process  in  Barbus  longiceps, 
Capoeta  damaseina  and  their  natural  sterile  hybrid  (Tele- 
ostei, Cyprinidae).     J.  Fish  Biol.  49:458-468. 

Susca,  V.,  M.  Deflorio,  A.  Corriero,  C.  R.  Bridges,  and 
G.  De  Metrio. 

2000.  Sexual  maturation  in  the  bluefin  tuna  (Thunnus 
thynnus  I  from  the  central  Mediterranean  Sea.  In  Pro- 
ceedings of  the  6th  international  symposium  on  the 
reproductive  physiology  offish  iB.  Noberg,  O.  S.  Kjesbu, 
G.  L.  Taranger.  E.  Andersson,  and  S.  O.  Stefansson,  eds.). 
p.  105.  Fish  Symp.  99.  Department  of  Fisheries  and 
Marine  Biology.  Univ.  Bergen,  Bergen,  Norway. 
Susca,  V.,  A.  Corriero,  C.  R.  Bridges,  and  G.  De  Metrio. 

2001a.     Study  of  the  sexual  maturity  of  female  bluefin  tuna: 
purification  and  characterization  of  vitellogenin  and  its 
use  in  an  enzyme-linked  immunosorbent  assay.     J.  Fish 
Biol.  58:815-831. 
Susca,  V.,  A.  Corriero,  M.  Deflorio,  C.  R.  Bridges,  and  G.  De  Metrio. 
2001b.     New  results  on  the  reproductive  biology  of  the  blue- 
fin tuna  [Thunnus  thynnus)  in  the  Mediterranean.     Col. 
Vol.  Sci.  Pap.  ICCAT  52:745-751. 
Taylor,  R.  G,  H.  J.  Grier,  and  J.  A.  Whittington. 

1998.     Spawning  rhythms  of  common  snook  in  Florida. 
J.  Fish  Biol.  53:502-520. 
Yuen,  H.  S.  H. 

1955.     Maturity  and  fecundity  of  bigeye  tuna  in  the  Pacific. 
U.S.  Fish  and  Wildlife  Service  Special  Scientific  Report: 
Fisheries  150,  30  p.     Washington,  DC. 
Zar,  J.  H. 

1996.  Biostatistical  analysis.  662  p.  Prentice-Hall.  Upper 
Saddle  River,  NJ. 


418 


Abstract-From  1995  to  1998.  we  col- 
lected female  black  rockfish  (Sebastes 
mt'lanops)  off  Oregon  in  order  to 
describe  their  basic  reproductive  life 
history  and  determine  age-specific 
fecundity  and  temporal  patterns  in 
parturition.  Female  black  rockfish  had 
a  50^  probability  of  being  mature  at 
394  mm  fork  length  and  7.5  years-of- 
age.  The  proportion  of  mature  fish  age 
10  or  older  significantly  decreased  each 
year  of  this  study,  from  0.511  in  1996 
to  0.145  in  1998.  Parturition  occurred 
between  mid-January  and  mid-March, 
and  peaked  in  February.  We  observed  a 
trend  of  older  females  extruding  larvae 
earlier  in  the  spawning  season  and 
of  younger  fish  primarily  responsible 
for  larval  production  during  the  later 
part  of  the  season.  There  were  dif- 
ferences in  absolute  fecundity  at  age 
between  female  black  rockfish  with 
prefertilization  oocytes  and  female 
black  rockfish  with  fertilized  eggs; 
fertilized-egg  fecundity  estimates  were 
considered  superior.  The  likelihood 
of  yolked  oocytes  reaching  the  devel- 
oping embryo  stage  increased  with 
maternal  age.  Absolute  fecundity  esti- 
mates ( based  on  fertilized  eggs)  ranged 
from  299,302  embryos  for  a  6-year-old 
female  to  948,152  embryos  for  a  16- 
year-old  female.  Relative  fecundity 
(based  on  fertilized  eggs  I  increased 
with  age  from  374  eggs/g  for  fish  age  6 
to  549  eggs/g  for  fish  age  16. 


Maturity,  ovarian  cycle,  fecundity,  and 
age-specific  parturition  of  black  rockfish 
(Sebastes  melanops) 

Stephen  J.  Bobko 
Steven  A.  Berkeley 

Department  of  Fisheries  and  Wildlife 

Hatfield  Marine  Science  Center 

2030  SE  Marine  Science  Drive 

Oregon  State  University 

Newport,  Oregon  97365 

Present  address  (for  S.  A.  Berkeley,  contact  author):  Long  Marine  Laboratory 

University  of  California,  Santa  Cruz 

100  Shaffer  Rd 

Santa  Cruz,  California  95060 

E-mail  address  (for  S  A  Berkeley,  contact  author):  stevenab@cats.ucsc.edu 


Manuscript  submitted  13  March  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
30  March  2004  bj  the  Scientific  Editor. 

Fish   Bull.  102:418-429(2004). 


Many  fish  species  in  the  North  Pacific 
have  a  long  reproductively  active  life 
span,  which  increases  the  likelihood 
of  producing  offspring  during  peri- 
ods of  favorable  environmental  condi- 
tions. This  bet  hedging  reproductive 
strategy  reduces  the  impact  of  envi- 
ronmental variation  on  reproductive 
success  (Goodman,  1984;  Leaman 
and  Beamish,  1984;  Schultz,  1989). 
In  species  with  age-structured  spawn- 
ing schedules,  a  broad  age  distribu- 
tion will  maximize  the  length  of  the 
spawning  season.  The  more  protracted 
the  reproductive  period,  the  greater 
the  likelihood  that  some  spawning 
will  occur  during  conditions  favorable 
for  larval  survival  (Lambert,  1990). 
Age-related  differences  in  the  timing 
of  spawning  have  been  observed  in 
many  fishes;  usually  larger,  older  fish 
spawn  earlier  (Simpson,  1959;  Bage- 
nal,  1971;  Berkeley  and  Houde,  1978; 
Shepherd  and  Grimes,  1984;  Lambert, 
1987),  but  in  some  cases  younger  fish 
spawn  earlier  in  the  season  (Hutch- 
ings  and  Myers,  1993). 

Age  truncation,  an  inevitable  result 
of  fishing,  can  increase  recruitment 
variability  by  reducing  the  length  of 
the  spawning  season  or  by  selectively- 
removing  older,  more  fit  individuals 
from  the  population.  Factors  that 
might  affect  individual  reproductive 
success  include  the  number  of  eggs 
produced,  the  quality  of  eggs  (e.g., 
yolk  or  oil  globule  volume),  and  the 
size  or  health  of  eggs  and  larvae.  Off 


the  coast  of  Oregon,  widow  rockfish 
(Sebastes  entomelas)  have  exhibited 
increased  absolute  fecundity,  and 
more  importantly  have  increased  rela- 
tive fecundity,  with  age  (Boehlert  et 
al.,  1982).  Individual  populations  of 
shortbelly  rockfish  (Sebastes  jordani), 
have  been  found  to  produce  larvae 
with  differing  lipid  and  protein  com- 
positions and  consequently  potentially 
differing  rates  of  survival  (MacFar- 
lane  and  Norton.  1999).  Zastrow  et 
al.  (1989)  reported  that  striped  bass 
eggs  stripped  from  wild  fish  increase 
in  quality  with  maternal  age  due  to 
increased  amounts  of  proteins  and 
lipids,  although  relative  concentra- 
tions remain  unchanged. 

Black  rockfish  ( Sebastes  melanops), 
like  most  other  rockfish,  are  long- 
lived,  moderately  fecund  livebearers 
with  long  reproductive  life  spans.  Al- 
though their  longevity  and  low  rate 
of  natural  mortality  is  presumed  to 
be  an  adaptation  to  allow  success- 
ful reproduction  over  their  lifespan 
despite  long  periods  between  favor- 
able environmental  conditions,  it  al- 
so makes  them  more  susceptible  to 
overexploitation.  The  objective  of  our 
research  presented  in  the  present 
article  is  twofold.  First,  we  describe 
the  basic  reproductive  life  history 
of  black  rockfish.  with  an  emphasis 
on  the  ovarian  developmental  cycle 
and  maturity  schedule.  Second,  we 
investigate  age-specific  fecundity  and 
temporal  patterns  in  parturition  and 


Bobko  and  Berkeley:  Maturity,  ovarian  cycle,  fecundity,  and  parturition  of  Sebastes  melanops 


419 


discuss  their  effect  on  reproductive  success  in  a 
population  undergoing  truncation  of  the  upper  end 
of  its  age  distribution. 


Materials  and  methods 

We  collected  female  black  rockfish  during  the 
months  of  peak  female  reproductive  development, 
November  through  March,  for  three  successive 
years  from  1995-96  through  1997-98.  Female  black 
rockfish  were  primarily  obtained  from  recreational 
charter  boat  landings  in  Newport,  Depoe  Bay,  and 
Charleston,  Oregon,  in  addition  to  some  fish  from 
commercial  landings  from  Port  Orford,  Oregon 
(Fig.  1).  We  also  collected  fish  by  rod  and  reel  and 
spearfishing.  When  possible,  the  sex  of  all  avail- 
able black  rockfish  was  determined,  and  females 
were  staged  as  immature  or  mature.  Immature 
females  were  measured  (FL),  and  mature  females 
were  returned  to  the  laboratory.  On  extremely  busy 
days  when  numerous  charter  boats  were  fishing, 
all  mature  females  were  collected,  but  immature 
fish  were  not  measured.  In  total  we  collected  1643 
female  black  rockfish.  Immediately  upon  return 
to  the  laboratory,  we  recorded  fork  length,  total 
weight  when  possible  (most  samples  from  charter 
boats  were  carcasses  only),  liver  weight,  and  ovary 
weight.  Ovaries  were  assigned  a  maturity  stage 
based  on  macroscopic  appearance  and  preserved  in 
10%  buffered  formalin.  We  initially  followed  the  gross 
maturity  stage  scheme  of  Nichol  and  Pikitch  (19941  for 
darkblotched  rockfish  ( Sebastes  crameri)  but  ultimately- 
abandoned  their  classification  of  maturity  stages  in  favor 
of  the  simplified  maturity  stages  reported  by  Gunderson 
et  al.  (1980)  (Table  1).  Sagittal  otoliths  were  removed 
and  stored  dry  for  age  determination.  All  aging  was  done 
by  an  expert  age  reader  from  the  Oregon  Department  of 
Fish  and  Wildlife  (ODFW),  who  used  the  break-and-burn 
technique  (Beamish  and  Chilton,  1982).  Ten  percent  of 
the  otoliths  were  randomly  selected  for  a  second  read- 
ing to  ensure  consistency  in  interpretation  of  annuli. 
It  should  be  noted  that  black  rockfish  ages  have  not 
been  validated.  However,  ages  have  been  validated  for 
yellowtail  rockfish  (Sebastes  flavidus),  a  closely  related 
species  (Leaman  and  Nagtegaal,  1987),  by  using  otoliths; 
moreover,  the  break-and-burn  aging  method  is  widely 
accepted  as  valid  for  aging  rockfish  (MacLellan,  1997), 
and  ages  thus  derived  are  routinely  used  in  rockfish 
stock  assessments.  Because  our  sample  included  all 
mature  females  that  we  encountered,  we  used  these  data 
to  estimate  the  age  distribution  of  mature  females  in 
each  time  period,  and  the  age  distribution  of  parturition 
during  each  time  interval. 

Histological  preparations  were  made  from  the  ova- 
ries of  175  females  collected  monthly  from  March  1996 
through  March  1997  to  track  seasonal  ovarian  devel- 
opment. Females  collected  in  March  1996  and  Novem- 
ber 1996  through  March  1997  were  from  our  regular 
sampling  program,  whereas  fish  collected  from  April 


45°  00'N  " 


42°  50'  ' 


124°  30'  120°00'W 

Figure  1 

Map  of  the  Oregon  coast  showing  the  study  area  where  black 
rockfish  (Sebastes  melanops)  were  collected. 

through  October  1996  were  obtained  from  Newport 
recreational  charter  boat  landings.  Females  were  ran- 
domly selected  from  each  maturity  stage  observed  each 
month  and  from  as  wide  a  range  of  ages  as  available 
(Table  2).  Ovaries  were  embedded  in  paraffin,  sectioned 
at  4-5  jjm,  and  stained  with  gill-3  haematoxylin  and 
eosin  y  solution. 

We  determined  stage-specific  fecundity  in  black  rock- 
fish for  females  with  unfertilized  yolked  oocytes  (?i=184) 
and  fertilized  eggs  (n  =  85).  Postfertilization  ovaries  were 
very  fragile  and  tended  to  rupture  easily  and  release 
embryos  under  the  slightest  pressure.  Consequently, 
for  estimating  fecundity  for  these  stages,  we  used  only 
fish  collected  by  ourselves  so  that  we  were  certain  that 
no  eggs  or  larvae  had  been  released  during  capture. 
To  ensure  that  no  eggs  were  lost  after  capture,  these 
fish  were  immediately  placed  into  plastic  bags  in  order 
to  retain  any  eggs  that  might  be  extruded  before  the 
ovary  could  be  processed.  Ovaries  were  processed  fol- 
lowing procedures  modified  from  Lowerre-Barbieri  and 
Barbieri  (1993)  to  separate  eggs  and  embryos  from 
connective  tissue.  Briefly,  fixed  ovaries  were  manually 
manipulated  and  rinsed  with  water  through  a  1-mm 
square  mesh  sieve,  which  retained  most  of  the  con- 
nective tissue,  into  another  sieve  with  0.75-mm  mesh. 
Ovary  connective  tissue  was  retained  in  the  coarse 
sieve,  and  freed  eggs  were  collected  in  the  fine-mesh 
sieve.  Freed  eggs  were  patted  dry,  weighed  (nearest 
0.1  g),  and  three  subsamples  were  collected,  weighed 
(nearest  0.001  g),  and  placed  in  10%  buffered  formalin. 


420 


Fishery  Bulletin  102(3) 


Table  1 

Macroscopic  and  histological  descriptions  of  stages  used  to 


describe  female  black  rockfish  maturity. 


Maturity  stage 


Macroscopic  description 


Histological  description 


1     Immature 


Small  and  translucent  ovary,  pink  during  months 
without  sexual  activity  and  yellowish  (except  for 
very  small  fish)  during  months  with  reproductive 
activity. 


2  Vitellogenesis  Ovary  firm  and  yellow  or  occasionally  cream  in 
color.  Large  range  of  size,  but  all  with  visible  opaque 
eggs. 


3     Fertilization 


4     Eyed  larvae 


5     Spent 


6    Resting 


Eggs  are  golden  and  translucent.  Ovary  extremely 
large  in  relation  to  body  cavity.  Ovary  wall  thin  and 
easily  torn. 

Eyes  of  developing  embryos  visible,  giving  ovary 
an  overall  greyish  color.  Ovary  fills  a  large 
portion  of  the  body  cavity. 

Ovary  flaccid,  purplish-red  in  color.  Eyed  larvae 
may  still  be  visible. 

Ovary  again  firm  and  pink  in  color.  Black  spots  may 
be  visible. 


Oocyte  cytoplasm  intensely  basophilic.  Densely 
packed  oogonial  nests  and  developing  oocytes,  with 
larger  oocytes  containing  small  clear  vesicles. 


Oogonia  and  developing  oocytes  still  visible,  but 
ovary  dominated  by  large  oocytes  with  numerous 
small  red-staining  yolk  globules. 

Fertilized  eggs  ovulated  and  found  within  the 
ovarian  cavity.  Eggs  have  a  single  pink-staining 
yolk  mass  and  clear  oil  droplet. 

Presence  of  developing  larvae  with  black  pigmented 
eyes.  Yolk  mass  absorbed  in  late-stage  larvae,  but 
oil  droplet  usually  present. 

Early-stage  oocytes  loosely  associated.  Extensive 
network  of  blood  vessels.  Possibility  of  encountering 
residual  larvae. 

Similar  appearance  to  immature  fish.  Ovary  wall 
slightly  thicker  in  early  summer. 


Table  2 

Monthly  ranges  for  age,  length,  and  maturity  stage  of 
black  rockfish  collected   off  Oregon   from   March   1996 
through  March  1997  for  histological  analysis. 

Month 

Age  (yr 
range 

FL(mm) 
range 

Maturity  stage 
range 

n 

March 

7-25 

375-510 

1.4-6 

10 

April 

7-18 

364-447 

1.5-6 

12 

May 

7-13 

340-465 

1  and  6 

15 

June 

5-13 

349-432 

1  and  6 

15 

July 

5-13 

360-475 

1  and  6 

14 

August 

5-11 

357-493 

1-2,6 

15 

September 

6-16 

366-488 

1  and  2 

12 

October 

5-16 

357-420 

1  and  2 

11 

November 

5-11 

355-434 

1  and  2 

16 

December 

5-14 

365-439 

1  and  2 

10 

January 

6-17 

369-473 

1-4 

16 

February 

7-17 

378-464 

1-5 

17 

March 

6-13 

380-467 

1,5-6 

12 

AF  =  EW 


The  number  of  ova  in  the  subsamples  were  counted  and 
absolute  fecundity  was  estimated  by  using  the  following 
algorithm: 


issc 


ssw 


where  AF  =  absolute  fecundity,  or  the  total  number  of 
eggs  per  female; 
EW  =  weight  of  rinsed  eggs  (or  larvae); 
SSCt   =  subsample  count  i,  where  j=l  to  3;  and 
SSWt   =  subsample  weight  i,  where  £=1  to  3. 

Relative  fecundity  (RF),  based  on  gonad-free  somatic 
weight  was  estimated  by 

RF  =  —  —. 

TW-GW 

where  AF  =  absolute  fecundity,  or  the  total  number  of 
eggs  per  female; 
TW  =  total  weight:  and 
GW  =  gonad  weight. 

For  our  analyses  of  fecundity,  we  used  only  fish  in  which 
the  number  of  eggs  or  larvae  estimated  from  the  three 
subsamples  had  coefficients  of  variation  less  than  or 
equal  to  59;,  and  for  prefertilization  eggs  we  used  only 
females  with  average  egg  diameters  of  at  least  450  jj 
to  ensure  inclusion  of  all  developing  oocytes.  Only  one 
cohort  of  developing  oocytes  is  present  in  the  ovary  of 


Bobko  and  Berkeley:  Maturity,  ovarian  cycle,  fecundity,  and  parturition  of  Sebastes  melanops 


421 


5n  -i 

Maturity  stage  2  •  Vitellogenic 

40  - 

Age  8 

Collected  February  1996 

30  - 

10  - 
o  - 

i            i 

' 1 

L     1 

1 — i — 

L     i          i          i          i          i 

50 

40 
30 
20 
10 
0 


0.50  0  55  0  60  0  65  0  70  0  75  0.80  0.85  0.90  0.95   1.00 

Maturity  stage  3  -  Fertilized 

Age  8 

Collected  February  1996 


-[ — ' — i — ' — i 1 1 — — ' — — i — — 

0  50     0  55     0  60     0.65     0.70     0  75     0  80     0  85     0  90     0  95      100 

Oocyte  or  egg  diameter  (mm) 

Figure  2 

Prefertilization-  and  fertilized-egg-diameter  frequency  distributions  show- 
ing a  single  mode  of  developing  oocytes  at  both  developmental  stages. 


black  rockfish,  either  during  development  (stage  2)  or 
after  fertilization  (stage  3)  (Fig.  2).  Analysis  of  covari- 
ance  (ANCOVA)  was  used  to  test  for  annual  effects  in 
the  relationship  between  prefertilization  fecundity  and 
age  and  a  maturity-stage  effect  (prefertilization  vs. 
fertilized-egg  development  stages)  on  both  absolute  and 
relative  fecundity  at  age.  We  also  used  ANCOVA  to  test 
for  a  maturity  stage  effect  in  the  relationship  between 
absolute  fecundity  and  fork  length.  All  ANCOVA  analy- 
ses were  conducted  by  using  multiple  linear  regression 
with  the  function  lm  in  S-PLUS  2000  (MathSoft.  Inc.. 
Seattle,  WA). 

To  predict  the  probability  of  a  female  black  rockfish 
being  mature  based  on  its  fork  length,  we  fitted  our 
maturity-at-length  data  to  a  logistic  regression.  Dur- 
ing those  months  without  reproductive  activity,  late 
spring  through  early  fall,  it  was  difficult  to  distinguish 
between  immature  and  mature-resting  ovaries.  Conse- 
quently, only  those  females  collected  during  the  peak 
months  of  reproductive  development  and  from  sampling 
events  where  all  fish,  mature  and  immature,  were  col- 
lected were  included  in  our  analysis.  Binary  maturity 
observations  (0=immature,  l=mature)  and  fork  length 
were  fitted  to  a  logistic  model  by  using  the  function 
glm,  family  =  binomial  of  S-PLUS  (S-PLUS  2000).  The 
model  used  was 


S0+/Jl«. 


mFL)=P(Y  =  \\FL)-- 


\  +  e 


00+010.  ■ 


where  P(Y=1\FL) 


probability  of  female  black  rockfish 
being  mature  at  size  FL;  and 


/30  and  P{  =  regression  coefficients  for  the  inter- 
cept and  fork  length,  respectively. 

For  functional  purposes,  the  response  variable  was 
interpreted  as  the  percentage  of  female  black  rockfish 
mature  at  length.  Assuming  this  relationship  of  fork 
length  to  maturity  had  not  changed  over  time,  we  ap- 
plied our  logistic  model  to  fork-length  data  from  random 
sampling  conducted  by  ODFW  during  the  summers  of 
1992-2000  to  calculate  the  percentage  of  female  black 
rockfish  caught  by  the  recreational  fishery  off  Newport 
that  were  mature  in  each  year. 

Fork  length-at-age  data  for  female  black  rockfish  were 
fitted  with  the  von  Bertalanffy  growth  function  (VB- 
GF)  by  using  the  nonlinear  function  nisi )  in  S-PLUS 
2000.  Age  at  50%  maturity  was  calculated  by  using 
our  estimate  of  length  at  509i  maturity  and  a  VBGF 
rearranged  to  the  form 


^U',  '()    ~*~  ' 


where   t 


L 


|9j  =  age  at  50%  maturity; 
,,=  =  asymptotic  length; 
k  =  Brody  growth  parameter; 
t0  =  age  at  zero  length;  and 
._  =  length  at  50%  maturity. 


Timing  of  parturition  was  estimated  by  microscopi- 
cally determining  embryo  development  stages  for  all 
females  with  fertilized  eggs  following  Yamada  and 


422 


Fishery  Bulletin  102(3) 


Kusakari's  (1991)  stages  of  embryonic  development  for 
kurosoi  (Sebastes  schlegeli)  modified  to  reflect  the  gesta- 
tion period  of  37  days  for  black  rockfish  (Boehlert  and 
Yoklavich,  1984).  Gestation  period  is  likely  to  vary  with 
water  temperature.  In  determining  gestation  period, 
Boehlert  and  Yoklavich  (1984)  held  black  rockfish  in 
the  laboratory  at  9-11  :,C.  Mean  water  temperatures 
in  our  study  area  during  the  period  of  egg  and  larval 
development  (December-April)  were  10.9°,  10.1°,  and 
11.4°C  in  1995-96,  1996-97,  and  1997-98,  respectively 
(http://co-ops.nos.noaa.gov/data).  Even  in  the  strong  El 
Nino  year  of  1997-98,  nearshore  water  temperature 
during  the  winter  larval  development  period  was  only 
slightly  outside  this  range.  Therefore,  we  assumed  a 
37-day  gestation  period  for  all  years  of  our  study.  Us- 
ing the  Boehlert  and  Yoklavich  (1984)  equation;  (stage 
duration  =  0.0452 xstage1  090)  we  solved  for  duration  at 
each  stage  by  adding  5  days  to  account  for  the  time 
between  hatching  (stage  32)  and  parturition  (also  from 
Boehlert  and  Yoklavich,  1984).  To  calculate  the  time 
until  parturition  for  each  stage,  we  subtracted  the  pre- 
vious stage  durations  from  the  total  gestation  period  of 
37.  For  example,  at  stage  1,  parturition  would  occur  in 
37  days.  At  stage  2,  parturition  would  take  place  in  37 
days  -  stage-1  duration  (-2  days)  =  35  days. 

For  each  year  of  our  study,  estimated  parturition 
dates  for  all  females  in  our  sample  were  grouped  into 
one-week  time  intervals  and  further  subdivided  into 
age  categories:  6-8:  9-11;  12-14;  and  >15.  These  num- 
bers were  then  multiplied  by  the  appropriate  value  for 
age-class-specific  fecundity  based  on  fertilized  eggs 
(Table  3)  to  estimate  relative  spawning  output  by  week 
for  each  age  class. 


Results 

Ovarian  development 

Black  rockfish  off  Oregon  exhibited  group-synchronous 
oocyte  development;  and  females  extruded  only  one 
brood  of  larvae  per  year  (Fig.  2).  Based  on  our  observa- 
tions of  ovarian  development  from  all  three  years  of 
this  study,  parturition  took  place  from  mid-January 
through  mid-March  and  peaked  in  February.  Following 
parturition,  unextruded  larvae  were  quickly  resorbed 
and  the  ovary  lost  much  of  its  vascularization.  From 
April  through  early  August  ovaries  were  in  a  resting 
state  and  contained  oogonial  nests  and  slightly  larger 
oocytes  with  a  basophilic  cytoplasm  and  a  maximum 
diameter  of  50  jjl.  Also  present  at  this  time  were  develop- 
ing oocytes  ranging  from  50  to  150  n  in  diameter  with 
small  lipid  vacuoles  surrounding  the  nuclear  membrane. 
Yolk  deposition  (vitellogenesis)  began  in  late  August 
and  was  observed  through  the  third  week  of  February. 
In  the  final  stages  of  vitellogenesis.  the  largest  oocytes 
were  approximately  700  n  in  diameter  and  had  numer- 
ous oil  vacuoles  and  yolk  globules  throughout  the  cyto- 
plasm. The  first  female  with  fertilized  eggs  (stage  3)  was 
observed  during  the  second  week  of  January,  and  stage-3 


Table  3 

Age  group-specific  absolute  fecundity  (based  on  fertilized 
eggs  l  and  age  distribution  of  mature  females  as  a  percent- 
age of  all  mature  females,  used  to  estimate  larval  pro- 
duction. Calculated  from  data  pooled  from  1996  through 
1998. 


Age  group 
(yr) 

Absolute 
fecundity' 

Percentage  of  all  mature 

females  represented 

by  each  age  group 

6-8 

9-11 
12-14 
15  and  older 

364,183.5 
558,837.1 
753,490.7 
948,144.3 

42.19 
38.48 
13.94 
5.39 

'  Absolute  fecundity  for  each  age  group  is  the  estimated  fecundity 
(based  on  fertilized  eggs)  for  ages  7.  10,  13,  and  16.  respectively. 


females  were  observed  until  the  third  week  of  February. 
Recently  fertilized  eggs  were  approximately  850  ,u  in 
diameter.  The  period  of  parturition  as  indicated  by  the 
occurrence  of  ovaries  containing  eyed  larvae  extended 
from  the  second  week  in  January  through  the  second 
week  of  March.  Spent  females  were  first  collected  during 
the  last  week  of  January  and  were  most  frequently  col- 
lected in  late  February  and  early  March. 

Sexual  maturity 

Parameter  values  for  the  length-maturity  logistic  model 
were  ft,  =  -26.73  and  ^  =  0.068.  The  smallest  mature 
female  black  rockfish  we  observed  was  345  mm;  all 
individuals  were  mature  by  450  mm.  Fifty  percent  of 
females  were  estimated  to  be  mature  at  394  mm  fork 
length  (Fig.  3).  As  reflected  in  our  length-maturity  logis- 
tic model,  there  was  a  decreasing  trend  in  the  percent 
maturity  for  female  black  rockfish  in  recreational  land- 
ings from  ODFW  collections  from  1992  through  2000 
(Fig.  4).  The  von  Bertalanffy  parameter  estimates  for 
female  black  rockfish  were  Lre  =  442  mm,  k  =  0.33,  t0  = 
0.75  (Fig.  5).  Using  these  estimates,  along  with  the  fork 
length  at  509c  maturity,  we  estimated  the  age  at  50% 
maturity  for  female  black  rockfish  to  be  7.5  years.  The 
median  age  of  mature  females  decreased  in  each  col- 
lection year  from  10  years  in  1996  to  9  in  1997  and  to 
7  years  in  1998.  In  addition,  we  observed  a  significant 
decrease  in  the  proportion  of  mature  fish  age  10  or  older 
over  the  three  years  of  our  study  (Pearson's  ^2  =  52.4, 
df=2,  P<0.001).  The  proportions  decreased  from  0.511 
in  1996,  to  0.318  in  1997,  and  0.145  in  1998. 

Fecundity 

Absolute  fecundity  for  prefertilization  female  black 
rockfish  ranged  from  482,528  oocytes  for  a  5-year-old 
female  to  998,050  oocytes  for  a  19-year-old  female. 
The  results  of  ANCOVA  (Table  4)  over  a  common  age 


Bobko  and  Berkeley:  Maturity,  ovarian  cycle,  fecundity,  and  parturition  of  Sebastes  melanops 


423 


100 

e-26  731+0  068-FL 

•^ — • — • — • 

P(Y= 

UFL)  = 

•  / 

1    +  g-26  731+0  068-FL 

80 

•     /                • 
/            • 

60 

- 

•/ 

40 

- 

20 

- 

*^% 

U 

I 

I                              I                              I 

250 


300 


350  400 

Fork  length  (mm) 


450 


500 


Figure  3 

Logistic  regression  model  for  the  estimated  percentage  of  sexually  mature 
female  black  rockfish  as  a  function  of  fork  length,  with  associated  observed 
percent  mature  at  10-mm  length  intervals. 


range  showed  no  evidence  of  differences  in 
slopes  among  the  years  1996-98  (P=0.161). 
ANCOVA  also  showed  no  significant  difference 
in  elevations  (P=  0.632),  indicating  no  annual 
effect  and  allowing  one  model  to  be  fitted  to 
the  pooled  data  (Fig.  6).  Absolute  fecundity 
for  females  with  fertilized  eggs  ranged  from 
299,302  embryos  for  a  6-year-old  to  948.152 
embryos  for  a  16-year-old.  Because  of  the  low 
number  of  females  with  developing  embryos 
collected  in  1996  and  1998,  19  and  4  females, 
respectively,  and  based  on  the  results  of  prefer- 
tilization  females,  all  data  were  pooled  and 
fitted  with  one  model  (Fig.  7).  Although  we 
were  able  to  pool  the  data  for  all  years  for 
fecundity-age  regressions  for  both  prefertiliza- 
tion  females  and  fertilized  females,  there  was 
evidence  of  interaction  (i.e.,  unequal  slopes) 
between  stage-specific  absolute  fecundity  and 
age  (2-tailed  t-test,  P=0.020)  requiring  sepa- 
rate linear  regressions  to  be  fitted  to  the  data 
(Fig.  8). 

Similar  to  the  ANCOVA  results  for  abso- 
lute fecundity,  there  were  no  differences  in 
slopes  or  elevations  for  relative  fecundity  for 
prefertilization  females  for  the  years  1996-98 
(Table  4).  Again,  based  on  the  results  of  the 
ANCOVA  for  prefertilization  females  and  due 
to  the  low  number  of  fertilized  females  collected 
and  1998,  all  relative  fecundity  data  for  femal 
fertilized  eggs  were  pooled.  Unlike  the  results 


• 

50" 

• 

• 

40" 

• 

• 

• 

• 

30" 

1 

1 

• 

• 
1 

1992 


1994 


1996 
Year 


1998 


2000 


Figure  4 

Estimated  percent  maturity  for  recreationally  landed  female  black 
rockfish  from  Newport,  Oregon,  based  on  our  logistic  regression  model 
of  fork  length  on  maturity.  Data  were  from  regular  random  summer 
port  sampling  conducted  by  the  Oregon  Department  of  Fish  and  Wild- 
life from  1992  through  2000. 


in  1996  relation  between  absolute  fecundity  and  age  there  was 

es  with  no  evidence  of  interaction  (i.e.,  unequal  slopes)  between 

for  the  stage-specific  relative  fecundity  and  age  (2-sided  f-test 


424 


Fishery  Bulletin  102(3) 


Table  4 

Results  of  analyses  of  covariance  testing  for  differences  in 

slopes  and 

elevations  of  annua 

1  absolute  fecundity-age  relati 

on  and 

annual  relative  fecundity-age  relation.  Response  variables 

=  AF  and  RF.  treatment  factors 

=  year,  and  covariate 

=  age. 

Source  of  variation 

df 

Sum  of  squares 

Mean  square 

F 

P 

Absolute  fecundity  i  based  on  prefertilization  oocytes) 

Equality  of  slopes 

2 

105.347 

52,674 

1.85 

0.161 

Error 

160 

4,564.320 

28,527 

Equality  of  elevation 

2 

168,544 

84,272 

0.29 

0.747 

Error 

162 

3,459,048 

288,254 

Relative  fecundity  i based  on  prefertilization  oocytes) 

Equality  of  slopes 

2 

259.04 

129.52 

0.58 

0.559 

Error 

160 

35,452.22 

221.58 

Equality  of  elevation 

2 

799.52 

399.76 

1.81 

0.166 

Error 

162 

35.711.26 

220.44 

300 


200  " 


100 


L  =  442.02  •(  1  -e1"  °  33  '(a9e"(  °  75 1") 


10  15  20 

Age  (years) 


Figure  5 

Fork  length  at  age  fitted  to  the  von  Bertalanffy  growth  model  for  female 
black  rockfish. 


(2-tailed  i-test,  P<0.0001,)  of  a  stage  effect 
(i.e.,  unequal  elevations)  which  necessitated 
that  the  data  be  fitted  with  a  parallel-line 
multiple  linear  regression  model  (Fig.  10 1. 

Temporal  patterns  in  parturition 

From  1996  through  1998  we  estimated  rela- 
tive larval  production  for  four  age  groups: 
6-8;  9-11;  12-14:  and  15  years  and  older 
(Fig.  11).  In  each  year  parturition  took  place 
from  mid-January  until  mid-March,  and 
older,  larger  fish  extruded  larvae  earlier 
than  younger  fish.  In  1996  and  1997,  the 
9-11  year-old  fish  dominated  larval  pro- 
duction, responsible  for  60. lQ  and  49. 6^  of 
all  larvae  extruded,  respectively  (Table  5). 
In  1998  age  6-8  fish  produced  the  largest 
percentage  of  larvae  (65.3%).  In  all  years, 
relative  larval  production  was  lowest  for  the 
oldest  age  group  (15+),  declining  to  near  0 
by  1998. 


Discussion 


P=  0.096).  There  was,  however,  strong  evidence  (2-tailed 
/-test,  P<0.001,)  of  a  stage  effect  li.e.,  unequal  elevations) 
which  necessitated  that  the  data  be  fitted  with  a  paral- 
lel-lines multiple  linear  regression  model  (Fig.  9). 

Absolute  fecundity  for  prefertilization  female  black 
rockfish  ranged  from  443.671  oocytes  for  a  381-mm-FL 
female  to  1,135,457  oocytes  for  a  495-mm-FL  female. 
For  fertilized  females,  absolute  fecundity  ranged  from 
283,618  oocytes  for  a  381-mm-FL  female  to  1,073,356 
oocytes  for  a  510-mm-FL  female.  The  results  of  AN- 
COVA  over  a  common  size  range  showed  no  evidence  of 
differences  in  slopes  between  maturity  stages  (2-sided 
/-test  P=0.206).  There  was,  however,  strong  evidence 


Ovarian  development  for  black  rockfish  in  Oregon  was 
similar  to  the  developmental  cycles  reported  for  other 
rockfish  species  (Moser,  1967;  Bowers,  1992;  Nichol  and 
Pikitch,  1994)  with  the  exception  of  seasonal  timing 
and  stage  duration.  Females  underwent  vitellogenesis 
for  up  to  six  months  before  fertilization,  which  occurred 
from  December  through  February.  In  all  three  years, 
parturition  off  the  Oregon  coast  occurred  between  mid- 
January  and  mid-March  and  peaked  in  February.  Wyllie 
Echeverria  (1987)  observed  similar  timing  for  parturi- 
tion of  black  rockfish  off  north-central  California,  with  a 
peak  in  February  but  with  parturition  occurring  through 
May. 


Bobko  and  Berkeley:  Maturity,  ovarian  cycle,  fecundity,  and  parturition  of  Sebastes  melanops 


425 


All  female  black  rockfish.  except  the 
smallest  immature  females,  followed  a  sea- 
sonal cycle  in  which  their  ovaries  developed 
an  orange  coloring  during  the  months  of 
reproductive  activity — a  pattern  observed 
in  olive  rockfish  (Love  and  Westphal,  1981). 
Similarly,  Nichol  and  Pikitch  (1994)  ob- 
served darkblotched  rockfish  undergoing 
an  "immature  cycling"  and  even  assigned 
these  fish  a  maturity  stage.  After  the  re- 
productive season,  the  ovaries  of  immature 
black  rockfish  once  again  became  pale  pink 
in  color.  Because  these  fish  were  function- 
ally immature  and  there  was  no  way  to 
project  when  they  would  become  sexually 
mature,  they  were  combined  with  those 
small,  young  females  undergoing  no  sea- 
sonal ovarian  development  and  were  staged 
as  immature. 

Our  estimate  of  fork  length  at  50%  ma- 
turity for  female  black  rockfish  off  Oregon 
was  similar  to  the  400  mm  estimate  re- 
ported for  north-central  California  females 
(Wyllie  Echeverria,  1987),  but  lower  than 
the  estimate  of  422  mm  from  Washington 
(Wallace  and  Tagart,  1994).  Our  estimated 
age  at  50%  maturity  of  7.5  years  was  simi- 
lar to  the  estimates  of  7.9  and  7  years  from 
Washington  and  north-central  California, 
respectively.  McClure  (1982)  reported  that 
over  50%  of  examined  female  black  rockfish 
collected  off  Depoe  Bay,  Oregon,  were  ma- 
ture by  age  six.  The  difference  between  our 
estimate  and  McClure's  was  most  likely  due 
to  using  whole  otoliths  to  age  fish,  which 
resulted  in  underestimates  of  age,  and 
to  assigning  maturity  stages  only  during 
summer  months,  which  we  have  already  de- 
scribed as  problematic.  Both  absolute  and 
relative  fecundity  increased  with  age  for 
female  black  rockfish  in  Oregon  waters, 
although  there  was  a  great  deal  of  varia- 
tion not  accounted  for  by  age.  The  low  r2 
values  for  absolute  fecundity  regressions 
for  pre-  and  postfertilization  females  (0.25 
and  0.45  respectively)  are  due  largely  to 
the  relatively  poor  correspondence  between 
age  and  size  (Fig.  5).  Black  rockfish,  like 
many  slow  growing,  long-lived  fish  grow 
slowly  after  sexual  maturity.  The  rate  of 
growth  during  their  first  few  years  can  be 
quite  variable  depending  on  oceanographic 
conditions  and  food  availability.  As  a  re- 
sult, young  fish  can  be  as  large  or  larger 
than  much  older  fish  (Fig.  5).  Length  is  a 
better  predictor  of  fecundity  than  age  as  judged  by  the 
goodness-of-fit  of  the  multiple  linear  regression  model 
(Fig.  10;  /-2=0.70). 

An  increase  in  absolute  fecundity  with  age  was  ob- 
served in  both  prefertilization  and  postfertilization 


Absolute  fecuodity  (-10,000) 

ro             -p*             en             cd            o             ro             £* 
o            o            o            o            o            o            o 

1996 
•       1997 
o       1998 

A 
•                                                             A                  A 

%      % 

• 

0 

o                           • 

.     :  •  '        »   ^- 

°      •      *      1 

.    1    °    • 

o       •       -       .          ^S^ 

«               o               i 

1  *  • 

'    i 

*               4F(prefertilization)  =  298.413  +  36.823  age 
P<0.001 
r2  =  0.23 

U               I                    I                    I                    I                    I                    I                    1                    1                    I 

4              6              8             10            12            14            16            18            20 
Age  (years) 

Figure  6 

Scatter  plot  of  black  rockfish  absolute  fecundity  (AFl  (based  on  pre- 
fertilization eggs)  on  age  by  year  (1996-98)  with  a  fitted  regression  line 
from  pooled  data. 

140" 

_    120" 

o 

o 

o 

°    100- 
1       80- 

CJ 

§       60- 

o 

_o 

<       40- 

20" 

1996 
•       1997 
o       1998 

• 
• 

•     ! 

i 

s  : 

• 

• 

• 
• 

: 

•                                                                              • 

:  y^  • 

/     •      8 

i   i   : 

t      /^(fertilized)  =  -90,008  +  64,885  age 
P<  0.001 
r2  =  0,45 

0            '                               ll                                               ii 

4              6              8             10            12            14            16            18            20 

Age  (years) 

Figure  7 

Scatter  plot  of  black  rockfish  absolute  fecundity  (AF)  (based  on  fertilized 
eggs)  on  age  by  year  (1996-98)  with  a  fitted  regression  line  from  pooled 
data. 

females,  but  they  occurred  at  different  rates.  As  il- 
lustrated in  Figure  8,  the  absolute  fecundity  for  a  post- 
fertilization  6-year-old  black  rockfish  was  only  58% 
of  the  estimated  absolute  fecundity  for  a  prefertiliza- 
tion fish  of  the  same  age.  By  age  15  absolute  fecundity 


426 


Fishery  Bulletin  102(3) 


140 


120 


o 
o 

°.    100 
o 

~      80 

C 
13 
O 

•2      60 

£ 
3 
o 
£       40 

< 

20 


-* —   prefertllization 
-o—    fertilized 


10  12  14 

Age  (years) 


16 


20 


Figure  8 

Separate-lines  regressions  fitted  to  absolute  fecundity  (based  on  pre- 
fertilization  and  fertilized  eggs)  on  age  for  black  rockfish  in  Oregon. 


800" 

— ■ —   prefertllization 

■-o—    fertilized 

•                     • 

!           i 

•     •     • 
• 

5     600" 

O) 

cn 

03 

••Is                 •                          • 

0 

>. 

o 

§      400- 

•   f   \   :   s 

o 

_Q) 

.   '   •   i 

ro 

•                    8                    ° 

£     200- 

o                                            o             0 

RF  =  375.7  +  17.5-age  -  106.5'stage 
o                     P<  0.001 
^  =  0.27 

0              I                     i                      i 

6                    8                   10                  12                  14 

16 

Age  (years) 

Figure  9 

Parallel-lines  model  fitted  to  relative  fecundity  I  based  on  prefertiliza- 

tion  and  fertilized  eggs)  on  age  for  black  rockfish  in  Oregon. 

estimates  for  fertilized  and  prefertilization  females  were 
approximately  equal.  Yolked  oocytes  from  older  females 
were  more  successful  in  reaching  the  developing  embryo 
stage.  This  may  be  attributed  to  higher  rates  of  fertil- 
ization, greater  viability  of  embryos,  or  a  combination 
of  both  in  older  female  black  rockfish.  Regardless  of 
the  mechanism  there  should  have  been  signs  of  greater 
atresia  in  the  ovaries  of  young  fish,  which  we  did  not 


observe  in  our  histological  preparations.  This  may  have 
been  due  to  rapid  resorption  of  unfertilized  oocytes  or 
an  artifact  of  the  fragile  nature  of  fertilized  ovaries, 
which  made  it  difficult  to  obtain  representative  histo- 
logical preparations.  Nevertheless,  these  results  sug- 
gest that  fecundity  in  black  rockfish  is  best  described 
after  fertilization,  but  care  must  be  taken  to  minimize 
embryo  loss.  These  results  also  suggest  that  current 


Bobko  and  Berkeley:  Maturity,  ovarian  cycle,  fecundity,  and  parturition  of  Sebastes  melanops 


427 


estimates  of  reproductive  potential,  in  which  fe- 
cundity for  prefertilization  females  is  used,  may 
overestimate  actual  larval  production  because 
an  increasing  proportion  of  the  stock  consists  of 
young  fish. 

We  observed  a  recurring  trend  of  older,  larger 
fish  extruding  larvae  earlier  in  the  reproductive 
season  and  larval  output  being  increasingly  domi- 
nated by  younger  and  younger  fish.  Eldridge  et  al. 
(1991)  reported  that  larger  (and  most  likely  older) 
yellowtail  rockfish  {Sebastes  flavidus)  spawned 
earlier  in  the  season  than  smaller  fish — a  pat- 
tern also  reported  for  darkblotched  rockfish  (S. 
crameri)  by  Nichol  and  Pikitch  (1994).  Reduced 
food  availability  has  been  suggested  as  a  poten- 
tial cause  for  delayed  reproduction  in  Sebastes  for 
smaller,  younger  individuals  with  high  metabolic 
requirements  for  somatic  growth  (Larson,  1991). 
We  feel  that  limiting  the  amount  of  energy  that 
can  be  spent  on  reproductive  development  would 
cause  lower  fecundity  or  reduced  yolk  content,  but 
not  necessarily  a  delay  in  reproductive  develop- 
ment that  would  result  in  suboptimal  timing  of 
parturition. 

Stock  assessments  rarely  consider  changes  in 
population  age  composition  resulting  from  the 
removal  of  older  age  classes  except  to  the  extent 
that  total  egg  and  larval  production  is  reduced. 
The  decreasing  representation  of  mature  female 
black  rockfish  age  10  and  older  in  the  three  years 
of  our  study  indicates  that  age  truncation  is  oc- 
curring in  black  rockfish  in  Oregon.  This  trun- 
cation not  only  removes  biomass  and  potential 
larval  production,  but  truncation  of  the  upper 
end  of  the  age  distribution  eliminates  mature 
females  with  higher  fecundity  per  individual,  a 
greater  success  in  carrying  eggs  through  to  the 
larval  stage,  and  an  age  group  that  extends  the 
overall  parturition  season.  Further  research  is 
necessary  to  explore  the  controlling  mechanisms 
of  differential  reproductive  success  with  age  and 
to  determine  how  best  to  incorporate  these  find- 
ings into  stock  assessment  models. 


Acknowledgments 

Many  individuals  contributed  to  the  completion  of 
this  study.  Tom  Rippetoe  provided  many  hours  of 
expert  assistance  in  all  aspects  of  this  research. 
Bob  Mikus  aged  all  of  the  adult  black  rockfish 
for  this  study.  We  thank  him  and  the  Oregon 
Department  of  Fisheries  and  Wildlife  for  all  their 
support.  We  thank  Dan  Detman  for  his  help  and 
time  in  collecting  black  rockfish.  We  also  thank 
Brock  McLeod,  Jason  Castillo,  David  Stewart,  and 
Michael  Hogansen  for  all  their  help.  We  are  espe- 
cially grateful  for  all  those  unpaid  volunteers  who 
helped  with  fieldwork:  Joe  O'Malley,  Mark  Amend, 
Bill  Pinnix,  Wolfe  Wagman,  and  Pat  McDonald. 


160" 

/1F  =  -1 .888,811   +6.122FL  -160,053  stage 

Absolute  fecundity  (xio.000) 

o                    o                    o 

r2  =  0  70 

.          .      .*>  o 

•  •           1   ^f 

..    .  :  '••"tt:    .•  °.    ° 

fi*.\C?  ,"8  .  =  °  •  ° 

o     ° 

•  • 

•• 

°^  • . 

•  o 

o 

o 

— • —    prefertilization 
o       fertilized 

365             390             415             440             465             490             515 

Fork  length  (mm) 

Figure  10 

Parallel-lines  model  fitted  to  absolute  fecundity  (based  on  prefer- 

tilization and  fertilized  eggs)  on  age  for  black  rockfish  in  Oregon. 

I  I  Age  6  -  8 

^M  Age  9  -  1 1 

I  I  Age  12-14 

^mm  Age  15  + 


^     ^    <e?    <eP    ^    <eP    ^    ^    ^ 

«£>'     &      K*'    ^'     </      t^'     #'    ^'     ^' 

Week-month 

Figure  11 

Percent  relative  larval  production  estimated  from  observations 
of  larval  development  and  age-group-specific  absolute  fecundity 
(based  on  fertilized  eggs)  for  all  mature  females  belonging  to  each 
age  group  of  black  rockfish  collected  in  Oregon  during  1996-98. 


428 


Fishery  Bulletin  102(3) 


Table  5 

Age  group-specific  relative  larval  production  for  1996-98 
for  female  black  rockfish  off  Oregon. 

Percentage  of  relative  larval  production 


Age  group 


1996 


1997 


1998 


6-8 

9-11 
12-14 
15  and  older 


26.4% 
60.1% 
11.3% 

2.2% 


43.1% 

49.6% 

5.8% 

1.5% 


65.3% 

32.9% 

1.8% 

0.0% 


We  thank  the  charter  boat  operators  in  Depoe  Bay. 
Newport,  and  Charleston,  Oregon,  for  their  kindness, 
patience,  and  cooperation.  Only  through  their  assistance 
was  this  research  possible.  This  research  was  partially 
supported  by  the  NOAA  Office  of  Sea  Grant  and  Extra- 
mural Programs,  U.S.  Department  of  Commerce,  under 
grant  numbers  NA36RG0451  (project  no.  R/OPF-46), 
and  by  appropriations  made  by  the  Oregon  State  legisla- 
ture. Additional  funding  was  provided  through  Hatfield 
Marine  Science  Center  scholarships:  the  1996  Barbara 
Schwantes  Memorial  Fellowship;  the  1997  Mamie  L. 
Markham  Endowment  Award;  and  the  1998  Bill  Wick 
Marine  Fisheries  Award. 


Literature  cited 

Bagenal,  T.  B. 

1971.     The  interrelation  of  the  size  offish  eggs,  the  date 
of  spawning  and  the  production  cycle.     J.  Fish  Biol. 
3:207-219. 
Beamish,  R.  J.,  and  D.  E.  Chilton. 

1982.     Preliminary  evaluation  of  a  method  to  determine 
the  age  of  sablefish  (Anoplopoma  fimbria).     Can.  J. 
Fish.  Aquat.  Sci.  39:277-287. 
Berkeley,  S.  A.,  and  E.  D.  Houde. 

1978.     Biology  of  two  exploited  species  of  halfbeaks, 
Hemiramphus  brasiliensis  and  H.  balao  from  southeast 
Florida.     Bull.  Mar.  Sci.  28:624-644. 
Boehlert,  G.  W.,  W.  H.  Barss,  and  P.  B.  Lamberson. 

1982.     Fecundity  of  the  widow  rockfish,  Sebostes  entome- 
las,  off  the  coast  of  Oregon.     Fish.  Bull.  4:881-884. 
Boehlert,  G.  W.,  and  M.  M.  Yoklavich. 

1984.     Reproduction,  embryonic  energetics,  and  the  mater- 
nal-fetal relationship  in  the  viviparous  genus  Sebastes 
(Pisces:  Scorpaenidae).     Biol.  Bull.  167:354-370. 
Bowers,  M.  J. 

1992.     Annual  reproductive  cycle  of  oocytes  and  embryos 
of  yellowtail    rockfish,    Sebastes   flavidus    (family 
Scorpaenidael.     Fish.  Bull.  90:231-242. 
Eldridge,  M.  B.,  J.  A.  Whipple,  M.  J.  Bowers,  B.  M.  Jarvis,  and 
J.  Gold. 

1991.     Reproductive  performance  of  yellowtail  rockfish. 
Sebastes  flavidus.     Environ.  Biol.  Fish.  30:91-102. 
Goodman  D. 

1984.     Risk  spreading  as  an  adaptive  strategy  in  iteropar- 
ous  life  histories.     Theoret.  Popul.  Biol.  25:1-20. 


Gunderson,  D.  R.,  P.  Callahan,  and  B.  Goiney. 

1980.  Maturation  and  fecundity  of  four  species  of 
Sebastes.     Mar.  Fish.  Rev.  42l3-4):l. 

Hutchings.  J.  A.,  and  R.  A.  Myers. 

1993.  Effect  of  age  on  the  seasonality  of  maturation  and 
spawning  of  Atlantic  cod,  Gadus  morhua,  in  the  north- 
west Atlantic.     Can.  J.  Fish.  Aquat.  Sci.  50:2468-2474. 

Lambert,  T.  C. 

1987.  Duration  and  intensity  of  spawning  in  herring 
Clupea  harengus  as  related  to  the  age  structure  of  the 
mature  population.  Mar.  Ecol.  Prog.  Ser.  39:209- 
220. 

Lambert,  T.  C. 

1990.  The  effect  of  population  structure  on  recruitment 
in  herring.     J.  Cons.  Int.  Explor.  Mer.  47:249-255. 

Larson.  R.  J. 

1991.  Seasonal  cycles  of  reserves  in  relation  to  reproduc- 
tion in  Sebastes.     Environ.  Biol.  Fish.  30:57-70. 

Leaman,  B.  M..  and  R.  J.  Beamish. 

1984.     Ecological  and  management  implications  of  longev- 
ity in  some  northeast  Pacific  groundfishes.     Int.  North 
Pac.  Fish.  Comm.  Bull.  42:85-97. 
Leaman,  B.  M.,  and  D.  A.  Nagtegaal. 

1987.     Age  validation  and  revised  natural  mortality  rate 
for  yellowtail  rockfish.     Trans.  Am.  Fish.  Soc.  116(2): 
171-175. 
Love,  M.  S.,  and  W.  V.  Westphal. 

1981.  Growth,  reproduction,  and  food  habits  of  olive  rock- 
fish. Sebastes  serranoides,  off  central  California.  Fish. 
Bull.  79:533-545. 

Lowerre-Barbieri,  S.  K..  and  L.  R.  Barbieri. 

1993.  A  new  method  of  oocyte  separation  and  preservation 
for  fish  reproduction  studies.     Fish.  Bull.  91:165-170. 

MacFarlane,  R.  B.,  and  E.  C.  Norton. 

1999.  Nutritional  dynamics  during  embryonic  develop- 
ment in  the  viviparous  genus  Sebastes  and  their  applica- 
tion to  the  assessment  of  reproductive  success.  Fish. 
Bull.  97:  273-281. 

Maclellan,  S.  E. 

1997.  How  to  age  rockfish  (Sebastes  I  using  S.  alutus  as  an 
example — the  otolith  burnts  (sic)  section  technique.  Ca- 
nadian Tech.  Rep.  Fish.  Aquat.  Sci.  42  p. 

McClure,  R.  E. 

1982.  Neritic  reef  fishes  off  central  Oregon:  aspects  of 
life  histories  and  the  recreational  fishery.  M.S.  thesis, 
94  p.     Oregon  State  Univ.,  Corvallis,  OR. 

Moser,  H.  G. 

1967.     Seasonal  histological  changes  in  the  gonads  of 
Sebastodes  paucispinis  (Ayres),  and  ovoviviparous  teleost 
(family  Scorpaenidae).     J.  Morph.  123:329-353. 
Nichol,  D.  G.,  and  E.  K.  Pikitch. 

1994.  Reproduction  of  darkblotched  rockfish  off  the 
Oregon  coast.     Trans.  Am.  Fish.  Soc.  123:469-481. 

Schultz,  D.  L. 

1989.     The  evolution  of  phenotypic  variance  with  itero- 
parity.     Evolution  43:473-475. 
Shepherd.  G.  R.,  and  C.  B.  Grimes. 

1984.     Reproduction  of  weakfish.  Cynoscion   regalis, 
in  the  New  York  Bight  and  evidence  for  geographi- 
cally specific  life  history  characteristics.     Fish.  Bull. 
82(3):501-511. 
Simpson,  A.  C. 

1959.     The  spawning  of  plaice  in  the  North  Sea.     Fish. 
Invest.  M.A.F.F.  London  (Ser.  2)  22:1-111. 
Wallace,  F.  W.,  and  J.  V.  Tagart. 

1994.     Status  of  the  coastal  black  rockfish  stocks  in  Wash- 


Bobko  and  Berkeley:  Maturity,  ovarian  cycle,  fecundity,  and  parturition  of  Sebastes  me/anops 


429 


ington  and  northern  Oregon  in  1994.  Appendix  F  in 
Status  of  the  Pacific  coast  groundfish  fishery  through 
1994  and  recommended  acceptable  biological  catches 
for  1995.  Pacific  Fisheries  Management  Council,  Port- 
land, OR. 
Wyllie  Echeverria,  T. 

1987.     Thirty-four  species  of  rockfishes:  Maturity  and 
seasonality  of  reproduction.     Fish.  Bull.  85:826-832. 


Yamada,  J.,  and  M.  Kusakari. 

1991.     Staging  and  time  course  of  embryonic  develop- 
ment in  kurosoi,  Sebastes  schlegeli.     Environ.  Biol. 
Fish.  30:103-110. 
Zastrow,  C.  E.,  E.  D.  Houde,  and  E.  H.  Saunders. 

1989.  Quality  of  striped  bass,  Morone  saxatilis,  eggs  in 
relation  to  river  source  and  female  weight.  Rapp.  P.-V. 
Cons.  Int.  Explor.  Mer  191:34-42. 


430 


Abstract— Octopuses  are  commonly 
taken  as  bycatch  in  many  trap  fisher- 
ies for  spiny  lobsters  I  Decapoda:  Pal- 
muridae)  and  can  cause  significant 
levels  of  within-trap  lobster  mortality. 
This  article  describes  spatiotempo- 
ral  patterns  for  Maori  octopus  i  Octo- 
pus maorum  >  catch  rates  and  rock 
lobster  (Jasus  edwardsii)  mortality 
rates  and  examines  factors  that  are 
associated  with  within-trap  lobster 
mortality  in  the  South  Australian 
rock  lobster  fishery  (SARLF).  Since 
1983,  between  38,000  and  119,000 
octopuses  per  annum  have  been 
taken  in  SARLF  traps.  Catch  rates 
have  fluctuated  between  2.2  and  6.2 
octopus/100  trap-lifts  each  day.  There 
is  no  evidence  to  suggest  that  catch 
rates  have  declined  or  that  this  level 
of  bycatch  is  unsustainable.  Over  the 
last  five  years,  approximately  240,000 
lobsters  per  annum  have  been  killed 
in  traps,  representing  ~4%  of  the  total 
catch.  Field  studies  show  that  over 
98%  of  within-trap  lobster  mortal- 
ity is  attributable  to  octopus  pre- 
dation.  Lobster  mortality  rates  are 
positively  correlated  with  the  catch 
rates  of  octopus.  The  highest  octo- 
pus catch  rates  and  lobster  mortality 
rates  are  recorded  during  summer 
and  in  the  more  productive  southern 
zone  of  the  fishery.  In  the  southern 
zone,  within-trap  lobster  mortality 
rates  have  increased  in  recent  years, 
apparently  in  response  to  the  increase 
in  the  number  of  lobsters  in  traps 
and  the  resultant  increase  in  the 
probability  of  octopus  encountering 
traps  containing  one  or  more  lobsters. 
Lobster  mortality  rates  are  also  posi- 
tively correlated  with  soak-times  in 
the  southern  zone  fishery  and  with 
lobster  size.  Minimizing  trap  soak- 
times  is  one  method  currently  avail- 
able for  reducing  lobster  mortality 
rates.  More  significant  reductions  in 
the  rates  of  within-trap  lobster  mor- 
tality may  require  a  change  in  the 
design  of  lobster  traps. 


Maori  octopus  (Octopus  maorum)  bycatch  and 
southern  rock  lobster  (Jasus  edwardsii)  mortality 
in  the  South  Australian  rock  lobster  fishery 

Daniel  J.  Brock 

South  Australian  Research  and  Development  Institute  (Aquatic  Sciences) 

2  Hamra  Ave. 

West  Beach,  South  Australia  5024,  Australia 

Present  address:  Department  of  Soil  and  Water 

Adelaide  University 

Adelaide,  South  Australia  5005,  Australia 
E-mail  address:  Brock. Daniel  a1  saugov.sa.gov  au 

Timothy  M.  Ward 

South  Australian  Research  and  Development  Institute  (Aquatic  Sciences) 

2  Hamra  Ave. 

West  Beach,  South  Australia  5024,  Australia 


Manuscript  submitted  28  April  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
2  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:430-440  (2004). 


Fishing  traps  are  used  throughout  the 
world  to  target  a  wide  range  of  crusta- 
ceans, fishes,  and  cephalopods.  Com- 
mercial trap  fisheries,  especially  those 
for  decapod  crustaceans,  are  often  the 
most  valuable  fisheries  within  a  region 
(Phillips  et  al.,  1994).  Traps  are  gen- 
erally considered  to  be  an  efficient 
and  benign  form  of  fishing  because 
they  appear  to  cause  relatively  minor 
damage  to  benthic  habitats,  can  be 
designed  to  target  particular  species 
and  size  ranges,  and  produce  live 
catches  in  good  condition  while  mini- 
mizing bycatch  (Miller,  1990). 

There  are  49  species  of  spiny  lob- 
sters (Decapoda:  Palinuridae)  world- 
wide, 33  of  which  support  commercial 
trap  fisheries.  The  largest  of  these 
are  in  Cuba,  South  Africa,  Mexico, 
Australia,  and  New  Zealand  (Wil- 
liams, 1988).  The  main  trap  fisher- 
ies in  Australia  are  for  western  rock 
lobster  {Panulirus  cygnus)  in  Western 
Australia  and  southern  rock  lobster 
(Jasus  edwardsii)  along  the  southern 
coastline.  Octopuses  constitute  a  sig- 
nificant component  of  the  bycatch  in 
both  fisheries  (Joll1;  Knight  et  al.2)  . 

In  South  Australia,  J.  edwardsii 
supports  the  State's  most  valuable 
commercial  fishery.  Octopus  maorum 
is  a  significant  bycatch  species  and 
is  thought  to  be  the  major  cause  of 
lobster  mortality  in  traps  (Prescott 
et  al.3). 


Although  the  octopus  bycatch  of  the 
South  Australian  rock  lobster  fishery 
(SARLF)  is  saleable,  the  commercial 
value  of  this  product  does  not  offset 
the  value  of  the  large  number  of  lob- 
sters that  are  killed  in  traps  by  octo- 
pus. Many  fishermen  are  convinced 
that  incidental  mortality  of  octopus 
resulting  from  lobster  fishing  acts 
to  control  octopus  numbers  and  that 
if  these  rates  were  reduced,  octopus 
abundance  and  levels  of  within  trap 
predation  would  increase. 

Despite  the  prevalence  of  octopus 
bycatch  in  lobster  fisheries,  there 
have  been  only  a  few  studies  on  the 
interaction  between  octopus  and  lob- 


1  Joll,  L.  1977.  The  predation  of  trap- 
caught  western  rock  lobster  {Panulirus 
Longipes  cygnus)  by  octopus.  Depart- 
ment of  Fisheries  and  Wildlife,  Western 
Australia,  Report  29,  58  p.  (Available 
from  Department  of  Fisheries,  168-170 
St  George's  Terrace.  Perth,  Western  Aus- 
tralia, 6000.] 

-  Knight,  M.  A.,  A.  Tsolos,  and  A.  M. 
Doonan.  2000.  South  Australian 
fisheries  and  aquaculture  information 
and  statistics  report.  Research  Report 
Series  49,  69  p.  [Available  from  SARD] 
Aquatic  Science,  2  Hamra  Avenue,  West 
Beach,  South  Australia  5022.] 

:!  Prescott,  J..  R.  McGarvey,  Y.  Xiao,  and  D. 
Casement.  1999.  Rock  lobster.  South 
Australian  Fisheries  Assessment  Series 
99/04,  35  p.  [Available  from  SARDI 
Aquatic  Science,  2  Hamra  Avenue,  West 
Beach,  South  Australia  5022.1 


Brock  and  Ward:  Octopus  bycatch  and  lobster  mortality  in  the  South  Australian  rock  lobster  fishery 


431 


Figure  1 

Map  of  the  marine  fishing  areas  (MFAs)  of  the  South  Australian  rock  lobster  fishery. 
Shading  shows  the  MFAs  that  were  considered  in  this  study  and  where  most  fishing  effort  is 
concentrated. 


sters  in  traps  (Joll1).  Furthermore,  there  is  a  paucity  of 
quantitative  data  on  the  impact  of  fishing  on  octopus 
populations,  the  proportion  of  lobster  mortality  that 
is  attributable  to  octopus  predation,  or  the  long-term 
economic  and  ecological  effects  that  octopus-induced 
mortality  may  have  on  lobster  fisheries. 

In  this  study,  we  examined  the  interaction  between 
O.  maorum  and  J.  edwardsii  in  the  South  Australian 
rock  lobster  fishery  (SARLF).  The  objectives  were  1)  to 
determine  the  number  of  lobsters  and  octopus  caught 
and  the  number  of  lobsters  killed  in  traps  each  year  in 
the  fishery;  2)  to  describe  the  interannual  and  seasonal 
patterns  in  lobster  catch  rate  (CPUEL),  octopus  catch 
rate  (CPUE0),  and  lobster  mortality  rate  (ML);  3)  to 
examine  some  factors  that  may  affect  lobster  mortal- 
ity rates;  4)  to  estimate  what  proportion  of  the  lobster 
mortality  is  attributable  to  octopus  predation;  and  5)  to 
determine  whether  the  rate  of  lobster  mortality  through 
octopus  predation  in  traps  is  size  dependent. 


Materials  and  methods 

South  Australian  rock  lobster  fishery 

The  SARLF  is  divided  into  a  northern  zone  (NZ)  and 
a  southern  zone  (SZ),  each  of  which  is  further  divided 
into  marine  fishing  areas  (MFAs)  for  statistical  purposes 
(Fig.l).  There  are  68  and  183  fishermen  licensed  to  oper- 


ate in  the  NZ  and  SZ  respectively.  The  fishing  season 
extends  from  November  to  May  in  the  NZ  and  October 
to  April  in  the  SZ.  A  quota  management  system  was 
introduced  in  the  SZ  in  1993,  whereas  the  NZ  is  man- 
aged by  gear  restrictions  and  temporal  closures. 

Total  annual  catch  and  effort  for  the  SARLF 

Catch  and  effort  data  are  recorded  on  a  daily  basis  by  all 
individual  fishermen.  Since  1983.  a  standardized  logbook 
for  recording  catch  and  effort  has  been  used  across  the 
fishery.  Data  provided  by  fishermen  include  MFA  fished, 
average  depth  fished,  number  of  trap-lifts,  number  and 
total  weight  of  live  lobsters,  number  of  dead  lobsters,  and 
number  and  total  weight  of  octopus.  This  information 
is  stored  in  a  South  Australian  rock  lobster  database 
that  is  managed  by  the  South  Australian  Research  and 
Development  Institute,  Aquatic  Sciences. 

Interannual  and  seasonal  patterns 
in  CPUEL,  CPUEQ,  and  ML 

Although  commercial  fishing  for  lobsters  occurs  along 
most  of  the  South  Australian  coastline,  the  majority  of 
effort  is  concentrated  in  only  a  few  MFAs.  In  the  NZ 
over  the  last  5  years  about  72%  of  total  trap-lifts  were 
made  in  MFAs  15,  28,  39,  40,  and  49.  In  the  SZ  over  the 
same  period  95%  of  trap-lifts  were  made  in  MFAs  51, 
55,  56,  and  58  (Fig.  1). 


432 


Fishery  Bulletin  102(3) 


Data  from  the  database  were  used  to  calculate  catch 
rates  of  lobsters  (CPUEL),  octopus  (CPUE0),  and  MLon 
an  annual  and  monthly  basis  for  the  nine  major  MFAs 
listed  above.  Catch  rates  from  these  MFAs  for  each 
fisherman  were  calculated  according  to  the  formula: 
catch  rate  =  catch  number/itrap-liftslday).  Annual  and 
seasonal  trends  in  CPUEj ,  CPUE0,  and  ML  were  cal- 
culated for  each  zone  and  MFA. 

Factors  that  affect  within-trap  lobster  mortality 

Potential  factors  that  affect  within-trap  lobster  mortality 
were  analysed  by  using  a  general  linear  model  (type-3 
sums  of  squares)  under  the  assumption  that  the  number 
of  dead  lobsters  follows  a  log-normal  distribution. 

The  number  of  dead  lobsters/trap-lift/day/license 
(with  a  ln+1  transformation)  was  used  as  the  measure 
of  lobster  mortality.  A  model  of  the  following  structure 
was  used  to  examine  factors  that  affect  the  numbers 
of  dead  lobster: 

Dead  lobster  =  License  +  MFA  +  Month  +  Year 

+  Effort  +  Depth  +  Octopus  +  Lobster  catch 

+  Soak-time  +  I  License  xYear)  +  I  License  xMonth) 

+  (YearxMonth)  +  (YearxMFA)  +  iSoak-timexYear) 

+  <  Soak-time  xMonth). 

In  the  model.  License  represents  an  individual  fisher- 
man, MFA  is  the  marine  fishing  area.  Month  accounts 
for  seasonal  variation  and  Year  accounts  for  interannual 
variation.  Effort  is  the  number  of  trap-lifts/license  each 
day,  Depth  is  the  average  depth  fished  by  each  License 
on  a  particular  day.  Octopus  and  Lobster  are  the  respec- 
tive daily  catches/license,  and  Soak-time  is  the  number 
of  days  that  the  traps  remained  in  the  water  since  the 
previous  trap-lift. 

The  interaction  terms  License xYear  and  License x 
Month  account  for  variations  in  the  catch  characteris- 
tics of  the  individual  licenses  over  time  that  result  from 
changes  in  fishing  practises  and  efficiency  associated 
with  different  boats,  license  holders,  and  skippers.  The 
interaction  terms  YearxMonth  and  YearxMFA  account 
for  variation  in  the  population  dynamics  of  octopus  and 
lobster  over  time  in  different  locations  that  could  result 
in  differential  trends  in  lobster  mortality.  The  inter- 
action terms  Soak-time  xYear  and  Soak-time  xMonth 
reflects  the  change  in  general  fishing  strategies  over 
time.  In  quota-managed  fisheries  the  average  soak-time 
will  be  affected  by  a  number  of  factors,  for  example, 
that  may  include  price,  weather,  and  the  fishermen's 
perceived  ability  to  catch  their  quota. 

The  analysis  was  run  separately  for  the  SZ  (/?  =  493,629 
traps)  and  NZ  (ra=155,628  traps)  because  the  respective 
zones  have  different  fishing  seasons  and  management 
structures.  The  relationship  between  the  number  of 
dead  lobsters  and  the  factors  depth,  soak-time, and  num- 
ber of  octopuses  and  lobsters  were  presented  graphically 
by  the  equation: 

Lobsters  killed  in  traps  <*.  factor  a, 


where  a  =  the  parameter  estimated  by  use  of  the  model. 

Source  of  lobster  mortality  and  size-dependent  mortality 

A  sampling  program  was  conducted  on  three  commer- 
cial vessels  from  the  SZ  during  the  2001-02  fishing 
season.  Five  days  were  spent  on  each  vessel.  All  lobsters 
caught  were  measured  (carapace  length,  mm),  and  the 
sex  (male  or  female),  maturity  (mature  or  immature), 
status  (dead  or  alive),  and  cause  of  death  (octopus  or 
other)  were  recorded. 

The  method  used  to  distinguish  between  lobsters 
killed  by  octopus  or  other  means  followed  that  of  Joll.1 
This  suitability  of  this  approach  was  confirmed  through 
examination  of  the  carcasses  of  over  one  hundred  lob- 
sters killed  by  octopus  in  aquarium  trials  (Brock  et 
al.4).  Lobsters  with  shells  that  were  partly  or  completely 
separated  at  the  juncture  between  abdomen  and  cepha- 
lothorax  but  were  otherwise  undamaged  were  deemed 
to  have  been  killed  by  octopuses,  whereas  lobsters  with 
shells  without  this  separation  and  with  evidence  of  bite 
marks  were  deemed  to  have  been  eaten  by  other  preda- 
tors (fish  or  cuttlefish). 

Anecdotal  evidence  from  fishermen  suggests  that 
larger  lobsters  are  more  susceptible  to  predation  than 
smaller  ones.  The  effect  of  CL  on  the  probability  of 
mortality  was  examined  separately  for  males  and  fe- 
males by  generalized  linear  modeling.  The  probability 
of  mortality  at  a  given  size  was  modeled  with  a  logistic 
equation  of  the  form: 

P{sex,  CL)  =  l/(l+e-,a+6CL,), 

where  P(sex,  CL)  =  the  probability  of  a  lobster  of  a 
given  sex  at  carapace  length  CL 
being  dead;  and 
a  and  b  are  parameters  to  be  estimated. 


Results 

Estimation  of  total  lobster  catch,  octopus  bycatch, 
and  lobster  mortality 

In  1999,  there  were  1.6  million  trap-lifts  in  the  SARLF, 
and  70%  of  this  total  effort  was  in  the  SZ  (Fig.  2).  The 
number  of  traps-lifts  in  the  SZ  declined  from  2.2  mil- 
lion in  1983  to  1.2  million  in  1999  (Fig.  2Ai.  In  contrast, 
fishing  effort  in  the  NZ  remained  relatively  consistent 
with  406,000  trap-lifts  in  1983  and  480,  000  trap-lifts 
in  1999  (Fig  2B). 

The  total  annual  lobster  catch  has  generally  increased 
in  each  fishing  zone  since   1983  (Fig.  2,  A  and  B). 


'  Brock,  D.  J.,  T.  M.  Saunders,  and  T.  M.  Ward.  In  review.  A 
two-chambered  trap  with  potential  for  reducing  within-trap 
predation  by  octopus  on  rock  lobster.  Can.  J.  Fish.  Aquat. 
Sci.,  19  p.  lAvailable  from  SARDI  Aquatic  Science,  2  Hamra 
Avenue.  West  Beach,  South  Australia  5022.1 


Brock  and  Ward:  Octopus  bycatch  and  lobster  mortality  in  the  South  Australian  rock  lobster  fishery  433 


A 

7  -r 

6  |< 

5 

4 

3 

2 

1 

0 


Southern  zone 
I  lobster       — ♦ —  trap  lifts 


\ 


T  2'5 

--  2 

1 

05 

0 


B 

7 

6 

5  + 

4 

3 

2 

1 

0 


Northern  zone 
I  lobster        — ♦ —  trap  lifts 


2.5 
■■  2 
--  1.5 

1 


uxi 


-♦-♦-♦ 

iinlniiilli 


i"i"i"i"i"i"i"i"i"i"i"i"i' 


0.5 

0 


1983 


1987 


1991 


1995 


1999 


1983 


1987 


1991 


1995 


1999 


1983 


1987 


1991 


1995 


1999 


1983 


1987 


1991 


1995 


1999 


Season 


Figure  2 

Annua]  total  catch  and  effort  for  each  zone  of  the  South  Australian  rock  lobster  fishery  for  number  of 
live  lobsters  caught  (A  l,  number  of  trap-lifts  used  (B).  number  of  dead  lobsters  caught  (C),  and  number 
of  octopuses  caught  (D)  (Note:  change  of  scale  in  D). 


In  the  SZ,  the  annual  lobster  catch  rose  from  3.8  mil- 
lion lobsters  to  a  peak  of  6.4  million  lobsters  in  1991 
and  was  5.4  million  lobsters  in  1999  (Fig.  2A).  In  the 
NZ,  560,000  lobsters  were  taken  in  1983  compared  to 
850,000  in  1991  (Fig.  2B). 

The  total  annual  octopus  catch  varied  among  years  in 
both  zones,  but  between  70%  and  907c  of  the  total  octo- 
pus catch  were  landed  in  the  SZ  (Fig.  2).  The  total  num- 
ber of  octopus  ranged  from  36,000  in  1986  to  109,000 
in  1992  (Fig.  2C)  in  the  SZ,  and  from  4700  octopuses  in 
1985  to  11,200  in  1998  in  the  NZ  (Fig.  2D). 

In  1999,  over  226,000  lobsters  were  killed  in  traps  in 
the  SARLF  (Fig.  2).  Since  1983,  the  mean  proportion  of 
dead  lobsters  out  of  the  total  catch  has  been  approxi- 
mately 4%.  In  the  SZ,  the  number  of  lobsters  killed  in 
traps  has  generally  increased  from  118,000  in  1983  to 
196,000  in  1999;  a  peak  of  274,000  dead  lobsters  oc- 
curred in  1992  (Fig.  2C).  In  the  NZ,  there  has  also  been 
a  general  increase  in  the  number  of  lobster  killed  in 
traps  each  year;  24,000  dead  lobsters  were  recorded  in 
1983,  compared  to  31,000  in  1999  and  a  peak  of  39,000 
dead  lobsters  recorded  in  1998  (Fig.  2D). 


Interannual  and  seasonal  patterns  in 
CPUEL  CPUE0,and  ML 

Southern  zone  Mean  annual  CPUEL  in  the  SZ  increased 
from  175  to  466  lobsters/I  100  trap-lifts/day)  between 
1983  and  1999,  and  the  largest  increase  occurred 
between  1997  and  1999  (Fig.  3A).  Mean  annual  CPUE0 
ranged  from  1.8  to  6.2  octopus/dOO  trap-lifts/day)  in 
1987  and  1992,  respectively  (Fig.  3C).  Mean  annual 
ML  rose  from  5  to  17  dead  lobster/(100  trap-lifts/day) 
between  1983  and  1999  (Fig.  3E).  Peaks  in  both  CPUE0 
and  ML  occurred  in  1985,  1992,  and  1995. 

Mean  monthly  CPUEL  declined  during  the  fishing  sea- 
son from  310  to  164  lobster/(100  trap-lifts/day)  between 
October  and  April  (Fig.  4A).  In  contrast,  mean  monthly 
CPUE0  increased  from  2.6  to  3.7  octopus/dOO  trap-lifts/ 
day)  between  October  and  December  and  declined  to  1.8 
octopus/dOO  trap-lifts/day)  in  April  (Fig.  4C).  Similarly 
mean  monthly  ML  increased  from  10.7  to  12.8  dead  lob- 
ster/(100  trap-lifts/day)  between  October  and  November 
and  declined  to  6.7  dead  lobster/)  100  trap-lifts/day)  in 
April  (Fig.  4E). 


434 


Fishery  Bulletin  102(3) 


A    lobster 

700  , 


Southern  zone 


Northern  zone 


maiahmiiii 


-i — I — I — I — I — I — I — I — I — I — I — I — i — I — r- 


-1 — I — I — I — I — I — I — I — I — I — I — I — I — I — I — I — I 
1983    1987    1991     1995    1999     1983    1987    1991    1995    1999 


(_  octopus 
10  -, 


L)  octopus 
10 


— i — i — i — i — i — i — i — i — i — i — i — i — i — i — i — i 
1983    1987    1991    1995    1999 


6  -I 
4 
2 
0 


H#fe^%ftft 


t.    dead  lobster 
30 


Q-     25 
8      20 

£      15 

V> 

2      10 
•o 

TO 
<D 
T3  5 

6 

2         0 


T 1 1 1 1 1 1 1 1 1 1 1 I 1 1— I 1 

1983    1987    1991    1995    1999 


T 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 

1983    1987    1991    1995    1999 


r  dead  lobster 
30  -, 

25  - 

20 

15 

10 

5 

0 


-tfrftttt 


T — I 1 — I 1 — I 1 1 1 1 1 1 1 — 1 1 — I 1 

1983         1987         1991         1995         1999 


Season 


Figure  3 

Annual  catch  rates  in  each  fishing  zone  for  lobsters  (CPUE,  I  (A  and  B),  octopus  (CPUE, ,i 
(C  and  Dl,  and  dead  lobsters  (M[  J  iE  and  F).  Error  bars  are  ±SD  of  mean. 


Since  1983,  mean  annual  CPUEL  has  increased  in  all 
MFAs,  and  has  been  consistently  higher  in  MFAs  56 
and  58  than  in  other  areas  (Fig.  5A).  CPUE()  has  varied 
among  years  but  has  followed  similar  trends  in  different 
MFAs  with  consistent  peaks  in  1993  (Fig  5C).  Prior  to 
1992,  ML  was  similar  among  MFAs  but  after  1992  was 
generally  highest  in  MFAs  56  and  58  (Fig.  5E).  ML  has 
increased  over  time  in  all  MFAs. 


Northern  zone  In  the  NZ,  mean  annual  CPUEL  rose 
from  135  to  179  lobsters/(  100  trap-lifts/day)  between 
1983  and  1991,  decreased  to  138  lobsters/1 100  trap-lifts/ 
day)  in  1993  and  then  rose  again  to  177  lobsters/I  100 
trap-lifts/day)  in  1999  (Fig  3B).  CPUE0  ranged  between 


1.0  and  2.4  octopus/l  100  trap-lifts/day)  in  1987  and  1993 
respectively  (Fig.  3D).  ML  ranged  from  5.0  to  7.3  dead 
lobsters/(100  trap-lifts/day  I  in  1983  and  1988.  respec- 
tively (Fig.  3F). 

Mean  monthly  CPUEL  declined  from  196  to  88  lob- 
ster/100 trap-lifts/day  between  November  and  May 
(Fig.  4B).  Mean  monthly  CPUE0  was  reasonably  constant 
at  between  1.43  and  1.7  octopus/100  trap-lifts/day  for 
the  first  five  months  of  the  season  before  a  decline  to  1.0 
octopus/100  trap-lifts/day  in  May  (Fig.  4D>.  The  mean 
monthly  M,  declined  from  7.5  to  3.4  dead  lobster/100 
trap-lifts/day  between  November  and  May  (Fig.  4Fi. 

Since  1983,  mean  annual  CPUEL  has  been  relatively 
low  and  stable  in  MFAs  15,  28,  and  40  but  has  been 


Brock  and  Ward:  Octopus  bycatch  and  lobster  mortality  in  the  South  Australian  rock  lobster  fishery 


435 


A     lobster  Southern  zone 

600 

500 


-     400 
o 

o 

s;    300 

w 

CD 

«      200 


Oct     Nov    Dec     Jan     Feb     Mar     Apr 


octopus 


in 

%  6 

CL 

ra  5  - 

i     *■ 

~S  3 

CD 

en  o 

=3  ^ 

a. 

2         1 

o 

o 

0 


Oct     Nov    Dec    Jan     Feb    Mar    Apr 

k     dead  lobster 
20 


15 


1 1 1 1 1 1 1 

Oct     Nov    Dec    Jan     Feb    Mar    Apr 


B     lobster  Northern  zone 

600  -, 
500 
400 
300 
200  - 
100 
0 


D 


— i 1 1 1 1 1 1 

Nov    Dec    Jan     Feb    Mar    Apr    May 
octopus 


6 
5 
4 

3  - 
2 

1 

0 


1 1 1 1 1 1 1 

Nov    Dec     Jan     Feb    Mar     Apr    May 


15  - 

10  ■ 

5  - 

0  - 

— i 1 1 1 

Nov    Dec     Jan     Feb     Mar     Apr    May 


Month 


Figure  4 

Mean  monthly  catch  rates  in  each  fishing  zone  for  lobsters  lCPUEL)  (A  and  Bl,  octopus 
(CPUE0)  (C  and  Di,  and  dead  lobsters  (ML)(E  and  F).  Error  bars  are  ±SD  of  mean. 


higher  and  more  variable  in  MFAs  39  and  49  (Fig.  5B). 
There  were  large  interannual  fluctuations  in  CPUE(1 
in  each  MFA,  and  these  trends  were  similar  among 
MFAs  (Fig.  5D).  ML  was  highest  in  MFA  40,  where  a 
maximum  of  12.5  dead  lobsters/l  100  traps  lifts)  was  re- 
corded 1998  and  lowest  in  MFA  15  where  the  maximum 
was  5.2  dead  lobsters/100  trap-lifts  in  1997  (Fig.  5F). 
No  clear  long-term  trends  in  ML  were  apparent  in  any 
MFA. 

Factors  that  affect  within-trap  lobster  mortality 

Based  on  the  mean  square  values,  the  number  of  octo- 
pus had  the  greatest  effect  on  lobster  mortality  in  both 
zones  (Table  1,  A  and  B).  The  number  of  dead  lobsters 


increased  with  both  octopus  and  lobster  catches  and 
with  soak-time  and  decreased  as  depth  increased  (Figs. 
6  and  7).  Based  on  the  relative  size  of  the  mean  square 
values,  the  factor  with  the  greatest  effect  on  the  number 
of  dead  lobsters  in  the  SZ  was  the  number  of  octopus 
caught,  followed  by  soak-time,  number  of  lobsters  caught, 
and  depth.  In  the  NZ,  the  number  of  octopus  caught  was 
also  the  most  important  factor,  followed  by  the  number 
of  lobsters  caught,  depth,  and  soak-time. 

Source  of  lobster  mortality  and  size-dependent  mortality 

A  total  of  3627  lobsters  from  635  trap-lifts  were  mea- 
sured. In  the  sample  there  were  212  lobsters  killed  in 
traps  of  which  207  (98%)  were  killed  by  octopus  and  5 


436 


Fishery  Bulletin  102(3) 


Southern  zone 


-i — i — i — i — i — i 
1995  1999 


octopus 


1983 
li,     dead  lobster 


- 1 — i — i — i — i — i — 
1987  1991 


— i — r — i — i — r- 
1983         1987 


1991 


-MFA  51 
-MFA  56 


1995 

-MFA  55 
MFA  58 


1999 


B  lobster  Northern  zone 
600 
500 
400  - 
300 
200 
100 

0 


1983  1987 

D     octopus 


-1 — I — I — 
1991 


1995 


-T 1 1 1 1 1 1 1 T 

1983         1987         1991  1995 

>_MFA  15        g     MFA  28        a     MFA  39 
MFA  40     _*_MFA  49 


Figure  5 

Annual  catch  rates  of  the  major  MFAs  in  each  fishing  zone  for  lobsters  (CPUEL)  (A  and  Bi. 
octopus  (CPUE0i  i C  and  D),  and  dead  lobsters  (MLl  (E  andF). 


by  other  predators.  The  mean  CL  of  dead  male  lobsters 
was  greater  than  live  males  (120  ±21.1(SD)  vs.  110  ±18.3 
(SD)  mm,  P<0.001).  There  was  no  significant  difference 
in  the  mean  size  of  live  and  dead  female  lobsters.  For 
both  sexes  the  probability  of  mortality  increased  with 
size  according  to  the  following  relationships: 

P  (ML,  males)  =  i/i+e-(-5.04+o.02CL) 
P  <ML,  females)  =  m+e-l-*.i&+o.oiCL)m 

Above  100  mm  CL,  the  probability  of  mortality  increased 
more  sharply  for  male  lobsters  than  for  female  lobsters 
(Fig.  8). 


Discussion 

Logbook  data  from  the  SARLF  show  that  over  the  last 
five  years  approximately  240,000  lobsters  have  been 


killed  in  traps  each  year.  Although  there  are  numerous 
predators  of  trapped  lobsters — such  as  seals,  conger  eels, 
and  several  species  of  finfish — the  impacts  of  these  taxa 
appear  to  be  minor  compared  to  the  effects  of  predation 
by  O.  maorum.  The  field-sampling  program  conducted  in 
the  SZ  in  2001-02  suggested  that  over  98%  of  within- 
trap  mortality  was  attributable  to  O.  maorum.  Although 
the  sampling  program  was  spatially  and  temporally 
restricted,  this  finding,  in  conjunction  with  the  strong 
correlations  between  annual,  seasonal  and  spatial  trends 
in  the  CPUE0  and  M, ,  clearly  demonstrates  that  O. 
maorum  is  the  major  predator  of  lobsters  in  SARLF 
traps. 

The  results  of  this  study  suggest  that  about  4</r  of  the 
total  annual  catch  of  the  SARLF  is  lost  to  predation 
by  O.  maorum  in  traps.  Mortality  rates  attributable  to 
octopuses  in  other  Australian  lobster  fisheries  range 
from  \%  in  the  Western  Australian  fishery  for  Panulirus 
cygnus1  (O.  tetricus)  to  5%  in  the  Tasmania!)  fishery  for 


Brock  and  Ward:  Octopus  bycatch  and  lobster  mortality  in  the  South  Australian  rock  lobster  fishery 


437 


Table  T 

Results  of  the  general  linear  model  of  factors  that  affect  lobster  mortality  (all  data  log  transformed)  for  (A)  southern  zone 
(r2=0.62),  (B)  northern  zone  (r2=0.38l. 


Source 


df 


Squares 


Mean  square 


.F-value 


P>F 


Model 
Error 
Corrected  total 


5319 
483.961 
489,280 


314,536 
194,140 

508,677 


59.13 
0.401 


147.41 


;0.0001 


Source 


df 


Type-3  SS 


Mean  square 


F-  value 


P>F 


License 

MFA 

Year 

MFAxYear 

Month 

Effort 

Lobster  catch 

Depth 

Octopus 

Soak-time 

Soak-time  x  Year 

License  x  Year 

License  x  Month 

Year  x  Month 

Soak-time  *  Month 


245 
3 

17 
51 
6 
1 
1 
1 
1 
1 

17 

3415 

1460 

94 

6 


49158.1 

9.6 

2494.6 

233.0 

2830.7 

229.6 

6728.7 

1335.0 

35930.5 

6842.1 

286.9 

53019.8 

5900.2 

3760.4 

310.2 


200.6 

3.2 

146.7 

4.7 

471.8 

229.6 

6728.7 

1335.0 

35930.5 

6842.1 

16.9 

15.5 

4.0 

40.0 

51.7 


500.2 

8.0 

365.8 

11.6 

1176.1 

572.4 

16773.4 

3327.9 

89568.8 

17056.1 

42.1 

38.7 

10.1 

99.7 

128.9 


<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 


H 


Source 


df 


Squares 


Mean  square 


F-value 


P>F 


Model 
Error 
Corrected  total 


2159 
148,731 
150,890 


39,217 

64,713 

103,931 


18.17 
0.435 


41.75 


<0.0001 


Source 


df 


Type-3  SS 


Mean  square 


F-value 


P>F 


License 

MFA 

Year 

MFAxYear 

Month 

Effort 

Lobster  catch 

Depth 

Octopus 

Soak-time 

Soak-time  x  Year 

License  x  Year 

License  x  Month 

Year  x  Month 

Soak-time  x  Month 


95 
4 

17 

68 
7 
1 
1 
1 
1 
1 

17 

1287 

553 

95 
6 


3361.5 
174.8 
241.2 
175.4 
317.3 
27.1 

1299.7 
391.3 

6305.1 
210.8 
117.7 

7275.6 

1170.7 

275.9 

2.8 


35.4 

43.7 

14.2 

2.6 

45.3 

27.1 

1299.7 

391.3 

6305.1 

210.8 

6.9 

5.7 

2.1 

2.9 

0.5 


81.3 

100.4 

32.6 

5.9 

104.2 

62.3 

2987.2 

899.3 

14491.0 

484.4 

15.9 

13.0 

4.9 

6.7 

1.1 


<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
<0.0001 
0.3743 


438 


Fishery  Bulletin  102(3) 


A    depth                                                                                D     soak  time 

1  ♦                                                                                      41 

0  8  ■ 
0.6  - 

3  - 
* 

2- 

*  • 

» 

</>       0  4- 

***■"  '''»»•  |  M  |  [J,                                   1    " 

♦ 

0) 

3       0.2  - 

■DO                     50                   100                  150                          0                      5                      10                    15 

-g                                   Depth  (m)                                                              Soak  time  (days) 

o 

o      O    octopus                                                                            D    lobster 

a        5  -, 

10  i                                      im 

CO 

a>         4  - 

„♦»♦♦*                        8. 

^^^^ 

QC 

3  - 

♦  ♦** 

yS^ 

2  - 

♦  *# 

f 

♦ 

♦■ 

1   - 

♦                                                                                                  2  -, 

0                 5                10              15              20                            0              200            400            600            800 

No.  of  octopuses                                                           No.  of  lobsters 

Figure  6 

The  relative  number  of  dead  lobster  as  a  function  of  lAl  depth,  iBi  soak-time,  (C)  number  of 

octopus,  and  (D)  no.  of  lobsters  for  the  southern  zone.  Error  bars  are  ±SD  of  mean. 

<D 

or. 


1  -< 

A     depth 

0  8  - 

♦ 

0.6  - 

* 
* 

0.4  - 

"*"~* 

0.2  - 

50  100  150 

Depth  (m) 


octopus 


5 

4  - 
3 
2 

1  H 


.♦♦ 


♦  ♦ 


*♦< 


- 1 1 1 1 

5  10  15  20 


No.  of  octopuses 


4  -, 
3.5 

3  - 
2.5  - 

2 
1.5  - 

1 
0  5 

0 


B 


soak  time 


5  10 

Soak  time  (days) 


15 


I)    lobster 
10  -, 


100  200  300 

No.  of  lobsters 


400 


Figure  7 

The  relative  number  of  dead  lobster  as  a  function  of  (A)  depth.  (Bl  soak-time,  (C)  number  of 
octopus,  and  i  Di  number  of  lobsters  for  the  northern  zone.  Error  bars  are  ±  SD  of  mean. 


Brock  and  Ward:  Octopus  bycatch  and  lobster  mortality  In  the  South  Australian  rock  lobster  fishery  439 


0.4  -I 

0.35  • 

1        °3' 

/ 

2      0  25- 
o 

°-        0  2  ■ 

>. 

1       0  15- 

t: 

o          0  1- 
0  05  - 

male    j$ 
J2jMlsfrm*F*F^^               female 

60            80           100          120          140          160          180         200 

Carapace  length  (mm) 

Figure  8 

Size-dependent  mortality  of  lobsters  with  respect  to 

sex. 

J.  edwardsii  (O.  maorum),  (Gardener5),  and  a  localized 
study  in  the  New  Zealand  fisheries  for  J.  edwardsii 
(O.  maorum)  found  the  proportion  of  the  lobster  catch 
killed  by  octopus  to  be  as  high  as  10%  (Ritchie6).  The 
estimates  of  lobster  mortality  from  these  other  studies 
should  be  treated  with  caution  however  because  the  cur- 
rent study  is  the  only  one  that  documents  within-trap 
lobster  mortality  from  a  fishery-wide  data  set. 

The  general  linear  modeling  approach  that  we  used  to 
determine  the  factors  associated  with  ML  has  some  limi- 
tations. For  example,  the  logbook  data  for  the  SARLF, 
like  the  monitoring  data  for  most  other  fisheries,  are 
not  completely  independent,  and  interdependence  among 
observations  can  bias  estimates  of  parameters.  Simi- 
larly, some  of  the  factors  in  the  model,  notably  CPUEL 
and  CPUE0  are  partially  correlated.  In  addition,  the 
large  number  of  observations  and  degrees  of  freedom 
tend  to  make  most  factors  significant.  We  considered 
all  of  these  issues  when  interpreting  the  results  of  the 
analyses  and  used  the  mean  square  (MS)  values  to  rank 
the  importance  of  factors. 

In  both  zones,  inter-  and  intra-annual  fluctuations 
in  ML  largely  reflect  the  effects  of  CPUE0  and  CPUEL. 
The  broad  trends  in  annual  CPUE0  have  largely  cor- 
responded to  those  for  M,  with  peaks  in  both  generally 
synchronous  in  both  fishing  zones.  In  the  SZ,  the  gen- 
eral increase  in  ML  since  1983  appears  to  result  from 
the  increase  in  CPUE,  which  has  more  than  doubled 
over  this  period.  This  assessment  is  supported  by  catch- 
rate  data  from  individual  MFAs.  The  two  MFAs  in  the 
SZ  that  have  had  the  greatest  increases  in  CPUEL  over 
the  last  5  years  (56  and  58)  have  also  had  the  highest 


5  Gardener,  C.  2002.  Personal  commun.  Tasmanian  Aqua- 
culture  and  Fisheries  Institute,  Private  Bag  49,  Hobart, 
Tasmania  7001. 

6  Ritchie.  L.  D.  1972.  Octopus  predation  on  trap-caught  rock 
lobster — Hokianga  area,  N.Z.  September-October  1972.  New 
Zealand  Marine  Department.  Fisheries  Technical  Report  81, 
40  p.  (Available  from  Ministry  of  Fisheries,  101-103  The 
Terrace,  Wellington,  New  Zealand,  1020.] 


corresponding  increase  in  ML.  Increases  in  CPUEL  are 
likely  to  elevate  ML  by  both  increasing  the  probability 
of  octopus  encountering  traps  containing  lobsters  and 
the  number  of  lobsters  in  traps  entered  by  octopus. 

However,  ML  is  also  positively  correlated  with  soak- 
time,  especially  in  the  SZ.  This  finding  is  consistent 
with  patterns  observed  in  the  New  Zealand  fishery  for 
J.  edwardsii6  and  reflects  the  increased  opportunities 
for  octopus  predation  when  pots  containing  lobsters 
remain  in  the  water  for  longer  periods.  In  the  SZ,  fish- 
ermen return  to  port  each  day  and  choose  to  fish  or  not 
to  fish  each  day  according  to  factors  such  as  weather 
and  price;  therefore,  although  a  24-h  soak  period  is  still 
most  common,  soak  times  can  range  from  one  to  five 
days.  In  the  NZ,  fishermen  remain  at  sea  for  extended 
periods  and  consequently  soak  times  longer  than  24 
hours  are  rare. 

There  was  considerable  variation  in  the  fishery  data, 
especially  in  the  southern  zone.  It  is  likely  that  much 
of  this  variation  is  related  to  the  large  geographical 
extent  of  the  fishery  as  opposed  to  fishing  practises. 
Across  the  fishery  lobster  growth  rates  and  subsequent 
catch  rates  vary  greatly  (McGarvey  et  al.,  1999a).  For 
example,  since  1991,  the  CPUEL  in  MFAs  56  and  58 
have  been  twice  those  of  MFAs  51  and  55  in  the  SZ. 
Although  the  variation  in  CPUE0  between  the  zones 
has  been  similar,  the  higher  variability  in  CPUEL  in 
the  SZ  is  reflected  in  the  variation  in  ML. 

Data  spanning  17  years  and  covering  about  50.000 
km2  represent  one  of  the  few  long-term  and  large-scale 
data  sets  on  the  distribution  and  abundance  of  an  octo- 
pus species  (Hernandez-Garcia  et  al.,  1998:  Quetglas  et 
al.,  1998).  The  paucity  of  octopus  studies  on  these  scales 
reflects  the  logistical  constraints  of  fishery-independent 
surveys  of  octopus  populations  and  the  poor  and  incon- 
sistent methods  generally  used  to  record  fishery  catch 
and  effort  data  (Boyle  and  Boletsky,  1996).  The  few  data 
that  are  available  on  the  distribution  patterns  of  octo- 
pus have  been  obtained  mainly  from  small  commercial 
fisheries  and  CPUE(,  has  been  included  as  a  measure 
of  relative  abundance  (Defeo  and  Castilla,  1998;  Her- 
nandez-Garcia et  al.,  1998).  This  approach  has  proven 
useful,  but  several  potential  biases  must  be  considered 
when  CPUE0  data  are  being  interpreted:  these  include 

1)  changes  in  fishing  methods  and  efficiency  over  time; 

2)  the  distribution  pattern  (e.g.,  random  or  aggregated) 
of  the  species  under  consideration;  and  3)  spatiotempo- 
ral  fluctuations  in  catchability  (Richards  and  Schnute, 
1986;  Rose  and  Kulka,  1999).  There  are  several  reasons 
why  the  data  from  the  SARLF  may  provide  a  useful 
measure  of  the  relative  abundance  of  octopus  over  these 
spatial  and  temporal  scales.  Most  importantly,  the  ba- 
sic unit  of  effort  in  the  fishery,  the  trap,  has  remained 
unchanged  since  1983.  Furthermore,  although  O.  mao- 
rum is  retained  as  bycatch  and  kills  J.  edwardsii  in 
traps,  it  is  neither  targeted  nor  avoided  by  fishermen, 
and  fishing  effort  is  thus  relatively  independent  of  its 
distribution  patterns  because  the  economic  effects  of 
both  the  sale  of  octopus  bycatch  and  the  costs  of  lobster 
predation  are  relatively  small  compared  to  the  primary 


440 


Fishery  Bulletin  102(3) 


economic  driving  force  for  the  fishery,  the  lobster  catch 
rates,  and  because  the  catch  rates  of  octopus  are  dif- 
ficult to  predict.  In  addition,  O.  maorum  is  a  solitary 
animal  that  tends  to  be  dispersed  randomly  throughout 
areas  of  suitable  habitat  (Mather  et  al.,  1985). 

The  higher  total  catches  and  catch  rates  of  both  lob- 
ster and  octopus  in  the  SZ,  compared  to  the  NZ,  prob- 
ably reflect  the  more  extensive  reef  habitat  and  more 
intense  nutrient-enrichment  upwelling  in  this  portion 
of  the  SARLF  (Lewis,  1981).  There  have  been  large 
interannual  fluctuations  in  CPUE0  in  both  zones  since 
1983.  Such  fluctuations  in  population  size  are  common 
among  other  cephalopods,  especially  squid,  and  may 
result  from  life  history  strategies  that  are  characterized 
by  rapid  growth,  short  lifespan  (<two  years)  and  almost 
universal  mortality  after  a  single  spawning  event  (Boyle 
and  Boletsky,  1996).  Despite  these  fluctuations,  CPUE0 
has  not  declined  noticeably  in  any  MFA  since  1983, 
which  suggests  that  octopus  mortality  from  fishing  is 
consistent  with  little  impact  on  octopus  populations 
since  the  advent  of  fishing.  This  observation  and  the 
poor  relationship  between  octopus  catches  and  effort  re- 
fute the  belief  of  some  SARLF  fishermen  that  incidental 
fishing  mortality  acts  to  control  octopus  abundance. 

This  study,  however,  did  confirm  the  view  of  fisher- 
men that  larger  lobsters  are  killed  more  commonly  by 
octopus  than  are  smaller  ones.  This  effect  was  most 
evident  for  male  lobsters,  which  grow  to  larger  sizes 
than  females.  There  could  be  several  reasons  for  the 
size-dependent  mortality  rates  for  rock  lobsters.  For 
example,  octopus  could  actively  select  larger  prey,  or 
large  lobsters  could  be  captured  more  easily  than  small 
lobsters  in  traps  by  octopus.  Because  large  lobsters  can 
be  worth  more  and  produce  more  eggs  than  smaller 
lobsters,  the  increased  mortality  rates  of  large  lobsters 
suggest  that  the  total  economic  and  ecological  impacts 
of  octopus  predation  in  the  SARLF  are  greater  than 
indicated  by  the  absolute  number  of  lobsters  killed. 

Octopus  predation  of  lobsters  in  traps  is  a  significant 
problem  in  the  SARLF.  However,  the  economic  effects 
vary  between  the  zones.  In  the  quota-managed  SZ,  ad- 
ditional lobsters  are  harvested  to  replace  those  killed 
in  traps,  which  increases  the  time  and  costs  of  catch 
quotas,  and  imposes  an  impact  on  lobster  abundance. 
In  the  input-controlled  NZ,  where  there  is  no  direct 
restriction  on  the  quantity  of  lobsters  taken,  lobsters 
killed  in  traps  represent  both  a  direct  economic  loss 
and  an  impact  on  lobster  abundance. 

Like  most  fisheries  for  spiny  lobsters,  the  SARLF  is 
close  to  being  fully  exploited.  Reducing  rates  of  octopus 
predation  provides  one  option  available  for  increas- 
ing the  value  of  the  fishery.  Some  minor  reductions  in 
lobster  mortality  may  be  achieved  by  minimizing  soak- 
times,  especially  in  the  SZ.  More  significant  reductions 
in  the  rates  of  within-trap  lobster  mortality  may  be 
achieved  by  redesigning  lobster  traps  (Brock  et  al.4). 


Acknowledgments 

This  research  was  funded  jointly  by  the  Fisheries 
Research  and  Development  Corporation,  South  Austra- 
lian Rock  Lobster  Advisory  Council,  South  Australian 
Research  and  Development  Institute  (Aquatic  Sciences), 
and  the  University  of  Adelaide.  The  authors  thank  Thor 
Saunders  for  his  support  and  assistance  during  the 
research.  Jim  Prescott  provided  advice  for  extracting 
data  from  the  South  Australian  Rock  Lobster  Database. 
Yongshun  Xiao  provided  statistical  advice  and  assisted 
with  the  numerical  modeling. 


Literature  cited 

Boyle,  R.  R,  and  S.  V.  Boletsky. 

1996.     Cephalopod  populations:  definition  and  dynam- 
ics.    Phil.  Trans.  R.  Soc.  Lond.  351:985-1002. 
Defeo,  O.,  and  J.  C.  Castilla. 

1998.     Harvesting  and  economic  patterns  in  the  artisanal 
Octopus  mimus  (Cephalopodal  fishery  in  a  northern 
Chile  cove.     Fish.  Res.  38:121-130. 
Hernandez-Garcia,  V.,  J.  L.  Hernandez-Lopez,  and  J.  J.  Castro. 
1998.     The  octopus  lOctopus  vulgaris)  in  the  small-scale 
trap  fishery  off  the  Canary   Islands  (Central-East 
Atlantic).     Fish.  Res.  35:183-189. 
Lewis,  R.  K. 

1981.     Seasonal  upwelling  along  the  south-eastern  coast- 
line of  South  Australia.     Aust.  J.  Mar.  Freshw.  Res. 
32:843-854. 
Mather.  J.  A.,  S.  Resler,  and  J.  Cosgrove. 

1985.  Activity  and  movement  patterns  of  Octopus 
dofleini.     Mar.  Behav.  Physiol.  11:301-314. 

McGarvey,  R.,  G.  Ferguson,  and  J.  H.  Prescott. 

1999a.     Spatial  variation  in  mean  growth  rates  of  rock  lob- 
ster, Jasus  edwardsii,  in  South  Australian  waters.     Mar. 
Freshw.  Res.  50:333-342. 
Phillips.  B.  F.,  J.  S.  Cobb,  and  J.  Kittaka. 

1994.     Spiny  lobster  management,  550  p.     Blackwell 
Scientific,  London. 
Quetglas,  A.,  F.  Alemany,  A.  Carbonell,  P.  Merella,  and 
P.  Sanchez. 

1998.  Biology  and  fishery  of  Octopus  vulgaris  Cuvier, 
1797,  caught  by  trawlers  in  Mallorca  (Balearic  Sea, 
Western  Mediterranean).     Fish.  Res.  36:237-249. 

Richards.  L.  J.,  and  L.  J.  Schnute. 

1986.  An  experimental  and  statistical  approach  to  the 
question:  Is  CPUE  an  index  of  abundance?  Can.  J. 
Fish.  Aquat.  Sci.  43:1214-1227. 

Rose,  G.  A.,  and  D.  W.  Kulka. 

1999.  Hyperaggregation  offish  and  fisheries:  how  catch- 
per-unit-effort  increased  as  the  northern  cod  iGadus 
morhua)  declined.  Can.  J.  Fish.  Aquat.  Sci.  56(suppl. 
11:118-127. 

Williams.  A.  B. 

1988.  Lobsters  of  the  world — an  illustrated  guide,  186 
p.     Osprey  Books.  New  York,  NY. 


441 


Abstract  — Management  of  coastal 
species  of  small  cetaceans  is  often 
impeded  by  a  lack  of  robust  estimates 
of  their  abundance.  In  the  Austral 
summers  of  1997-98,  1998-99,  and 
1999-2000  we  conducted  line-transect 
surveys  of  Hector's  dolphin  iCepha- 
lorhynchus  hectori)  abundance  off  the 
north,  east,  and  south  coasts  of  the 
South  Island  of  New  Zealand.  Survey 
methods  were  modified  for  the  use  of 
a  15-m  sailing  catamaran,  which  was 
equipped  with  a  collapsible  sighting 
platform  giving  observers  an  eye- 
height  of  6  m.  Eighty-six  percent  of 
2061  km  of  survey  effort  was  allo- 
cated to  inshore  waters  (4  nautical 
miles  [nmi]  or  7.4  km  from  shore), 
and  the  remainder  to  offshore  waters 
(4-10  nmi  or  7.4-18.5  km  from  shore). 
Transects  were  placed  at  45°  to  the 
shore  and  spaced  apart  by  1,  2,  4,  or  8 
nmi  according  to  pre-existing  data  on 
dolphin  density.  Survey  effort  within 
strata  was  uniform.  Detection  func- 
tions for  sheltered  waters  and  open 
coasts  were  fitted  separately  for  each 
survey.  The  effect  of  attraction  of  dol- 
phins to  the  survey  vessel  and  the 
fraction  of  dolphins  missed  on  the 
trackline  were  assessed  with  simul- 
taneous boat  and  helicopter  surveys 
in  January  1999.  Hector's  dolphin 
abundance  in  the  coastal  zone  to  4 
nmi  offshore  was  calculated  at  1880 
individuals  (CV=15.7%,  log-normal 
95%  01  =  1384-2554).  These  surveys 
are  the  first  line-transect  surveys  for 
cetaceans  in  New  Zealand's  coastal 
waters. 


Small-boat  surveys  for  coastal  dolphins: 
line-transect  surveys  for  Hector's  dolphins 
(Cephalorhynchus  hectori) 


Stephen  Dawson1 
Elisabeth  Slooten2 
Sam  DuFresne' 
Paul  Wade3 
Deanna  Clement2 

1  Department  of  Marine  Science 
University  of  Otago 
340  Castle  Street 
Dunedm,  New  Zealand 
E-mail  address  stevedawsoofastonebow  otago  acnz 

•"Zoology  Department 
University  of  Otago 
340  Castle  Street 
Dunedm,  New  Zealand 

3  National  Marine  Mammal  Laboratory 
National  Marine  Fisheries  Service 
7600  Sand  Point  Way  NE 
Seattle,  Washington  98115 


Manuscript  submitted  27  April  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
3  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  201:441-451  (2004). 


Several  international  workshops  on 
cetacean  bycatch  problems  have  stated 
that  a  key  impediment  to  the  conser- 
vation of  coastal  and  riverine  small 
cetaceans  is  the  lack  of  quantitative 
data  on  abundance  (e.g.,  IWC,  1994). 
An  important  reason  for  this  lack  of 
data  is  that  line-transect  surveys  are 
often  conducted  from  large  (>50  m) 
vessels  (e.g.  Barlow,  1988)  and  hence 
are  extremely  expensive  ($US  10,000/ 
day).  Such  costs  usually  put  high-qual- 
ity surveys  such  as  those  conducted 
for  harbor  porpoise  in  the  U.S.  (e.g., 
Carretta  et  al„  2001)  beyond  the  reach 
of  less  affluent  nations.  The  need  for 
abundance  estimates  is  especially 
great  for  the  coastal  and  riverine  spe- 
cies found  in  Asia,  Africa,  Australasia, 
and  South  America  (Table  1).  Several 
of  these  species  have  apparently  small 
populations  and  restricted  distribu- 
tions, and  all  suffer  from  being  taken 
as  bycatch  in  fishing  gear,  principally 
in  gill  nets  (IWC,  1994).  In  addition,  it 
is  difficult  or  impossible  for  large  ves- 
sels to  work  close  to  shore,  in  shallow 
waters,  where  some  of  these  species 
are  most  common. 


The  work  described  in  this  contri- 
bution had  two  aims:  1)  to  adapt  ship- 
based  line-transect  methods  (e.g., 
Barlow,  1988)  to  a  15-m  catamaran, 
and  2 )  to  provide  an  updated  estimate 
of  the  abundance  of  Hector's  dolphin 
(Cephalorhynchus  hectori).  Hector's 
dolphin,  a  small  delphinid  found 
only  in  the  inshore  waters  of  New 
Zealand,  is  subject  to  bycatch  in  gill 
nets  throughout  its  range  (Dawson  et 
al.,  2001).  At  least  in  the  Canterbury 
region,  and  off  the  North  Island  west 
coast,  recent  catch  levels  are  clearly 
unsustainable  (Dawson  and  Slooten, 
1993;  Martien  et  al.,  1999;  Slooten  et 
al.,  2000;  Dawson  et  al..  2001).  Stud- 
ies of  mt-DNA  indicate  that  the  very 
small  North  Island  population  is  dis- 
tinct and  that  there  are  at  least  three 
separate  populations  in  South  Island 
waters  (Pichler  et  al.,  1998;  Pichler 
and  Baker,  2000;  see  also  Baker  et 
al.,  2002).  At  the  time  of  the  present 
study  the  only  quantitative  population 
estimate  was  from  a  strip-transect 
survey  conducted  in  1984-85  (Daw- 
son and  Slooten,  1988),  in  which  the 
offshore  distribution,  as  well  as  the 


442 


Fishery  Bulletin  102(3) 


proportion  of  dolphins  detected  within  the  strip,  was 
estimated.  A  current,  more  robust  estimate  is  needed 
for  management.  This  study  describes  line-transect 
boat  surveys  conducted  to  estimate  Hector's  dolphin 
abundance  on  the  north,  east,  and  south  coasts  of  the 
South  Island  of  New  Zealand. 


Figure  1 

Photograph  of  the  observer  platform  on  the  catamaran  Cat 


Materials  and  methods 

Vessel  choice  and  field  methods 

Displacement  catamarans  are  inherently  suitable  for 
inshore  surveys  because  of  their  resistance  to  rolling 
and  their  ability  to  sustain  reasonably  high  cruising 
speeds  with  modest  power.  We  based  our 
surveys  from  a  15.3-m  sailing  catama- 
ran (RV  Catalyst),  which  is  powered  by 
two  50-hp  diesel  engines,  and  cruises 
at  9-10  knots  while  using  <10  liters  of 
fuel  per  hour.  We  fitted  a  collapsible 
aluminum  sighting  platform  (~6  m  eye 
height;  Fig.  1)  to  increase  the  resolution 
with  which  observers  could  measure  the 
downward  angles  to  sightings  (see  Lerc- 
zak  and  Hobbs,  1998.  for  details)  and  to 
allow  the  observers  to  see  animals  far- 
ther away.  The  surveys  were  conducted 
with  a  crew  of  six  (five  observers,  one 
skipper). 

Three  people  stood  on  the  platform  at 
any  given  time;  one  scanned  the  surface 
waters  to  the  right  of  the  platform,  and 
the  other  scanned  to  the  left,  and  a  third 
person  (the  recorder)  recorded  sightings 
dictated  by  the  observers.  Sightings 
made  by  the  recorder  were  not  used  in 
our  analyses  because  his  or  her  sight- 
\Jj   ^  ing  effort  was  unavoidably  uneven  (the 

recorder  could  not  make  sightings  while 
recording  another  sighting).  The  record- 
er did  not  point  out  sightings  to  observ- 
ers. Observers  and  data  recorder  rotated 


alyst 


Table  1 

Examples  of  coastal  and  riverine  species  of  special  conservation  concern. 

Common  name 

Scientific  name 

Habitat 

Vaquita 

Phoeoena  sinus 

Northern  Gulf  of  California 

Chilean  dolphin 

Cephalorhynchus  eutropia 

Inshore  coastal  Chile 

Hector's  dolphin 

<  'ephalorhynchus  hectori 

Inshore  coastal  New  Zealand 

Commerson's  dolphin 

Cephalorhynchus  commersoni 

Inshore  coastal  Chile,  Argentina,  Falkland  Is,  Kerguelen  Is. 

Heaviside's  dolphin 

( 'ephalorhynchus  heavisidii 

Inshore  coastal  South  Africa  and  Namibia 

Peale's  dolphin 

Lagenorhynchus  australis 

Coastal  Chile.  Argentina,  Falkland  Is. 

Finless  porpoise 

Neoph  ocoena  phocaenoides 

Coastal  and  riverine  Asia  and  Indonesia 

Indo-Pacific  humpbacked 

Sousa  chinensis 

Inshore  tropical  and  estuarine  habitats  in  western  Pacific 

dolphins 

and  Indo  Pacific 

Burmeister's  porpoise 

Phoeoena  spinipinnis 

Coastal  Chile.  Argentina,  Uruguay.  Brazil 

Franciscana 

Pontoporia  blainvillei 

Coastal  Brazil  and  Argentina 

Indus  river  dolphin 

Platanista  minor 

Indus  River 

Ganges  river  dolphin 

Platanista  gangetica 

Ganges,  tiramaputra.  Karnphuli,  Meghna  rivers 

Boto 

Inia  geoffrensis 

Amazon  River 

Tucuxi 

Solatia  fluviatilus 

Coastal  and  estuarine  Atlantic  Central  and  South  America 

Dawson  et  al.:  Line-transect  surveys  of  Cepha/orhynchus  hectori 


443 


every  30  minutes  to  avoid  fatigue.  Although  Hector's 
dolphins  are  easily  identified  from  other  species,  and 
group  size  is  typically  small  (usually  2-8;  Dawson  and 
Slooten,  1988),  in  order  to  maintain  even  sighting  effort 
on  both  sides  of  the  trackline,  observers  did  not  confer 
during  a  sighting.  Sighting  information  was  entered  into 
a  custom-written  program  on  a  Hewlett-Packard  200LX 
palmtop  computer  on  the  sighting  platform.  Data  record- 
ed included  horizontal  sighting  angle,  downward  angle 
to  sighting  (in  reticles),  species,  group  size,  orientation 
of  the  animals  when  first  sighted,  depth,  Beaufort  sea 
state,  swell  height,  glare,  GPS  fix,  date,  and  time.  The 
program  also  recorded  survey  effort  by  storing  a  GPS  fix 
every  60  seconds.  Weather  conditions  were  recorded  at 
the  start  of  field  effort,  and  whenever  they  changed. 

Observers  used  reticle-  and  compass-equipped  Fujinon 
7x50  (WPC-XL)  binoculars  to  make  sightings  and  to 
measure  the  downward  angle  from  the  land,  or  horizon, 
to  the  sighting.  If  the  former,  the  corresponding  dis- 
tance to  land  was  measured  with  RADAR  (Furuno  1720 
model),  or,  if  within  a  few  hundred  meters  of  shore, 
with  a  Bushnell  lightspeed  laser  rangefinder  (tested 
accuracy  ±1  m  from  12  to  800  m).  We  calibrated  the  ac- 
curacy of  the  RADAR  by  comparison  with  transit  fixes 
and  laser  rangefinder  measurements.  Sighting  angles 
were  recorded  by  using  angle  boards  (see  Buckland  et 
al.,  1993)  in  the  first  season,  and  thereafter  with  the 
compasses  in  the  binoculars.  There  were  no  ferrous 
metals  or  significant  electrical  fields  within  6  m  of  the 
sighting  platform. 

Navigation  was  facilitated  by  the  use  of  a  Cetrek  343  GPS 
chartplotter  with  digitized  C-MAP  charts  onto  which 
transect  waypoints  were  plotted.  Depths  were  measured 
with  a  JRC  JFV-850  echosounder  (at  200  kHz). 

At  the  start  of  each  survey,  several  days  were  spent 
training  observers  at  Banks  Peninsula,  where  sighting 
rates  are  high.  Training  continued  until  we  gained 
about  100  sightings  (data  gathered  in  this  period  were 
not  used  in  the  analyses).  An  observer  manual  (avail- 
able from  authors)  specified  scanning  behavior  and 
recording  methods.  To  ensure  a  wide  shoulder  on  the 
histograms  of  perpendicular  sighting  distances,  observ- 
ers were  instructed  to  concentrate  their  effort  within 
45°  of  the  trackline  and  to  spend  less  time  searching 
out  to  90°.  Observers  spent  about  85%  of  the  time  scan- 
ning with  binoculars.  Regular  scans  with  the  naked  eye 
minimized  fatigue  and  reduced  the  chance  of  missing 
groups  close  to  the  boat.  To  promote  consistency,  observ- 
ers were  asked  to  re-read  the  manual  at  least  once  a 
week  throughout  the  survey. 

While  the  survey  was  underway,  exploratory  data 
analyses  were  undertaken  to  assess  data  quality.  These 
analyses  showed  that  in  the  early  stages  of  the  first  sur- 
vey, observers  were  rounding  angles  of  sightings  close  to 
the  trackline  to  zero.  The  use  of  the  angle  boards  was 
modified  to  minimize  this  problem,  and  they  were  not 
used  in  subsequent  surveys.  The  data  from  these  early 
lines  were  discarded  and  the  survey  lines  repeated. 

Survey  effort  was  restricted  to  sea  conditions  of  Beau- 
fort 3  or  less  and  swell  heights  of  <2  meters.  Transect 


lines  were  run  down-swell  and  down-sun  to  minimize 
pitching  and  effects  of  glare.  Deviations  of  up  to  10°  from 
the  intended  course  were  made  if  needed  to  further  re- 
duce pitching  or  glare.  The  inshore  end  of  each  line  was 
surveyed  to  just  outside  the  surf  zone  on  open  coasts, 
or  until  a  2  m  depth  was  reached,  or  to  within  50  m 
of  rocky  shores.  All  surveys  were  conducted  in  passing 
mode  to  minimize  the  extent  of  vessel  attraction. 

Line-transect  data  were  collected  in  three  surveys  in 
three  consecutive  summer  seasons,  each  focussing  on 
a  particular  coastal  area  (Fig  2;  Motunau  to  Timaru, 
5  January-21  February  1998;  Timaru  to  Long  Point, 
9  December  1998-16  February  1999;  Farewell  Spit  to 
Motunau,  17  December  1999-28  January  2000). 

Survey  design 

In  order  to  obtain  a  clear  picture  of  density  and  to  mini- 
mize variance  in  encounter  rate,  Buckland  et  al.  (1993) 
recommend  placing  transects  across  known  density 
gradients.  Because  short-distance,  alongshore  move- 
ments are  well-known  for  Hector's  dolphins  (Slooten  and 
Dawson,  1994:  Brager  et  al.,  2002)  and  the  dolphins' 
density  declines  sharply  with  distance  offshore  (Dawson 
and  Slooten,  1988),  transects  were  placed  at  45°  to  the 
coast.  On  curved  coastlines  (within  strata)  we  divided 
the  coastline  into  blocks,  drew  an  imaginary  baseline 
along  the  coast,  and  placed  lines  at  45°  to  that  baseline. 
The  starting  point  of  the  first  line  along  the  baseline 
was  decided  randomly;  thereafter  lines  were  spaced  at 
regular  intervals  according  to  the  sampling  intensity 
required  in  that  stratum  (Fig.  2).  Within  harbors  we 
placed  lines  at  45°  to  an  imaginary  line  down  the  center 
of  the  harbor  (Fig.  3).  The  aim  of  this  scheme  was  to 
ensure  that,  within  a  stratum,  any  one  point  had  the 
same  chance  of  being  sampled  as  any  other. 

Survey  effort  was  stratified  according  to  existing  data 
on  distribution,  obvious  habitat  differences,  and  areas 
of  intrinsic  management  interest.  In  summer,  very  few 
Hector's  dolphins  are  seen  beyond  four  nmi  from  shore 
(Dawson  and  Slooten,  1988);  therefore  most  sampling 
effort  was  placed  in  this  inshore  zone  (i.e.  45°  lines  at 
2-,  4-,  or  8-nmi  spacings,  approximately  proportional  to 
density  as  determined  from  previous  surveys).  Within 
harbors,  transect  spacings  were  either  one  or  two  nau- 
tical miles.  In  the  offshore  zone  (from  4  to  10  nmi)  we 
expected  very  low  densities,  and  therefore  used  sparse 
transect  spacing  (-30  nmi  apart).  It  was  not  our  inten- 
tion to  estimate  density  in  this  offshore  zone.  A  subse- 
quent aerial  survey  was  found  to  be  better  suited  for 
this  purpose  (Rayment  et  al.1). 

Our  goal  was  to  estimate  effective  half  strip  width 
(ESW)  separately  for  strata  with  different  exposure 
to  wind  and  swell.  Hence,  in  each  survey  we  aimed  to 
gain  sufficient  sightings  to  estimate  ESW  separately  for 
harbors  or  protected  waters,  and  open  coasts.  To  reach 


1  Rayment,  W.,  E.  Slooten,  and  S.  M.  Dawson.  2003.  Unpubl. 
data.  Department  of  Marine  Science,  Univ.  Otago,  P.O.  Box 
56,  Dunedin,  New  Zealand. 


444 


Fishery  Bulletin  102(3) 


172  55'S 

N 

/        /, 

t 

X//     -'  •  "' 

>       '^\\    / 

^  -M-- 

^ 

(\\    o 

'>&°x£     W 
„ov-  * 

Akaroa 

(^  yp%%y\ 

lx^fv>^ 

^    *#3? 

(^NS.     o^Jy 

J3_5P!§  _>_   -\febO    - .2*C  - 



x/^ 

/\\  /      /A 

^  J^-^j, 

6  It      it 

\J\l\ 

- 

wrr 

-_T-           ;  ■ 

1  km 

Figure  2 

Map  of  New  Zealand's  South  Island,  showing  transect  lines  and  sightings 
of  Hector's  dophins  (dots)  1997-2000. 


Figure  3 

Example  of  transect  layout  in  harbors  1 1997- 
98  Akaroa  Harbor  transect  lines  and  sight- 
ings, showing  three  replicates  I. 


Buckland  et  al.'s  (1993)  target  of  60-80  detections  for 
robust  ESW  estimation,  in  the  1997-98  survey  we  con- 
ducted replicate  surveys  (with  a  new  set  of  lines  each 
time)  in  the  harbors  and  bays  stratum  (e.g.,  Fig.  3). 
Low  sighting  rates  in  the  area  surveyed  in  1999-2000 
would  have  required  unrealistic  effort  levels  to  reach 
this  target;  therefore  we  gained  extra  sightings  from 
areas  with  the  same  exposure  but  higher  sighting  rates 
(e.g.,  data  used  to  calculate  ESW  for  the  Marlborough 
Sounds  were  supplemented  by  data  gathered  in  Akaroa 
Harbour  by  the  same  observers,  in  the  same  summer). 
Hence  different  sample  sizes  were  available  to  estimate 
density  and  ESW  (Table  2).  Because  observers  changed 
between  surveys,  we  did  not  pool  sightings  across  years 
for  estimating  ESW.  Strata  areas  were  measured  from 
nautical  charts  with  a  digital  planimeter. 


Data  analysis 

Within  each  stratum,  Hector's  dolphin  abundance  (Ns) 
was  estimated  as  (Buckland  et  al.,  1993): 

Ns=^^.  (1) 

2LESW 

where  A  =  size  of  the  study  area; 
n    =  number  of  groups  seen; 
S  =  expected  group  size; 
L  =  length  of  transect  line  surveyed,  and 
ESW  =  the  effective  half  strip  width. 

Because  there  was  no  significant  relationship  between 
group  size  and  detection  distance,  expected  group  size 
was  estimated  as  a  simple  mean  group  size. 


Dawson  et  al.:  Line-transect  surveys  of  Cephalorhynchus  hecton 


445 


Table  2 

Survev  effort  by  stratum. 

Number  of  sightings  is  the  total  number  made  before 

truncation 

and  quality 

auditing 

see  "Vessel 

choice  and  field  methods"). 

Survev  effort 

No.  of 

Sightings 

Survey  zone 

Stratum 

(km) 

sightings 

per  km 

Motunau  to  Timaru 

Banks  Peninsula  harbors  and  bavs 

223 

89 

0.399 

(1997-98) 

Banks  Peninsula  Marine  Mammal  Sanctuary  (BPMMSi 

265 

66 

0.249 

(excluding  open  coasts) 

<4  nmi  offshore,  to  the  north  and  south  of  BPMMS 

174 

21 

0.121 

Offshore  (4-10  nmi ) 

89 

4 

0.045 

Timaru  to  Long  Point 

Timaru-Long  Point  (excluding  Te  Waewae  Bay) 

336 

13 

0.04 

(1998-99) 

Te  Waewae  Bay 

101 

14 

0.14 

Offshore  (4-10  nmi) 

106 

0 

0 

Motunau  to  Farewell  Spit 

Farewell  Spit-Stephens  Island 

120 

0 

0 

(1999-20001 

Marlborough  Sounds  (including  Queen  Charlotte  Sound i 

205 

3 

0.015 

Cape  Koamaru-Port  Underwood 

68 

0 

0 

Cloudy  Bay  and  Clifford  Bay 

89 

13 

0.146 

Cape  Campbell-Motunau 

192 

0 

0.026 

Offshore  (4-10  nmi  I 

93 

2 

0.022 

Using  the  program  Distance  3.5  (Research  Unit  for 
Wildlife  Population  Assessment,  University  of  St.  An- 
drews, UK),  we  fitted  detection  functions  to  perpendicu- 
lar distance  data  to  estimate  ESW  (note  that  this  value 
is  derived  directly  from  f(0)).  Akaike's  information  crite- 
rion (AIC)  was  used  to  select  among  models  fitted  to  the 
data.  Models  and  adjustments  were  the  following:  haz- 
ard/cosine, hazard/polynomial,  half-normal/hermite,  half- 
normal/cosine,  uniform/cosine  (Buckland  et  al.,  1993). 
Following  Buckland  et  al.  (1993),  perpendicular  sighting 
distances  were  truncated  to  eliminate  the  farthest  5%  of 
sightings  and  binned  manually  for  /10)  estimation. 

The  coefficient  of  variation  (CV)  for  the  abundance  es- 
timate was  calculated  from  the  coefficients  of  variation 
of  each  variable  element  in  Equation  1  above  (Buckland 
et  al.,  1993): 


CV(N5)=JcV-{)])  +  CV2(S)  +  CV2[ESW). 


(2) 


The  CV(n)  was  estimated  empirically  as  recommended 
by  Buckland  et  al.  (1993): 


CV{n)  = 


vardi) 


(3) 


(4) 


where  var(«)  =  /.£/,<«,  II,-n  I  L):  I  ik-\). 

lj    =  the  length  of  transect  line  i; 

nj  =  the  number  of  sightings  on  transect  i;  and 

k    -  number  of  transect  lines. 


CV(S)  was  estimated  from  the  standard  error  of  the 
mean  group  size.  CVlESW)  was  estimated  with  the 
bootstrapping  option  in  Distance  3.5  software.  This 
process  incorporates  uncertainty  in  model  fitting  and 
model  selection  (Buckland  et  al.,  1993). 


Measuring  the  effect  of  attraction 

Conventional  line-transect  estimates  can  be  biased  as 
a  result  of  responsive  movement  of  the  target  species 
and  animals  on  or  near  the  trackline  being  missed  by 
observers  (Buckland  et  al.,  1993).  Buckland  and  Turnock 
(1992)  presented  a  method  using  co-ordinated  boat  and 
helicopter  surveys  to  quantify  and  adjust  for  the  com- 
bined effects  of  responsive  movements  of  dolphins  to  the 
boat  and  to  eliminate  the  bias  from  observers  failing  to 
see  animals  on  or  near  the  trackline.  Their  approach  is 
better  suited  to  the  restricted  space  available  on  small 
boats  than  to  a  dual-platform  approach  (Palka  and  Ham- 
mond, 2001).  Additionally,  sightings  can  be  made  much 
farther  ahead  (reducing  the  possibility  that  the  animals 
have  already  responded),  and  the  two  sighting  teams 
are  totally  isolated  from  each  other.  For  these  reasons 
we  adapted  Buckland  and  Turnock's  (1992)  approach  in 
our  trials  of  1998-99. 

Simultaneous  boat-and-helicopter  surveys  were  car- 
ried out  to  the  south  of  Banks  Peninsula,  predominantly 
between  Birdlings  Flat  and  the  mouth  of  the  Rakaia 
River.  This  area  was  chosen  because  it  displayed  rep- 
resentative and  varying  densities. 

A  Robinson  R22  helicopter  with  pilot  and  one  observer 
(ES)  followed  a  zig-zag  flight  path  approximately  1.5  km 
in  front  of  the  boat,  traveling  out  to  1000  m  on  either 
side  of  the  vessel's  trackline  at  a  height  of  500  ft  ( 152  m) 
(Fig.  4).  To  aid  the  process  of  tracking  sightings  from 
the  air,  sighting  positions  were  marked  with  Rhodamine 
dye  bombs.2  The  position  of  the  helicopter  in  relation 


Dye  bombs  consisted  of  a  tablespoon  of  Rhodamine  dye  in  a 
paper  cup  2/3  filled  with  sand.  An  additional  (empty)  paper 
cup  was  taped  upside  down  on  top  of  the  first  cup  with 
paper-based  masking  tape.  On  impact  the  two  cups  broke 
apart,  releasing  the  sand+dye  mix  into  the  water. 


446 


Fishery  Bulletin  102(3) 


Figure  4 

Schematic  diagram  of  simultaneous  helicopter-and-boat 
surveys  for  Hector's  dolphins  south  of  Bank  Peninsula, 
South  Island,  New  Zealand. 


to  the  boat  was  determined  with  the  boat's  RADAR. 
The  absolute  position  of  the  boat  was  determined  to 
an  accuracy  of  2-5  m  by  differential  GPS  (Trimble 
GeoExplorer;  postprocessed).  Distances  to  land  were 
obtained  at  the  time  of  sighting  with  RADAR  or  during 
analysis  by  using  GIS  coastline  data  and  the  computer 
program  "SDR  Map"  (Trimble  Navigation,  Christchurch, 
New  Zealand). 

Boat  observers  followed  our  standard  sighting  pro- 
cedures (see  above).  On  most  occasions  the  helicopter 
was  outside  the  field  of  view  of  the  observers'  binoculars 
because  the  observers  were  scanning  the  water  surface, 
and  the  helicopter  was  well  above  what  the  observers 
could  sec  through  the  binoculars.  When  it  was  within 
their  view,  observers  made  a  conscious  effort  to  remain 
unbiased  by  the  movements  of  the  helicopter.  On  mak- 
ing a  sighting,  the  helicopter  observer  informed  an 
independent  observer  located  in  the  cabin  (observers 
on  the  platform  could  not  hear  communications  from 
the  helicopter  observer  and  vice  versa).  The  helicopter 
then  hovered  briefly  above  the  sighting  while  a  range 


and  bearing  in  relation  to  the  boat  was  taken  by  RA- 
DAR. The  helicopter  then  ceased  hovering  but  tracked 
the  group  of  dolphins  either  until  the  boat  observers 
had  sighted  the  group,  or  the  group  had  passed  abeam 
of  the  boat.  A  second  range  and  bearing  were  then 
taken.  Sightings  lost  by  the  helicopter  observer  during 
tracking  were  discarded  in  our  analyses.  The  indepen- 
dent observer,  in  liaison  with  the  helicopter  observer 
and  boat  observers,  determined  whether  the  sighting 
was  a  duplicate  (i.e.,  made  by  both  helicopter  and  boat 
observers)  by  using  information  on  location  and  group 
size.  These  decisions  were  checked  again  in  analysis  by 
inspection  of  plotted  locations  of  sightings  made  from 
either  platform  or  both  platforms. 

Following  the  approach  of  Buckland  and  Turnock 
(1992),  let 

gjy)  =  the  probability  that  a  group  detected  from  the 
helicopter  at  perpendicular  distance  y  from  the 
trackline  of  the  ship  is  subsequently  detected 
from  the  ship; 

fs(y)  =  SsW'H'  with ,»  =  |M  "< y)<b 

(area  under  helicopter  detection  function ). 

w  -  truncation  distance  for  perpendicular  distances 

y; 
nh   =  number  of  helicopter  detections; 
ns  =  number  of  ship  detections; 

nhs  =  number  of  detections  made  from  both  platforms 
(duplicate  detections); 
f/Jy1  =  probability  density  function  fitted  to  helicopter 

detection  distances; 
f'hJy'   =  probability  density  function  fitted  to  duplicate 
detection  distances  as  recorded  from  the  heli- 
copter; 
fix)  =  probability  density  function  fitted  to  perpen- 
dicular distances  recorded  from  the  ship; 
L  =  length  of  transect  line. 

A  conventional  estimate  of  density  of  groups,  assuming 
no  responsive  movement  andglO)  =  1  (all  animals  on  the 
trackline  seen  with  certainty)  is  calculated  as 


IK 


nj(0) 
2L 


(5) 


A  corrected  estimate,  allowing  for  responsive  movement 
and  including  an  estimate  of  g(0)  is  given  by 


where     /s(0)  =  — 


IX 


gs(0) 


»,./,(()) 
214(0)' 


\™gAy)d) 

Ja 


",/ 1 )  I 


(6) 

(7) 

(8) 


Dawson  et  al.:  Line-transect  surveys  of  Cephalorhynchus  hectori 


447 


A  correction  factor  for  abundance  estimates  of  Hector's 
dolphin  groups  can  be  estimated  by 


c  =  DL  ID,. 


(9) 


Using  Distance  3.5,  we  fitted  a  half-normal  model  with 
cosine  adjustments  to  estimate  /10).  The  half-normal 
model  was  fitted  to  helicopter  data  to  estimate  /",,(0)  and 
the  uniform  model  with  cosine  adjustments  was  used  to 
estimate  fhs(0)-  All  were  selected  by  using  AIC.  Potential 
model  choices  were  the  following:  hazard/cosine,  hazard/ 
polynomial,  half-normal/cosine,  half-normal/hermite 
and  uniform/cosine  (Buckland  et  al.,  1993).  Truncation 
distance  was  640  m  for  boat  sightings,  and  1000  m  for 
helicopter  and  duplicate  sightings.  To  ensure  that  only 
high-quality  data  were  used  to  estimate  effective  half 
search  widths,  sightings  for  which  range  (radial  distance) 
was  estimated  by  eye  and  those  made  during  Beaufort 
sea  state  >2  were  removed  before  f(0)  estimation. 

The  error  for  the  correction  factor  (c)  was  estimated 
by  bootstrapping  on  transect  lines  and  applying  the 
estimation  procedure  to  each  of  199  bootstrap  data  sets. 
The  standard  deviation  of  the  bootstrap  estimates  was 
used  as  the  standard  error  of  c. 

Ideally,  the  correction  factor  would  be  estimated  sepa- 
rately for  each  survey  from  separate  sets  of  boat-and-he- 
licopter  trials  conducted  in  areas  of  representative  den- 
sity. Financial  and  logistical  constraints  prevented  this; 
therefore  the  correction  factor  estimated  in  1998-99  was 
applied  to  each  of  the  line-transect  surveys  reported  in 
the  present  study.  We  note  that  this  is  not  uncommon 
(e.g.,  Carretta  et  al.,  2001). 

Unbiased  abundance  estimates  were  calculated  by 


N, 


"■N, 


(10) 


The  CVs  of  the  corrected  abundance  estimates  (NL,) 
were  calculated  with  the  following  equation  (Turnock 
et  al„  1995): 


CV(NU)  =  JCV2(£)  +  CV2(NS), 


where  CV(c) 


5£(r) 


(11) 


(12) 


Upper  (Nuc)  and  lower  (NLC)  95%  confidence  inter- 
vals for  Nv  were  calculated  by  using  the  Satterthwaite 
degrees  of  freedom  procedure  outlined  in  Buckland  et  al. 
(1993).  This  procedure  assumes  a  log-normal  distribu- 
tion of  Nc,  using 


NLC  =  NL.  I C,  and 
Nuc  =  NUC, 


where  C  =  exp  \  r,  ( 0.025 )    log,    1  + 


-[CV(N,  >]" 


(13) 


(14) 


The  Satterthwaite  degrees  of  freedom  (df)  for  corrected 
abundance  estimate  confidence  intervals  were  calcu- 
lated by 


Table  3 

Summary   of  variables   required   for   corr 

ection   factor 

(boat-and-helicopter  trials) 

Parameter 

Estimate 

Length  of  transect.  L  (km) 

308 

Truncation  distance,  w  (km) 

1.0 

Number  of  helicopter  detections,  nh 

58 

Number  of  ship  detections,  ns 

126 

Number  of  duplicate  detections,  nhs 

33 

ESW  of  helicopter  (km) 

0.532 

ESW  for  duplicates  ( km  I 

0.342 

Apparent  ESW  of  boat  (kmi 

0.268 

Apparent  density  estimate  ( groups/km2 1 

0.7631 

Corrected  density  estimate  (groups/km2) 

0.3839 

Boat  detection  probability  "near"  trackline 

0.8861 

Correction  factor  (c) 

0.5032 

Standard  error,  SE(c) 

0.0912 

df= 


CV\N,  ) 


Cl'V)     CV\NS) 


(15) 


tf-1 


dfs 


where  B  is  the  number  of  bootstrap  samples,  and  dfs  is 
the  Satterthwaite  degrees  of  freedom  for  the  uncorrected 
abundance  estimate,  Ns  (see  Buckland  et  al.,  1993). 

The  CV  of  combined  abundance  estimates  (Nai)  was 
computed  by 


SEUouih 


J{SE\Nm 


)  +  SE-(N,,)  +  ...  +  SE-iN. 


«)}• 


and 


SEiloiah 
CV{total)  =  -^- 

N,  (total) 


(16) 


(17) 


Results 

The  three  line-transect  surveys  covered  2061  km  of  tran- 
sect, and  231  sightings  were  used  to  estimate  density 
(Table  2).  Sighting  rates  were  highest  around  Banks 
Peninsula  (Table  3). 

The  simultaneous  boat-and-helicopter  surveys  indi- 
cated that  boat  observers  missed  11.4%  of  the  dolphins 
on  the  trackline,  but  that  strong  responsive  movement 
towards  the  boat  resulted  in  apparent  densities  twice  as 
high  as  they  normally  would  be  (Table  3).  If  the  observ- 
ers' attention  was  drawn  to  dolphin  groups  by  the  posi- 
tion of  the  helicopter,  the  results  of  these  trials  would 
be  biased.  This  is  unlikely,  however,  because  several 
groups  sighted  by  the  helicopter  observer  subsequently 
passed  within  200  m  of  the  boat  and  were  not  seen  by 
observers.  We  saw  no  evidence  that  the  dolphins  were 
affected  by  the  helicopter. 

Detection  functions  for  boat-and-helicopter  sightings 
(Fig.  5,  C  and  D)  are  relatively  smooth  in  comparison 


448 


Fishery  Bulletin  102(3) 


A 


0        50      100      150     200      250     300     350     400     450     500     550     600     650 


1  0 
08 
06 
04 
02 

n 

a' 

a 

-v^^ 

0  100         200         300         400         500         600         700        800         900        1000 


0  50        100       150      200      250      300      350      400      450      500      550      600 


0  50        100      150      200      250      300      350      400      450      500      550      600 


E 


100    700    300    400    S00    600    700 


SO    100   ISO   200   2S0   300   3S0   400   4S0   500 


Perpendicular  distance  (m) 

Figure  5 

Histograms  of  perpendicular  sighting  distances,  and  their  fitted  detection  functions  as  used  to 
estimate  effective  strip  width.  /;  =  number  of  sightings.  The  fitted  model  (hazard,  cosine,  uniform, 
or  half  normal)  and  any  adjustments  to  it  (cosine  or  none)  are  given  in  brackets.  (A)  1997-98  har- 
bors and  bays  (n  =  71;  hazard/cosine);  (B)  1997-98  open  coasts  U?=75;  uniform/cosine);  (C)  1998-99 
open  coasts  (re=121;  half-normal/cosine);  (D)  1998-99  helicopter  sightings  (/i  =  58;  half-normal  i;  (Ei 
1998-99  duplicate  sightings  (n  =  33;  uniform/cosine);  (F>  1999-2000  harbors  and  sounds  (ra=70; 
hazard/cosinei;  (G)  1999-2000  open  coasts  (n  =  89;  uniform/cosine). 


Dawson  et  al.:  Line-transect  surveys  of  Cepha/orhynchus  hecton 


449 


with  those  presented  in  Turnock  et  al.  (1995).  The  de- 
tection function  for  the  duplicate  sightings  (Fig.  5E)  was 
more  difficult  to  fit.  Given  the  restricted  sample  size  of 
duplicates  (n=33),  this  result  is  not  unexpected. 

In  the  1998-99  Timaru  to  Long  Point  and  1999-2000 
Motunau  to  Farewell  Spit  surveys,  robust  estimation 
of  ESW  was  facilitated  by  addition  of  extra  sightings 
gained  under  similar  sighting  conditions  at  Banks 
Peninsula  (Fig.  5,  C,  F,  G).  None  of  the  three  surveys 
showed  significant  evidence  of  larger  groups  being  seen 
farther  away.  A  broad  pattern  of  abundance  declining 
to  the  north  and  south  of  the  Timaru-Banks  Peninsula 
area  is  evident  (Fig  2,  Table  2).  We  made  six  sightings 
on  288  km  of  offshore  lines  (4-10  nmi  offshore),  con- 
firming that  densities  in  this  zone  are  low. 

Information  on  sea  state  is  usually  collected  dur- 
ing boat  line-transect  surveys  and  sometimes  used  to 
poststratify  data  (e.g..  Barlow,  1995).  In  our  study  this 
was  not  advantageous,  for  three  reasons.  1)  We  avoided 
collecting  data  in  conditions  with  whitecaps;  therefore 
only  a  few  sightings  were  collected  in  Beaufort  3.  Hence 
variance  estimates  for  this  Beaufort  state  are  large.  2) 
Differences  among  Beaufort  states  for  key  parameters 
such  as  sighting  rate,  average  group  size,  and  effective 
strip  width  were  small  and  showed  overlapping  confi- 
dence intervals  (we  concede  that  statistical  power  is 
low  because  of  reason  1  stated  above).  Note  that  data 
were  pooled  in  the  same  way  as  for  ESW  estimation. 
3)  Stratification  by  Beaufort  state  does  not  produce 
abundance  estimates  that  match  the  zones  of  intrinsic 
management  interest  (e.g.,  Banks  Peninsula  Marine 
Mammal  Sanctuary;  Dawson  and  Slooten,  1993). 


Discussion 

The  catamaran  survey  platform  was  a  near-ideal  vessel 
for  close  inshore  surveys.  The  sighting  platform  (Fig.  1) 
was  a  relatively  inexpensive  modification  (-US$2000) 
that  could  be  dismantled  in  about  10  minutes  to  allow 
sailing.  The  vessel's  minimal  draught  allowed  coverage 
of  very  shallow  areas,  which  are  an  important  part  of  the 
distribution  of  Hector's  dolphin  and  many  other  inshore 
cetaceans.  Although  catamarans  are  inherently  resistant 
to  rolling,  pitching  can  be  a  problem  when  motoring 
into  a  head  sea  or  swell.  We  minimized  this  pitching  by 
arranging  lines  so  they  could  be  run  down-swell.  The  45° 
placement  of  lines  facilitated  this  reduction  in  pitching 
because  it  provided  two  alternative  sets  of  lines  (at  90° 
to  one  another).  Further,  these  could  be  run  inshore  or 
offshore,  allowing  a  choice  of  four  options. 

A  significant  advantage  of  vessels  with  low  running 
costs  is  that  the  cost  of  training  is  low.  We  could  af- 
ford to  spend  7-10  days  training  before  each  survey. 
Further,  waiting  for  weather  to  improve  is  inexpensive; 
therefore  one  does  not  need  to  gather  data  in  marginal 
sighting  conditions. 

Estimated  abundances  (Table  4)  were  not  significantly 
different  from  those  estimated  in  the  1984-85  strip 
transect  survey.  Recent  mark-recapture  estimates  of 


dolphin  abundance  at  Banks  Peninsula  in  1996,  based 
on  photo-ID  data,  differed  from  the  line-transect  es- 
timate for  this  area  by  less  than  6%  (Gormley,  2002; 
Jolly-Seber  model  allowing  different  capture  probabili- 
ties between  first  and  subsequent  captures). 

Our  surveys  confirmed  previous  work  showing  the 
patchy  nature  of  Hector's  dolphin  distribution  (Dawson 
and  Slooten,  1988).  Research  at  Banks  Peninsula  on 
the  alongshore  range  of  individually  identified  dolphins 
has  shown  a  mean  alongshore  range  of  about  31  km 
(SE  =  2.43;  Brager  et  al.,  2002).  Despite  wide-ranging 
surveys  over  13  years,  the  most  extreme  sightings  of 
any  individual  were  106  km  apart.  These  data  indicate 
very  high  site  fidelity  and  indicate  that  even  small-scale 
discontinuities  in  distribution  may  be  long  lasting.  Lack 
of  extensive  movement  along-shore,  and  hence  limited 
contact  with  neighboring  populations,  is  likely  to  be 
the  mechanism  by  which  Hector's  dolphin  has  become 
segregated  into  genetically  distinct  populations  (Pichler 
et  al.,  1998;  Pichler  and  Baker,  2000). 

The  new  abundance  data,  in  combination  with  the 
genetic  data  indicating  segregation  of  Hector's  dolphin 
into  four  populations  (Pichler  and  Baker,  2000)  and 
modeling  work  indicating  that  the  species  is  in  decline 
in  most  of  its  range  owing  to  bycatch  in  gill  nets  (Mar- 
tien  et  al.,  1999;  Slooten  et  al.,  2000),  underscore  the 
urgent  need  for  better  information  on  bycatch  rates. 

Despite  strong  evidence  of  bycatch  throughout  the 
species'  range,  observer  coverage  sufficient  to  estimate 
bycatch  has  been  achieved  only  in  one  area  (Canter- 
bury) for  one  fishing  season  (1997-98;  Baird  and  Brad- 
ford, 2000).  During  this  season  six  Hector's  dolphins 
were  observed  entangled  in  commercial  gill  nets  (a 
further  two  were  caught  but  released  alive),  resulting 
in  a  bycatch  estimate  of  17  individuals  (Starr3).  One 
mortality  was  observed  in  a  trawl  net,  but  very  low 
observer  coverage  prevented  any  calculations  of  overall 
trawl  bycatch  (Baird  and  Bradford,  2000).  No  attempt 
was  made  to  assess  bycatch  in  recreational  gillnetting 
during  this  period,  but  during  a  more  recent  summer 
(2000-01)  five  Hector's  dolphin  mortalities  occurred  in 
gill  nets  that  were  probably  set  by  recreational  fish- 
ermen (Department  of  Conservation  and  Ministry  of 
Fisheries,  2001).  It  is  not  reasonable  to  assume  that 
all  mortalities  in  recreational  gillnets  are  detected.  In 
our  opinion  it  is  likely  that  combined  commercial  and 
recreational  gillnet  bycatch  off  Canterbury  is  about 
15-30  individuals  per  year. 

Hector's  dolphin  abundance  on  the  north,  east,  and 
south  coasts  of  the  South  Island  estimated  from  the  sur- 
veys reported  in  the  present  study  is  1880  individuals 
(CV=15.7%).  Hector's  dolphins  are  more  common  on  the 


Starr,  P.  2000.  Comments  on  "Estimation  of  the  total 
bycatch  of  Hector's  dolphins  (Cephalorhynchus  hectori)  from 
the  inshore  trawl  and  setnet  fisheries  off  the  east  coast  of 
the  South  Island  in  the  1997-98  fishing  year."  Unpublished 
paper  presented  to  Conservation  Services  Levy  Working 
Group,  28  p.  Department  of  Conservation,  P.O.  Box  10-420 
Wellington.  New  Zealand. 


450 


Fishery  Bulletin  102(3) 


Table  4 

Corrected  abundance  estimates  (only  strata  with  sightings  are  listed  I 

.  Number  of  sighti 

ngs  represents  only  those  made  in  that 

stratum  and  used  for  density  estimation.  The  number 

rsed  for  estimating 

effective  half  strip  width. (ESW)  differs  because  it 

includes  sightings  from  extra  transects  in  areas  of  similar  exposure  ar 

d  transects  repeal 

ed  on  the  same  day  ( 

rnd  hence 

not  true 

replicates  for  the  purposes 

of  estimating  density). 

No.  of 

ESWim) 

N(c) 

Lower 

Upper 

Survey  zone 

Stratum 

sightings 

(n,CV%) 

rv. 

95%  CI 

95%  CI 

Motunau  to  Timaru 

Akaroa  harbor 

56 

275 

62 

32 

121 

(1997-981 

(71,22.6) 

(33.9) 

Other  harbors  and  bays 

8 

275 
171,22.6) 

14 
(67.5) 

3 

79 

Banks  Peninsula  Marine  M 

ammal  Sanctuary 

62 

261 

821 

535 

1258 

iBPMMS)  (excluding  harbors  and  bays) 

(75,  10.3) 

(22.1) 

<4  nmi  offshore,  to  the  north  and  south  of  BPMMS      19 

261 

300 

133 

679 

175,  10.3) 

(36.5) 

Timaru  to  Long  Point 

Timaru-Long  Point  (exclud 

ng  Te  Waewae  Bay 

13 

268 

310 

201 

478 

(1998-99) 

(121,  10.5) 

(28.4) 

Te  Waewae  Bay 

14 

268 
(121,  10.5) 

89 
(32.4) 

36 

218 

Motunau  to  Farewell  Spit 

Queen  Charlotte  Sound 

3 

214 

20 

4 

111) 

(1999-2000) 

(70,20.2) 

(100.5) 

Cloudy  and  Clifford  Bay 

13 

277 
(89,6.1) 

162 
(55.4) 

56 

474 

Cape  Campbell-Motunau 

5 

277 
(89,6.1) 

102 
(55.2) 

34 

305 

Total 

1880 
(21.3) 

1246 

2843 

South  Island  west  coast,  where  an  aerial  survey  of  simi- 
lar design  resulted  in  an  estimate  of  5388  (CV=20.6<*; 
Slooten  et  al.,  in  press).  Thus  Hector's  dolphin  abun- 
dance in  South  Island  waters  is  estimated  at  7268  in- 
dividuals (CV=15.8'7r ).  The  North  Island  subspecies  of 
Hector's  dolphin,  now  considered  critically  endangered 
(IUCN4)  remains  to  be  surveyed  quantitatively. 

The  new  abundance  estimates  provide  an  empirical 
basis  from  which  to  calculate  levels  of  take  that  would 
still  allow  the  currently  depleted  populations  to  recover 
(e.g.,  Wade,  1998).  These  levels  of  take  should  be  seen 
as  short-term  targets  for  bycatch  reduction  in  gill  and 
trawl  nets.  For  the  management  of  Hector's  dolphin  to 
be  put  on  a  rational  basis,  a  more  comprehensive  and 
wide-ranging  assessment  of  bycatch,  including  statisti- 
cally robust  observer  programs  in  coastal  fisheries,  is 
urgently  needed. 


nificant  contributions  to  equipment  used  in  the  survey- 
were  made  by  the  New  Zealand  Whale  and  Dolphin 
Trust  and  the  University  of  Otago.  We  are  very  grateful 
for  the  hard  work  put  in  by  the  other  observers:  Laszlo 
Kiss,  Nadja  Schneyer,  Gail  Dickie,  Niki  Alcock,  Lesley 
Douglas,  James  Holborow,  Ellie  Dickson,  Guen  Jones, 
Will  Rayment,  and  Dan  Cairney.  Jay  Barlow,  Jeff  Laake, 
Anne  York,  and  Debbie  Palka  shared  their  thoughts  on 
survey  design  and  field  methods.  David  Fletcher  helped 
with  aspects  of  variance  estimation.  Akaroa  Harbour 
Cruises  provided  much  appreciated  field  support.  Daryl 
Coup  wrote  the  sightings  program  we  used  to  collect 
data  in  the  field.  Otago  University's  Department  of 
Surveying  very  helpfully  provided  the  GPS  base-station 
data  for  postprocessing  our  GPS  fixes. 


Literature  cited 


Acknowledgments 

These  surveys  were  funded  principally  by  Department  of 
Conservation  Contracts  SCO  3072,  3074,  and  3075.  Sig- 


4  IUCN.  2000.  IUCN  red  list  of  threatened  species.  In- 
tel national  Union  for  Conservation  of  Nature  and  Natural 
Resources.  Species  Survival  Commission.     [Available  at 

www.redlist.oru  I 


Baird.  S.  J.,  and  E.  Bradford. 

2000.  Estimation  of  the  total  bycatch  of  Hector's  dolphins 
Wcphalurhynchiis  hectori)  from  the  inshore  trawl  and 
setnel  fisheries  off  the  east  coast  of  the  South  Island  in 
the  L997-98  fishing  year.  28  p.  Published  client  report 
on  contract  3024.  funded  by  Conservation  Services  Levy, 
Department  of  Conservation.  Wellington.  Web  site: 
http://csl.doc.govt.nz/CSL3024.pdf.  [Accessed  8  March 
2001.1 


Dawson  et  al.:  Line-transect  surveys  of  Cepholorhynchus  hectori 


451 


Baker,  A.  N..  A.  N.  H.  Smith,  and  F.  B.  Pichler. 

2002.     Geographical  variation  in  Hector's  dolphin:  recogni- 
tion of  a  new  subspecies  of  Cephalorhynchus  hectori.     J 
Roy.  Soc.  N.Z.  32:713-727. 
Barlow,  J. 

1988.     Harbor  porpoise,  Phocoena  phocoena,  abundance 
estimation  for  California,  Oregon,  and  Washington:  I. 
Ship  surveys.     Fish.  Bull.  86:417-432. 
1995.     The  abundance  of  cetaceans  in  California  waters. 
Part  1:  Ship  surveys  in  summer  and  fall  of  1991.     Fish. 
Bull.  93:1-14. 
Brager,  S.,  S.  M.  Dawson,  E.  Slooten,  S.  Smith,  G.  S.  Stone,  and 
A.  Yoshinaga. 

2002.     Site  fidelity  and  along-shore  range  in  Hector's  dol- 
phin, an  endangered  marine  dolphin  from  New  Zea- 
land.    Biol.  Conserv.  108:-287. 
Buckland,  S.  T..  D.  R.  Anderson,  K.  P.  Burnham,  and 
J.  L.  Laake 

1993.     Distance  sampling:  estimating  abundance  of  bio- 
logical populations,  446  p.     Chapman  and  Hall,  New 
York,  NY. 
Buckland,  S.  T.,  and  B.  J.  Turnock. 

1992.  A  robust  line  transect  method.  Biometrics  48: 
901-909. 

Carretta,  J.  V.,  B.  L.  Taylor,  and  S.  J.  Chivers. 

2001.     Abundance  and  depth  distribution  of  harbor  por- 
poise i Phocoena  phocoena)  in  northern  California  de- 
termined from  a  1995  ship  survey.     Fish.  Bull.  99:29-39. 
Dawson,  S.  M.,  and  E.  Slooten. 

1988.  Hector's  Dolphin  Cephalorhynchus  hectori:  distri- 
bution and  abundance.  Rep.  Int.  Whal.  Comm.  (special 
issue)  9:315-324. 

1993.  Conservation  of  Hector's  dolphins:  the  case  and 
process  which  led  to  establishment  of  the  Banks  Pen- 
insula Marine  Mammal  Sanctuary.  Aquat.  Cons. 
3:207-221. 

Dawson,  S.  M.,  E..  Slooten,  F  Pichler,  K.  Russell,  and 
C.  S.  Baker. 

2001.     North  Island  Hector's  dolphins  are  threatened  with 
extinction.     Mar.  Mamm.  Sci.  17:366-371. 
Department  of  Conservation  and  Ministry  of  Fisheries. 

2001.  Hector's  dolphin  and  setnetting  Canterbury  Area: 
a  paper  for  public  comment,  9  p.  [Available  from  Min- 
istry of  Fisheries,  Private  Bag  1926,  Dunedin,  New 
Zealand. I 

Gormley,  A. 

2002.  Mark-recapture  abundance  estimation  in  marine 


mammals.     M.S.  thesis.,  105  p.     Univ.  Otago,  Dunedin, 

New  Zealand. 
IWC  (International  Whaling  Commission). 

1994.     Report  of  the  International  Whaling  Commission 

workshop  and  symposium  on  the  mortality  of  cetaceans  in 

passive  fishing  nets  and  traps.     Rep.  Int.  Whal.  Comm. 

(special  issueil5.  629  p. 
Lerczak,  J.  A.,  and  R.  O.  Hobbs. 

1998.  Calculating  sighting  distances  from  angular  read- 
ings during  shipboard  aerial,  and  shore-based  marine 
mammal  surveys.     Mar.  Mamm.  Sci.  14:590-599. 

Martien.  K.  K.,  B.  L.  Taylor,  E.  Slooten,  and  S.  M.  Dawson. 

1999.  A  sensitivity  analysis  to  guide  research  and  man- 
agement for  Hector's  dolphin.  Biol.  Conserv.  90:183- 
191. 

Palka,  D.  L.,  and  P.  S.  Hammond. 

2001.     Accounting  for  responsive  movement  in  line  transect 

estimates  of  abundance.     Can.  J.  Fish.  Aquat.  Sci. 

58(41:777-787. 
Pichler,  F,  and  C.  S.  Baker. 

2000.  Loss  of  genetic  diversity  in  the  endemic  Hector's 
dolphin  due  to  fisheries-related  mortality.  Proc.  Roy. 
Soc.  (Londi,  Series  B.  267:97-102. 

Pichler,  F.  B.,  S.  M.  Dawson,  E.  Slooten,  and  C.  S.  Baker. 

1998.  Geographic  isolation  of  Hector's  dolphin  popula- 
tions described  by  mitochondrial  DNA  sequences.  Cons. 
Biol.  12:676-682. 

Slooten,  E..  and  S.  M.  Dawson. 

1994.  Hector's  Dolphin.  In  Handbook  of  marine  mam- 
mals, vol  V  ( Delphinidae  and  Phocoenidae  M  S.  H  Ridgway 
and  R.  Harrison,  eds.),  p.  311-333.  Academic  Press. 
New  York,  NY. 

Slooten.  E..  S.  M.  Dawson,  and  W.  J.  Rayment. 

In  press.     Aerial  surveys  for  coastal  dolphins:  Abundance 
of  Hector's  dolphins  off  the  South  Island  west  coast, 
New  Zealand.     Mar.  Mamm.  Sci.  20(3  I. 
Slooten.  E..  D.  Fletcher,  and  B.  L.  Taylor. 

2000.     Accounting  for  uncertainty  in  risk  assessment: 
Case  study  of  Hector's  dolphin  mortality  due  to  gillnet 
entanglement.     Cons.  Biol.  14:1264-1270. 
Turnock,  B.  J..  S.  T.  Buckland.  and  G.  C.  Boucher. 

1995.  Population  abundance  of  Dall's  porpoise  iPhocoe- 
noides  dalli)  in  the  Western  North  Pacific  Ocean.  Rep. 
Int.  Whal.  Comm.  (special  issue)  16:381-397. 

Wade,  P. 

1998.  Calculating  limits  to  the  allowable  human-caused 
mortality  of  cetaceans  and  pinnipeds.  Mar.  Mamm. 
Sci.  14:1-37. 


452 


Abstract— A  developmental  series  of 
larval  and  pelagic  juvenile  pygmy 
rockfish  (Sebastes  wilsoni)  from  cen- 
tral California  is  illustrated  and 
described.  Sebastes  wilsoni  is  a  non- 
commercially.  but  ecologically,  impor- 
tant rockfish,  and  the  ability  to  dif- 
ferentiate its  young  stages  will  aid 
researchers  in  population  abundance 
studies.  Pigment  patterns,  meristic 
characters,  morphometric  measure- 
ments, and  head  spination  were 
recorded  from  specimens  that  ranged 
from  8.1  to  34.4  mm  in  standard 
length.  Larvae  were  identified  ini- 
tially by  meristic  characters  and  the 
absence  of  ventral  and  lateral  midline 
pigment.  Pelagic  juveniles  developed 
a  prominent  pigment  pattern  of  three 
body  bars  that  did  not  extend  to  the 
ventral  surface.  Species  identifica- 
tion was  confirmed  subsequently  by 
using  mitochondrial  sequence  data 
of  four  representative  specimens  of 
various  sizes.  As  determined  from  the 
examination  of  otoliths,  the  growth 
rate  of  larval  and  pelagic  juvenile 
pygmy  rockfish  was  0.28  mm/day, 
which  is  relatively  slow  in  compari- 
son to  the  growth  rate  of  other  spe- 
cies of  Sebastes.  These  data  will  aid 
researchers  in  determining  species 
abundance. 


Description  and  growth  of  larval  and  pelagic 
juvenile  pygmy  rockfish  (Sebastes  wilsoni) 
(family  Sebastidae) 


Thomas  E.  Laidig 

Keith  M.  Sakuma 

Santa  Cruz  Laboratory 

Southwest  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

110  Shaffer  Rd 

Santa  Cruz,  California  95060 

E-mail  address:  torn  laidigiS'noaa. gov 

Jason  A.  Stannard 

La  Jolla  Laboratory 

Southwest  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

P.  O.  Box  271 

La  Jolla,  California  92038 


Manuscript  submitted  9  -June  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 

25  February  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:452-463(20041. 


Rockfishes  (genus  Sebastes)  form  a 
diverse  group  comprising  at  least  72 
species  occurring  in  the  northeast- 
ern Pacific  (Love  et  al.,  2002).  Many 
of  these  species  represent  a  substan- 
tial portion  of  the  groundfish  fishery 
off  the  west  coast  of  North  America, 
accounting  for  20%  of  the  groundfish 
landings  in  California  in  2000  (Pacific 
Fishery  Management  Council,  2000). 
A  few  species  are  relatively  abun- 
dant but  are  not  harvested  because 
of  their  small  size.  These  species  play 
vital  roles  in  the  community  ecology, 
including  providing  prey  for  the  larger, 
commercially  important  species.  The 
pygmy  rockfish  (Sebastes  wilsoni) 
having  a  maximum  size  of  23  cm  total 
length,  is  among  these  small  species 
(Love  et  al.,  2002).  Pygmy  rockfish  are 
common  over  sediment  and  rocky  sea- 
floor  habitats  at  a  depth  of  30-274  m 
(Stein  et  al.,  1992;  Yoklavich  et  al., 
2000).  Stein  et  al.  (1992)  observed 
that  pygmy  rockfish  were  by  far  the 
most  abundant  fish  species  off  Heceta 
Bank,  Oregon,  and  Love  et  al.  (1996) 
reported  "clouds"  of  pygmy  rockfish 
mixed  with  two  other  small  species, 
squarespot  rockfish  (S.  hopkinsi  )  and 
halfbanded  rockfish  (S.  semicinctus) 
off  southern  California.  In  Soquel 
Canyon  in  central  California,  pygmy 
rockfish  dominated  fish  assemblages 


in  rock-boulder  habitat  at  75-175  m 
(Yoklavich  et  al.,  2000). 

Accurate  identification  of  larval 
stages  is  critical.  Biomass  of  rock- 
fish populations  can  be  estimated 
from  larval  production  (Ralston  et 
al.,  2003)  and  larval  and  juvenile 
abundance  studies  (Moser  and  But- 
ler, 1987;  Hunter  and  Lo,  1993  I.  If 
the  larval  and  juvenile  rockfish  ana- 
lyzed in  these  studies  are  not  correct- 
ly identified,  it  could  lead  to  either 
over-  or  underestimates  of  biomass 
or  recruitment  potential  of  a  popula- 
tion. Identification  of  young  stages 
of  Sebastes  has  been  accomplished 
through  rearing  studies  and  through 
descriptions  based  on  developmental 
series  of  field-caught  specimens  of 
various  sizes  (Matarese  et  al.,  1989; 
Moser,  1996).  Otolith  morphologies 
have  been  useful  in  discerning  some 
Sebastes  species  (Laidig  and  Ralston, 
1995;  Stransky,  2001).  Recently,  mo- 
lecular methods  have  proven  to  be  an 
effective  tool  for  the  identification  of 
Sebastes  larvae  (Seeb  and  Kendall, 
1991;  Rocha-Olivares,  1998;  Rocha- 
Olivares  et  al„  2000). 

In  this  study,  we  identify  and  de- 
scribe the  larvae  and  pelagic  juveniles 
of  pygmy  rockfish  based  on  morpho- 
metries and  pigmentation  patterns, 
and  estimate  age  and  growth  at  two 


Laidig  et  al.:  Descriptions  and  growth  of  larval  and  juvenile  Sebastes  wilsoni 


453 


developmental  stages.  Further,  we  examine  otolith  ra- 
dius at  time  of  larval  extrusion  to  separate  pygmy  rock- 
fish  from  other  similarly  pigmented  Sebastes  specimens. 
We  also  use  mitochondrial  DNA  (mtDNA)  sequence 
data  to  identify  four  putative  pygmy  rockfish  specimens 
representing  a  continuum  of  late-larval  through  pelagic 
juvenile  stages.  The  molecular  results  are  used  to  con- 
firm identifications  based  on  morphological,  meristic, 
and  pigmentation  characters  and  to  assure  that  the 
assembled  developmental  series  is  monospecific. 


Methods 

Specimen  collection 

Specimens  of  larval  and  pelagic  juvenile  pygmy  rockfish 
were  obtained  from  research  cruises  conducted  aboard 
the  NOAA  RV  David  Starr  Jordan  off  central  California. 
Specimens  were  collected  in  midwater  (5-30  m)  from 
mid-May  to  mid-June,  1990-92,  between  Bodega  Bay 
(north  of  San  Francisco)  and  Cypress  Point  (south  of 
Monterey  Bay)  by  using  a  26  mx26  m  modified  Cobb 
midwater  trawl  (12.7-mm  stretched-mesh  codend  liner). 
Specimens  also  were  collected  during  early  March, 
1992-93,  between  Salt  Point  (north  of  San  Francisco) 
and  Cypress  Point  with  a  5  mx5  m  modified  Isaacs-Kidd 
(MIK)  frame  trawl  with  2-mm  net  mesh  and  0.505-mm 
mesh  codend.  Specimens  from  the  Cobb  trawl  were  frozen 
and  specimens  from  the  MIK  frame  trawl  were  preserved 
in  95%  ethanol  for  later  analysis. 

Meristics,  morphometries,  and  body  pigmentation 

We  examined  pigmentation  patterns  and  physical  char- 
acteristics of  122  pygmy  rockfish  larvae  and  pelagic 
juveniles.  Standard  length  (SL)  was  measured  for  each 
individual  and  sizes  ranged  from  8.1  to  34.4  mm.  Speci- 
mens greater  than  19.9  mm  were  identified  by  using 
meristic  characters  (Chen.  1986;  Matarese  et  al..  1989; 
Moreland  and  Reilly,  1991;  and  Laroche1),  and  pigment 
patterns  were  recorded.  Specimens  less  than  20  mm 
were  identified  initially  from  pigment  patterns  from  a 
series  starting  with  the  smallest  (8.1  mm  SL)  identifi- 
able individuals  with  complete  fin-ray  counts.  Counts 
of  dorsal-,  anal-,  and  pectoral-fin  rays,  and  the  number 
of  gill  rakers  on  the  first  arch  were  recorded  whenever 
possible  and  subsequently  used  in  identifications.  Gill 
raker  counts  were  obtained  only  from  fish  larger  than 

15  mm  SL. 

We  measured  snout-to-anus  length,  head  length, 
snout  length,  eye  diameter,  body  depth  at  the  pectoral 
fin  base,  body  depth  at  anus,  and  pectoral-fin  length  on 

16  specimens  ranging  from  8.1  to  29.6  mm  SL,  follow- 
ing Richardson  and  Laroche  (1979).  Head  spination  was 
examined  on  thirty-three  specimens  (8.1  to  29.6  mm 


1  Laroche,  W.  A.  1987.  Guide  to  larval  and  juvenile  rock- 
fishes  (Sebastes)  of  North  America.  Unpubl.  manuscript, 
311  p.     Box  216,  Enosburg  Falls,  VT  05450. 


SL)  that  were  stained  with  alizarin  red-s.  Terminol- 
ogy for  head  spination  follows  Richardson  and  Laroche 
(1979).  In  the  following  descriptions,  larval  and  juvenile 
lengths  always  refer  to  SL  and  pigmentation  always 
refers  to  melanin. 

Otolith  examination 

Sagittal  otoliths  were  removed  from  61  larval  and  pelagic 
juvenile  pygmy  rockfish  (8.1-34.4  mm  SL),  and  growth 
increments  were  counted  beginning  at  the  first  incre- 
ment after  the  extrusion  check  (the  mark  in  the  otolith 
formed  when  the  larvae  are  released  from  their  mother) 
by  using  a  compound  microscope  at  lOOOx  magnifica- 
tion (see  Laidig  et  al.,  1991).  No  validation  of  the  these 
growth  increments  was  performed  during  the  present 
study,  and  none  has  been  conducted  by  other  research- 
ers. However,  we  assumed  that  these  growth  increment 
counts  corresponded  to  daily  ages  based  on  validation 
of  daily  growth  increments  in  other  co-occurring  rock- 
fishes,  namely  shortbelly  rockfish,  S.  jordani  (Laidig  et 
al.,  1991),  black  rockfish,  S.  melanops  (Yoklavich  and 
Boehlert,  1987),  bocaccio,  S.  paucispinis,  chilipepper, 
S.  goodei,  widow  rockfish,  S.  entomelas,  and  yellowtail 
rockfish,  S.  flavidus  (Woodbury  and  Ralston,  1991).  The 
radius  of  the  otolith  was  measured  from  the  primordium 
to  the  postrostral  edge  of  the  extrusion  check  for  com- 
parison with  similar  measurements  from  other  Sebastes 
spp.  (as  reported  in  Laidig  and  Ralston,  1995).  Transfor- 
mation from  the  larval  stage  to  the  pelagic  juvenile  stage 
was  ascertained  by  the  presence  of  accessory  primordia 
(Laidig  et  al.,  1991;  Lee  and  Kim,  2000). 

Molecular  confirmation 

Total  genomic  DNA  was  isolated  from  skeletal  muscle 
tissue  of  four  larval  and  juvenile  putative  pygmy  rock- 
fish specimens  by  using  a  CTAB  and  phenol-chloro- 
form-isoamyl  alcohol  protocol  ( Winnepenninckx  et  al., 
1993;  Hillis  et  al.,  1996).  These  four  specimens  ranged 
in  length  from  15  to  27  mm  and  had  pigment  patterns 
similar  to  the  fish  identified  as  pygmy  rockfish  in  the 
present  tudy.  Polymerase  chain  reaction  (PCR)  ampli- 
fications and  sequencing  of  partial  mitochondrial  DNA 
regions  (cytochrome  b  [cyt-6]  and  control  region  [CR]) 
followed  the  methods  of  Rocha-Olivares  et  al.  (1999a, 
1999b).  PCR  products  were  verified  on  29c  agarose  gels 
and  purified  by  using  a  QIAquick™  PCR  Cleanup  Kit 
(Qiagen,  Inc.,  Valencia,  CA)  following  manufacturer 
protocols.  Complementary  strand  sequence  data  were 
generated  by  using  ABI  PRISM' M  DyeDeoxy™  termina- 
tor cycle  sequence  chemistry  on  an  automated  sequencer 
(Applied  Biosystems,  Model  377,  Foster  City,  CA). 

Cytochrome  b  sequence  data  (750  base  pairs)  from 
the  four  specimens  were  aligned  with  (previously  gen- 
erated) orthologous  sequences  from  119  individuals 
representing  61  species  of  Sebastes  (Rocha-Olivares  et 
al.,  1999b).  Species  identifications,  based  on  cyt-fo  data, 
were  made  by  using  distance-based  cluster  analyses  in 
PAUP  v4.0b2  (Phylogenetic  Analysis  Using  Parsimony, 


454 


Fishery  Bulletin  102(3) 


version  4,  Sunderland,  MA)  and  pairwise  comparisons 
of  sequence  divergence  (i.e.,  the  number  of  nucleotide 
differences  between  two  individuals  expressed  as  a  per- 
centage). A  secondary  data  set,  which  included  an  ad- 


Table  1 

Frequency  of  occurrence  ( number  of  fish  i 
and  pectoral-fin  ray  counts,  and  gill  ra 
122  pygmy  rockfish  tSebastes  wilsoni). 

of  dorsal-,  anal-, 
ker  counts  from 

Character 

Count 

Frequency  of 
occurrence 

Percent 
occurrence 

Dorsal-fin  rays 

12 

8 

7 

13 

110 

91 

14 

3 

2 

Anal-fin  rays 

5 

2 

2 

6 

116 

95 

7 

4 

3 

Pectoral-fin  rays 

16 

5 

5 

17 

92 

90 

18 

5 

5 

Gill  rakers 

36 

11 

14 

37 

15 

18 

38 

22 

28 

39 

20 

25 

40 

10 

13 

41 

1 

1 

42 

1 

1 

ditional  450  base  pairs  of  control  region  sequence,  was 
generated  for  the  four  undetermined  specimens  and  for 
a  subgroup  of  known  reference  species  with  low  levels 
of  sequence  divergence  from  the  four  putative  pygmy 
rockfish  specimens  ( Puget  Sound  rockfish  [S.  empha- 
eus],  redstripe  rockfish  [S.  proriger],  harlequin  rockfish 
IS.  variegatus].  sharpchin  rockfish  [S.  zaeentrus],  and 
pygmy  rockfish).  Species  identifications,  based  on  this 
extended  (cyt-fe  +  CR)  data  subset,  followed  analyses 
described  above. 


Results 

General  development 

All  122  fish  had  completed  notochord  flexion  and  pos- 
sessed a  full  complement  of  segmented  fin  rays  by  8.1 
mm.  The  mode  for  dorsal-fin  ray  counts  was  13,  for 
anal-fin  rays  6,  and  for  pectoral-fin  rays  17  (Table  1). 
The  mode  for  gill  raker  counts  was  38,  and  the  range 
was  36-42.  Anal-  and  dorsal-fin  spines  began  to  develop 
between  9.1  and  14.0  mm.  Lateral  line  pores  began  to 
develop  at  29  mm,  although  a  full  complement  (37  to  46 
pores)  was  not  reached  in  our  specimens.  Morphometric 
measurements  were  taken  from  16  individual  pygmy 
rockfish  of  8.1-29.6  mm  (Table  2). 

Head  spination 

At  8.1  mm,  the  postocular,  parietal,  nuchal,  inferior  post- 
temporal,  supracleithral,  superior  opercular,  preopercu- 
lars  (with  the  exception  of  the  2nd  anterior),  and  1st  and 


Table  2 

Morphometric  measurements  (in  mm)  fr 

om  16  individual 

;  of  pygmy  rockfish  tSebastes  wilsoni). 

Snout-anus 

Head 

Snout 

Eye 

Body  depth 

Body  depth 

Pectoral-fin 

SL 

length 

length 

length 

diameter 

at  pectoral  base 

at  anus 

length 

8.1 

5.2 

3.3 

0.8 

1.3 

2.8 

2.3 

1.5 

9.0 

5.3 

3.0 

1.0 

1.5 

3.0 

2.3 

1.7 

10.8 

6.5 

4.2 

1.2 

1.7 

3.5 

2.8 

2.3 

12.1 

7.0 

4.3 

1.3 

2.0 

3.8 

2.8 

2.5 

12.8 

7.7 

4.8 

1.3 

2.2 

3.7 

3.0 

2.5 

14.2 

8.3 

4.7 

1.3 

2.2 

4.3 

3.5 

3.3 

15.2 

9.2 

5.7 

1.7 

2.0 

4.2 

3.5 

3.5 

16.2 

9.7 

5.2 

1.5 

2.0 

4.5 

3.8 

3.8 

17.5 

10.8 

6.2 

2.0 

2.3 

5.0 

3.8 

4.2 

18.6 

11.5 

6.0 

2.0 

2.7 

5.5 

4.3 

5.0 

20.7 

12.7 

6.2 

2.0 

2.7 

6.2 

5.2 

5.0 

22.3 

14.2 

6.7 

2.2 

2.8 

6.8 

5.7 

6.2 

23.8 

14.5 

6.7 

2.0 

3.2 

7.3 

6.3 

5.9 

24.3 

14.2 

7.0 

2.2 

2.7 

7.0 

5.8 

6.3 

28.9 

17.0 

8.0 

2.6 

3.3 

8.5 

7.3 

7.0 

29.6 

16.9 

7.8 

2.6 

3.5 

8.7 

7.5 

6.9 

Laidig  et  al.:  Descriptions  and  growth  of  larval  and  juvenile  Sebastes  wilsoni 


455 


2nd  inferior  and  1st  superior  infraorbital  spines  were 
present  (Table  3).  The  nasal,  pterotic,  and  4th  superior 
infraorbital  first  appeared  at  10.8  mm.  At  12.1  mm,  the 
inferior  opercular  spine  became  evident.  Between  14.2 
and  17.5  mm,  the  preocular,  tympanic,  superior  postem- 
poral,  2nd  anterior  preopercular.  and  3rd  superior  infraor- 
bital spines  formed.  After  17.5  mm,  no  further  changes 
in  head  spination  were  noted.  The  supraocular,  coronal, 
3rd  inferior  infraorbital,  and  2nd  superior  infraorbital 
spines  did  not  occur  on  any  of  the  fish  examined. 

Body  pigmentation 

At  9.1  mm,  pygmy  rockfish  had  no  pigment  along  the 
lateral  and  ventral  body  surfaces  (Fig.  1A,  Table  4). 
Pigment  was  heavy  on  the  top  of  the  head  and  present 


on  the  operculum.  The  dorsal  midline  surface  had  a 
few  melanophores  under  the  soft  dorsal  fin.  The  ante- 
rior lower  jaw  was  pigmented  on  the  tip  and  one  to  two 
melanophores  were  present  on  each  side  of  the  snout 
near  the  tip.  Pigment  also  was  present  at  the  base  and 
on  the  distal  half  of  the  pectoral  fin. 

By  14.0  mm,  dense  pigment  along  the  dorsal  midline 
stretched  from  the  caudal  peduncle  to  the  first  dorsal-fin 
spine  (Fig.  IB,  Table  4).  The  only  lateral  pigment  on  the 
body  consisted  of  a  few  melanophores  along  the  midline 
near  the  caudal  peduncle.  The  ventral  surface,  includ- 
ing the  anal  and  pelvic  fins,  remained  unpigmented. 
Pigment  on  the  pectoral  fins  had  mostly  disappeared  by 
11.0  mm  and  was  rarely  observed  at  14.0  mm.  Opercu- 
lar and  head  pigment  increased  in  density  by  14.0  mm. 
Pigment  along  the  lower  edge  of  the  orbit  began  to 


Table  3 

Development  of  head  spi 

aes  in 

individua 

pygmy  rockfish  iSebastc 

•  wilsoni 

).  "1' 

means  spines  present  and  "0"  means  spines 

absent. 

Spines 

Standard  length  (mm) 

8.1 

9.0 

10.8      12.1 

12.8 

14.2 

15.2 

16.2      17.5 

18.6 

20.7     22.3     23.8     24.3     28.9     29.6 

Nasal 

0 

0 

1 

1 

1 

1 

1 

1 

111111 

Preocular 

0 

0 

0 

0 

0 

0 

1 

1 

111111 

Supraocular 

0 

0 

0 

0 

0 

0 

0 

0 

0 

0 

0           0          0          0           0          0 

Postocular 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

Coronal 

0 

0 

0 

0 

0 

0 

0 

0 

0 

0 

0           0           0           0           0          0 

Tympanic 

0 

0 

0 

0 

0 

0 

0 

1 

1 

111111 

Parietal 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

Nuchal 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

Pterotic 

0 

0 

1 

1 

1 

1 

1 

1 

111111 

Posttemporal 

Superior 

0 

0 

0 

0 

0 

0 

1 

1 

111111 

Inferior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

Supracleithral 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

Opercular 

Superior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

Inferior 

0 

0 

0 

1 

1 

1 

1 

1 

Preopercular 

1st  anterior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

2nd  anterior 

0 

0 

0 

0 

0 

1 

1 

1 

111111 

3rd  anterior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

1st  posterior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

2nd  posterior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

3rd  posterior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

4th  posterior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

5th  posterior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

Infraorbital 

1st  inferior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

2nd  inferior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

3rd  inferior 

0 

0 

0 

0 

0 

0 

0 

0 

0           0          0          0          0          0 

1st  superior 

1 

1 

1 

1 

1 

1 

1 

1 

111111 

2nd  superior 

0 

0 

0 

0 

0 

0 

0 

0 

0           0           0           0           0           0 

3rd  superior 

0 

0 

0 

0 

0 

0 

1 

1 

111111 

4th  superior 

0 

0 

1 

1 

1 

1 

1 

1 

111111 

456 


Fishery  Bulletin  102(3) 


Figure  1 

Developmental  series  of  pygmy  rockfish  [Sebastes  wilsoni)  (drawn  by  authors).  (A)  9.1-mm- 
SL  larva;  (B)  14.0-mm-SL  larva;  iC)  17.5-mm-SL  larva;  (D)  23.0-mm-SL  pelagic  juvenile; 
I E )  28.6-mm-SL  pelagic  juvenile.  Arrow  marks  ventral  end  of  anterior  body  bar;  I  F  )  34.4- 
mm-SL  pelagic  juvenile.  Note  that  not  all  head  spines  are  included  in  the  illustrations. 


form  in  some  specimens  by  12.0  mm  and  was  visible  in 
most  specimens  by  14.0  mm  (Fig.  IB).  Snout  pigment 
was  represented  by  one  or  four  melanophores.  Anterior 
lower  jaw  pigment  was  heavy  and  confined  to  the  tip 
of  the  jaw. 

By  17.5  mm,  the  dorsal  midline  pigment  had  become 
much  darker  and  denser  (Fig.  IC,  Table  4)  and  extended 
from  the  caudal  fin  to  the  head  region,  except  for  a  gap 
where  the  nape  pigment  was  beginning  to  form.  All 
fins  were  unpigmented.  Pigment  along  the  ventral  body 


midline  began  to  form  at  17.5  mm,  with  a  few  postanal 
melanophores.  Lateral  midline  pigment  formed  in  two 
locations.  Melanophores  near  the  peduncle  increased 
anteriorly,  and  pigment  began  forming  dorsal  to  the  gut 
cavity  and  increased  posteriorly  toward  the  peduncle. 
A  body  bar  began  to  form  on  the  lateral  surface  above 
the  pectoral  fin  between  the  spinous  dorsal  fin  and  the 
anterior  lateral  midline  pigment.  Opercular,  eye,  and 
head  pigment  all  increased  in  density.  Melanophores  on 
the  snout  also  became  more  prevalent  between  the  tip  of 


Laidig  et  al.:  Descriptions  and  growth  of  larval  and  |uvenile  Sebastes  wilsoni 


457 


Figure  1  (continued) 


the  upper  jaw  and  top  of  the  head.  Pigment  on  the  tip 
of  the  lower  jaw  spread  posteriorly  and  became  denser 
than  in  smaller  specimens. 

When  the  pelagic  juveniles  reached  23.0  mm,  the 
dorsal  midline  pigment  was  a  dark  strip  extending  from 
the  caudal  fin  to  the  head  (Fig.  ID,  Table  4).  Nape  pig- 
ment almost  merged  with  the  head  pigment,  except  for 
a  small  unpigmented  area  below  the  insertion  of  the 
parietal  and  nuchal  spines.  Hypural  pigment  was  pres- 
ent distally  in  all  individuals  at  this  size  and  at  larger 


sizes  (n  =  30).  Anterior  and  posterior  lateral  midline  pig- 
ment merged  to  form  a  continuous  line  along  the  body. 
The  number  of  melanophores  increased  on  the  ventral 
body  surface,  and  a  few  melanophores  were  present  at 
the  anal-fin  ray  bases  and  along  the  ventral  midbody 
posterior  to  the  anal  fin.  The  anterior  body  bar  broad- 
ened and  became  more  defined.  A  few  melanophores 
on  the  flanks  under  the  soft  dorsal  fin  began  to  form 
a  midbody  bar.  A  third  body  bar  began  to  form  on  the 
caudal  peduncle.  Pigment  on  the  operculum,  top  of  the 


458 


Fishery  Bulletin  102(3) 


2 


5  f 


J3  .i 

+*  TO 
g.| 

^  II 

l  g    I    5 

n  -S  «  -5 

J  u    3    J 

_•  a. 

S 

£ 


CO 


ho 
C 


II  CO 

cu  £ 

>-,  o 

a  <- 

£  z 

TO  < 


fc   J3  — 


co    co 


O    _c    -Q 


£  5 

£  <D 

i  CJ 

<=.  £ 

i-H  — 

CD  m 

O  CO 

_  o 

T3  s- 

0J  O 

ac  -o 

2  ii 

>  K 

co  D 

to  O 

CO  -7 

£  W 

CO  - 

9| 

.2  o 

g  £ 

to  -" 

--  c*- 

CD  o 


o  ii 

co  03 

II  Q 

S  S 


0) 

•~ 

-i 

o.  -a 

on 

a» 

i- 

m 

S- 

a 
o 

b 

B 
> 

- 

<2  £ 

t.    o 

3  -= 

eft     . 

>>.2 
-a   C 

O     OJ 
CO    co 

£  II 

>  z 

ii  < 

«  s  « 

4s  lo  2. 


»  3 

>>   o 


£3   S 


■g  H 

a  >< 

CO       CO 


s  • 


CC 


£     -    Oh 


■°  D 

B  D 

g  < 

o  O 

ii  ,S  " 

CC 

vi  _  a 

-      Q)       W 

»    «     X 

£  -c    co 

Ik  s 

a  o    u 

a  -a 
co    " 
= 
n 
W 


iS  o 
cu 
— 


ii 


OJC     - 

°-  5 

c_    a 

O      II 

g    1-5 


<  a. 

2;  . 

.  c 

en  ^ 

5  — 

o  CO 

^  CO 

-*  k 

Q;  O 

CU  "C 


:l. 


3     " 
=  X 

.5  <C 
£.5  13  S-S 


OOOOOOOOOOOOO 
OOOOOOOOOOOOO 


oppoo"oi>ooo 

OOOOOOOi-Jr-ir-i 


OOOOOOOOOOOOOOOOOOOpl>t--t-'t-; 

oooooooooooooooooooooooo 


OOOOOOOOOOOOOOOOOIOOOOOOOO 
OOOOOOOOOOOOOOOOOO'-H'— •'-'•— ''— '■— '-—, 


ooooooooocoaot>oi>pooppooooopoo 

0000000000001^0'-^'-*^H'--'•-,'-,'--|'-,'— ,'-,'--,'-,'-, 


ooooooooooooococoocooooooooo 

OOOOOOOOOOCJOOOO'-'O'— 1»— '•— !■— t      >— I      i-H      iH I      i-H 


ooooooooooo-— imcomooocooooooo© 

ddddddo'dddo'oddo^— '^^^*-H--''-,'-''-<'-'^ 


ooooi^t^^c^ccooooopoopooococ-^c^ooo 
r^rH^HOooooodddooooodddddooT-itHTH 

ooooooooooooooooooooooooooo 

OOOOOOO<NC0-^^-C0OU5iT3OOOOOOOOOOpO 
OOOOOOIOI>C1GOOOOOOOOOOOOOOOOOO 

dddddddddd•-''-*-^•-,'-,*- ,--,'-,'-t'~,'-,'-,,-H'-H'-,,-l'~' 


OOOOOOCOPO*-*COCDCQOOOCOOOOOpOOOO 
000000000000*-^      0"o^-(^H^H^H^t-(^— t— > 


ooooooooooooooooooooo 


ooooooooocoaotopooooooooooooooo 

ddddddoddddd-HO^'-'-''-'— '—'^^— |^H^H— '^ 


ooooooooooopooooooopc^coc^t^ooo 
odddddddoddddddodo'o'dddoo^'-^'-1 


oooooooooooooooopoopoopopoo 


oooooooooooooooooooopooooop 


ooooc^i-'Oioocricxiooooppppoppoppooo 
dddoooddod'-<^'-*'-''-,^'-,'-,'-H,-,'-,^H— <--t'-i^H— < 

ooooooooooooooooooppoppooop 

C^WtDO^t^XCDC^X^I^C^^^^LO^^C^COCrJCOCOCO^'-H 
rriffrO^HNCO-^mcDt^MO^Of-iC^COTpiOCOC^OOOiOi-tWC 


Laidig  et  al.:  Descriptions  and  growth  of  larval  and  |uvenile  Sebastes  wtlsoni 


459 


head,  and  snout  also  increased  in  density,  and  pigment 
formed  posteriorly  along  the  upper  and  lower  jaws.  Me- 
lanophores  posteriorly  around  the  eye  socket  increased 
in  number.  The  fins  remained  unpigmented. 

Pygmy  rockfish  28.6  mm  long  had  dorsal  pigment 
that  stretched  continuously  from  the  jaws  to  the  caudal 
fin  (Fig.  IE.  Table  4).  Pigmentation  was  heavy  along 
the  dorsal  midline,  head,  and  nape.  Snout  pigmenta- 
tion also  intensified.  More  melanophores  were  present 
on  the  hypural  margin.  Pigment  along  the  ventral  body 
surface  darkened,  especially  in  the  area  posterior  to 
the  anal  fin.  More  melanophores  were  observed  at  the 
anal-fin  ray  articulations  than  on  smaller  specimens. 
The  three  body  bars  increased  in  width  and  length  and 
were  better  defined  than  on  smaller  specimens.  The  bar 
on  the  caudal  peduncle  began  to  exhibit  a  rectangular 
shape  that  is  characteristic  of  the  juvenile  stage.  The 
midbody  bar  also  took  on  a  rectangular  shape,  although 
the  dorsal  half  was  indented.  The  midbody  bar  and  the 
caudal  peduncle  bars  did  not  reach  the  ventral  midline. 
The  anterior  body  bar  extended  from  the  spinous  dorsal 
fin  to  the  vent  (see  arrow  Fig.  IE).  Anteriorly,  the  bar 
formed  a  more  or  less  rectangular  pattern  on  the  dorsal 
half  of  the  body  above  the  pectoral  fin.  In  general,  the 
lateral  body  surface  became  more  heavily  pigmented, 
especially  on  the  dorsal  half.  The  lateral  midbody  pig- 
ment line  began  to  be  incorporated  into  the  body  bars. 
Opercular  pigment  became  denser  and  merged  with 
the  nape  pigment.  The  area  anterior  to  the  nape  and 
operculum  was  less  pigmented  than  the  surrounding 
areas.  Pigment  along  the  posteroventral  portion  of  the 
orbit  became  denser  than  in  smaller  specimens.  A  cheek 
bar  began  to  form  ventral  to  the  eye  (as  evidenced  by 
the  two  melanophores  in  Fig.  IE ).  Melanophores  formed 
along  the  ventral  surface  of  the  lower  jaw  and  covered 
the  lateral  surface  of  the  upper  jaw.  Pigment  began  to 
develop  on  the  membranes  of  the  spinous  dorsal  fin, 
typically  with  some  unpigmented  areas  between  the 
dorsal  fin  pigment  and  the  dorsal  body  pigment. 

The  largest  specimen.  34.4  mm,  had  the  densest  and 
most  distinctive  pigmentation  (Fig.  IF,  Table  4).  Pig- 
ment was  present  on  most  of  the  body.  Along  the  dorsal 
surface,  the  pigment  formed  a  complete  line  from  the 
tip  of  the  upper  jaw  to  the  caudal  fin.  The  number  of 
melanophores  increased  along  the  hypural  region,  the 
postanal  ventral  midline,  and  at  the  anal-fin  articula- 
tions. The  mid-  and  caudal  body  bars  were  rectangular 
and  still  did  not  reach  the  ventral  midline,  leaving  an 
unpigmented  ventrolateral  area.  The  anterior  body  bar 
comprised  heavy  pigment  extending  posteriorly  between 
dorsal-fin  spines  VIII-XI  and  the  vent,  a  lighter  area 
just  anterior  to  this,  and  another  heavily  pigmented 
area  stretching  from  about  dorsal-fin  spines  III— VI 
almost  to  the  middle  of  the  gut  cavity.  Anterior  to  this 
bar  was  an  area  of  mottled  pigmentation.  Pigment  was 
visible  just  anterior  to  the  base  of  the  pectoral  fin. 
Pigment  covered  both  the  spinous  and  soft  dorsal  fins, 
except  along  the  distal  edge.  All  other  fins  remained 
unpigmented.  Opercular  pigment  was  dense  and  merged 
with  the  nape  pigment,  but  these  were  separated  from 


40 

35 
E 
£    30 

_SL  =  0.28(AGE)-5A5 

_c 

■  ms 

°>    25 

S     20 

~o 

a     15 

55 

10 

■             / 

20          40          60          80         100        120        140        160 

Age  (days) 

Figure  2 

Standard  length  and  age  of  pygmy  rockfish  i  Sebastes 

wilsoni)  in  =  60 1.  Solid  line  indicates  predicted  values 

from  linear  model. 

the  head  and  eye  pigment  by  an  area  of  low  pigment 
density.  Two  cheek  bars  radiated  from  the  lower  margin 
of  the  orbit.  Pigment  occurred  along  both  jaws  and  cov- 
ered the  snout  and  ventral  portion  of  the  lower  jaw. 

Otolith  examination 

A  linear  relationship  between  standard  length  and  age 
(as  estimated  from  otolith  increment  counts)  resulted  in 
a  good  estimate  of  growth  of  pygmy  rockfish  (slope  =  0.28 
mm/d;  intercept=-5.15  mm;  r2=0.91;  ?i  =  60;  Fig.  2).  The 
radius  of  the  extrusion  check  ranged  from  9.5  to  11.0 
^m,  averaging  10.5  /jm  (SD  =  0.29;  «  =  60).  Accessory  pri- 
mordia  first  appeared  in  a  19.8-mm  specimen  and  were 
observed  in  otoliths  from  all  larger  specimens.  Based  on 
this  character,  transition  from  larval  to  pelagic  juvenile 
stage  occurs  at  around  20  mm  SL. 

Molecular  confirmation 

Interspecific  levels  of  divergence,  calculated  among  adult 
reference  species,  ranged  from  0.13%  (rougheye  rockfish 
[S.  aleutianus]  vs.  shortraker  rockfish  [S.  borealis])  to 
9.7%  (black  rockfish  [S.  inermis]  vs.  bocaccio)  with  an 
average  of  4.1%.  Two  of  the  specimens  (FT2  and  FT3; 
Fig.  3A)  were  identical  to  one  of  the  adult  pygmy  rockfish 
references  (i.e.  0%  sequence  divergence  i  and  differed 
from  the  other  verified  adult  pygmy  rockfish  by  a  single 
nucleotide  substitution  (0.13%  seq.  div. ).  The  remain- 
ing two  specimens  (FT1  and  FT4;  Fig.  3A)  also  were 
most  similar  to  both  adult  pygmy  rockfish  references 
(0.13-0.40%  seq.  div.). 

Although  all  four  specimens  were  most  similar  to 
pygmy  rockfish  based  on  cytochrome  b  data,  only  a 
small  number  of  nucleotide  differences  separated  them 
from  Puget  Sound,  redstripe,  harlequin,  and  sharpchin 
rockfish  (0.27-1.87%;  Fig.  3A).  A  secondary  data  subset 
that  included  control  region  sequence  (cyt-6+CR)  yield- 
ed concordant  results;  all  four  larval  specimens  were 


460 


Fishery  Bulletin  102(3) 


3.5- 

A   Cyt-£> 

B     Cyt-to  +  CR 

3- 

• 
•                                       • 

• 

X  emphaeus  (BU5) 

2.5- 

X  emphaeus  (BU6) 
•  pronger  (BS4) 

>g 

□  vanegatus  (BS5) 

0) 

□  vanegatus  (BS7) 

=       2- 

O  zacentrus  (BQ7) 

• 

A  iv/fcoro(BP11) 

0 

o 

A  wilsoni  (BT1 1 ) 

> 

s 

-o 

•               • 

8    1.5- 

X 
H 

<D 

s                                          □ 

D 

X 

CT 

X                 H 

Q> 

w        1- 

□ 

D                 X 

X 

D 

x                                         o 

A 

o              X              x              □ 

A 

0.5- 

D                 O                 O                 H 

A 
A                                                                  A 

H                 D                 □                 * 

A 

a                 H                 H                 * 

A 

AAA 

0- 

1                              1                              t                             1 

12           3           4           12           3           4 

Larval  and  juvenile  specimens 

Figure  3 

Percent  sequence  divergence  based  on  the  number  of  nucleotide  differences  in  iAi 

cytochome  6  and  (Bi 

cytochrome  b  +  control  region  between  the  four  putative  pygmy  rockfish  I  Sebastes  wilsoni)  specimens 

iFT1-FT4i  and  five  closely  related  reference  species  of  Sebastes,  including  two  a 

dult  S.  wilsoni.  Per- 

cent  sequence  divergence  was  calculated  as  the  number  of  nucleotide  differences 

over  750  base  pairs 

(cyt-6)  and  1200  base  pairs  (cyt-6+CRi. 

most  similar  to  pygmy  rockfish  (0.25-0.83%;  Fig.  3B). 
Increased  levels  of  interspecific  nucleotide  variation, 
attributable  to  the  faster  evolving  control  region,  re- 
sulted in  more  pronounced  differences  between  the  four 
specimens  and  the  other  species  of  Sebastes  within  the 
subset  (range:  0.83-3.00%;  Fig.  3B).  Additionally,  a 
distance-based  analysis  (UPGMA)  of  haplotypes  (cyt- 
fr+CR)  clustered  all  four  specimens  with  pygmy  rockfish 
reference  material. 


Discussion 

Postflexion  larval  pygmy  rockfish  can  be  identified 
through  a  combination  of  pigment  and  meristic  char- 
acters. At  approximately  8-10  mm,  the  larval  pigment 
pattern  is  similar  to  only  four  of  the  30  Sebastes  spe- 
cies illustrated  in  the  literature  that  occur  within  our 
geographic  area  (Matarese  et  al.,  1989;  Moser,  1996; 
Laroche1):  yellowtail  (S.  flavidus),  blue  (S.  mystinus), 
canary  (S.  pinniger),  and  sharpchin  rockfish.  Yellowtail 
and  blue  rockfish  can  be  separated  from  pygmy  rockfish 
because  they  exhibit  ventral  body  and  hypural  pigment 
at  this  size — pigment  that  does  not  show  up  in  pygmy 
rockfish  until  approximately  14  and  15  mm,  respectively. 


In  canary  rockfish,  the  presence  of  ventral  body  pigment 
and  dorsal  midline  pigment  posterior  to  the  soft  dorsal 
fin  (instead  of  at  the  base  of  the  soft  dorsal-fin  rays  as  in 
pygmy  rockfish)  can  help  differentiate  this  species  from 
pygmy  rockfish.  Pigmentation  patterns  of  sharpchin 
rockfish  are  very  similar  to  pygmy  rockfish  at  10  mm; 
however,  sharpchin  rockfish  retain  pigmented  pelvic  fins 
until  12.7  mm  (Laroche  and  Richardson.  1981).  Counts  of 
anal-fin  rays  often  can  be  used  to  differentiate  these  two 
species  because  pygmy  rockfish  have  six  rays  and  sharp- 
chin have  rockfish  seven  rays  (Chen,  1986;  Matarese  et 
al.,  1989;  Moreland  and  Reilly,  1991;  Laroche1!.  There 
is  a  small  overlap  in  anal-fin  ray  counts  (approximately 
1'  i  I,  and,  because  of  this,  100%  certainty  of  identification 
cannot  be  reached  by  anal-fin  ray  counts  alone.  There- 
fore, in  order  to  increase  confidence  in  identifications,  a 
combination  of  pigmentation  and  fin-ray  counts  should 
be  employed.  After  approximately  15  mm.  a  full  comple- 
ment of  fin  rays  and  gill  rakers  typically  is  present  and 
can  be  used  in  combination  with  pigmentation  patterns 
to  differentiate  pygmy  rockfish  from  most  other  rockfish 
species.  In  these  late-stage  larvae,  only  three  species 
(yellowtail.  black  (S.  melanops),  and  blue  rockfish)  have 
a  pigment  pattern  that  could  be  confused  with  pygmy 
rockfish  (Matarese  et  al.,  1989;  Moser,  1996;  Laroche1). 


Laidig  et  al.:  Descriptions  and  growth  of  larval  and  juvenile  Sebostes  wilsoni 


461 


but  these  patterns  can  be  easily  separated  by  using 
meristic  characters. 

Pelagic  juvenile  pygmy  rockfish  have  a  distinctive 
pigment  pattern  consisting  of  three  body  bars  that  can 
be  used  to  discriminate  this  species  from  other  Sebastes 
species.  Yellowtail,  halfbanded,  and  redstripe  rockfish 
are  the  only  species  that  have  a  similar  three-barred 
pigment  pattern  (Matarese  et  al.,  1989;  Moser,  1996; 
Laroche1).  Yellowtail  rockfish  can  be  distinguished  by 
the  lack  of  cheek  bars  and  the  presence  of  body  bars 
that  extend  all  the  way  to  the  ventral  surface.  Also,  in 
yellowtail  rockfish,  the  body  bars  form  at  a  larger  size 
than  in  pygmy  rockfish.  In  halfbanded  rockfish,  the 
most  anterior  body  bar  is  more  densely  pigmented  than 
the  other  bars  and  typically  forms  a  diamond  shape. 
The  caudal  body  bar  is  much  wider  and  covers  the  en- 
tire peduncle.  Redstripe  rockfish  are  the  most  similar 
and  are  difficult  to  separate  from  pygmy  rockfish  by 
using  pigmentation  alone.  However,  these  two  species 
can  be  separated  with  greater  than  90%  certainty  by 
using  meristic  counts.  Pygmy  rockfish  have  a  mean 
anal-fin  ray  count  of  6  (95%  from  the  present  study,  and 
93%  from  Laroche1),  whereas  redstripe  rockfish  have  an 
average  of  7  anal-fin  rays  (100%  from  Chen,  1986;  97% 
from  Laroche1). 

It  should  be  noted  that  the  only  illustration  of  pygmy 
rockfish  prior  to  our  study  was  a  35.0-mm  pelagic  juve- 
nile by  Laroche,1  which  showed  several  pigment  differ- 
ences from  our  specimens  of  equivalent  size.  Laroche's 
illustrated  specimen  had  only  faint  body  barring,  no 
cheek  bars,  and  no  ventral  pigment,  whereas  all  our 
specimens  had  prominent  body  barring,  at  least  one 
cheek  bar,  and  ventral  pigment  along  the  anal-fin  ar- 
ticulations. At  this  time  we  cannot  determine  whether 
these  differences  were  due  to  geographic  variability  in 
pigment  patterns  (Laroche's  specimen  probably  was 
collected  farther  north  than  all  of  our  specimens),  or 
a  misidentification  of  the  original  specimen  illustrated 
by  Laroche.1 

The  identification  of  larval  and  pelagic  juvenile  pygmy 
rockfish  used  in  our  study  was  confirmed  by  using  DNA 
sequence  analyses.  Previous  molecular  identifications 
and  subsequent  descriptions  of  juvenile  starry  rockfish 
(S.  constellatus)  and  swordspine  rockfish  (S.  ensifer) 
also  were  based  on  mitochondrial  cytochrome  b  data 
(Rocha-Olivares  et  al.,  2000).  In  our  study,  orthologous 
cytochrome  b  sequence  was  sufficient  for  identifica- 
tion purposes,  particularly  for  those  specimens  exhibit- 
ing exact  haplotype  matches  to  reference  adult  pygmy 
rockfish  (e.g.,  FT2/FT3:  0.0%  sequence  divergence). 
Relatively  low  levels  of  interspecific  genetic  variation 
occurred  between  larval  specimens  and  several  refer- 
ence species  (pygmy,  sharpchin.  harlequin,  and  Puget 
Sound  rockfish,  and,  to  a  lesser  extent,  redstripe  rock- 
fish). Rocha-Olivares  et  al.  (1999a)  used  control  region 
sequence,  in  addition  to  cytochrome  b,  to  resolve  phy- 
logenetic  relationships  among  recently  diverged  species 
of  the  Sebastes  subgenus  Sebastomus.  In  the  present 
study,  the  control  region  sequence  was  used  to  increase 
divergence  levels  between  species  and  to  aid  in  insur- 


ing correct  molecular  identifications  of  specimens  FT1 
and  FT4.  Species  assignment  to  pygmy  rockfish  was 
supported  by  the  smallest  divergence  (based  on  cyt-6 
and  cyt-6+CR)  from  reference  pygmy  rockfish  compared 
with  the  other  Sebastes  species. 

Larval  and  juvenile  pygmy  rockfish  can  also  be  sepa- 
rated from  other  Sebastes  species  by  comparing  the 
radius  of  the  extrusion  check  on  their  otoliths.  Of  the 
fourteen  other  Sebastes  species  or  species  complexes 
with  measured  otolith  extrusion  check  radii  ( Laidig  and 
Ralston,  1995;  Laidig  et  al.,  1996;  Laidig  and  Sakuma, 
1998),  only  four  species  and  two  complexes  have  radii 
close  to  the  average  extrusion  check  radii  for  pygmy 
rockfish  (10.5  nm,  SD  =  0.3).  Stripetail  rockfish  (S.  saxi- 
cola)  had  an  average  extrusion  check  radius  (11.6  (im, 
SD  =  0.5)  that  was  larger  than  the  largest  radius  for 
pygmy  rockfish  (11.0  (jm).  Quillback  rockfish  (S.  ma- 
liger)  had  an  average  extrusion  check  radius  of  9.1  jim 
(SD  =  0.1).  which  was  smaller  than  the  smallest  radius 
for  pygmy  rockfish  (9.5  p.m).  Species  with  extrusion 
check  radii  similar  to  pygmy  rockfish  were  kelp  rockfish 
(S.  atrovirens)  at  10.6  ^m  (SD  =  0.2),  blue  rockfish  at 
10.9  /./m  (SD  =  1.1),  and  the  copper  rockfish  (S.  caurinus, 
extrusion  check  radius  =  10.5  /jm;  SD  =  0.4)  and  gopher 
rockfish  (S.  carnatus,  extrusion  check  radius  =  10.6  ,um; 
SD  =  0.3)  complexes  (see  Laidig  et  al.,  1996.  for  complex 
definitions).  Of  these  species,  the  only  one  that  would  be 
confused  with  pygmy  rockfish,  by  pigmentation  alone, 
would  be  blue  rockfish  at  small  sizes.  However,  pygmy 
rockfish  and  blue  rockfish  are  easily  separated  by  using 
meristic  characters. 

Growth  rates  of  larval  rockfish  generally  are  slow 
during  the  first  month  of  life  and  increase  thereafter 
(Laidig  et  al.,  1991;  Sakuma  and  Laidig,  1995;  Laidig 
et  al.,  1996).  Because  the  youngest  fish  in  our  study 
was  estimated  to  be  40  days  old,  our  linear  model  can 
not  be  used  to  estimate  early  larval  growth  rates.  For 
pygmy  rockfish  older  than  40  days,  the  growth  rate  of 
0.28  mm/day  was  somewhat  slower  than  that  observed 
for  other  Sebastes.  Woodbury  and  Ralston  (1991)  found 
that,  for  fish  older  than  40  days,  growth  rates  varied 
from  0.30  for  widow  rockfish  (S.  entomelas)  to  0.97  mm/ 
day  for  bocaccio.  Other  species  exhibiting  slightly  faster 
growth  rates  after  40  days  of  age  include  stripetail 
rockfish  (0.37  mm/day;  Laidig  et  al.,  1996),  grass  rock- 
fish (S.  rastrelliger;  0.36  mm/day;  Laidig  and  Sakuma, 
1998),  and  shortbelly  rockfish  (S.  jordani;  0.53  mm/day; 
Laidig  et  al.,  1991).  Yellowtail  rockfish  had  a  more 
similar  growth  rate,  ranging  from  0.19  to  0.46  mm/day 
(Woodbury  and  Ralston,  1991).  These  differences  in 
growth  may  reflect  genetic  variability  or  responses  to 
environmental  variables.  Woodbury  and  Ralston  (1991) 
suggested  that  annual  variability  in  growth  rates  of 
juvenile  rockfish  was  related  to  year-to-year  changes 
in  environmental  conditions,  especially  temperature. 
Boehlert  (1981)  determined  that  temperature  greatly 
affected  growth  rate  of  young  splitnose  rockfish  (S.  dip- 
loproa)  in  the  laboratory.  Boehlert  and  Yoklavich  (1983) 
observed  slower  growth  rates  for  black  rockfish  in  colder 
temperatures.  Lenarz  et  al.  (1991)  analyzed  the  vertical 


462 


Fishery  Bulletin  102(3) 


distribution  of  late  larval  and  pelagic  juvenile  rockfish 
and  determined  that  pygmy  rockfish  were  present  on 
average  in  deeper,  colder  water  than  that  favored  by 
other  rockfish  species.  This  spatial  separation  of  pelagic 
juvenile  pygmy  rockfish  and  other  Sebastes  spp.  may 
explain  the  slower  growth  observed  in  pygmy  rockfish. 


Acknowledgments 

We  would  like  to  thank  the  scientists  and  crew  from 
the  Southwest  Fisheries  Science  Center  (SWFSC)  who 
collected  the  samples  aboard  the  NOAA  RV  David  Starr 
Jordan.  We  thank  Geoff  Moser  and  Bill  Watson  (NOAA, 
SWFSC)  for  examining  some  of  our  pygmy  rockfish 
specimens.  Reference  sequences  of  Sebastes  were  gener- 
ated and  kindly  provided  by  personnel  at  the  Fisheries 
Resources  Division  of  the  SWFSC,  La  Jolla,  CA  (R.  D. 
Vetter,  A.  Rocha-Olivares,  B.  J.  Eitner,  C.  A.  Kimbrell, 
and  C.  Taylor).  In  addition,  we  thank  Mary  Yoklavich 
for  all  her  valuable  comments  and  all  the  reviewers  who 
contributed  to  this  manuscript. 


Literature  cited 

Boehlert.  G.  W. 

1981.     The  effects  of  photoperiod  and  temperature  on  the 
laboratory  growth  of  juvenile  Sebastes  diploproa  and 
a  comparison  with  growth  in  the  field.     Fish.  Bull. 
79:789-794. 
Boehlert,  G.  W„  and  M.  M.  Yoklavich. 

1983.     Effects  of  temperature,  ration,  and  fish  size  on 
growth  of  juvenile  black  rockfish,  Sebastes  melanops. 
Environ.  Biol.  Fish.  8:17-28. 
Chen,  L. 

1986.     Meristic  variation  in  Sebastes  iScorpaenidaei.  with 
an  analysis  of  character  association  and  bilateral  pattern 
and  their  significance  in  species  association.     NOAA 
Tech.  Rep.  NMFS-45,  17  p. 
Hillis,  D.  M.,  B.  K.  Mable,  A.  Larson,  S.  K.  Davis,  and 
E.  A.  Zimmer. 

1996.     Nucleic  acids  IV:  sequencing  and  cloning.     Chapter 
9  in  Molecular  systematics,  2nd  ed.  ID.  M.  Hillis.  C. 
Moritz,  and  B.  K.  Mable,  eds.l,  p.  321-381.     Sinauer 
Associates,  Sunderland,  MA. 
Hunter,  J.  R..  and  N.  C.  -H.  Lo. 

1993.     Ichthyoplankton  methods  for  estimating  fish  bio- 
mass  introduction  and  terminology.     Bull.  Mar.  Sci. 
53:723-727. 
Laidig.  T  E.,  and  S.  Ralston. 

1995.     The  potential  use  of  otolith  characters  in  iden- 
tifying larval  rockfish  (Sebastes  spp.).     Fish.  Bull. 
93:166-171. 
Laidig,  T.  E.,  S.  Ralston,  and  J.  R.  Bence. 

1991.     Dynamics  of  growth  in  tin-  early  life  history  of  short- 
belly  rockfish,  Sebastes  jordani.     Fish.  Bull.  89:611- 
621. 
Laidig,  T.  E..  and  K.  M.  Sakuma. 

1998.  Description  of  pelagic  larval  and  juvenile  grass 
rockfish,  Sebastes  rastrelliger  (Family  Scorpaenidaei. 
with  an  examination  of  age  and  growth  Fish  Bull. 
96:788-796. 


Laidig,  T.  E.,  K.  M.  Sakuma,  and  M.  M.  Nishimoto. 

1996.  Description  of  pelagic  larval  and  juvenile  strip- 
etail  rockfish,  Sebastes  saxicola  (family  Scorpaenidaei. 
with  an  examination  of  larval  growth.  Fish  Bull. 
94:289-299. 

Laroche,  W.  A.,  and  S.  L.  Richardson. 

1981.  Development  of  larvae  and  juveniles  of  the  rock- 
fishes  Sebastes  entomelas  and  S.  zacentrus  (Family 
Scorpaenidaei  and  occurrence  off  Oregon,  with  notes 
on  head  spines  of  S.  mystinus,  S.  flavidus,  and  S. 
melanops.     Fish.  Bull.  79:231-257. 

Lee,  T.-W..  and  G.-C.  Kim 

2000.  Microstructural  growth  in  otoliths  of  black  rock- 
fish, Sebastes  schlegeli,  from  prenatal  larval  to  early 
juvenile  stages.     Ichthyol.  Res.  47:335-341. 

Lenarz,  W.  H.,  R.  J.  Larson.  S.  Ralston. 

1991.  Depth  distributions  of  late  larvae  and  pelagic  juve- 
niles of  some  fishes  of  the  California  Current.  CalCOFI 
(California  Cooperative  Oceanic  Fisheries  Investiga- 
tions! Rep.  32:41-46. 

Love.  M. 

1996.  Probably  more  than  you  wanted  to  know  about 
the  fishes  of  the  Pacific  coast,  381  p.  Really  Big  Press. 
Santa  Barbara,  CA. 

Love.  M.,  M.  Yoklavich.  and  L.  Thorsteinson. 

2002.  The  rockfishes  of  the  northeast  Pacific,  405  p. 
Univ.  California  Press.  Berkeley,  CA. 

Matarese.  A.  C,  A.  W.  Kendall  Jr.,  D.  M.  Blood,  and 
B.  M.  Vinter. 

1989.     Laboratory  guide  to  early  life  history  stages  of 
northeast  Pacific  fishes.     NOAA  Tech.  Rep.  NMFS- 
80,  652  p. 
Moreland,  S.  L„  and  C.  A.  Reilly. 

1991.  Key  to  the  juvenile  rockfishes  of  central 
California.  In  Methods  used  to  identify  pelagic  juvenile 
rockfish  (genus  Sebastes)  occurring  along  the  coast  of 
central  California  (T.  E.  Laidig  and  P.  B.  Adams,  eds.l. 
p.  59-180.  NOAA  Tech.  Memo..  NOAA-TM-NMFS- 
SWFSC-166. 
Moser,  H.  G. 

1996.     The  early  stages  of  fishes  in  the  California  Cur- 
rent region.     Calif.  Coop.     Oceanic  Fish.  Invest.  Atlas 
no.  33,  1505  p. 
Moser,  H.  G..  and  J.  L.  Butler. 

1987.     Descriptions  of  reared  larvae  of  six  species  of 
Sebastes.     In  Widow  rockfish:  proceedings  of  a  work- 
shop, Tiburon,  California,  December  11-12,  1980  i  W   H 
Lenarz  and  D.  R.  Gunderson,  eds.l,  p.  19-29.     NOAA 
Tech.  Rep.  NMFS-48. 
Pacific  Fishery  Management  Council. 

2000.     Status  of  the  Pacific  coast  groundfish  fishery 
through  1999  and  recommended  acceptable  biological 
catches  for  2000,  165  p.     Pacific  Fishery  Management 
Council.  Portland.  OR. 
Ralston,  S.,  J.  R.  Bence,  M.  B.  Eldridge,  and  W.  H   Lenarz. 

2003.  An  approach  to  estimating  rockfish  biomass 
based  on  larval  production,  with  application  to  Sebastes 
jordani.     Fish.  Bull.  101:129-146. 

Richardson.  S.  L..  and  W.  A.  Laroche. 

1979.     Development  and  occurrence  of  larvae  and  juveniles 
of  the  rockfishes  Sebastes  erameri.  Sebastes  pinniger, 
and  Sebastes  helvomaculatus  '  Family  Scorpaenidae  i  <iff 
Oregon.     Fish.  Bull.  77:1-46 
Rocha-Olivares.  A. 

1998.  Multiplex  haplotype-specific  PCR:  a  new  approach 
for  species  identification  of  the  earh  life  sialics  of  rock- 


Laidig  et  al.:  Descriptions  and  growth  of  larval  and  juvenile  Sebastes  wi/soni 


463 


fishes  of  the  species-rich  genus  Sebastes  Cuvier.     J. 
Exp.  Mar.  Biol.  Ecol.  231:279-290. 

Rocha-Olivares,  A.,  C.  A.  Kimbrell,  B.  J.  Eitner,  and  R.  D.  Vetter. 
1999a.  Evolution  of  a  mitochondrial  cytochrome  b  gene 
sequence  in  the  species-rich  genus  Sebastes  (Teleostei, 
Scorpaenidael  and  its  utility  in  testing  the  monophyly 
of  the  subgenus  Sebastomus.  Mol.  Phylogenet.  Evol. 
11:426-440. 

Rocha-Olivares,  A.,  H.  G.  Moser,  and  J.  Stannard. 

2000.  Molecular  identification  and  description  of  juve- 
nile stages  of  the  rockfishes  Sebastes  eonstellatus  and 
Sebastes  ensifer.     Fish.  Bull.  98:353-363. 

Rocha-Olivares,  A..  R.  H.  Rosenblatt,  and  R.  D.  Vetter. 

1999b.  Molecular  evolution,  systematics,  and  zoogeog- 
raphy of  the  rockfish  subgenus  Sebastomus  (Sebastes, 
Scorpaenidael  based  on  mitochondrial  cytochrome  b 
and  control  region  sequences.  Mol.  Phylogenet.  Evol. 
11:441-458. 

Sakuma,  K.  M„  and  T.  E.  Laidig. 

1995.  Description  of  larval  and  pelagic  juvenile  chilipep- 
per,  Sebastes goodei  (family  Scorpaenidael,  with  an  exam- 
ination of  larval  growth.     Fish.  Bull.  93:721-731. 

Seeb,  L.  W.,  and  A.  W.  Kendall  Jr. 

1991.  Allozyme  polymorphisms  permit  the  identifi- 
cation of  larval  and  juvenile  rockfishes  of  the  genus 
Sebastes.     Environ.  Biol.  Fishes  30:191-201. 


Stein,  D.  L.,  B.  N.  Tissot,  M.  A.  Hixon,  and  W.  Barss. 

1992.  Fish-habitat  associations  on  a  deep  reef  at  the  edge 
of  the  Oregon  continental  shelf.  Fish.  Bull.  90:540- 
551. 

Stransky,  C. 

2001.     Preliminary  results  of  a  shape  analysis  of  redfish 
otoliths:  comparison  of  areas  and  species,  9  p.     North 
Atl.  Fish.  Org.  Sci.  Coun.  Res.  Doc.  01/14. 
Winnepenninckx.  B.,  T.  Backeljau,  and  R.  DeWachter. 

1993.  Extraction  of  high  molecular  weight  DNA  from 
mollusks.     Trends  Genet.  9:407. 

Woodbury,  D.  P.,  and  S.  Ralston. 

1991.     Interannual  variation  in  growth  rates  and  back-cal- 
culated birthdate  distributions  of  pelagic  juvenile  rock- 
fishes (Sebastes  spp.)  off  central  California  coast.     Fish. 
Bull.  89:523-533. 
Yoklavich,  M.  M„  and  G.  W.  Boehlert. 

1987.     Daily  growth  increments  in  otoliths  of  juvenile 
black  rockfish,  Sebastes  melanops:  An  evaluation  of 
autoradiography  as  a  method  of  validation.     Fish.  Bull. 
85:826-832. 
Yoklavich,  M.  M.,  H.  G.  Greene,  G.  M.  Cailliet,  D.  E.  Sullivan, 
R.  N.  Lea,  and  M.  S.  Love. 

2000.  Habitat  associations  of  deep-water  rockfishes  in  a 
submarine  canyon:  an  example  of  a  natural  refuge.  Fish. 
Bull.  98:625-641. 


464 


Abstract— A  critical  process  in  assess- 
ing the  impact  of  marine  sanctuaries 
on  fish  stocks  is  the  movement  offish 
out  into  surrounding  fished  areas. 
A  method  is  presented  for  estimat- 
ing the  yearly  rate  of  emigration  of 
animals  from  a  protected  ("no-take") 
zone.  Movement  rates  for  exploited 
populations  are  usually  inferred  from 
tag-recovery  studies,  where  tagged 
individuals  are  released  into  the  sea 
at  known  locations  and  their  location 
of  recapture  is  reported  by  fishermen. 
There  are  three  drawbacks,  however, 
with  this  method  of  estimating  move- 
ment rates:  1)  if  animals  are  tagged 
and  released  into  both  protected  and 
fished  areas,  movement  rates  will  be 
overestimated  if  the  prohibition  on 
recapturing  tagged  fish  later  from 
within  the  protected  area  is  not  made 
explicit;  2)  the  times  of  recapture  are 
random;  and  3)  an  unknown  propor- 
tion of  tagged  animals  are  recaptured 
but  not  reported  back  to  research- 
ers. An  estimation  method  is  pro- 
posed which  addresses  these  three 
drawbacks  of  tag-recovery  data.  An 
analytic  formula  and  an  associated 
double-hypergeometric  likelihood 
method  were  derived.  These  two 
estimators  of  emigration  rate  were 
applied  to  tag  recoveries  from  south- 
ern rock  lobsters  (Jasus  edwardsil) 
released  into  a  sanctuary  and  into 
its  surrounding  fished  area  in  South 
Australia. 


Estimating  the  emigration  rate  of  fish  stocks 
from  marine  sanctuaries  using  tag-recovery  data 

Richard  McGarvey 

Aquatic  Sciences,  South  Australian  Research  and  Development  Institute  (SARDI) 

2  Hamra  Avenue 

West  Beach,  South  Australia  5024,  Australia 

E-mail  address;  mcgarvey  nchardia  saugovsa  gov.au 


Manuscript  submitted  24  March  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
5  March  2004  by  the  Scientific  Editor. 
Fish.  Bull.  102:464-472(20lili 


Marine  sanctuaries,  also  known  as 
marine  protected  areas  (MPAs),  ma- 
rine reserves,  and  no-take  areas,  are 
being  widely  promoted  and  imple- 
mented. Important  for  assessing  the 
impact  of  these  "no-take"  sanctuaries 
(from  which  fishing  has  been  excluded  I 
on  exploited  populations  is  the  rate 
of  emigration  of  animals  out  into  the 
remaining  fished  habitat. 

The  most  widely  available  data  for 
estimating  movement  rates  of  com- 
mercially or  recreationally  exploited 
populations  are  those  from  tagged 
and  recovered  fish  (Hilborn,  1990). 
Animals  are  captured  alive,  a  visible 
numbered  tag  is  inserted  and  they 
are  released  back  into  the  wild.  Be- 
cause the  accuracy  of  tag-recovery 
studies  relies  on  fishermen  reporting 
recaptured  tags,  the  quality  of  tag-re- 
covery information  is  lower  than  that 
from  a  controlled  experiment. 

Tag-recovery  experiments  have 
three  limitations  for  estimating  move- 
ment rates  of  animals — the  first  two 
apply  to  most  tagged  populations,  the 
third  applies  specifically  to  emigra- 
tion from  sanctuaries:  1)  times  at 
large  (the  numbers  of  days  from  when 
each  animal  is  tagged  and  released  to 
when  it  is  subsequently  recaptured 
in  the  fishery)  are  highly  variable;  2) 
not  all  recaptured  tags  are  reported 
to  researchers  by  fishermen  and  this 
rate  of  tag  nonreporting  is  often  un- 
known; and  3)  tag  recoveries  cannot 
be  obtained  from  within  sanctuaries 
for  the  simple  reason  that  no  fishing 
is  allowed  there. 

If  this  last  asymmetry  (of  recap- 
tures from  the  sanctuary  coming  only 
from  tagged  animals  that  emigrate) 
is  not  accounted  for  in  the  estimation 
model,  then  the  emigration  rate  out  of 


the  sanctuary  will  be  overestimated. 
With  previous  movement  estimators, 
tag  releases  and  recaptures  from  all 
strata  have  been  assumed.  The  aim 
of  the  present  article  is  to  develop  an 
unbiased  estimator  of  emigration  rate 
from  no-take  areas  by  using  data  of 
tag  releases  both  into  the  sanctuary 
and  into  the  fished  zone  surrounding 
it,  but  where  recoveries  from  nonmov- 
ing  tagged  animals  are  only  possible 
from  the  fished  zone.  An  estimate  for 
the  recovery  rate  (proportion  of  fish 
recaptured  and  their  tags  reported)  in 
the  fished  zone  was  also  obtained. 


Materials  and  methods 

Tag-recovery  data 

The  data  used  to  estimate  the  emi- 
gration rate  from  Gleesons  Landing 
Lobster  Sanctuary  (Fig.  1)  are  tag 
recoveries  from  lobsters  tagged  and 
released  both  inside  the  sanctuary 
and  into  the  fished  zone  surrounding 
the  sanctuary.  A  large  South  Aus- 
tralian lobster  tagging  program  was 
undertaken  in  1993-96  throughout 
South  Australian  waters.  T-bar  tags 
(Hallprint,  Victor  Harbour,  South 
Australia)  were  inserted  into  the 
ventral  muscle  at  the  first  segment 
of  the  lobster  abdomen.  The  rate  of 
tag  shedding  was  estimated  from 
double  tags  at  between  6fr  and  1292 
per  year  (Xiao1)  and  is  incorporated 
in  the  recovery  rate. 


Xiao  Y.  2003.  Personal  commun. 
Aquatic  Sciences,  South  Australian 
Research  and  Development  Institute 
(SARDI),  P.O.  Box  120.  Henley  Beach, 
South  Australia  5022.  Australia. 


McGarvey:  Estimating  emigration  rates  from  marine  sanctuaries  using  tag-recovery  data 


465 


Gleesons  Sanctuary  and  surrounding  blocks 

| |  Yorke  Peninsula  and  outlying  blocks 

'-,-  |  Gleesons  Landing  Lobster  Sanctuary 
I  |  Block  33 

1  Block  40 


Yotke  Peninsula 


40 


39 


35  S 


Figure  1 

Location  of  Gleesons  Landing  Lobster  Sanctuary  (small  dark  area  on  the  bound- 
ary of  MFA  blocks  33  and  40)  along  the  west  coast  of  the  Yorke  Peninsula  in 
South  Australia. 


As  part  of  this  tagging  program  (Table  1),  3235  south- 
ern rock  lobsters  ijasus  edwardsii)  were  tagged  and 
released  into  the  "fished  zone"  surrounding  Gleesons 
Sanctuary,  namely  into  statistical  reporting  blocks  33 
and  40  (Fig.  1).  In  January  1994,  413  lobsters  were  cap- 
tured, tagged,  and  released  inside  the  Gleesons  Sanctu- 
ary. These  lobsters  were  predominantly  in  the  range  of 
80-120  mm  carapace  length  (CL),  around  the  size  of 
maturity  of  about  100  mm  CL;  more  lobsters  below  the 
legal  minimum  length  (98.5  mm  CL)  were  released  in 
the  fished  zone. 

Gleesons  Landing  Lobster  Sanctuary  (Fig.  1)  is  an 
area  where  lobster  fishing  has  been  prohibited  since 
1982.  It  lies  along  the  Yorke  Peninsula's  western  coast 
in  an  area  of  medium  to  low  lobster  catches.  In  width, 
this  sanctuary  extends  1-2  km  from  shore  to  seaward 
and  runs  7-8  km  north-south. 

Nearly  all  tag  recoveries  were  reported  by  commercial 
lobster  fishermen  who  noticed  tagged  lobsters  in  their 
catch  in  the  course  of  day-to-day  fishing  operations.  Tag 
recoveries  of  lobsters  released  into  both  the  sanctuary 
and  the  fished  zone  were,  therefore,  only  possible  from 


the  fished  zone.  GPS  coordinates,  date,  and  carapace 
length  were  recorded  for  all  tagged  and  recaptured 
lobsters.  Longer-range  movements  from  both  sanctu- 
ary and  fished  zone  were  directed  southwest  towards 
the  shelf  edge. 

Prescott  et  al.2  previously  described  qualitative  fea- 
tures of  the  movement  of  South  Australian  Jasus  ed- 
wardsii: 1)  nearly  all  longer-distance  movements  were 
directed  offshore  to  deeper  water  and  away  from  the 
coast;  2)  in  order  of  greater  to  lesser  average  distances 
moved,  were  i)  immature  females,  ii)  males,  and  iii) 
mature  or  egg-bearing  females,  for  nearly  all  five  South 
Australian  regions  analyzed;  3 1  movements  were  largely 
restricted  to  lobsters  in  a  specific  length  range  at  time 


2  Prescott,  J.,  R.  McGarvey,  G.  Ferguson,  and  M.  Lorkin. 
1998.  Population  dynamics  of  the  southern  rock  in  South 
Australian  waters.  Fisheries  Research  and  Development 
Corporation  of  Australia  Report  93/086,  p.  23-27.  Aquatic 
Sciences,  South  Australian  Research  and  Development  Insti- 
tute (SARDI),  P.O.  Box  120,  Henley  Beach,  South  Australia 
5022,  Australia. 


466 


Fishery  Bulletin  102(3) 


Table  1 

Tag-recovery  data  from  Gleesons  Landing  lobster  sanctuary  and  the  surrounding  fishing  zone  used  in  estimating  yearly  move- 
ment rates  of  southern  rock  lobsters. 


Data 


Variable 
name 


Observed  number 
of  lobsters 


Number  of  lobsters  tagged  and  released  into  the  sanctuary 

Number  of  lobsters  recovered  that  had  moved  (>3  km  I  from  the  sanctuary 

into  the  fishing  zone 
Number  of  lobsters  tagged  and  released  into  the  surrounding  fishing  zone 
Number  of  lobsters  recovered  that  had  moved  (>3  km)  within  the  fishing  zone 
Number  of  lobsters  ecovered  that  had  not  moved  (>3  km)  within  the  fishing  zone 


Nf 


NF 
NF 
NF 


413 

29 

3235 

89 

277 


of  tagging,  roughly  100-140  mm  CL  for  females,  and 
100-150  mm  CL  for  males,  with  a  noticeable  shift  to 
smaller  sizes  for  both  sexes  on  the  southeast  coast  of 
South  Australia  where  growth  and  thus  size  of  maturity 
are  known  to  be  lower;  4)  overall,  most  lobsters  in  the 
fished  areas  did  not  move  large  distances,  about  15% 
moving  more  than  5  km;  5)  two  areas  stood  out  as  being 
habitats  from  where  significant  movement  occurred,  the 
coastal  zone  off  the  Coorong  and  Yorke  Peninsula;  and 
6 )  for  Yorke  Peninsula,  higher  than  proportional  num- 
bers of  tagged  lobsters  that  moved  significant  distances 
were  tagged  and  released  inside  Gleesons  Sanctuary. 

In  the  present  study  study,  a  lobster  was  classified  as 
having  undergone  movement  if  its  measured  distance 
from  point  of  tagging  to  point  of  recapture  was  greater 
than  3  km.  This  definition  of  lobster  "movement"  was 
chosen  for  two  reasons.  1)  The  mean  width  of  MPA 
coastal  zone  to  be  protected  in  the  currently  proposed 
state  representative  system  is  assumed  to  be  5  km  wide; 
that  is,  it  is  assumed  that  sanctuary  areas  will  extend 
from  the  shore  outward  to  sea  across  the  full  3  nmi 
(which  is  about  5  km)  of  state  territorial  waters.  Thus, 
a  3-km  movement  would  represent  slightly  more  than 
the  mean  distance  needed  for  lobsters  to  leave  the  state- 
protected  territorial  waters  of  the  reserve  and  enter  wa- 
ters open  for  fishing.  This  assumption  is  strengthened 
by  the  knowledge  that  most  longer-range  movements  of 
South  Australian  rock  lobster  are  directed  from  inshore 
to  offshore.  2)  According  to  the  geographical  features 
of  the  present  study,  a  3-km  movement  seaward  from 
any  location  in  Gleesons  Landing  Sanctuary  would 
place  the  tagged  lobster  well  into  the  fished  zone,  i.e., 
it  would  constitute  a  movement  out  of  the  sanctuary.  Of 
sanctuary-tagged  lobsters,  4  of  33  recaptured  lobsters 
in  the  first  season  after  tagging  exited  the  reserve  but 
moved  less  than  3  km.  These  4  recaptured  lobsters  were 
excluded  from  the  data  set.  The  mean  distance  moved 
by  lobsters  from  the  sanctuary  was  37.4  km. 

Because  movement  of  South  Australian  lobsters  is 
directed  strongly  away  from  the  inshore  zone,  the  im- 
migration rate  of  lobsters  back  into  the  Gleesons  Land- 
ing Sanctuary  is  likely  to  be  quite  low.  Moreover,  Jasus 
edwardsii  seek  shelter  daily  and  remain  on  specific 


reefs  through  most  of  their  life  (MacDiarmid  et  al.  1991; 
Kelly  2001).  Long-distance  movements  occur  rarely 
more  than  once  in  a  lifetime.  Thus,  in  the  fishing  zone, 
where  there  is  a  continual  removal  of  adult  lobsters 
from  reef  habitat,  the  on-going  creation  of  new  shelter 
space  is  higher  than  in  the  sanctuary  and  thus  lobsters 
that  did  stray  inshore  into  the  sanctuary  would  be  less 
likely  to  find  shelter,  further  reducing  the  probability 
of  migration  into  the  sanctuary.  In  the  estimator  pre- 
sented below,  only  the  emigration  rate  (the  movement 
rate  out  of  the  sanctuary)  is  calculated. 

The  recapture  data  included  lobsters  at  large  for  a 
wide  range  of  times,  many  having  been  recaptured  lon- 
ger than  one  year  after  tag  release.  However,  to  estimate 
emigration  rate,  we  sought  the  proportion  of  lobsters 
emigrating  out  per  year.  Therefore,  subsets  of  recapture 
data  were  selected  that  had  a  mean  time  at  large  of  one 
year.  The  temporal  distributions  of  recaptured  lobsters 
showed  distinct  modes  around  1  year  at  large  (recap- 
tures between  0.5  and  1.5  years  at  large.  Fig.  2),  and  the 
number  of  recaptures  in  these  1-year  modes  were  used 
for  estimating  yearly  movement  rate  (Table  1). 

Some  tagged  and  released  lobsters  were  recaptured 
more  than  once.  For  these  lobsters,  the  single  recapture 
was  selected  and  used  for  which  the  time  at  large  was 
closest  to  one  full  year. 

Notation 

The  information  on  movement  in  each  set  of  tag  releases 
is  taken  to  be  binary:  each  recaptured  animal  is  clas- 
sified as  having  moved  or  as  having  not  moved  during 
its  approximately  1-year  time  at  large  (from  time  of  tag 
release  to  time  of  recapture). 

To  carry  out  the  movement-rate  estimation,  it  is  use- 
ful to  consider  the  complete  set  of  four  possible  outcomes 
for  each  tagged  and  released  animal:  1)  it  moved  and 
was  recovered  after  one  year  (denoted  M,R);  2)  it  did 
not  move  and  was  recovered  after  one  year  (NM.R);  3) 
it  moved  and  was  not  recovered  after  one  year  (M.NR); 
4)  it  did  not  move  and  was  not  recovered  after  one  year 
(NM.NR).  These  four  possible  recapture  outcomes  ap- 
plied to  animals  tagged  and  released  in  both  strata, 


McGarvey:  Estimating  emigration  rates  from  marine  sanctuaries  using  tag-recovery  data 


467 


80 
60 
50 
40 
20 


12 
10 
8 
6 
4 
2 


0J 


Fished  zone 


M 


}MM) 


0  0 

1  — JT-Th-n-i    ~ 


Sanctuary 


firm      \h 


m 


0  12  3  4  5  6 

Time  at  large  (years) 

Figure  2 

Histograms  over  time  at  large  (in  monthly  bins!  of  recapture 
numbers  from  the  fishing  zone  iMFA  blocks  33  and  40  i  and  from 
the  Gleesons  Landing  Lobster  Sanctuary.  The  diamond  mark- 
ers indicate  divisions  between  modes  at  0.5,  1.5.,  2.5,  etc.  years 
at  large;  recaptures  from  the  sanctuary  and  fishing  zone  that 
occurred  between  0.5  and  1.5  years  after  release  (between  the  black 
diamond  markers!  identify  the  subsets  of  data  used  to  estimate 
yearly  emigration  rate  from  the  sanctuary. 


inside  and  outside  the  sanctuary.  The  tag-recovery  data 
provided  direct  measures  for  only  three  of  these  eight 
possible  numbers  of  recaptures. 

We  define  "not  recovered"  to  include  both  tagged  ani- 
mals that  were  not  recaptured,  as  well  as  those  that 
were  recaptured  by  a  fisherman  but  whose  tag  informa- 
tion (notably  the  location  of  recapture)  was  not  reported 
back  to  researchers  and  therefore  was  not  included  in 
the  tag-recovery  database. 

The  movement-rate  estimate  is  given  in  terms  of  the 
following  data  inputs:  the  number  of  lobsters  tagged 
and  released  in  li  fished  and  2)  protected  zones,  and 
the  numbers  recovered  that  3)  moved  (>3  km)  or  4)  did 
not  move  from  the  fished  zone  over  one  year  after  tag- 
ging, and  the  5)  number  that  moved  (>3  km)  from  the 
sanctuary  in  one  year. 

Superscripts  'F'  and  'S'  denote  fished  zone  and  sanc- 
tuary, respectively,  for  the  location  of  tag  release.  Let 
N^MR  and  Nfj  R  denote  the  numbers  of  animals  that 
were  recovered  after  a  year  and  that  moved  or  that  did 
not  move  in  the  fished  zone.  From  animals  tagged  and 
released  inside  the  sanctuary,  only  the  number  that 
moved  and  were  recovered  (A^;/?)  is  available  as  an  un- 
biased measure.  In  addition,  we  know  the  total  number 
of  animals  originally  tagged  and  released  in  the  fished 
zone  and  sanctuary,  Nj.  and  N^.  Input  quantities  from 
the  tag-recovery  data  set  will  henceforth  be  indicated  by 
a  tilde  ("):  (iV® R  iVf.  N$MiRj N?  R_ N% }  (Table  1). 


Assumptions 

Three  assumptions  were  used  to  derive  an  emigration- 
rate  estimate:  1)  The  two  ways  to  define  an  estimate  for 
the  proportion  that  moved  within  the  fished  zone,  namely 
as  a  proportion  by  using  only  recapture  numbers,  and 
as  a  proportion  over  the  number  originally  tagged,  can 
be  set  equal.  2)  Recapture  probabilities  of  animals  that 
were  tagged  and  released  inside  the  sanctuary  and  that 
moved  are  assumed  to  equal  those  that  were  tagged  and 
released  into  the  fished  zone  and  that  also  moved.  (The 
first  two  assumptions  were  employed  explicitly  in  steps 
2  and  3  below.)  3)  A  third  assumption  is  implicit  in  step 
2,  specifically  in  the  recapture-conditioned  movement 
proportion  in  the  fished  zone  (.PjtfR,  Eq.  2):  recapture 
probabilities  of  animals  tagged  and  released  in  the 
fished  zone  that  moved  and  of  those  that  did  not  move 
are  assumed  to  be  equal.  Assumptions  2  and  3  would 
both  follow  from  assuming  equal  recapture  probabilities 
for  all  lobsters  in  the  fished  zone. 

Emigration  rate:  derivation  of  the  estimate  formula 

In  this  section,  an  emigration-rate  formula  is  derived.  It 
provides  a  closed-form  estimate  of  the  yearly  proportion 
of  lobsters  emigrating  out  of  the  sanctuary. 

The  proportion  of  animals  moving  can  be  estimated 
from  tag-recovery  data  in  two  ways,  namely  as  "tag- 


468 


Fishery  Bulletin  102(3) 


conditioned"  and  "recapture-conditioned"  proportions. 
A  tag-conditioned  movement  proportion  (Eq.  li  is  the 
total  number  of  lobsters  that  moved  (>3  km)  divided  by 
the  number  originally  tagged  and  released.  It  includes, 
in  the  numerator,  all  tagged  animals  that  moved,  both 
those  that  were  recovered,  as  well  as  those  that  were 
not  recovered.  With  a  recapture-conditioned  movement- 
rate  estimate  (e.g.,  Eq.  2),  only  counts  of  recaptured 
lobsters  are  used.  The  estimate  expresses  the  movement 
proportion  as  the  number  of  tagged  animals  that  were 
recaptured  and  that  also  moved  l>3  km)  divided  by  the 
total  number  recaptured.  These  two  definitions  for  the 
movement  proportion  will  be  used  to  derive  an  estima- 
tion formula  in  terms  of  the  five  data  inputs. 

Step  1  The  derivation  begins  by  writing  the  estimate 
for  proportion  of  lobsters  that  moved  (P|f)  in  tag-condi- 
tioned form: 


Ns     +  Ns 

pS  _  JV  MR  T        .U..YW 

N* 


(1) 


This  estimate  of  movement  rate  from  the  sanctuary  is 
based  on  a  tag-conditioned  proportion  because  we  have 
no  observations  of  recaptured  lobsters  from  the  sanctu- 
ary that  did  not  move  (no  unbiased  measure  of  NfjMM) 
which  a  recapture-conditioned  movement  proportion 
would  have  required.  However  we  did  have  information 
about  N§fNR,  the  nonrecovery  of  tagged  animals  that 
emigrate  from  the  sanctuary  into  the  fished  zone.  It  can 
be  estimated  (steps  2  and  3)  with  the  second  assumption 
that  recovery  rate  for  lobsters  moving  from  the  sanctu- 
ary equals  that  of  lobsters  moving  (>3  km)  within  the 
fished  zone. 


Substituting  Equations  2  and  3  into  Equation  4  and  solv- 
ing for  NF,  XR,  the  number  of  lobsters  that  moved  >3  km 
within  the  fished  zone  but  were  not  recovered,  yields 


NF      =  NF 


NF 


NF      +N 


1 


(5) 


Step  3  Assumption  2  permits  the  derivation  of  a  for- 
mula forWS  yR.  We  first  define  the  recovery  proportions 
of  animals  that  moved  within  the  fished  zone  (F)  as 


fF  = 

I  m 


NF 

'^  MM 


Mr        +MF 
1  v  U  .XR       A  *  M  M 


(6) 


and  from  the  sanctuary  (S)  as 


fs 

I  M 


Nt 


iV.U  XR  T        .V.fl 


(7i 


Assumption  2,  that  the  recovery  rate  (necessarily  in  the 
fished  zone)  for  animals  that  were  tagged  and  released  in 
the  sanctuary  and  that  moved  into  the  fished  zone  is  the 
same  as  for  animals  that  were  both  released  and  recap- 
tured after  moving  within  the  fished  zone  becomes 


fF  =  fs 

I  M        I  M  ■ 


(8) 


Substituting  Equations  6  and  7  into  Equation  8  and 
rearranging  terms,  we  have 


iy  M.XR 


Ns    (NF      +  Nf    ) 


N 


n; 


(9) 


Step  2  Under  assumption  1,  the  two  ways  in  which 
movement  proportion  in  the  fished  zone  can  be  defined 
(as  tag-  and  recapture-conditioned  proportions)  are 
equated.  For  fished  zone  releases,  the  recapture-condi- 
tioned Crc')  movement  proportion  is  written 


N* 


NF      +NF 


(2) 


For  the  recapture-conditioned  estimate  formula  (Eq.  2), 
all  three  quantities  on  the  right-hand  side  are  given  as 
data  inputs.  With  only  numbers  of  lobsters  recovered,  the 
formula  is,  in  this  sense,  conditional  on  recapture. 

The  tag-conditioned  I7c'»  proportion  of  lobsters  moving 
>3  km  of  those  released  in  the  fished  zone  is  written 


Af^    +NF 

,/•  ,,    _  lyM.R  TJVM,.Vfl 

NF 


The  first  assumption  is 


Din     _     pf.l 

rM        ~  rM 


(3) 


i4i 


Step  4  Substituting  Equation  5  into  Equation  9  and 
substituting  the  result  into  Equation  1  yields  a  closed- 
form  estimation  formula  for  the  quantity  we  seek,  the 
proportion  moving  from  the  sanctuary  in  one  year: 


Ps=- 


Nf 


■ArS 


/Vs'  ■iNF       +Nh     ) 


(10) 


Numerical  estimator:  double-hypergeometric  likelihood 
method 

A  likelihood  formulation  of  this  estimator  was  also 
constructed.  The  likelihood  function  describing  a  single 
tag-recapture  experiment  is  hypergeometric  (Seber, 
1982;  Rice,  1995)  because  sampling  is  without  replace- 
ment. The  set  of  possible  outcomes  from  each  of  the 
two  tagging  experiments  can  be  formulated  as  a  2x2 
contingency  table  for  the  experimental  populations  of  all 
lobsters  originally  tagged  and  released.  The  two  pairs 
of  outcomes  represented  in  each  contingency  table  are 
"moved"  or  "not  moved"  and  "recovered"  or  "not  recov- 
ered," yielding  the  four  possible  outcomes  from  both  sets 
of  tag  releases  (see  "Notation"  section). 


McGarvey:  Estimating  emigration  rates  from  marine  sanctuaries  using  tag-recovery  data 


469 


In  this  study  the  data  from  two  interacting  tag-recov- 
ery experiments  were  used  to  generate  an  estimate  of 
reserve  emigration  rate,  namely  of  lobsters  tagged  and 
released  into  the  sanctuary  and  into  the  fished  zone. 
Thus,  the  product  of  a  pair  of  linked  hypergeometric 
probability  mass  functions,  each  corresponding  to  a  2- 
way  contingency  table,  is  the  natural  form  of  the  likeli- 
hood function  for  Pfj. 

The  derivation  of  Equation  10  was  made  with  two  as- 
sumptions, namely  Equations  4  and  8.  Incorporated  in 
the  likelihood,  the  two  assumptions  constrain  the  eight 
recapture  numbers  in  the  contingency  tables.  In  the 
likelihood  formulation,  a  third  constraint  was  needed 
which  is  analogous  to  assumption  1  but  which  applies 
to  sanctuary  releases. 

The  derivation  for  constructing  this  likelihood  from 
a  pair  of  linked  hypergeometric  probability  functions 
will  proceed  by  1)  writing  out  the  "raw"  contingency 
tables  in  terms  of  the  eight  recapture  numbers  (TV), 
as  denoted  in  the  "Tag-recovery  data"  and  "Notation" 
sections,  2)  algebraically  re-expressing  the  elements 
of  the  tables  so  that  the  parameter  to  be  estimated  is 
explicit.  3)  imposing  the  three  constraints,  and  4)  writ- 
ing out  the  likelihood,  using  the  hypergeometric  form 
for  contingency  tables. 

For  the  lobsters  tagged  and  released  in  the  sanctuary, 
the  raw  contingency  table  is 


Recovered 

Not  recovered 

Totals 

Moved 

MS 
"  M.R 

MS 
"M.NR 

fjS       +  MS 
"M.R  *  "  M.NR 

Not  moved 

MS 

"nm.r 

MS 

"nm.nr 

NST- 

Totals 

MS         .    MS 

"  M.R  +  "NM.R 

NST- 

ftf. 

For  the  lobsters  tagged 

in  the  fished  zone 

Recovered 

Not  recovered 

Totals 

Moved 

NF 

"  Mil 

NF 
"m.nr 

"M.R  +  "M.NR 

Not  moved 

MF 
"NM.R 

NF 
"nm.nr 

Totals 

^M.R  +  ^NM.R 

NF- 

NF 

The  two  hypergeometric  probability  mass  functions 
(pmfsi  giving  the  model-predicted  proportion  of  lobsters 
that  moved  and  were  recovered,  based  on  the  two  con- 
tingency tables,  are  written  as 


*  Km)- 


N*     +N 

ly  M.R  T  ly  M.NR 

NF 


N* 


-(N      +N       ) 

ylyM.R  TJV  M.NR' 

NF 


\\ 


NF 


(12) 


N      +N 

ly  MR       *     NM.R  ) 


Because  the  goal  is  to  estimate  the  movement  propor- 
tion, Pft  (rather  than  any  specific  value  of  N),  this  pro- 
portion will  need  to  be  made  explicit  in  the  likelihood 
function  as  the  sole  freely  varying  parameter.  Substitut- 
ing from  the  definition  of  Py  (Eq.  1),  we  have 


m"        =  ps  .  AT*  _  MS 

lyM.NR        1M     iyT       lyM.R- 


(13) 


Substituting  for  all  occurrences  of  Nfj  NR,  Equation  11 
becomes 


p«o= 


PmK)(N^(1-P^^ 


Ni 


AT? 


N* 


Ns    +  Ns 

ly  M.R  T  ly  NM.R  ) 


(14) 


Writing  the  full  joint-likelihood  expression  formed  by 
the  product  of  the  two  hypergeometric  pmfs  gives 


L  = 


NF    +NF      V NF  -(NF    +Nr      ) 

ly  M.R  T  ly  M.NR        ly  T        yly  M.ff  T  ly  M.NR  ' 


NF 


NF 


' 

NF 

1 

NF    +NF 

{■ly  M  R       ly  NM.R  ) 

(Pm-Nt) 

(N*-Q.-F*j\ 

{     K.R     J 

Ns 

\          lyNM.R          ) 

(          Nt          } 

k 

K, 

+  NS 

?  T  JY  NM.R  ) 

As  formulated,  the  value  of  Nf!M  R  is  still  undeter- 
mined by  data  or  constraint.  A  third  constraint  is 
therefore  required.  As  with  assumption  1  for  the  fished 
zone  (Eq.  4),  we  apply  the  assumed  equivalence  of  tag- 
and  recapture-conditioned  proportions  to  the  sanctuary 
releases: 


ps.rc     »>-&     /(/Vs       +/Vi     )  =  P  ' 

rM  ly  M.R  '  yly  NM.R  T  ly  M.R'       *M 


-(MS      +MS        )/  fjs 

~Xly  MR  +  IV  M.NR"  lyT- 


P<Nm.r)  = 


MS       .    AfS 
ly  M.R  T  ly  M.NR 

Ns 


Ms  _  i  \r>     +  ms       , 

"  T       W,M.R  ^lyM.NR' 

Ns 


Ms      +MS 
ly  M.R  Tiv  NM.R  I 


In  this  application,  Nf,M  R  is  understood  as  the  number 
of  lobsters  that  would  have  been  taken  if  fishing  had 
1 1 1 1  not  been  excluded  from  the  sanctuary.  Solving  for  N^M  R 

yields  the  third  constraint. 


MS  -IMS        .MS)/iMS  ,JU*  )_MS 

ly  NM.R  ~yly  M.R    lyT  "  WV  M.R^ly  M.NR1      lyM,R' 


470 


Fishery  Bulletin  102(3) 


Table  2 

Intermediate  calculated  quantities  from  the  numerical  estimation.  The  equalities  of  P[jr 
and  2. 

=P£"  and /'(,=/'«  state 

assumptions  1 

Intermediate  quantity 

Variable  name 

Estimate 

The  proportions  of  lobsters  tagged  in  the  fishing  zone  that  moved  (>3  km  I; 
recapture-conditioned  <P[jr '")  or  tag-conditioned  iPy'0) 

pF.ri_pF.ti 
rM     ~rM 

0.243 

Number  of  lobsters  that  moved  but  were  not  recovered  in  the  fishing  zone 

ATF 

"  M.NR 

697.7 

Number  of  lobsters  that  moved  from  the  sanctuary  but  were  not  recovered 

Ns 

iv  M.NR 

227.3 

Number  of  lobsters  that  did  not  move  and  would  have  been  recovered  had 
there  been  equivalent  levels  of  harvesting  in  the  sanctuary 

MS 
JV. VM.fi 

17.7 

Recovery  proportions  (in  the  fishing  zone) — assumed  to  be  equal  for  lobsters 
that  moved  inside  the  fishing  zone  f[:  or  from  the  sanctuary 

fF-fS 

1  M~'M 

0.113 

without  which  this  numerical  estimator  did  not  con- 
verge. 

The  factorial  terms  in  the  binomial  coefficients  of 
Equations  12  and  14  are  defined  only  for  natural  num- 
bers. However,  in  numerical  minimization,  factorials 
must  be  replaced  with  continuously  varying  approxima- 
tions because  the  negative  log-likelihood  objective  func- 
tion is  minimized  by  using  numerical  derivatives.  The 
factorial  z!  was  extended  from  natural  numbers  to  the 
real  line  by  using  the  gamma  function,  T(z+1)  and  by 
using  an  asymptotic  approximation  formula  for  In  T(z) 
(Eq.  6.1.41  in  Abramowitz  and  Stegun,  1965): 


lnr<2)  = 

1 


InO 


1,    „  1 

z  +  -  ln(  2/r  )  + 

2  12z 


1 


1 


691 


360z! 

1 


(15) 


12602s     I68O27     11882!'     360360211      15621 


The  negative  log  likelihood  was  minimized  numeri- 
cally by  using  the  AD  Model  Builder  parameter  estima- 
tion software  (http://otter-rsch.com/admodel.htm). 


Results 

The  closed-form  estimator  for  the  proportion  of  lobsters 
that  moved  from  the  sanctuary  (P|j)  gave  an  estimate  of 
0.6206;  i.e.,  about  62%  of  the  lobsters  tagged  in  Gleesons 
Sanctuary  moved  out  in  one  year.  The  estimate  obtained 
numerically,  by  maximizing  the  double-hypergeometric 
likelihood,  yielded  a  value  of  0.6212. 

The  small  difference  between  the  analytic  and  nu- 
merical estimates  (0.09%)  is  presumably  due  to  the 
use  of  the  numerical  approximation  for  the  log-gamma 
function  by  the  expansion  of  Equation  15.  The  close 
agreement  suggests  that  the  error  introduced  by  that 
approximation  is  small. 

The  AD  Model  Builder  parameter  estimation  soft- 
ware allows  one  to  estimate  confidence  intervals  of  the 
movement-rate  estimate  in  two  ways:  asymptotically. 
as  diagonal  elements  of  the  covariance  matrix,  and  by 


6  - 

/"\ 

5  - 

/         \ 

4- 

/                    \ 

3  - 

/                                 \ 

2  - 

/                                      \ 

1  - 

J                         \ 

0- 

^^r^ 

0.4  0  5  0.6  0.7 

Estimate  values  for  Pf . 


0.8 


Figure  3 

Profile  likelihood  (solid  line)  and  asymptotic  normal 
approximation  (dashed  line)  for  the  likelihood  confidence 
range  about  the  estimate  of  PM. 


using  a  profile  likelihood.  Confidence  intervals  for  the 
emigration  rate  estimate  were  thus  obtained  numerical- 
ly from  the  hypergeometric  likelihood  by  using  both  the 
asymptotic  normal  approximation  and  an  exact  profile 
likelihood.  These  gave  95%  errors  of  21.2%  and  21.5% 
of  the  estimate,  respectively.  The  approximate  normal 
probability  density  function  and  the  profile  likelihood 
probability  density  function  were  also  plotted  (Fig.  3), 
yielding  close  agreement.  Asymptotic  confidence  inter- 
vals therefore  appear  satisfactory  for  emigration  propor- 
tion estimates  not  lying  near  the  bounds  of  0  and  1. 

Intermediate  calculation  results  (Table  2)  included 
the  recovery  rate  and  movement  rate  (>3  km)  within 
the  fished  zone. 

When  independent  estimates  of  exploitation  rate  are 
available,  typically  from  stock  assessment,  the  rate  of 
tag  reporting  can  be  calculated  from  the  tag-estimated 
recovery  rate.  The  exploitation  rate  (yearly  proportion  of 
legal-size  lobsters  harvested)  for  the  recapture  year  and 


McGarvey:  Estimating  emigration  rates  from  marine  sanctuaries  using  tag-recovery  data  471 


location  of  the  present  study  (the  1995  northern  zone 
rock  lobster  season)  was  estimated  to  be  26%  (Ward  et 
al.3)  by  using  total  yearly  effort  and  catches  by  weight 
and  number  and  a  vector  of  weights  at  age.  The  tag-re- 
covery rate  of  11.3%  (Table  2)  is  the  estimated  propor- 
tion of  tagged  lobsters  that  were  captured  and  for  which 
tags  were  reported.  Thus  the  estimated  tag-reporting 
rate  (of  those  recaptured)  is  0.113/0.26  =  43%.  If  tag 
shedding  and  natural  mortality  were  also  incorporated 
as  additional  causes  for  nonrecovery,  the  estimate  would 
fall  in  the  neighborhood  of  a  50%  tag-reporting  rate. 
This  estimated  level  of  tag-reporting  falls  within  the 
range  considered  probable  by  fishermen.  Thus,  the  re- 
covery-rate estimate  falls  within  a  plausible  range  of 
values,  adding  confidence  that  the  tag-recovery  data 
are  consistent  with  external  estimates  of  exploitation 
rate. 

Substantial  movement  of  Jasus  edwardsii  out  of  a  ma- 
rine sanctuary  was  previously  observed  in  New  Zealand 
(Kelly  and  MacDiarmid,  2003)  but  not  in  Tasmania 
(Gardner  and  Ziegler4).  Long-distance  movement  of 
this  genus  was  also  observed  in  New  Zealand  (Booth, 
1997)  but  was  much  less  common  in  Tasmanian  Jasus 
edwardsii  populations  (Gardner  et  al.,  in  press). 


Discussion 

The  emigration-rate  derivation  above  combined  recap- 
ture-and  tag-conditioned  movement  proportions.  Both 
ways  to  define  a  movement  rate  were  used  to  constrain 
the  range  of  solutions  for  both  analytic  and  numerical 
estimators.  Equating  these  two  definitions  for  movement 
proportion  reduced  the  degrees  of  freedom  by  1,  thereby 
circumventing  the  absence  of  a  count  of  recaptured  lob- 
sters from  within  the  fished  zone. 

Previous  estimators  of  movement  rates  among  spatial 
cells  from  tag-recovery  data  have  used  either  tag-  or 
recapture-conditioned  approaches.  Hilborn  (1990:  see 
also  Quinn  and  Deriso,  1999)  developed  a  tag-condi- 
tioned movement-rate  estimator.  This  estimator  gen- 
erally requires  prior  knowledge  of  the  tag  reporting 
rate.  Schwarz  et  al.  (1993)  employed  data  consisting 
of  simultaneous  tag  releases  and  recaptures  repeated 
over  a  number  of  years  at  the  same  time  each  year  to 
estimate  movement,  survival,  and  recovery  rates  in 
each  spatial  stratum.  Schwarz  et  al.  (1993)  presented 
a  general  formulation  for  modeling  this  multiple  yearly 


3  Ward,  T.  M.,  R.  McGarvey.  Y.  Xiao,  and  D.  J.  Brock. 
2002.  Northern  zone  rock  lobster  [Jasus  edwardsii)  fish- 
ery. South  Australian  Fisheries  Assessment  Series  Report 
2002/04b.  109  p.  Aquatic  Sciences,  South  Australian 
Research  and  Development  Institute  (SARDI):  RO.  Box  120, 
Henley  Beach,  South  Australia  5022.  Australia. 

4  Gardner,  C,  and  P.  Ziegler.  2001.  Are  catches  of  the  south- 
ern rock  lobster  Jasus  edwardsii  a  true  reflection  of  their 
abundance  underwater?  Tasmanian  Aquaculture  and  Fish- 
eries Institute  Final  Report.  TAFI  (Tasmanian  Aquaculture 
and  Fisheries  Institute),  University  of  Tasmania,  Private 
Bag  49,  Hobart  TAS  7001,  Australia. 


tag-recovery  data  set,  extending  a  series  of  estimators 
for  movement  and  survival  (Arnason,  1972,  1973),  and 
estimated  the  rate  of  tag  recovery.  Brownie  et  al.  (1993) 
generalized  the  estimator  of  Schwarz  et  al.  to  non-Mar- 
kovian  movement  rates.  McGarvey  and  Feenstra  (2002), 
following  Hilborn,  used  the  less  costly  and  more  com- 
monly available  single  tag-recovery  data  employed  in 
the  present  study  but  adopted  a  recapture-conditioned 
approach  for  estimating  yearly  movement  rates.  With 
"numbers  recaptured"  appearing  in  both  the  numerator 
and  denominator,  all  nonspatially  dependent  sources  of 
variation  (such  as  tag  reporting  and  shedding,  short- 
and  long-term  tag-induced  mortality,  and  natural  mor- 
tality) cancel  from  the  predicted  recapture-conditioned 
likelihood  proportions.  This  procedure  permits  a  cor- 
responding reduction  in  the  prior  information  required 
to  obtain  unbiased  movement  estimates. 

When  recapture  times  vary,  movement  estimation  is 
sensitive  to  spatial  differences  in  mortality  rate,  no- 
tably between  tag  and  recapture  cells.  Assuming  that 
the  nonreporting  rate  is  unknown,  mortality  can  be 
inferred  from  single  tag-release  information  only  impre- 
cisely, for  example  by  using  mean  tagged  time  at  large. 
For  this  reason  externally  obtained  mortality  estimates, 
typically  from  stock-assessment  models  using  fishery 
data,  can  be  usefully  combined  with  single  tag  recover- 
ies in  movement  estimation.  Hestbeck  (1995)  showed, 
when  survival  differs  by  cell,  that  ignoring  the  time  of 
movement  between  yearly  samples  could  bias  movement 
estimates.  McGarvey  and  Feenstra  (2002)  made  explicit 
the  variation  in  residence  time  and  thus  survival  in 
source  (tag-release)  and  destination  (recapture)  cells  for 
each  recaptured  animal.  By  using  prior  knowledge  of  a 
migration  season,  migration  source  cell  and  destination 
cell  residence  times  can  be  approximated  as  the  time 
from  the  date  of  tag  release  to  an  assumed  fixed  (yearly) 
date  of  movement,  and  from  that  date  to  the  date  of  re- 
capture. These  residence  times  are  used  in  exponential 
survival  factors  that  differ  spatially  given  externally- 
estimated  fishing  mortality  rates  in  each  cell. 

For  the  data  set  available  from  Gleesons  Landing,  all 
tagged  animals  were  released  during  the  peak  fishing 
season  (mid-summer).  Thus  recoveries  from  the  fol- 
lowing fishing  season  had  a  mean  and  mode  near  the 
desired  one-year-at-large.  In  future  tag-recovery  stud- 
ies, where  a  yearly  movement  rate  is  sought,  a  similar 
choice  for  timing  of  tag  releases,  namely  during  the 
season  of  highest  fishery  catches,  should  yield  a  peak  in 
recaptures  a  year  later.  Schwarz  et  al.  (1993)  employed 
this  strategy  with  their  multiple  yearly  tag-recovery 
data  sets. 

In  the  estimator  presented  above,  variations  in  ex- 
pected recovery  numbers  versus  time,  notably  due  to  sur- 
vival, were  neglected.  The  small  sample  (33  recoveries 
between  0.5  and  1.5  years  from  the  sanctuary)  and  lack 
of  recaptures  from  within  the  sanctuary  necessitated 
more  modest  estimation  goals.  Among  data  classes  avail- 
able for  movement  analysis,  notably  1)  multiple  yearly 
tag  recaptures  by  researchers  in  all  cells,  2)  multiple 
yearly  tag  recoveries  where  recapture  is  by  fishermen  (or 


472 


Fishery  Bulletin  102(3) 


hunters)  in  all  cells,  3)  single  tag  recoveries  by  fishermen 
in  all  cells,  and  4)  the  data  set  employed  in  the  present 
study  of  single-tag  recoveries  by  fishermen  in  one  of  two 
cells,  the  latter  represents  the  low  end  in  quality  and 
quantity  of  information  about  movement  and  survival. 

A  time-dependent  approach  could  theoretically  extend 
the  approach  of  McGarvey  and  Feenstra  (2002)  to  make 
explicit  the  residence  times  of  each  recaptured  indi- 
vidual in  the  fishing  zone  and  sanctuary,  respectively, 
and  thus  make  explicit  differences  in  the  predicted 
survival  rate  before  and  after  movement.  However  with- 
out prior  knowledge  of  when  movement  took  place  for 
each  recaptured  lobster,  a  modified  likelihood  method 
is  called  for,  requiring  integration  over  the  probable 
movement  times  between  tag  release  and  recapture. 
This  extension  of  residence-time-dependent  movement 
estimators  to  variable  times  of  movement  remains  a 
topic  for  future  research. 


Acknowledgments 

I  thank  Hugh  Possingham,  Andre  Punt,  and  two  review- 
ers for  comments  on  the  draft  manuscript.  Lobsters 
were  tagged  and  released  into  Gleesons  Landing  sanctu- 
ary by  Greg  Ferguson,  together  with  fishermen  Lenny 
and  Murray  Williams,  under  Fisheries  Research  and 
Development  Corporation  Project  93/086.  This  work  was 
supported  by  the  Australian  Fisheries  Research  and 
Development  Corporation  Project  No.  2000/195,  and  by 
the  South  Australian  rock  lobster  industry. 


Literature  cited 

Abramowitz,  M..  and  I.  R.  Stegun. 

1965.     Handbook  of  mathematical  functions,  p.  257.     Dover 
Publications,  New  York,  NY. 
Arnason,  A.  N. 

1972.  Parameter  estimates  from  mark-recapture  exper- 
iments on  two  populations  subject  to  migration  and 
death.     Res.  Popul.  Ecol.  (Kyoto)  13:97-113. 

1973.  The  estimation  of  population  size,  migration  rates, 
and  survival  in  a  stratified  population.  Res.  Popul. 
Ecol.  (Kyoto)  15:1-8. 


Booth,  J.  D. 

1997.     Long-distance  movements  in  Jasus  spp.  and  their 
role  in  larval  recruitment.     B.  Mar.  Sci.  61ill:  1 11— 
128. 
Brownie,  C,  J.  E.  Hines.  J.  D.  Nichols,  K.  H.  Pollock,  and 
J.  B.  Hestbeck. 

1993.     Capture-recapture  studies  for  multiple  strata  in- 
cluding non-Markovian  transitions.     Biometrics  49:1173- 
1187. 
Gardner,  C,  S.  Frusher,  M.  Haddon,  and  C.  Buxton. 

2003.     Movements  of  the  southern  rock  lobster  Jasus 
edwardsii  in  Tasmania,  Australia.     Bull.  Mar.  Sci.  73(3): 
653-671. 
Hestbeck,  J.  B. 

1995.     Bias  in  transition-specific  survival  and  movement 
probabilities  estimated  using  capture-recapture  data.     J. 
Appl.  Stat.  22:737-750. 
Hilborn,  R. 

1990.  Determination  offish  movement  patterns  from  tag 
recoveries  using  maximum  likelihood  estimators.  Can. 
J.  Fish.  Aquat.  Sci.  47:635-643. 

Kelly,  S. 

2001.  Temporal  variation  in  the  movement  of  the 
spiny  lobster  Jasus  edwardsii.  Mar.  Freshw.  Res. 
53(31:323-331. 

Kelly,  S.,  and  A.  B.  MacDiarmid. 

2003.     Movement  patterns  of  mature  spiny  lobsters,  Jasus 

edwardsii,  from  a  marine  reserve.     N.Z.  J.  Mar.  Freshw. 

Res.  37:149-158. 
MacDiarmid,  A.  B..  B.  Hickey,  and  R.  A.  Mailer. 

1991.  Daily  movement  patterns  of  the  spiny  lobster  Jasus 
edwardsii  (Hutton)  on  a  shallow  reef  in  northern  New 
Zealand.     J.  Exp.  Mar.  Biol.  Ecol.  147:185-205. 

McGarvey,  R.,  and  J.  E.  Feenstra. 

2002.  Estimating  rates  of  fish  movement  from  tag  recov- 
eries: conditioning  by  recapture.  Can.  J.  Fish.  Aquat. 
Sci.  59:1054-1064. 

Quinn,  T  J.  II,  and  R.  B.  Deriso. 

1999.     Quantitative  fish  dynamics,  p.  414-419.     Oxford 
Univ.  Press,  New  York.  NY. 
Rice,  J.  A. 

1995.     Mathematical  statistics  and  data  analysis,  p.  13-14, 
39-40.     Duxbury  Press,  Belmont  CA. 
Schwarz,  C.  J.,  J.  F.  Schweigert,  and  A.  N.  Arnason. 

1993.     Estimating  migration  rates  using  tag  recovery 
data.     Biometrics  49:177-193. 
Seber,  G.  A.  F. 

1982.  The  estimation  of  animal  abundance  and  related 
parameters,  2nd  ed.,  p.  59-60.     Griffin,  London. 


473 


Abstract— We  describe  reproduc- 
tive dynamics  of  female  spotted  sea- 
trout  (Cynoscion  nebulosus)  in  South 
Carolina  (SC).  Batch  fecundity  iBFl. 
spawning  frequency  (SF),  relative 
fecundity  (RF),  and  annual  fecundity 
(AF>  for  age  classes  1-3  were  esti- 
mated during  the  spawning  seasons  of 
1998,  1999,  and  2000.  Based  on  histo- 
logical evidence,  spawning  of  spotted 
seatrout  in  SC  was  determined  to  take 
place  from  late  April  through  early 
September.  Size  at  first  maturity  was 
248  mm  total  length  (TL);  50%  and 
100%  maturity  occurred  at  268  mm 
and  301  mm  TL,  respectively.  Batch 
fecundity  estimates  from  counts  of 
oocytes  in  final  maturation  varied 
significantly  among  year  classes.  One- 
year-old  spotted  seatrout  spawned  an 
average  of  145.452  oocytes  per  batch, 
whereas  fish  aged  2  and  3  had  a  mean 
BF  of  291,123  and  529,976  oocytes, 
respectively.  We  determined  monthly 
SF  from  the  inverse  of  the  proportion 
of  ovaries  with  postovulatory  follicles 
(POF)  less  than  24  hours  old  among 
mature  and  developing  females.  Over- 
all, spotted  seatrout  spawned  every 
4.4  days,  an  average  of  28  times 
during  the  season.  A  chronology  of 
POF  atresia  for  water  temperature 
>25°C  is  presented.  Length,  weight 
(ovary-free),  and  age  explained  67%, 
65%,  and  58%  of  the  variability  in 
BF,  respectively.  Neither  RF  (number 
of  oocytes/g  ovary-free  weight)  nor 
oocyte  diameter  varied  significantly 
with  age.  However.  RF  was  signifi- 
cantly greater  and  oocyte  diameter 
was  smaller  at  the  end  of  the  spawn- 
ing season.  Annual  fecundity  esti- 
mates were  approximately  3.2,  9.5, 
and  17.6  million  oocytes  for  each  age 
class,  respectively.  Spotted  seatrout 
ages  1-3  contributed  an  average  of 
29%,  39%,  and  21%  to  the  overall 
reproductive  effort  according  to  the 
relative  abundance  of  each  age  class. 
Ages  4  and  5  contributed  7%  and  4%, 
respectively,  according  to  predicted 
AF  values. 


Reproductive  dynamics  of  female  spotted  seatrout 
(Cynoscion  nebulosus)  in  South  Carolina* 


William  A.  Roumillat 

Marine  Resources  Research  Institute 

South  Carolina  Department  of  Natural  Resources 

217  Ft.  Johnson  Rd 

Charleston,  South  Carolina  29412 

E-mail  address  roumillatbiSrnrd.dnr.state.sc.us 

Myra  C.  Brouwer 

South  Atlantic  Fishery  Management  Council 
One  Southpark  Center,  suite  306 
Charleston,  South  Carolina  29407 


Manuscript  submitted  13  May  2002 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
19  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:473-487  (2004). 


The  spotted  seatrout  iCynoscion  neb- 
ulosus) is  an  estuarine-dependent 
member  of  the  family  Sciaenidae.  Spot- 
ted seatrout  are  year-round  residents 
of  estuaries  along  the  South  Atlantic 
coast  and  spawning  takes  place  inshore 
and  in  coastal  areas  (McMichael  and 
Peters,  1989;  Mercer1;  Luczkovich 
et.  al.2).  As  in  many  other  sciaenids, 
spawning  in  this  species  occurs  in 
the  evening  (Holt  et  al,  1985).  Male 
spotted  seatrout  have  the  capacity  to 
produce  "drumming"  sounds  that  are 
caused  by  the  contraction  of  the  swim- 
bladder  by  specialized  muscles  that 
are  seasonally  hypertrophied  from  the 
abdominal  hypaxialis  muscle  mass 
(Fish  and  Mowbray,  1970;  Mok  and 
Gilmore.  1983).  Direct  involvement  of 
sound  production  with  spawning  has 
been  shown  for  this  and  other  sciae- 
nids (Mok  and  Gilmore,  1983;  Saucier 
et  al.,  1992;  Saucier  and  Baltz,  1993; 
Luczkovich  et  al.2). 

We  have  collected  information  on 
the  spawning  behavior  of  spotted  sea- 
trout in  coastal  South  Carolina  since 
1990  (Saucier  et  al.,  1992;  Riekerk 
et  al.3).  Spawning  aggregations  were 
located  by  listening  for  drumming 
sounds  from  late  afternoon  until 
-2300  h  with  passive  hydrophone 
equipment.  Spawning  activity  was 
subsequently  verified  through  collec- 
tions of  newly  spawned  eggs  and  by 
the  rearing  of  the  larvae  in  the  labo- 
ratory (Saucier  et  al.,  1992). 

Spotted  seatrout  are  group-synchro- 
nous spawners  with  indeterminate  fe- 


cundity and  the  protracted  spawning 
season  extends  from  April  through 
September  along  the  South  Atlantic 
and  Gulf  of  Mexico  coasts  (Overstreet, 
1983;  Brown-Peterson  et  al..  1988; 
McMichael  and  Peters,  1989;  Saucier 
and  Baltz,  1993;  Brown-Peterson  and 
Warren,  2001;  Brown-  Peterson  et  al., 
2002;  Nieland  et  al..  2002,  Brown- 


*  Contribution  539  from  the  Marine 
Resources  Research  Institute  of  the 
South  Carolina  Department  of  Natu- 
ral Resources,  Charleston.  SC  29422- 
2559. 

1  Mercer.  L.  P.  1984.  A  biological  and 
fisheries  profile  of  spotted  seatrout. 
Cynoscion  nebulosus.  Special  Scien- 
tific Report  40,  87  p.  North  Carolina 
Department  of  Natural  Resources  and 
Community  Development,  Division  of 
Marine  Fisheries.  Morehead  City,  NC 
28577. 

2  Luczkovich,  J.  J..  H.  J.  Daniel  III  and  M. 
W.  Sprague.  1999.  Characterization  of 
critical  spawning  habitats  of  weakfish. 
spotted  seatrout  and  red  drum  in  Pamlico 
Sound  using  hydrophone  surveys.  Final 
report  and  annual  performance  report  F- 
62-2  and  F-62-2,  p  65-68.  North  Caro- 
lina Department  of  Environment  and 
Natural  Resources,  Division  of  Marine 
Fisheries,  Morehead  City,  NC  28557. 

3  Riekerk.  G.  H.  M„  S.  J.  Tyree,  and  W. 
A.  Roumillat.  1997.  Spawning  times 
and  locations  of  spotted  seatrout  in  the 
Charleston  Harbor  Estuarine  System 
from  acoustic  surveys.  21  p.  Final 
Report  to  Charleston  Harbor  Project, 
Bureau  of  Ocean  and  Coastal  Resources 
Management,  South  Carolina  Depart- 
ment of  Health  and  Environmental  Con- 
trol, 1362  McMillan  Ave.,  Charleston, 
SC  29405. 


474 


Fishery  Bulletin  102(3) 


Peterson,  2003;  Wenner  et  al.4).  As  in  other  indetermi- 
nate spawning  fish,  annual  fecundity  in  this  species  is 
determined  by  the  number  of  oocytes  released  during 
each  spawning  event  (batch  fecundity)  and  the  number 
of  spawning  events  occurring  during  the  course  of  the 
spawning  season  (spawning  frequency!.  Early  efforts  to 
estimate  fecundity  for  spotted  seatrout  did  not  take  into 
account  the  repetitive  nature  of  spawning  activities  in 
this  species  (Pearson,  1929;  Sundararaj  and  Suttkus, 
1962;  Overstreet,  1983)  and  only  recently  has  an  effort 
been  made  to  coordinate  batch  fecundities  with  spawn- 
ing frequencies  (Brown-Peterson  et  al.,  1988;  Brown- 
Peterson  and  Warren,  2001;  Nieland  et  al.,  2002).  This 
procedure  is  intuitively  necessary  to  estimate  the  re- 
productive output  for  an  entire  spawning  season  and 
is  made  even  more  useful  for  fisheries  management  if 
separated  by  size  class  or  age  cohort  within  a  popula- 
tion (Prager  et  al..  1987;  Goodyear,  1993;  Zhao  and 
Wenner5). 

An  important  component  of  assessment  for  manage- 
ment involves  determining  the  spawning  potential  ratio 
(SPR),  a  measure  of  the  effect  of  fishing  on  the  repro- 
ductive potential  of  a  stock  (Goodyear,  1993).  This  value 
is  usually  calculated  as  the  ratio  of  spawning  stock 
biomass  per  recruit  (SSBR)  in  the  presence  of  fishing 
mortality  (F)  to  the  SSBR  when  F  is  equal  to  zero  (Ga- 
briel et  al.,  1989;  Goodyear,  1993).  Spawning  potential 
ratio  is  currently  used  as  a  biological  reference  point 
for  definition  of  recruitment  overfishing  (i.e.,  Vaughan 
et  al.,  1992).  The  calculation  of  SPR  can  be  improved, 
however,  by  introducing  egg  production  into  the  model. 
Fecundity  is  a  much  better  predictor  of  reproductive  po- 
tential than  female  biomass.  Moreover,  SPR  calculations 
based  on  egg  production  may  be  more  sensitive  to  the 
size-age  composition  of  the  spawning  stock.  However, 
accurate  annual  fecundity  estimates  for  use  in  stock 
assessment  do  not  exist  for  this  or  many  other  species 
in  need  of  fisheries  management.  Therefore,  our  goal 
was  to  obtain  batch  fecundity  (BF),  spawning  frequency 
(SF),  and  annual  fecundity  (AF)  estimates  for  spotted 
seatrout  by  age  class. 


Materials  and  methods 

Data  to  address  the  main  objectives  of  this  study  were 
collected  from  late  April  through  early  September  1998- 


1  Wenner,  C.  A.,  W.  A.  Roumillat.  J.  E.  Moran  Jr.,  M.  B.  Maddox, 
L.  B.  Daniel  III,  and  J.  W.  Smith.  1990.  Investigations 
on  the  life  history  and  population  dynamics  of  marine  rec- 
reational lishes  in  South  Carolina:  part  1.  Final  Report 
F-37,  177  p.  Marine  Resources  Research  Institute,  Marine 
Resources  Division,  South  Carolina  Department  of  Natural 
Resources,  217  Ft.  Johnson  Rd.,  Charleston,  SC  29412 

s  Zhao,  B.,  and  C.  A.  Wenner.  1995.  Stock  assessment  and 
fishery  management  of  the  spotted  seatrout,  Cynoscion  nebu- 
losus,  on  the  South  Carolina  coast,  90  p.  Marine  Resources 
K'  i  arch  Institute,  Marine  Resources  Division,  South  Caro- 
lina Department  of  Natural  Resources,  217  Ft.  Johnson  Rd.. 
Charleston,  SC  29412. 


2000  as  part  of  a  long  term  monitoring  effort  11991-pres- 
ent)  to  assess  the  relative  abundance  of  age  classes  of 
recreationally  important  finfish  in  South  Carolina  estu- 
aries. The  study  followed  a  monthly  stratified  random 
sampling  design  in  three  estuarine  systems.  The  Cape 
Romain  system  comprised  two  strata;  Romain  Harbor 
and  northern  Bulls  Bay.  The  Charleston  Harbor  system 
contained  four  strata:  the  Wando,  Cooper,  and  Ashley 
Rivers,  and  Charleston  Harbor.  The  Ashepoo-Combahee- 
Edisto  (ACE)  Basin  system  comprised  a  single  stratum 
(Fig.  1).  The  number  of  sampling  sites  within  each  stra- 
tum ranged  from  23  to  30.  A  subset  of  12-14  sites  was 
randomly  selected  each  month.  Sampling  was  conducted 
only  during  the  daytime  ebbing  tide  (0700-1800  h), 
primarily  over  mud  and  oyster  shell  substrates  adjacent 
to  the  Spartina  alterniflora  marsh.  At  each  site,  we 
deployed  a  trammel  net  (182.8  m  long  by  2.4  m  deep; 
outer  walls:  17.8  cm  square  (35.6  cm  stretch];  inner  wall: 
3.2  cm  square  [6.4  cm  stretch])  from  a  rapidly  moving 
shallow  water  boat  in  an  arc  against  the  shoreline  at 
depths  ranging  from  0.5  to  2.0  m.  We  disturbed  the 
water  within  the  site  in  an  effort  to  frighten  fishes  into 
the  entrapment  gear.  We  then  hauled  the  trammel  net 
back  into  the  boat  and  removed  the  catch,  which  was 
kept  alive  in  a  70-liter  oxygenated  holding  tank.  Spot- 
ted seatrout  were  measured  for  total  length  (TL)  and 
standard  length  (SL)  and  a  subsample  offish  from  each 
effort  (5-10  individuals  for  each  20-mm  size  interval  per 
month)  were  sacrificed,  placed  on  ice,  and  transported  to 
the  laboratory  for  aging  and  reproductive  data. 

Specimens  were  processed  in  the  laboratory  2-12 
hours  after  capture.  We  recorded  standard  life-history 
parameters  (TL,  SL,  fish  weight,  gonad  weight,  sex,  and 
maturity)  for  each  specimen.  The  following  equation  was 
used  to  convert  lengths  when  necessary: 


TL  =  5.689  +  1.167ISL)  (r-  =  0.998) 


n  =  1191. 


We  removed  sagittal  otoliths  for  aging  and  preserved 
sections  (<2%  by  weight)  of  each  ovary  in  neutral  buff- 
ered formalin  for  histological  processing.  The  latter 
involved  standard  procedures  for  paraffin  embedding 
and  sectioning,  and  standard  hematoxylin  and  eosin-y 
staining  (Humason,  1972).  Histological  sections  were 
viewed  under  a  Nikon  Labophot  compound  microscope 
equipped  with  a  teaching  head  so  that  two  readers 
could  interpret  sections  simultaneously.  Maturity 
estimation  was  modified  from  that  of  Wenner  et  al.4 
(Table  1). 

Size  at  first  maturity  was  histologically  derived  by 
first  evidence  of  cortical  alveoli  stage  oocytes.  To  ar- 
rive at  estimates  of  50%  and  100%  maturity,  data  were 
subjected  to  PROBIT  analysis. 

Age  determination 

The  left  sagittae  were  marked  with  a  soft  lead  pencil 
through  the  core  and  embedded  in  epoxide  resin.  A 
transverse  section  (~0.5-mm  thick)  was  taken  through 
the  core  by  using  a  low-speed  saw  equipped  with  a  pair 


Roumillat  and  Brouwer:  Reproductive  dynamics  of  Cynoscion  nebulosus 


475 


80  30W  80  O'W  79  30'W 

Figure  1 

The  three  South  Carolina  estuarine  systems  (indicated  by  arrows)  where  C.  nebulosus  were  collected. 


Table  1 

Criteria  used  for  microscopic  staging  of  C.  nebulosus  ovaries.  FOM  =  final  oocyte  maturation;  POF  =  postovulatory  follicle. 


Stage 


Description 


Immature  Ovary  small  in  cross  section.  Early  stage  with  only  oogonia  evident;  later  stage  with  small 

(<0.08  mm)  primary  oocytes,  tightly  packed.  No  evidence  of  early  vitellogenesis. 

Developing  First  appearance  of  cortical  alveoli  stage  oocytes  through  late  vitellogenesis  but  no  evidence  of  early 

FOM  (lipid  and  yolk  globule  coalescence). 

Ripe  Ovary  containing  oocytes  demonstrating  FOM  (lipid  and  yolk  coalescence  through  hydration). 

Mature  with  day-0  POFs        Ovary  exhibits  POFs  <24  h  (see  Table  2 ).  Found  at  all  water  temperatures  throughout  the  spawning 

season. 
Mature  with  day-1  POFs        Ovary  exhibits  POFs  >24  h  consisting  of  closely  packed  granulosa  cells  (0.08-0.1  mml.  Only 

identified  in  water  <  25°C. 

Spent  Ovary  containing  alpha-  and  beta-stage  oocytic  atresia. 

Resting  Ovary  containing  small  primary  oocytes  and  oocytes  with  perinuclear  nucleoli  (<0.12  mm); 

usually  some  remnants  of  oocytic  atresia. 


of  diamond  wafering  blades.  The  resulting  section  was  was  assumed  for  the  process  of  aging.  Spotted  seatrout 

mounted  on  a  labeled  microscope  slide  and  examined  deposit  an  annulus  in  April  or  May  (Murphy  and  Taylor, 

with  a  Nikon  SMZ-U  microscope.  A  1  January  birth  date  1994;  Wenner  et  al.4).  Fish  in  the  first  three  months  of 


476 


Fishery  Bulletin  102(3) 


Table  2 

Criteria  used  for  microscopic  staging  of  C.  nebulosus  day-0  postovulatory  follicles  iPOFsl  in  water  temperatures  above  25°C. 
Measurements  represent  longest  axis  of  POFs. 


POF  chronology  fin  hours) 


Description 


0-4 

5-8 

9-12 

13-24 


Regular  arrangement  of  granulosa-cell  nuclei  proximal  to  the  basement  membrane  and 
obvious  multiple  layering  as  described  by  Hunter  and  Macewicz  1 1985 ).  200-300  ^m  i  Fig.  4Ai. 

Early  signs  of  atresia,  loss  of  the  obvious  layering,  hypertrophy  of  granulosa  cells,  and  a 
general  compaction  with  an  investment  of  blood  vessels.  180-250  jum  (Fig.  4B). 

Well-defined  lumen  separating  the  internal  granulosa  cells  from  the  outer  wall  of  granulosa 
cells  encompassed  by  theca.  150-200  jum  (Fig.  4C). 

Lumen  reduced  primarily  by  loss  of  granulosa  tissue  and  proximity  of  peripheral  layers.  130- 
175  ^m  (Fig.  4,  DandE). 


the  year  were  aged  by  the  addition  of  1  year  to  the  count 
of  the  number  of  annuli  on  the  thin  sections.  In  April  or 
May,  if  the  section  had  a  large  marginal  increment,  one 
was  also  added  to  the  annular  count.  If  the  marginal 
increment  was  small  or  if  the  ring  was  detectable  on  the 
edge  of  the  otolith  section,  age  was  equal  to  the  number 
of  annuli. 

Seasonality 

Spawning  season  for  spotted  seatrout  in  South  Carolina 
was  determined  by  using  two  techniques.  The  gonadoso- 
matic  index  (GSI)  was  calculated  as 

(GW/OFWT)  x  100, 

where  GW  =  gonad  weight  (g);  and 
OFWT  =  ovary-free  weight  (g). 

For  years  prior  to  this  study  (1991-97),  mean  monthly 
GSI  was  obtained  for  all  females  by  using  data  from 
the  South  Caroline  Department  of  Natural  Resources 
inshore  fisheries  archives  (Wenner6).  Reproductive  sea- 
sonality among  female  spotted  seatrout  throughout  the 
year  was  also  examined  by  using  histology  (Table  1). 

The  first  evidence  of  oocytes  in  final  oocyte  matura- 
tion (FOM)  as  evidenced  by  lipid  and  yolk  coalescence; 
Brown-Peterson  et  al.,  1988)  or  the  occurrence  of  post- 
ovulatory follicles  (POFs)  defined  the  beginning  of  the 
spawning  season.  To  determine  the  cessation  of  spawn- 
ing, the  percent  occurrence  of  females  in  spawning 
condition  (ripe  and  repeat  spawners)  and  those  in  post- 
spawning  condition  (spent  and  resting)  were  obtained 
for  the  months  of  August  and  September.  To  investigate 
the  condition  of  females,  we  examined  Fulton's  condition 
factor  (Ricker,  1975)  over  the  spawning  season  using 
linear  regression. 


6  Wenner,  C.  2002.  Unpubl.  data.  Marine  Resources  Re- 
search Institute,  Marine  Resources  Division,  South  Carolina 
Department  of  Natural  Resources,  217  Ft.  Johnson  Rd., 
i  harleston,  SC  29412. 


Spawning  frequency 

We  obtained  samples  for  spawning  frequency  (  SFi  deter- 
mination from  1  May  through  31  August  1998,  1999 
and  2000.  Although  samples  were  routinely  collected 
throughout  the  year,  only  from  early  May  through  late 
August  did  we  capture  enough  animals  in  the  appro- 
priate reproductive  state  for  SF  estimation.  Spawning 
frequency  was  calculated  as  either  the  inverse  of  the 
proportion  of  ovaries  with  day-0  POFs  (Hunter  and 
Macewicz,  1985;  Brown-Peterson  et  al.,  1988)  or  with 
oocytes  in  FOM  (Brown-Peterson  et  al.,  1988;  Liso- 
venko  and  Adrianov,  1991)  among  mature  and  develop- 
ing females. 

We  designated  two  distinct  morphological  features 
of  POFs  based  on  time  of  specimen  capture  and  water 
temperature.  We  interpreted  the  largest,  least  atro- 
phied POFs  to  be  <24  h  old  and  termed  them  "day-0" 
POFs  (Hunter  and  Macewicz,  1985).  The  presence 
of  day-0  POFs  in  the  ovary  indicated  that  spawning 
had  occurred  the  previous  night.  The  second  category 
comprised  smaller  POFs,  which  primarily  consisted  of 
closely  packed  granulosa  cells  determined  to  be  >24  h 
old. 

To  complete  the  chronology  of  POF  atresia  we  under- 
took round-the-clock  sampling  on  27-28  June  and  26 
July  2000.  During  these  efforts,  sampling  continued 
beyond  routine  hours  to  encompass  the  period  between 
dusk  and  dawn.  The  histological  samples  obtained  al- 
lowed for  the  calibration  of  criteria  used  to  age  POFs 
(Table  2).  To  determine  whether  SF  varied  among 
months  and  age  classes,  Kruskal-Wallis  tests  were 
used.  Because  both  factors  (month  and  age)  were  fixed 
(model  1).  it  was  not  possible  to  test  for  their  interac- 
tion by  using  a  two-way  parametric  ANOVA  without 
replication. 

As  a  result  of  targeting  fish  for  batch  fecundity  esti- 
mates (see  below),  we  had  available  numerous  specimens 
with  oocytes  in  FOM  with  which  to  establish  monthly 
SF.  However,  we  knew  that  these  specimens  were  dis- 
appearing from  our  shallow  sampling  sites  into  deeper 
spawning  areas  as  the  day  progressed  (Riekerk  et  al.3), 


Roumillat  and  Brouwer:  Reproductive  dynamics  of  Cynoscion  nebulosus 


All 


thus  potentially  adding  bias  to  our  SF  estimates.  Even 
though  estimates  of  SF  based  on  FOM  were  performed 
a  posteriori,  we  chose  to  report  them  strictly  for  com- 
parison to  other  studies  with  this  method.  Because 
sampling  for  females  exhibiting  FOM  was  accomplished 
in  a  directed  fashion,  statistical  comparisons  were  not 
attempted. 

Batch  fecundity  and  relative  fecundity 

Observations  taken  over  a  decade  of  sampling  the 
Charleston  Harbor  estuarine  system  showed  that  in 
females  captured  in  shallow  water  (<1.5  m)  during  the 
spawning  season,  FOM  began  at  about  1200  h  (Wenner6). 
Similarly,  Crabtree  and  Adams'  reported  FOM  begin- 
ning in  Florida  spotted  seatrout  at  about  mid-day.  Low- 
erre-Barbieri  et  al.8  found  hydrated  females  in  shallow 
water  in  the  vicinity  of  aggregations  of  drumming  males 
in  deeper  water.  We  also  speculated  that  from  mid- 
to  late  afternoon  hydrated  females  moved  along  the 
marsh  edge  toward  deeper  water  spawning  aggrega- 
tions (8-25  m).  Hydrophone  surveys  conducted  in  the 
Charleston  Harbor  area  over  several  years  (Riekerk  et 
al.3;  Wenner6)  indicated  that  noise  production  typically 
began  around  1800  h  and  ceased  around  2200  h.  Because 
this  behavior  has  been  associated  with  spawning  in  this 
and  other  sciaenids  (Mok  and  Gilmore,  1983;  Holt  et  al., 
1985;  Saucier  et  al.,  1992;  Saucier  and  Baltz,  1993),  we 
assumed  that  spawning  began  at  1800  h  and  stopped  at 
2200  h.  Thus  we  were  able  to  target  spotted  seatrout  in 
the  mid-  to  late  afternoon  specifically  to  capture  females 
with  oocytes  in  the  late  stages  of  FOM  for  batch  fecun- 
dity (BF)  estimation.  Because  we  have  consistently  iden- 
tified recently  spawned  females  in  shallow  areas,  they 
apparently  return  to  the  marsh  edge  where  they  once 
again  become  available  for  capture  with  our  sampling 
gear.  Our  stratified  random  sampling  of  estuarine  areas 
along  the  coast  (described  previously)  was  designed  to 
representatively  sample  these  recently  spawned  females 
for  SF  estimation. 

We  conducted  BF  sampling  during  two  consecutive 
afternoons  fortnightly  from  the  middle  of  April  through 
the  first  week  of  September  1998,  1999,  and  2000.  We 
deployed  a  trammel  net  from  a  shallow  water  boat  as 
described  above  at  preselected  sites  in  Charleston  Har- 
bor in  depths  ranging  from  1.0  to  1.5  meters  during  the 
afternoon  (1400-1800  h  EDT)  high  tide. 


7  Crabtree  R.  E.,  and  D.  H.  Adams.  1998.  Spawning  and 
fecundity  of  spotted  seatrout,  Cynoscion  nebulosus,  in  the 
Indian  River  Lagoon.  Florida.  In  Investigations  into  near- 
shore  and  estuarine  gamefish  abundance,  ecology,  and  life 
history  in  Florida,  p.  526-566.  Tech.  Rep.  for  Fed.  Aid  in 
Sport  Fish  Rest.  Act  Project  F-59.  Florida  Marine  Research 
Institute,  Department  of  Environmental  Protection,  100 
Eighth  Ave.  SE,  St.  Petersburg,  FL  33701. 

8  Lowerre-Barbieri,  S.  K.,  L.  R.  Barbieri,  and  J.  J.  Albers. 
1999.  Reproductive  parameters  needed  to  evaluate  recruit- 
ment overfishing  of  spotted  seatrout  in  the  southeastern 
U.S.  Final  report  to  the  Saltonstall-Kennedy  (S-K)  Grant 
Program  (grant  no.  NA77FD0074),  23  p. 


Restricting  our  sampling  to  the  hours  immediately 
preceding  the  evening  spawning  event  ensured  that 
those  females  preparing  to  spawn  were  available  for 
capture.  Male  spotted  seatrout,  identified  by  their 
drumming  sounds,  caught  during  this  targeted  effort 
were  measured  and  released  at  the  site  of  capture.  We 
supplemented  samples  for  BF  estimation  with  specimens 
from  local  sportfishing  tournaments  held  during  sum- 
mer months  in  the  Charleston  Harbor  area. 

We  processed  samples  in  the  laboratory  as  previously 
described.  If  ovaries  appeared  by  macroscopic  examina- 
tion to  contain  hydrated  oocytes,  they  were  fixed  in  10% 
buffered  seawater  formalin  for  potential  counts  (Hunter 
et  al.,  1985).  The  appropriateness  of  these  ovaries  for 
BF  counts  was  subsequently  determined  by  examining 
the  corresponding  histological  preparation. 

To  ensure  that  only  those  oocytes  destined  to  be  ovu- 
lated during  the  upcoming  spawning  event  were  counted, 
we  chose  to  use  only  those  oocytes  undergoing  FOM  that 
could  be  easily  separated  by  size  from  late  vitellogenic 
oocytes  (Nieland  et  al..  2002;  Lowerre-Barbieri  et  al.8). 
If  we  observed  numerous  recent  POFs  in  the  histologi- 
cal sample,  the  corresponding  whole  ovary  was  not  used 
for  oocyte  counts  (because  their  presence  indicated  that 
ovulation  had  occurred).  We  reweighed  ovaries  (approxi- 
mately 2  weeks  after  fixation)  to  the  nearest  0.01  g  and 
randomly  extracted  three  130-150  mg  aliquots  from 
eight  potential  locations  in  the  ovary  (each  lobe  was  par- 
titioned into  quarters  lengthwise).  We  stored  subsamples 
in  50%  isopropyl  and  counted  oocytes  under  a  Nikon 
SMZ-U  dissecting  microscope  at  12  x  magnification.  We 
counted  each  subsample  twice  by  using  a  Bogorov  tray 
and  a  hand-held  counter  and  conducted  a  third  count  if 
the  two  initial  counts  were  dissimilar  by  more  than  10%. 
We  used  the  mean  number  of  oocytes  in  each  subsample 
to  calculate  mean  oocyte  density  (number  of  oocytes  per 
gram  preserved  ovary  weight)  and  total  numbers  of  oo- 
cytes in  the  ovary.  We  compared  mean  oocyte  densities 
among  the  four  regions  of  each  ovarian  lobe  and  between 
the  two  lobes  by  using  a  two-way  analysis  of  variance 
(ANOVA).  Because  our  variances  were  heteroscedas- 
tic,  we  used  nonparametric  ANOVA  (Kruskal-Wallis  or 
ANOVA  on  ranks)  for  comparisons  of  mean  BF  among 
ages,  months,  and  years.  To  investigate  the  relationships 
between  BF  and  length,  somatic  weight  (ovary-free  body 
weight),  and  age,  we  used  linear  regression. 

Relative  fecundity  (RF)  was  calculated  as  the  num- 
ber of  oocytes  per  gram  somatic  weight  (ovary-free). 
To  select  samples  for  inclusion  in  RF  calculations,  we 
looked  for  the  presence  of  nuclear  migration  in  histo- 
logical preparations.  We  used  this  criterion  to  ensure 
that  oocytes  of  similar  morphological  dynamics  would 
be  used,  minimizing  the  potential  for  error.  We  used 
the  Kruskal-Wallis  test  to  investigate  the  effect  of  age 
on  RF.  Because  sample  sizes  were  quite  uneven  among 
months,  we  chose  to  compare  RF  between  the  beginning 
and  end  of  the  spawning  season  (May  and  August).  This 
comparison  was  done  by  using  a  Mann-Whitney  test. 
To  corroborate  any  trends  in  RF,  we  also  conducted 
diameter  measurements  on  the  preserved  (10%  buffered 


478 


Fishery  Bulletin  102(3) 


seawater  formalin)  oocytes.  We  used  a 
video  camera  mounted  on  a  Nikon  SMZ-U 
dissecting  microscope  and  coupled  to  a  PC 
equipped  with  a  frame-grabber  and  with 
OPTIMAS-  Image  Analysis  software  (ver- 
sion 6,  Media  Cybernetics,  Bothell,  WA). 
Two  readers  independently  measured  the 
diameter  of  approximately  30  preserved 
oocytes  in  each  of  three  subsamples  from 
27  ovaries.  To  test  for  uniformity  of  size 
throughout  the  ovary,  mean  oocyte  diame- 
ters were  compared  between  ovarian  lobes 
and  among  subsample  locations  within 
each  lobe  by  using  two-way  ANOVA.  We 
also  compared  mean  oocyte  diameters 
among  months  and  ages  by  using  two- 
way  ANOVA. 

Annual  fecundity 


Month 


Wiley  (1996)  demonstrated  that  spot- 
ted seatrout  in  South  Carolina  estuaries 
constitute  a  single  population.  Therefore, 
we  felt  justified  in  calculating  monthly 
egg  production  (MEP)  by  multiplying  the 
monthly  SF  (of  specimens  taken  along  the 
entire  coast)  by  the  mean  monthly  BF  (of  specimens  from 
Charleston  Harbor).  Because  not  all  age-1  female  trout 
were  mature  at  the  beginning  of  the  spawning  season, 
the  fraction  of  mature  age-1  females  obtained  from  previ- 
ous work  in  South  Carolina  (Wenner6)  was  used  to  refine 
the  MEP  estimate.  Because  the  latter  was  calculated  by 
using  SF  obtained  from  data  pooled  across  years,  any 
comparison  of  MEP  among  years  was  deemed  invalid. 
Kruskal-Wallis  tests  were  used  to  determine  whether 
MEP  varied  among  months  for  each  age  class. 

Monthly  MEP  estimates  were  summed  to  arrive  at 
an  annual  fecundity  (AF)  estimate  for  each  age  class. 
Because  the  majority  of  individuals  used  in  this  study 
were  aged  1-3,  AF  was  estimated  only  for  these  age 
classes.  We  used  linear  regression  to  investigate  the 
relationship  between  AF  and  age  and  thus  predict  AF 
for  spotted  seatrout  aged  4  and  5.  Using  these  predic- 
tions and  the  relative  abundance  of  each  age  class  in 
our  samples,  we  estimated  the  contribution  of  each  age 
class  to  the  annual  egg  production. 

All  statistical  analyses  were  conducted  with  the  Sta- 
tistical Package  for  the  Social  Sciences  (version  9.0, 
SPSS  Inc.,  Chicago,  ID.  The  level  of  significance  for 
all  tests  was  0.05. 


Results 

A  total  of  1038  spotted  seatrout  ranging  in  age  from 
1  to  5  was  collected  for  this  study.  Because  97%  of 
these  belonged  to  age  classes  1-3  we  report  reproductive 
parameters  only  for  these  ages.  We  examined  a  total  of 
941  mature  and  developing  females,  ranging  in  Length 
from  248  mm  to  542  mm  TL,  to  determine  spawning 


Figure  2 

Mean  monthly  gonadosomatic  index  iGSIl  for  spotted  seatrout  in  South 
Carolina  for  years  1991-2000  (circles).  Mean  water  temperature  for 
1991-2000  i triangles).  Error  bars  are  standard  errors,  n  =  1185. 


frequency  (569,  285,  and  87  for  ages  1-3,  respectively). 
Of  these,  135  specimens  (12  from  sportfishing  tourna- 
ments) were  used  to  conduct  oocyte  counts  (62,  52,  and 
21  for  ages  1-3.  respectively).  These  fish  ranged  in 
length  from  268  to  530  mm  TL.  Minimum  size  at  first 
maturity,  as  indicated  by  the  presence  of  cortical  alveoli 
stage  oocytes  in  histological  sections,  was  248  mm  TL. 
Size  at  50'/c  maturity  was  268  mm,  whereas  100';  matu- 
rity was  reached  at  301  mm  TL.  Condition  of  females,  as 
indicated  by  Fulton's  condition  factor,  diminished  over 
the  course  of  the  season  (P<0.01,  r2=0.24i. 

Seasonality 

Spawning  in  the  Charleston  Harbor  area  during  the 
study  period  began  in  mid  to  late  April  as  indicated  by 
the  presence  of  oocytes  in  late  FOM  or  POFs  in  histo- 
logical samples.  During  the  study  period,  mean  water 
temperatures  ranged  from  16°  to  34°C.  Highest  tem- 
peratures were  recorded  during  July  and  August  for  all 
three  years  of  the  study.  The  lowest  documented  water 
temperature  when  spawning  began  was  20°C.  Cessation 
of  spawning  occurred  when  water  temperature  was  28°C. 
Mean  monthly  GSI  for  spotted  seatrout  captured  along 
the  South  Carolina  coast  since  1991  (Fig.  2)  showed  a 
marked  increase  from  4.6  in  April  to  9.4  in  May.  Mean 
GSI  in  June  declined  to  6.3  and  remained  around  5.0  in 
July  and  August.  A  sharp  decline  was  noted  in  Septem- 
ber to  2.7,  the  lowest  level  for  the  season.  Overall,  mean 
gonadosomatic  index  (GSI)  values  followed  the  seasonal 
trend  in  water  temperature  (Fig.  2).  Percent  occurrence 
of  females  in  spawning  condition  as  evidenced  from 
histological  examination  declined  from  approximately 
87',  in  August  to  12',  in  September.  The  percentage  of 


Roumillat  and  Brouwer:  Reproductive  dynamics  of  Cynoscion  nebulosus 


479 


Spawning 

Sampling 
Day  0  POFs 

•            •           • 

m 

A            • 

•-          • 

1800       t 

2200 

0200 

0600 

t      1000         * 

1400 

t     1800 

Fig.  4  A 

Fig.  4B 

Fig.  4C 

Fig.  4D         Fig.  4E 

Sampling 
Day  1  POFs 

Fig.  4F 

A 

m                  A 

1800 

• 
2200 

0200 

0600 

Figure 

3 

1000 

1400 

1800 

Forty-eight-hour  chronological  time  line  indicating  C.  nebulosus  spawning  time  in  Charleston 
Harbor,  SC.  Postovulatory  follicles  (POFs)  observed  at  water  temperatures  >25  C  were  termed 
day-0  (<24  hours).  Day-1  POFs  (>24  hours)  were  observed  only  at  water  temperatures  <25°C. 
Time  of  capture  for  specimens  in  Figure  4  are  indicated. 

post-spawning  females  increased  from  99r  in  August  to 
91%  in  September.  Thus,  the  spawning  season  for  spot- 
ted seatrout  in  South  Carolina  extends  from  late  April 
through  early  September. 

Spawning  frequency 

Day-0  POFs  were  found  through  1800  h  of  the  day  fol- 
lowing a  spawning  event.  Day-1  POFs  were  first  observed 
in  our  routine  samples  when  they  were  36-37  hours  old 
(the  second  day  following  a  spawning  event)  only  when 
water  temperatures  were  below  25°C.  Day-1  POFs  were 
excluded  from  our  analysis  of  SF  because  they  did  not 
provide  evidence  of  a  previous  night's  spawning  event. 

Figure  3  illustrates  the  time  line  for  POF  atrophy  in 
spotted  seatrout  from  1-42  h  after  the  onset  of  spawn- 
ing at  1800  h.  Because  evidence  of  spawning  for  the  first 
12  h  was  documented  only  during  a  period  when  water 
temperatures  were  greater  than  25°C,  all  of  the  exam- 
ples shown  are  indicative  of  atrophy  in  warmer  tempera- 
tures (Fig.  4).  As  indicated  in  Table  2,  there  was  a  time- 
dependent  deterioration  of  POFs  such  that  only  those 
<24  h  were  detectable  at  water  temperatures  >25°C. 

Small  sample  sizes  prevented  calculation  of  monthly 
SF  for  each  age  class  by  year.  Therefore,  we  pooled 
data  for  all  three  years  of  this  study  to  obtain  a  single 
monthly  SF  estimate  by  age  class  (Tables  3  and  4). 
The  interaction  between  month  and  age  on  SF  could 
not  be  statistically  tested;  however,  age-3  fish  spawned 
more  frequently  than  younger  fish  (Kruskal-Wallis, 
P<0.05)  and  all  seatrout  spawned  more  frequently  in 
June  (Kruskal-Wallis,  P<0.05).  Peaks  in  SF  observed 
for  fish  ages  2  and  3  in  July  and  August,  respectively 
(Tables  3  and  4),  were  not  statistically  significant. 

Monthly  SF  values  based  on  the  occurrence  of  ova- 
ries containing  oocytes  in  FOM  are  also  presented  in 


Table  3 

Spawning  frequency  (SF)  expressed  as 

the  number  of 

spawnings  per  month  for 

C.  nebulosus  a 

ges  1-3  for  the 

spawning  seasons 

of  1998 

-2000.  Numbers  in  parenthe- 

ses  repr 

esent  days 

betweer 

spawnings.  n  -- 

=  number  offish 

in  a  sample.  FOM 

=  final  oocyte  maturation;  POF  =  post- 

ovulatory  follicle. 

Age 

SF 

SF 

(yr) 

Month 

n 

FOM  method 

POF  method 

1 

May 

89 

4.53(6.85) 

4.18(7.42) 

June 

166 

4.53(6.62) 

9.40(3.131 

July 

185 

4.68(6.62) 

6.54(4.74) 

August 

129 

6.26(4.95) 

4.57(6.79) 

Total 

569 

19.9(6.18) 

26.34(4.67) 

2 

May 

114 

11.5(2.78) 

6.80(4.56) 

June 

79 

5.70(5.26) 

7.60(3.95) 

July 

48 

0.65(47.62) 

9.04(3.43) 

August 

44 

9.87(3.14) 

6.34(4.89) 

Total 

285 

30.67(4.011 

29.36(4.19) 

3 

May 

46 

10.10(3.07) 

7.42(4.18) 

June 

23 

5.22(5.75) 

9.12(3.29) 

July 

10 

3.10(10.00) 

3.10(10.00) 

August 

8 

11.61  (2.67) 

11.61(2.67) 

Total 

87 

32.54(2.67) 

31.14(3.95) 

Overall 

941 

24.3(5.06) 

27.7(4.44) 

Table  3.  However,  statistical  comparisons  were  not  fea- 
sible because  of  the  nonrandom  collection  of  specimens. 
Overall  SF  was  estimated  to  be  once  every  4.4  days  and 
once  every  5.1  days  with  the  POF  and  FOM  methods, 
respectively. 


480 


Fishery  Bulletin  102(3) 


l\»«. 


•*•?•*►.    . 


Ml  W 


•  V 


100  um 

Figure  4 

Photomicrographs  of  C.  nebulosus  postovulatory  follicles  (POFs)  showing  chronology  of  atresia 
at  water  temperatures  >25°C.  (A)  0-4  hours  after  spawning.  (B)  5-8  hours  after  spawning. 
<C)  9-12  hours  after  spawning,  ID,  E,  F>  13-24  hours  after  spawning. 


Batch  fecundity 

As  expected,  we  found  a  significant  difference  in  mean 
BF  among  age  classes  (ANOVA  on  ranks,  P<0.05).  Age-1 


spotted  seatrout  produced  an  average  of  145,452  oocytes 
per  hatch  spawned.  Fish  aged  2  and  3  spawned  an  aver- 
age of  291,123  and  529,976  oocytes  per  batch,  respec- 
tively. Therefore,  mean  BF  was  compared  among  months 


Roumillat  and  Brouwer:  Reproductive  dynamics  of  Cynosaon  nebulosus 


481 


Table  4 

Fecundity  parameters  for  C.  nebulosus  ages  1-3  from  South  Carolina  estuaries 

BF  =  batch  fecundity  in  numbe 

•s  of  oocytes; 

SF  =  spawning  frequency  based  on  the  postovulatory  follicle  (POF)  method  and 

expressed  as  the  nun 

tber  of  spawnin 

is  per  month : 

MEP  =  monthly  egg  production  =  (BF  •  SFi'i  mature.  Annual  fecundity  is 

the 

sum  of  mean  month 

y  MEP  values 

ibr  each  vear 

class  and  represents  the  total  number  of  oocytes  produced  by  any  given  female  from  1  May  to  31  Aug 

jst.  Numbers  ir 

parentheses 

indicate  sample  sizes. 

Age  (yrl        Month                                                                              Mean  BF 

SF 

'i  mature 

Mean  MEP 

1                    May                                                                                117.760(12) 

4.18(89) 

78.6 

386,897 

June                                                                               135,403(161 

9.40(166) 

94.0 

1,196,418 

July                                                                                 141,237(16) 

6.54(185) 

97.0 

895,978 

August                                                                           176.594(18) 

4.57(129) 

100 

807,035 

Annual  fecundity=3,286,328  oocytes 

2                    May                                                                                280,724(34) 

6.80(114) 

100 

1,908,926 

June                                                                               307.322(101 

7.60(79) 

100 

2,335,650 

July                                                                                370.170(1) 

9.04(48) 

100 

3,346,337 

August                                                                      307,195(7) 

6.34(44) 

100 

1,947,620 

Annual  fecundity=9,538,533  oocytes 

3                    May                                                                                487,475(131 

7.42(461 

100 

3.617,061 

June                                                                               519,630(4) 

9.12(23) 

100 

4,739,027 

July                                                                                765.911(2) 

3.1(10) 

100 

2,374,325 

August                                                                           590,994(2) 

11.61(8) 

100 

6,861,439 

Annual  fecundity=  17.591,852  oocytes 

Table  5 

Monthly  relative  fecundity  (number  of  oocytes 
1998-2000.  SD  =  standard  deviation. 

/grams  ovary-free  weight)  for  C.  nebu 

losus 

ages  1- 

-3  for  the  spawning  seasons 

Month 

Mean 

Minimum 

Maximum 

SD                                    n 

May 

518.6 

223.9 

976.1 

146.2                                 46 

June 

603.2 

205.7 

1306.1 

241.8                                 20 

July 

820.9 

662.2 

1314.4 

279.0                                 5 

August 

693.6 

397.3 

1021.8 

207.9                               12 

and  years  for  each  age  class  separately  (Table  4).  There 
were  no  significant  interannual  or  monthly  variations  in 
mean  BF  for  any  of  the  age  classes  (age-1:  P=0.59,  ;;  =  62; 
age-2:  P=0.17,  n=52;  age-3:  P=0.07,  n=21).  However,  BF 
analysis  for  age-2  fish  excluded  the  month  of  July  because 
only  one  two-year-old  specimen  was  captured  that  month 
during  the  study  period.  We  investigated  the  relationship 
between  BF  and  total  length  by  using  linear  regression 
analysis.  After  pooling  data  across  years,  we  found  that 
total  length  explained  67%  of  the  variability  in  spotted 
seatrout  BF  (Fig.  5A).  Batch  fecundity  showed  a  similarly 
strong  relationship  to  female  somatic  (ovary-free)  weight 
(Fig.  5B)  but  did  not  relate  to  age  as  strongly  (Fig.  5C). 
The  equations  below  describe  these  relationships: 


BF=  2179.65ITL)  -  520597 
BF  =  530.60IOFWT)  +  18537.77 
BF  =  169398. 21( Age)  -  30956.33 


(r2=0.67)  P<0.001 
(r2=0.65)  P<0.001 
(r2=0.58)     P<0.001. 


Mean  MEP  was  significantly  different  among  months 
for  age-1  spotted  seatrout  (Kruskal-Wallis,  P<0.05). 
Age-1  fish  spawned  the  least  number  of  oocytes  in  May 
and  most  in  June  (Table  4).  Statistical  comparisons 
among  months  for  ages  2  and  3  were  inconclusive. 

Relative  fecundity 

Relative  fecundity  among  83  spotted  seatrout  ages 
1-3  ranged  from  224  oocytes  to  1314  oocytes/g  OFWT 
(Table  5).  Age  did  not  have  an  effect  on  relative  fecun- 
dity (Kruskal-Wallis,  P=0.75).  We  found  that  spotted 
seatrout  in  South  Carolina  produced  significantly  more 
oocytes  per  gram  ovary-free  weight  at  the  end  than  at 
the  beginning  of  the  spawning  season  (Mann  Whit- 
ney, P<0.05).  Mean  oocyte  diameters  did  not  vary  sig- 
nificantly between  ovarian  lobes  or  among  locations 
within  each  lobe  (ANOVA,  P=0.28).  A  comparison  among 


482 


Fishery  Bulletin  102(3) 


CJ 

O 

1200  - 

BF=  2179.65(71)-  520597                                                                         A 

X 

f  =  0.67.  P<0.001                                                     • 

CO 

0) 

n  =  134 

1000  - 

o 

o 

o 

800  - 

•                 / 

6 

%                                                  S^ 

c 

%                 s^ 

>. 

•                     *    /^ 

c 

600  - 

••  ••  */ 

3 

* '  s/9 

400  ■ 

•  >  x^rT 

o 

CO 

••  u*wK 

200  ■ 

CO 

2 

-     •  T"* 

2! 

1 —    n               i                        i 1 1                        i 

0                  300                  350                  400                  450                  500                  550 

Total  length  (mm) 

O 

1400  ■ 

B 

*"" 

SF=  530  60(O/=WT)  +  1853777 

Ul 

1200  • 

f  =  0  65.  P<0.001 

<u 

n=133                                                         • 

%. 

o 

o 
o 

1000  ■ 

o 

6 

800  ■ 

• 

c 

• 

>. 

•                         ■" 

■6 

600  ■ 

'       •'•      '        %l^ 

^ 

o 

•   ^^^"^ 

J3) 

00 

n 

c 
ro 

CD 

400  - 
200  - 

' — i r 1 1 1 i 

200                 400                 600                 800                1000               1200 

Ovary-free  body  weight  (g) 

o 

1200  - 

c 

BF=  169398.21(age)  -  3095633                                                                    ^ 

X 

?  =  0.58.  P<0  001                                • 

n  =  134 

>. 

1000  - 

o 

o 

"o 

800  - 

• 

6 

• 

^ 

• 

>> 

i 

t5 

600  ■ 

1 

o 

a> 

400  ■ 

o 

1 

ro 

200  • 

• 
• 

c 

ro 

5 

• 

•                      • 

1                                               2                                              3 

Age 

Figure  5 

Re 

at  Kinship  between  batch  fecundity  iBF)  and  total  length 

(TL)  lA),  between  BF  and  ovary-free  weight  (OFWTi  (Bi,  and 

bel 

ween  BF  and  age  (Cl  for  C.  nebulosus  ages  1-3.  Linear  re- 

gression  on  data  pooled  for  spawning  seasons  1998-2000. 

months  and  ages  revealed  that  age  had  no  effect 
on  oocyte  diameter  (ANOVA,  P=0.82).  However, 
the  effect  of  month  corroborated  the  pattern  of 
increasing  RF  as  the  spawning  season  progressed: 
oocytes  were  significantly  smaller  at  the  end  of 
the  season  (ANOVA,  P<0.05). 

Annual  fecundity 

Annual  fecundity  estimates  (summation  of  MEP) 
were  approximately  3.2  million.  9.5  million,  and 
17.6  million  oocytes  for  each  age  class,  respec- 
tively (Table  4).  The  equation  below  describes  the 
relationship  between  AF  and  age: 

AF  =  7152762(Age)  -  4166620  (r2=0.99>  P<0.05. 

From  this  relationship,  the  predicted  AF  for 
ages  4  and  5  were  24,444,430  and  31,597,190  oo- 
cytes, respectively.  We  expanded  AF  in  relation  to 
the  abundance  of  each  age  class  in  our  standard 
random  samples  for  the  three  years  of  the  study. 
We  estimated  that  the  overall  average  contribu- 
tion from  age-1  fish  to  the  reproductive  output 
for  the  season  was  approximately  29% .  whereas 
fish  aged  2  and  3  contributed  39%  and  21',  of 
oocytes,  respectively.  Ages  4-5  comprised  less 
than  3%  of  specimens  sampled  and  contributed 
7%  and  4%  based  on  predicted  AF  values. 


Discussion 

Studies  on  the  reproductive  biology  of  Cynoscion 
nebulosus  have  established  group-synchrony  and 
indeterminate  fecundity  for  this  species  through- 
out its  range  (i.e.  Brown-Peterson  et  al.,  1988; 
Brown-Peterson  and  Warren,  2001;  Nieland  et 
al.,  2002;  Mercer1  and  references  therein).  Fish 
with  these  features  release  gametes  in  several 
batches  over  a  protracted  spawning  season  and 
annual  fecundity  is  not  fixed  prior  to  the  onset  of 
spawning  (Wallace  and  Selman.  1981). 

Based  on  mtDNA  variation  among  spotted  sea- 
trout,  the  existence  of  two  populations,  one  in 
the  Gulf  of  Mexico  and  one  in  the  South  Atlantic, 
was  established  by  Gold  et  al.  (1999).  However, 
variations  in  reproductive  parameters  have  been 
suggested  among  geographic  locations  within  the 
Gulf  of  Mexico  (Brown-Peterson  et  al.,  2002). 
Wiley  (1996)  suggested  that  spotted  seatrout  com- 
prise a  single  stock  in  South  Carolina:  therefore 
reproductive  parameters  presented  in  the  present 
study  should  be  applicable  only  to  the  spotted 
seatrout  population  inhabiting  coastal  waters  of 
this  state.  Further  studies  should  be  conducted 
to  evaluate  the  applicability  of  these  parameters 
to  the  entire  southeast  coast. 

Other  investigators  (Brown-Peterson  et  al., 
1988;  Wieting,  1989;  Brown-Peterson  and  War- 


Roumillat  and  Brouwer:  Reproductive  dynamics  of  Cynosc/on  nebu/osus 


483 


ren,  2001;  Nieland  et  al.,  2002;  Lowerre-Barbieri  et  al.8) 
have  used  the  gonadosomatic  index  (GSI)  to  delineate 
the  spawning  season  in  spotted  seatrout.  Even  though 
the  GSI  provided  a  good  approximation  of  the  spawning 
season,  histological  data  alone  provided  more  precise 
evidence.  Spotted  seatrout  in  South  Carolina  began 
spawning  near  the  end  of  April  of  each  year  and  ceased 
by  early  September.  Similarly,  Lowerre-Barbieri  et  al.8 
reported  that  the  spawning  season  for  spotted  seatrout 
in  Georgia  extended  from  late  April  to  mid-September. 
We  found  histological  evidence  of  initial  spawning  in 
specimens  captured  in  20°C  water,  although  approxi- 
mately 75%  of  spawning  occurred  when  ambient  water 
temperatures  were  greater  than  25°C.  In  laboratory 
experiments,  Brown-Peterson  et  al.  (1988)  found  no  suc- 
cessful spawning  in  water  below  23°C  but  pointed  out 
that  others  (McMichael  and  Peters,  1989)  found  eggs 
and  larvae  in  20.4°C  water. 

We  found  that  females  became  mature  approximately 
one  full  year  after  their  birth.  A  female  born  in  May  of 
one  year  would  be  reproductively  active  in  May  of  the 
following  year.  Females  born  later  in  the  season  would 
not  be  mature  as  the  same  successive  season  began; 
therefore,  not  all  one-year-old  females  were  mature 
when  the  spawning  season  began  in  May,  but  became 
mature  before  that  season  ended.  This  maturity  sched- 
ule has  also  been  reported  for  spotted  seatrout  in  Loui- 
siana (Nieland  et  al.,  2002).  However,  Lowerre-Barbieri 
et  al.8  found  that  all  one-year-old  females  were  mature 
in  coastal  Georgia.  A  limited  sample  size  or  habitat 
segregation  of  mature  and  immature  trout  (Lowerre- 
Barbieri  et  al.8)  may  have  contributed  to  their  result. 

The  size  at  first  maturity  for  spotted  seatrout  in  this 
study  was  248  mm  TL.  This  size  is  comparable  to  what 
others  have  reported  in  other  areas  of  the  species'  range 
(Brown-Peterson  et  al.,  1988;  Brown  Peterson  and  War- 
ren, 2001;  Nieland  et  al.,  2002;  Mercer1  and  references 
therein;  Lowerre-Barbieri  et  al.8).  Our  estimate  of  size 
at  50%  maturity  (268  mm  TL)  was  larger  than  what 
Nieland  et  al.  (2002)  reported  for  100%  mature  trout 
in  Louisiana  (250  mm  TL).  However,  Nieland  et  al.'s 
(2002)  statement  that  animals  are  100%  mature  at  250 
mm  TL,  does  not  agree  with  the  growth  equation  they 
report  for  female  trout  when  age  =  1.  Because  we  found 
size  at  100%  maturity  among  female  spotted  seatrout 
in  South  Carolina  to  be  about  300  mm  TL,  we  wonder 
whether  Nieland  et  al.'s  (2002)  growth  equation  for 
female  TL  was  meant  to  represent  SL.  Were  this  the 
case,  they  might  have  offered  a  different  rationale  for 
size  at  maturity  among  trout  in  Louisiana. 

Brown-Peterson  et  al.  (1988)  and  Brown-Peterson 
and  Warren  (2001)  reported  size  at  100%  maturity  of 
356  mm  and  309  mm  TL  (using  the  SL-TL  conversion 
found  in  our  "Methods:  section)  for  spotted  seatrout  in 
Texas  and  Mississippi,  respectively.  Brown-Peterson  et 
al.  (1988),  however,  chose  a  combination  of  gears  that 
may  not  have  sampled  the  trout  population  in  Texas 
representatively  for  size-at-maturity  estimation.  In  Mis- 
sissippi, Brown-Peterson  and  Warren  (2001)  used  a 
more  appropriate  gear  for  capture  of  late  juvenile  and 


early  adult  fish.  Our  estimate  of  size  at  lOO1*  maturity 
was  quite  similar  to  theirs. 

Spawning  frequency 

Determining  the  number  of  multiple  spawning  events 
during  a  single  season  for  individual  fish  has  been  prob- 
lematic. Initially,  there  was  little  understanding  of  the 
reproductive  dynamics  of  spotted  seatrout,  and  BF  esti- 
mates were  reported  to  represent  the  output  for  a  whole 
season  (Pearson,  1929;  Sundararaj  and  Suttkus,  1962; 
Overstreet,  1983).  Hunter  et  al.  (1985)  and  Hunter  and 
Macewicz  (1985)  developed  techniques  to  overcome  these 
limitations  by  providing  protocols  for  the  use  of  hydrated 
oocytes  in  determining  BF  and  SF  among  group-syn- 
chronous species. 

To  use  the  techniques  of  Hunter  (1985)  and  Hunter 
and  Macewicz  (1985)  appropriately,  it  is  critical  to  obtain 
a  representative  sample  of  the  spawning  population. 
DeMartini  and  Fountain  (1981)  and  Lisovenko  and  Adri- 
anov  (1991)  maintained  that  the  relative  occurrence  of 
hydrated  oocytes  (as  determined  macroscopically)  was  an 
effective  measurement  of  SF  when  the  spawning  popula- 
tion was  sampled  representatively.  However,  when  sam- 
pling a  species  that  spawns  in  aggregations  at  specific 
geographic  locations,  as  do  many  of  the  sciaenids,  it  is 
inherently  impossible  to  obtain  a  statistically  representa- 
tive sample  of  the  spawning  population  for  SF  estimation 
based  on  FOM.  Because  the  window  of  opportunity  is 
temporally  and  spatially  constrained,  obtaining  a  sample 
that  includes  all  sizes  and  ages  involved  is  not  feasible; 
the  only  choice  in  this  situation  is  to  sample  in  a  directed 
fashion.  This  was  the  sampling  strategy  used  to  target 
females  for  BF  counts;  the  majority  of  the  animals  cap- 
tured whose  oocytes  evidenced  FOM  were  obtained  in  a 
nonrandom  fashion.  Additionally,  we  assumed  that  fishes 
demonstrating  FOM  were  moving  toward  deeper  water 
spawning  aggregations  and  away  from  our  capture  gear. 
For  these  reasons,  we  felt  that  our  SF  estimates  based 
on  the  proportion  of  females  with  oocytes  in  FOM  were 
biased  and  we  excluded  them  from  AF  estimation.  This 
is  an  important  matter  to  keep  in  mind  when  comparing 
frequencies  of  spawning  based  on  different  methods. 

Because  obtaining  representative  numbers  of  ani- 
mals with  late-maturing  oocytes  is  not  often  feasible, 
researchers  have  relied  on  the  relative  abundance  of 
postovulatory  follicles  (POFs)  to  calculate  SF  (Brown- 
Peterson  et  al.,  1988;  Brown-Peterson  and  Warren,  2001; 
Nieland  et  al.,  2002;  Lowerre-Barbieri  et  al.8).  The  POF 
method  lacks  the  limitations  (described  above)  of  the 
FOM  method.  Because  the  method  we  chose  allowed 
us  to  sample  all  sizes  and  ages  of  fish  in  the  estuary, 
obtaining  representative  numbers  of  animals  with  POFs 
was  accomplished  effectively.  Therefore,  we  felt  that  our 
estimates  of  SF  based  on  the  POF  method  were  more 
precise  and  we  chose  to  use  them  in  deriving  AF. 

The  POF  method  depends  on  the  ability  to  assess  the 
disappearance  of  these  structures.  Hunter  and  Macewicz 
(1985)  systematically  sampled  captive  spawning  ancho- 
vies to  develop  histological  criteria  for  POF  atrophy  in 


484 


Fishery  Bulletin  102(3) 


19°C  water.  Their  criteria  have  been  used  by  others 
to  estimate  rates  of  POF  atrophy  in  other  species  and 
thereby  determine  the  percentage  of  a  population  un- 
dergoing spawning  over  a  discrete  time  period  (Brown- 
Peterson  et  al,  1988;  Fitzhugh  et  al.,  1993;  Taylor  et  al., 
1998;  Macchi  and  Acha,  2000;  Brown-Peterson  and  War- 
ren, 2001;  Nieland  et  al.,  2002).  However,  even  though 
it  has  been  demonstrated  that  the  rate  of  POF  atresia 
depends  largely  on  ambient  water  temperature  (Fitzhugh 
and  Hettler,  1995),  few  (Brown-Peterson  et  al.,  1988; 
Macchi  and  Acha,  2000;  Nieland  et  al,  2002)  have  taken 
this  into  account  when  establishing  the  age  of  POFs  for 
SF  estimations.  Our  diurnal  sampling  of  reproductively 
active  spotted  seatrout  during  warm  water  conditions  en- 
abled us  to  establish  criteria  to  accurately  estimate  the 
age  of  POFs  throughout  the  spawning  season.  Further- 
more, we  verified  our  assessments  by  sampling  around 
the  clock  on  two  occasions  to  collect  fish  over  the  time 
period  immediately  following  a  spawning  event. 

Spotted  seatrout  ages  1-3  in  SC  spawned  less  fre- 
quently than  those  from  the  Indian  River  Lagoon,  Flor- 
ida (Crabtree  and  Adams7)  but  both  studies  showed 
that  older  fish  spawned  more  frequently  than  younger 
animals.  Our  estimates  for  spotted  seatrout  aged  1-3 
were  4.7,  4.2,  and  4  days,  respectively.  Trout  in  these 
age  classes  in  Florida  were  reported  to  spawn  once  ev- 
ery 4,  2.8,  and  2.5  days,  respectively.  These  differences 
probably  not  only  reflect  the  distinct  biological  environ- 
ments of  each  region  but  also  indicate  potential  discrep- 
ancies in  aging  methods.  No  age-specific  estimates  of 
SF  are  available  for  other  areas  in  the  species'  range. 
Brown-Peterson  and  Warren  (2001)  found  SF  among 
spotted  seatrout  in  Biloxi  Bay,  MS,  to  be  significantly 
lower  than  that  of  fish  inhabiting  the  other  two  areas 
included  in  their  study.  They  suggested  that  Biloxi  Bay 
was  a  less  conducive  spawning  habitat  because  of  sev- 
eral factors,  including  shoreline  development  and  a 
reduced  amount  of  aquatic  vegetation.  However,  because 
we  found  that  SF  varied  significantly  among  age  classes 
(age-3  fish  spawned  more  frequently),  the  relative  age 
composition  of  fish  sampled  by  Brown-Peterson  and 
Warren  (2001)  in  the  three  estuaries  might  also  have 
played  a  critical  role  in  the  determination  of  SF. 

Batch  fecundity 

The  best  approach  for  estimating  BF  is  to  use  only  oocytes 
in  FOM  (Hunter  et  al.,  1985;  Brown-Peterson  et  al.,  1988; 
Brown-Peterson  and  Warren,  2001;  Brown-Peterson  et 
al.,  2002;  Nieland  et  al.,  2002;  Lowerre-Barbieri  et  al.8). 
When  it  is  not  possible  to  obtain  these,  BF  estimations 
can  and  have  been  carried  out  in  some  species  by  using 
the  largest  vitellogenic  oocytes  (Overstreet,  1983;  Hunter 
et  al.,  1985;  Wieting  1989).  These  efforts  have  the  poten- 
i  ill  of  being  less  accurate  because  isolating  those  oocytes 
destined  to  be  spawned  is  difficult  if  the  latter  have 
not  yet  reached  final  maturation  (Nieland  et  al.,  2002). 
Inevitably  this  scenario  would  result  in  a  nonmeasurable 
overestimation  of  female  reproductive  output.  Brown- 
Peterson  et  al.  (1988)  and  Brown-Peterson  and  Warren 


(2001)  used  a  modification  of  this  approach  to  estimate 
BF  of  spotted  seatrout  in  Texas  and  Mississippi,  respec- 
tively. However,  even  though  the  potential  existed  for 
overestimating  BF,  their  estimates  fell  well  below  those 
presented  in  the  present  study,  as  did  those  presented 
by  Nieland  et  al.  (2002)  for  spotted  seatrout  ages  2-4  in 
Barataria  Bay,  Louisiana.  Mean  BF  for  ages  1-3  (170 
thousand,  226  thousand,  and  274  thousand  oocytes, 
respectively)  spotted  seatrout  in  Indian  River  Lagoon, 
Florida  (Crabtree  and  Adams7),  also  differed  from  those 
reported  here.  Our  estimate  took  into  account  that  not 
all  age-1  females  were  mature  at  the  beginning  of  the 
season.  Crabtree  and  Adams,7  however,  did  not  adjust 
their  estimate  to  reflect  this  discrepancy.  Moreover,  due 
to  differences  in  aging  methods,  their  age-1  and  2  cohorts 
possibly  included  ages  2  and  3,  respectively.  In  addition, 
in  the  Florida  study  as  well  as  in  ours,  relatively  few 
numbers  of  older  specimens  were  examined. 

The  relationships  between  BF  and  length,  weight,  and 
age  in  the  present  study  were  significant  and  predictive. 
Of  these,  TL  exhibited  the  most  predictive  relationship. 
This  fact  may  explain  why  age-1  and  age-2  spotted 
seatrout  in  Georgia  had  mean  BFs  considerably  higher 
than  ours  (175  thousand  and  407  thousand,  respectively; 
Lowerre-Barbieri  et  al.8):  the  size  ranges  for  age-1  and 
age-2  in  the  Georgia  study  were  greater  than  ours.  To- 
tal length  seems  to  be  the  most  reliable  predictor  of  BF 
among  spotted  seatrout  in  Georgia  and  SC  (Lowerre- 
Barbieri  et  al.,8  this  study)  and  in  Louisiana  (Nieland 
et  al.,  2002).  However,  Crabtree  and  Adams7  found  that 
BF  related  best  to  ovary-free  weight  among  spotted 
seatrout  in  Florida.  We  found  ovary-free  weight  to  be 
the  second  best  predictor  of  BF.  Overall,  it  appeared 
that  TL  and  ovary-free  weight  were  better  predictors  of 
BF  than  age  for  this  species  (Brown-Peterson,  2003). 

As  with  SF,  monthly  egg  production  (MEP)  estimates 
for  SC  spotted  seatrout  varied  throughout  the  season. 
Because  BF  was  not  significantly  different  among 
months  for  any  of  our  age  classes,  the  variation  in 
MEP  resulted  directly  from  the  frequency  of  spawning. 
Monthly  egg  production  estimates  for  age-1  fish  were 
lowest  in  May  and  highest  in  June  because  SF  was 
lowest  in  May  and  highest  in  June.  Spawning  frequency 
is  a  critical  reproductive  parameter  because  it  seems 
to  dictate  annual  reproductive  output  (DeMartini  and 
Fountain,  1981;  Brown-Peterson  and  Warren.  2001; 
Crabtree  and  Adams7);  therefore,  SF  should  be  carefully 
considered,  particularly  for  managed  species. 

Relative  fecundity 

We  found  that  relative  fecundity  (RF),  the  number  of 
oocytes  per  gram  of  somatic  weight,  did  not  show  a  sig- 
nificant relationship  with  female  size.  This  finding  was 
expected  because  dividing  fecundity  by  ovary-free  weight 
standardizes  the  values  independently  of  size.  However, 
this  finding  was  in  contrast  to  that  of  Brown-Peterson 
and  Warren  (2001).  They  collected  specimens  during 
the  morning  only,  whereas  we  sampled  ours  throughout 
the  day.  This  procedure  allowed  us  to  examine  ovaries 


Roumillat  and  Brouwer:  Reproductive  dynamics  of  Cynosc/on  nebulosus 


485 


over  the  entire  range  of  maturation  and  to  select  only 
those  clearly  showing  nuclear  migration  (based  on  his- 
tological observations)  ensuring  that  only  oocytes  in  the 
same  phase  of  FOM  (Brown-Peterson  et  al.,  1988)  were 
included  in  RF  calculations.  If  sampling  is  conducted 
during  a  time  period  that  is  not  close  to  active  spawning 
(i.e.,  when  oocytes  are  in  different  phases  of  FOM),  then 
the  number  of  oocytes  per  gram  may  be  miscalculated. 

As  with  BF,  our  RF  estimates  were  higher  than  those 
reported  for  seatrout  in  the  Gulf  of  Mexico  (Brown- 
Peterson  et  al.,  1988;  Brown-Peterson  and  Warren, 
2001),  although  spotted  seatrout  reproductive  para- 
meters appeared  to  vary  considerably  even  within  the 
Gulf  of  Mexico  (Brown-Peterson  et  al.,  2002).  This  was 
attributed  to  differential  environmental  conditions  or 
food  availability  (or  to  both)  (Brown-Peterson  and  War- 
ren, 2001;  Brown-Peterson  et  al.,  2002).  The  significant 
seasonal  increase  in  RF  that  we  observed  for  spotted 
seatrout  in  South  Carolina,  however,  has  not  been  re- 
ported elsewhere.  Brown-Peterson  et  al.  (1988)  found  no 
differences  in  mean  monthly  RF  among  spotted  seatrout 
in  Texas.  Brown-Peterson  and  Warren  (2001)  found  sig- 
nificantly higher  RF  values  in  June  than  in  August.  In 
both  instances,  however,  a  small  sample  size  may  have 
biased  their  results. 

Comparisons  of  mean  oocyte  diameters  among  months 
related  the  increase  in  RF  to  a  general  decrease  in 
oocyte  size  over  the  course  of  the  season.  This  phenom- 
enon is  widespread  among  marine  pelagic  spawners, 
and  scientists  have  put  forth  several  explanations  to 
account  for  it  (see  Chambers,  1997).  Bagenal  (1971)  sug- 
gested that  egg  size  decreased  over  the  spawning  sea- 
son owing  to  concurrent  increased  food  availability  for 
larvae.  Others  have  suggested  an  inverse  relationship 
between  temperature  and  egg  size  (Ware,  1975;  Woot- 
ton,  1994;  Miller  et  al.,  1995)  or  a  seasonal  decrease  in 
egg  size  that  is  correlated  to  the  condition  of  spawning 
females  (DeMartini  and  Fountain,  1981;  Chambers  and 
Waiwood,  1996).  The  latter  seems  to  apply  to  spotted 
seatrout  in  this  study  because  a  diminishing  trend 
through  the  spawning  season  was  observed  in  the  con- 
dition factor  of  females. 

Annual  fecundity 

Brown-Peterson  (2003)  presented  AF  estimates  for  spot- 
ted seatrout  throughout  their  range.  Our  estimates  were 
substantially  below  those  for  spotted  seatrout  in  Indian 
River  Lagoon  (Crabtree  and  Adams7)  but  approximated 
those  of  Lowerre-Barbieri  et  al.8  for  trout  in  Georgia.  A 
possible  reason  for  the  higher  values  in  Florida  was  the 
more  protacted  spawning  season  in  that  area  (50  days 
longer).  No  comparisons  of  AF  estimates  presented  in 
this  study  and  those  of  spotted  seatrout  in  the  Gulf  of 
Mexico  (Brown-Peterson,  2003)  were  made  because  they 
were  not  specific  to  age  classes. 

The  main  impetus  behind  the  present  study  was  to 
establish  annual  fecundity  (AF)  estimates  by  age  class. 
We  found  that  age-1  through  age-3  spotted  seatrout 
occurred  abundantly  in  SC  estuaries  and  that  each  of 


these  age  cohorts  showed  unique  fecundity  dynamics. 
The  AF  for  an  average  age-1  fish  was  one-third  that  of 
age-2  (-3.28  million  vs.  9.5  million).  One  year-old  fish, 
however,  constituted  the  majority  offish  in  our  samples; 
their  abundance  was  twice  that  of  2-year-olds  and  seven 
times  that  of  3-year-old  fish.  Even  though  the  average 
age-3  trout  produced  almost  twice  as  many  oocytes  dur- 
ing the  season  (17.5  million)  as  the  average  age-2  fish, 
their  reduced  abundance  in  our  estuaries  made  their 
overall  contribution  only  half  that  of  2  year-olds.  Ages 
4  and  5  were  estimated  to  produce  approximately  24.4 
million  and  31.6  million  oocytes  per  female,  respective- 
ly; however,  the  oocyte  production  by  the  predominant 
age  groups  overshadowed  theirs.  When  analyzed  in  re- 
lation to  the  occurrence  of  the  other  age  classes  in  our 
estuaries,  age-2  fish  contributed  the  greatest  number  of 
fertilizable  oocytes  to  the  environment  (39%). 

Reliable  fecundities  based  on  age  and  on  length  are 
optimal  for  stock  assessment  models  (Williams9).  This 
study  provided  AF  estimates  for  three  age  classes  that 
can  be  used  in  age-based  models  for  the  spotted  seat- 
rout population  in  South  Carolina.  Annual  fecundity 
estimates  based  on  length,  however,  have  not  been  at- 
tempted even  though  length  appears  to  be  the  best  pre- 
dictor of  fecundity  in  spotted  seatrout  (see  references  in 
Brown-Peterson,  2003).  Further  analyses  to  investigate 
the  relationship  between  egg  production  and  fish  length 
for  each  month  of  the  spawning  season  would  allow  for 
more  precise  management  efforts  based  on  individual 
length-based  estimates  of  AF. 


Acknowledgments 

We  thank  members  of  the  Inshore  Fisheries  Section  of 
the  South  Carolina  Department  of  Natural  Resources  for 
assisting  in  field  data  collection  throughout  this  study 
(C.  Wenner,  J.  Archambault,  H.  von  Kolnitz,  W.  Hegler, 
E.  Levesque,  L.  Goss,  C.  McDonough,  C.  Johnson,  A. 
Palmer).  C.  Wenner,  H.  von  Kolnitz,  and  E.  Levesque 
conducted  age  assessments.  Histological  processing  was 
provided  by  C.  McDonough,  R.  Evitt,  A.  Palmer,  and 
W.  Hegler.  Assistance  with  oocyte  counts  was  provided 
by  C.  McDonough,  T.  Piper,  K.  Maynard,  and  R.  Evitt. 
Data  management  was  coordinated  by  J.  Archambault, 
C.  Wenner,  E.  Levesque,  and  three  anonymous  reviewers 
provided  helpful  suggestions  on  the  manuscript.  Funding 
for  this  study  was  provided  by  the  National  Marine  Fish- 
eries Service  under  MARFIN  grant  no.  NA77FF0550. 


Literature  cited 

Bagenal.  T.  B. 

1971.  The  interrelation  of  the  size  offish  eggs,  the  date 
of  spawning,  and  the  production  cycle.  J.  Fish.  Biol. 
3:207-219. 


;l  Williams,  E.  2003.  Personal  commun.  National  Marine 
Fisheries  Service,  101  Pivers  Island,  Rd.,  Beaufort,  NC 
28516. 


486 


Fishery  Bulletin  102(3) 


Brown-Peterson,  N.  J. 

2003.     The  reproductive  biology  of  spotted  seatrout.     In 

Biology  of  the  spotted  seatrout  (S.  A.  Bortone,  ed.i  p. 

99-133.     CRC  Press,  Boca  Raton,  FL. 
Brown-Peterson,  N.  J.,  P.  Thomas,  and  C.  Arnold. 

1988.  Reproductive  biology  of  the  spotted  seatrout, 
Cynoscion  nebulosus,  in  South  Texas.  Fish.  Bull.  86:3 
73-387. 

Brown-Peterson,  N.  J.,  M.  S.  Peterson,  D.  L.  Nieland, 
M.  D.  Murphy,  R.  G.  Taylor,  and  J.  R.  Warren. 

2001.  Reproductive  biology  of  female  spotted  seatrout, 
Cynoscion  nebulosus,  in  the  Gulf  of  Mexico:  differences 
among  estuaries?     Environ.  Biol.  Fish.  63:405-415. 

Brown-Peterson,  N.  J.,  and  J.  R.  Warren. 

2002.  The  reproductive  biology  of  spotted  seatrout, 
Cynoscion  nebulosus,  along  the  Mississippi  Gulf  coast. 
Gulf.  Mex.  Sci.  19:61-73. 

Chambers,  R.  C. 

1997.     Environmental  influences  on  egg  and  propagule 
size  in  marine  fishes.     In  Early  life  history  and  recruit- 
ment in  fish  populations  (R.  C.  Chambers  and  A.  E.  Trip- 
pel,  eds.),  p.  93-102.     Chapman  and  Hall,  London. 
Chambers,  R.  C,  and  K.  G.  Waiwood. 

1996.     Maternal  and  seasonal  differences  in  egg  sizes  and 
spawning  characteristics  of  captive  Atlantic  cod,  Gadus 
morhua.     Can.  J.  Fish.  Aquat.  Sci.  53:1986-2003. 
DeMartini,  E.  E.,  and  R.  K.  Fountain. 

1981.     Ovarian  cycling  frequency  and  batch  fecundity  in 
the  queenfish,  Seriphus  politus:  attributes  representa- 
tive of  spawning  fish.     Fish.  Bull.  79:547-560. 
Fish,  M.  P.,  and  W.  H.  Mowbray. 

1970.     Sounds  of  western  North  Atlantic  fishes.  A  refer- 
ence file  of  biological  underwater  sounds.     The  Johns 
Hopkins  Press,  Baltimore  and  London. 
Fitzhugh,  G.  R.,  and  W.  F.  Hettler. 

1995.     Temperature   influence   on   postovulatory   fol- 
licle degeneration  in  Atlantic  menhaden.  Brevoortia 
tyrannus.     Fish.  Bull.  93:568-572. 
Fitzhugh,  G.  R.,  B.  A.  Thompson,  and  T  G.  Snider. 

1993.     Ovarian   development,   fecundity   and   spawn- 
ing frequency   of  black  drum,  Pogonias  cromis.    In 
Louisiana.     Fish.  Bull.  91:244-253. 
Gabriel,  W.  L.,  M.  P.  Sissenwine,  and  W.  J.  Overholtz. 

1989.  Analysis  of  spawning  stock  biomass  per  recruit: 
an  example  for  Georges  Bank  haddock.  N.  Am.  J.  Fish. 
Manag.  9:383-391. 

Gold,  J.  R.,  L.  R.  Richardson,  and  C.  Furman. 

1999.     Mitochondrial   DNA  diversity   and   population 
structure  of  spotted  seatrout  iCynoscion  nebulosus)  in 
coastal  waters  of  the  Southeastern  LTnited  States.     Gulf 
Mex.  Sci.  1:40-50. 
Goodyear,  C.  P. 

1993.  Spawning  stock  biomass  per  recruit  in  fisheries 
management:  foundation  and  current  use.  In  Risk 
evaluation  and  biological  reference  points  for  fisheries 
management  (S.  J.  Smith,  J.  J.  Hunt,  and  D.  Rivard. 
eds.),  p.  67-81.  Can.  Spec.  Publ.  Fish.  Aquat.  Sci. 
120. 
Holt,  G.  J.,  S.  A.  Holt,  and  C.  R.  Arnold. 

1985.     Diel  periodicity  of  spawning  in  sciaenids.     Mar. 
Ecol.  Prog.  Ser.  27:1-7. 
Humason,  G.  L. 

1972.     Animal  tissue  techniques.     W.  H.  Freeman  and 
Co.  San  Francisco  and  London. 
Hunter,  J.  B.,  N.  C.  H.  Lo,  and  R.  J.  H.  Leong. 

1985.     Batch  fecundity  in  multiple  spawning  fishes,     In 


An  egg  production  method  for  estimating  spawning  bio- 
mass of  pelagic  fish:  application  to  the  northern  anchovy 
Engraulis  mordax  (R.  Lasker,  ed.i,  p.  67-77.  NOAA 
Tech.  Rep.  NMFS  36. 

Hunter,  J.  R.,  and  B.  J.  Macewicz. 

1985.  Measurement  of  spawning  frequency  in  multiple 
spawning  fishes.  In  An  egg  production  method  for  esti- 
mating spawning  biomass  of  pelagic  fish:  application 
to  the  northern  anchovy  Engraulis  mordax  (R.  Lasker, 
ed.i,  p.  79-94.     NOAA  Tech.  Rep.  NMFS  36. 

Lisovenko.  L.  A.,  and  D.  P.  Andrianov. 

1991.  Determination  of  absolute  fecundity  of  intermittently 
spawning  fishes.     Voprosy  ikhtiologii  31:631-641. 

Macchi,  G.  J.,  and  E.  M.  Acha. 

2000.     Spawning  frequency  and  batch  fecundity  of  Bra- 
zilian menhaden,  Brevoortia  aurea,  in  the  Rio  de  la 
Plata  estuary  off  Argentina  and  Uruguay.     Fish.  Bull. 
98:283-289. 
McMichael,  R.  H.,  and  K.  M.  Peters. 

1989.     Early  life-history  of  spotted  seatrout.  Cynoscion  neb- 
ulosus (Pisces:  Sciaenidae),  in  Tampa  Bay.  Florida.     Es- 
tuaries 12:98-110. 
Miller,  T  J..  T.  Herra,  and  W.  C.  Leggett  . 

1995.     An  individual-based  analysis  of  the  variability 
of  eggs  and  their  newly  hatched  larvae  of  Atlantic  cod 
(Gadus  morhua)  on  the  Scotian  Shelf.     Can.  J.  Fish. 
Aquat.  Sci.  52:1083-1093. 
Mok,  H.  K.,  and  R.  G.  Gilmore. 

1983.     Analysis  of  sound  production  in  estuarine  aggre- 
gations of  Pogonias  cromis,  Bairdiella  chrysoura,  and 
Cynoscion  nebulosus  (Sciaenidae  i.     Bull.  Inst.  Zool.. 
Acad.  Sinica  (Taipei)  22:157-186. 
Murphy,  M.  D..  and  R.  G.  Taylor. 

1994.     Age,  growth  and  mortality  of  spotted  seatrout  in 
Florida  waters.     Trans.  Am.  Fish.  Soc.  123:482-497. 
Nieland,  D.  L..  R.  G.  Thomas,  and  C.  A.  Wilson. 

2002.     Age.  growth  and  reproduction  of  spotted  seatrout 
in  Barataria  Bay,  Louisiana.     Trans.  Am.  Fish.  Soc. 
131:245-259. 
Overstreet,  R.  M. 

1983.     Aspects  of  the  biology  of  the  spotted  seatrout. 
Cynoscion  nebulosus,  in  Mississippi.     Gulf  Res.  Rep. 
Suppl.  1:1-43. 
Pearson,  J.  C. 

1929.     Natural  history  and  conservation  of  the  redfish  and 
other  commercial  sciaenids  on  the  Texas  coast.     Bull. 
U.S.  Bur.  Fish.  44:129-214. 
Prager,  M.  H..  J.  F  O'Brien,  and  S.  B.  Saila. 

1987.     Using  lifetime  fecundity  to  compare  management 
strategies:  a  case  history  for  striped  bass.     Am.  J.  Fish. 
Manag.  7:403-409. 
Ricker,  W.  E. 

1975.     Computation  and  interpretation  of  biological  sta- 
tistics offish  populations.     Bull.  Fish.  Res.  Board  Can. 
191:1-382. 
Saucier,  M.  H.,  and  D.  M.  Baltz. 

1993.     Spawning  site  selection  by  spotted  seatrout, 
Cynoscion  nebulosus.  and  black  drum,  Pogonias  cromis. 
in  Louisiana.     Environ.  Biol.  Fish.  36:257-272. 
Saucier,  M.  H.,  D.  M.  Baltz.  and  W.  A.  Roumillat 

1992.  Hydrophone  identification  of  spawning  sites  of 
spotted  seatrout  Cynoscion  nebulosus  (Osteiehthys:  Sci- 
aenidae l  near  Charleston,  South  Carolina.  N.  Gulf 
Sci.  12:14 1-145. 

Sundararaj,  B.  I.,  and  R.  D.  Suttkus. 

1962.     Fecundity  of  the  spotted  seatrout.  Cynoscion  nebulo- 


Roumillat  and  Brouwer:  Reproductive  dynamics  of  Cynosaon  nebulosus 


487 


sus  iCuvier),  from  Lake  Borgne  Area,  Louisiana.     Trans. 
Am.  Fish.  Soc.  91:84-88. 
Taylor,  R.  G.,  H.  J.  Grier,  and  J.  A.  Whittington. 

1998.     Spawning  rhythms  of  common  snook  in  Florida.     J. 
Fish  Biol.  53:502-520. 
Vaughan,  D.  S.,  G.  R.  Huntsman,  C.  S.  Manooch  III, 
F.  C.  Rohde,  and  G.  F.  Ulrich. 

1992.     Population  characteristics  of  the  red  porgy,  Pag- 
rus  pagrus,  stock  off  the  Carolinas.     Bull.  Mar.  Sci.  50: 
1-20. 
Wallace,  R.  A.,  and  K.  Selman. 

1981.     Cellular  and  dynamic  aspects  of  oocyte  growth  in 
teleosts.     Am.  Zool.  21:325-343. 
Ware,  D.  M. 

1975.  Relation  between  egg  size,  growth,  and  natural  mor- 
tality of  larval  fish.  J.  Fish.  Res.  Board  Can.  32: 
2503-512. 


Wieting,  D.  S. 

1989.     Age,  growth  and  fecundity  of  spotted  seatrout 
(Cynoscion  nebulosus)  in  Louisiana.     M.S.  thesis,  94 
p.  Louisiana  State  Univ..  Baton  Rouge,  LA. 
Wiley,  B.  A. 

1996.     Use  of  polymorphic  microsatellites  to  detect  popu- 
lation structure  in  spotted  seatrout,  Cynoscion  nebu- 
losus (Cuvier).     M.S.  thesis,  42  p.     Univ.  Charleston, 
Charleston,  SC. 
Wootton,  R.  J. 

1994.     Life  histories  as  sampling  devices:  optimum  egg 
size  in  pelagic  fishes.     J.  Fish  Biol.  45:1067-1077. 


488 


Abstract— Dungeness  crabs  {Cancer 
magister)  were  sampled  with  commer- 
cial pots  and  counted  by  scuba  divers 
on  benthic  transects  at  eight  sites 
near  Glacier  Bay,  Alaska.  Catch  per 
unit  of  effort  (CPUE)  from  pots  was 
compared  to  the  density  estimates 
from  dives  to  evaluate  the  bias  and 
power  of  the  two  techniques.  Yearly 
sampling  was  conducted  in  two  sea- 
sons: April  and  September,  from  1992 
to  2000.  Male  CPUE  estimates  from 
pots  were  significantly  lower  in  April 
than  in  the  following  September;  a 
step-wise  regression  demonstrated 
that  season  accounted  for  more  of 
the  variation  in  male  CPUE  than 
did  temperature.  In  both  April  and 
September,  pot  sampling  was  signifi- 
cantly biased  against  females.  When 
females  were  categorized  as  oviger- 
ous  and  nonovigerous,  it  was  clear 
that  ovigerous  females  accounted  for 
the  majority  of  the  bias  because  pots 
were  not  biased  against  nonovigerous 
females.  We  compared  the  power  of 
pots  and  dive  transects  in  detecting 
trends  in  populations  and  found  that 
pots  had  much  higher  power  than  dive 
transects.  Despite  their  low  power,  the 
dive  transects  were  very  useful  for 
detecting  bias  in  our  pot  sampling  and 
in  identifying  the  optimal  times  of 
year  to  sample  so  that  pot  bias  could 
be  avoided. 


Estimating  Dungeness  crab  (Cancer  magister) 
abundance:  crab  pots  and  dive  transects  compared 


S.  James  Taggart 

Glacier  Bay  Field  Station 

Alaska  Science  Center 

U.S.  Geological  Survey 

3100  National  Park  Rd. 

Juneau,  Alaska  99801 

E-mail  address   |im_taggart(S!usgsgov 


Charles  E.  O'Clair 

National  Marine  Fisheries  Service 
Auke  Bay  Laboratory 
11305  Glacier  Highway 
Juneau,  Alaska  99801 


Thomas  C.  Shirley 

Juneau  Center,  School  of  Fisheries  &  Ocean  Sciences 
University  of  Alaska  Fairbanks 
11120  Glacier  Highway 
Juneau,  Alaska  99801 


Jennifer  Mondragon 

Glacier  Bay  Field  Station 
Alaska  Science  Center 
U.S.  Geological  Survey 
3100  National  Park  Rd. 
Juneau,  AK  99801 


Manuscript  submitted  13  March  2000 
to  Scient  ific  Editor's  Office. 

Manuscript  approved  for  publication 
25  March  2004  by  the  Scientific  Kditor. 

Fish   Bull.  102:488-497  (2004) 


Reliable  population  assessments  are 
fundamental  to  the  management  and 
conservation  of  commercially  har- 
vested crabs.  Many  crab  populations 
are  sampled  with  commercial  crab 
pots  to  estimate  population  trends,  to 
set  harvest  quotas,  or  to  differentiate 
natural  population  fluctuations  caused 
by  anthropogenic  changes  to  the  eco- 
system. Pots  are  used,  for  example,  to 
assess  the  population  status  of  blue 
crabs,  Callinectes  sapidus  (Abbe  and 
Stagg,  1996),  red  king  crabs,  Para- 
lithodes  camtschaticus  (Zheng  et  al., 
1993),  snow  crabs,  Chionoecetes  opilio 
(Dawe  et  al.,  1996),  and  southern  king 
crabs,  Lithodes  santolla  i  Wyngaard 
and  Iorio.  1996). 

The  Dungeness  crab  (Cancer  ma- 
gister) fishery  began  in  southeastern 
Alaska  in  1916  and  has  been  charac- 
terized by  large  fluctuations  on  an- 


nual and  decadal  scales  tOrensanz 
et  al.,  1998).  Large  variation  in  the 
Dungeness  crab  harvest  is  not  unique 
to  Alaska;  similar  fluctuations  have 
been  documented  in  California  and 
their  causes  are  the  subject  of  an  on- 
going debate  (Higgins  et  al.,  1997a. 
1997b).  It  is  not  clear  whether  the 
processes  that  cause  fluctuations  in 
California  are  the  same  as  those  re- 
sponsible for  oscillations  in  Dunge- 
ness crab  abundance  in  Alaska. 

Most  of  the  Dungeness  crab  fisher- 
ies in  Alaska  are  managed  by  regu- 
lating the  size  and  sex  of  the  crabs 
caught,  and,  in  some  places,  the  sea- 
son of  the  harvest.  In  southeastern 
Alaska,  legal  harvest  is  restricted  to 
males  with  a  carapace  width  greater 
than  or  equal  to  165  mm  (excluding 
the  10th  anteriolateral  spines)  and  the 
season  is  timed  to  avoid  sensitive  life 


Taggart  et  al.:  Estimating  abundance  of  Cancer  magister 


489 


58°30'N. 


^s  20' 


t r 

135°50'  135  4n\\ 

Figure  1 

Map  of  study  area  showing  eight  study  sites  in  or  near  Glacier  Bay  in  southeastern 
Alaska. 


history  periods  such  as  mating  and  molting  (Kruse, 
1993;  Orensanz  et  al.,  1998).  Pre-  and  postseason  stock 
assessment  surveys  using  crab  pots  were  initiated  in 
southeastern  Alaska  in  2000  (Rumble  and  Bishop, 
2002).  The  purpose  of  the  latter  management  strategy 
is  to  assess  the  abundance  of  legal-size  males  before  the 
fishing  season,  to  estimate  harvest  rates,  to  define  the 
timing  of  male  and  female  mating  and  molting  and  to 
determine  growth  rate  by  tagging  crabs. 

The  usefulness  of  surveys  with  pots  for  Dungeness 
crab  population  assessment,  however,  depends  on  the 
accuracy  of  these  surveys  in  measuring  population  pa- 
rameters. Factors  that  can  bias  catch  per  unit  of  effort 
(CPUE)  and  size-frequency  estimates  for  Dungeness 
crabs  are  pot  soak-time  (Miller,  1974;  High,  1976;  Got- 
shall,  1978;  Smith  and  Jamieson,  1989);  freshness  of 
bait  (High,  1976;  Smith  and  Jamieson,  1989);  pot  design 
(Miller,  1974;  High,  1976;  Smith  and  Jamieson,  1989); 
and  agonistic  interactions  among  conspecifics  inside 
and  at  the  entrance  of  pots  (Caddy,  1979;  Smith  and 
Jamieson,  1989).  Smith  and  Jamieson  (1989)  developed 
a  standardized  model  to  compensate  for  the  effect  of 
agonistic  interactions,  age  of  bait,  and  escapement. 
They  also  concluded  that  researchers  could  minimize 
these  biases  by  measuring  CPUE  with  standardized 
surveys  with  short  soak  times.  These  studies  measured 
sampling  bias  with  pots  by  comparing  catch  in  pots 
among  various  experimental  treatments.  Opportunities 
for  comparing  surveys  with  pots  to  direct  measures  of 
abundance  are  rare.  In  our  study,  we  compared  the  bias 
and  power  of  CPUE  estimates  from  surveys  with  pots  to 


independent  measures  of  abundance  conducted  by  scuba 
divers  on  benthic  dive  transects. 


Methods 

Study  area 

The  study  area  included  eight  sites  in  southeastern 
Alaska,  near  Glacier  Bay:  North  Beardslee  Islands 
(58°33'N  135°54'W),  South  Beardslee  Islands  (58°30'N 
135°55'W),  Berg  Bay  (58°31'N  136°13'W),  Bartlett  Cove 
(58°27'N  135°53'W),  Gustavus  Flats  (58°23'N  135°43'W), 
Secret  Bay  (58°29'N  135°56'W),  inner  Dundas  Bay 
(58°27'N  136°31'W),  and  outer  Dundas  Bay  (58°21'N 
136°18'W)  (Fig.  1).  All  study  sites  were  located  within 
Glacier  Bay  National  Park  and  Preserve,  with  the  excep- 
tion of  Gustavus  Flats,  which  was  located  adjacent  to  the 
Park  boundary  in  Icy  Strait. 

Glacier  Bay  is  a  large  (1312  km2)  glacial  fjord  system 
with  high  sedimentation  rates  of  clay-silt  particles  from 
streams  and  tidewater  glaciers  (Cowan  et  al.,  1988). 
The  primarily  unconsolidated  rocky  coastline  is  highly 
convoluted  with  numerous  small  bays.  Dungeness  crabs 
can  be  found  throughout  Glacier  Bay;  however  the  ma- 
jority of  the  population  are  found  in  the  lower  40  km  of 
the  estuary  where  our  sites  were  located  (Taggart  et  al., 
2003 ).  The  shallow  water  in  and  around  our  study  sites 
was  primarily  characterized  by  mud  bottom,  but  sand, 
pebble,  cobble,  and  shell  substrates  were  also  common 
(Scheding  et  al.,  2001). 


490 


Fishery  Bulletin  102(3) 


Table  1 

Sampling  dates  for  yearly 
Sample  size  (n)  is  listed  for 

spring  and  fall  pot  and  dive  surveys  of  Dungeness  crabs  (Cancer  magister) 
pots  and  dives  for  each  sampling  event. 

in 

Glacier  Bay,  Al 

iska. 

Year 

Spring 

sampling 

Fall 

sampl 

ing 

Pots 

it 

Dives 

n 

Pots 

n 

Dives 

n 

1992 

7-12  April 

248 

7-12  April 

69 

17-22  Sept 

250 

17-22  Sep. 

75 

1993 

20-27  April 

350 

20-27  April 

105 

23-28  Sept 

250 

23-28  Sep. 

75 

1994 

20-27  April 

350 

23  April-1  May 

105 

13-18  Sept 

249 

13-18  Sept. 

75 

1995 

19-26  April 

350 

23  April-1  May 

105 

9-14  Sept 

236 

15-19  Sept. 

75 

1996 

15-21  April 

350 

22-28  April 

105 

13-18  Sept 

242 

19-23  Sept. 

73 

1997 

17-22  April 

300 

23-28  April 

115 

14-19  Sept 

298 

20-25  Sept. 

120 

1998 

— 

— 

— 

— 

9-14  Sept 

296 

16-21  Sept. 

91 

1999 

— 

— 

— 

— 

9-14  Sept 

299 

17-22  Sept. 

107 

2000 

— 

— 

— 

— 

9-14  Sept 

297 

18-23  Sept. 

60 

Sampling  dates 

Sampling  was  conducted  biannually,  in  April  and  Sep- 
tember, from  1992  to  1997  and  annually,  in  September, 
from  1998  to  2000  (Table  1).  The  spring  and  fall  sam- 
pling periods  were  selected  to  coincide  with  crab  life 
history  events  and  to  avoid  sampling  during  commercial 
fishery  operations.  April  sampling  was  scheduled  to 
occur  before  larval  hatching  in  May- June  (Shirley  et 
al.,  1987)  and  before  the  summer  commercial  fishing 
season  from  15  June  to  15  August.  September  sampling 
began  after  the  end  of  the  fishing  season  (15  August) 
and  ended  before  the  beginning  of  the  winter  harvest 
(1  October  to  30  November). 

During  1992,  the  study  sites  were  sampled  with  pots 
(referred  to  as  "pot  sampling")  and  by  divers  (referred  to 
as  "dive  sampling")  concurrently  (Table  1).  In  1993  and 
1994,  sampling  was  conducted  on  nearby  study  sites 
and  the  dive  sampling  usually  one  day  ahead  of  the 
pot  sampling.  For  logistical  reasons,  starting  in  1995. 
we  separated  the  pot  sampling  and  the  dive-transect 
sampling  into  two  separate  research  cruises.  The  pot 
sampling  was  conducted  on  the  first  cruise  and  the 
dive  sampling  occurred  on  the  second  cruise;  pot  and 
dive  sampling  were  separated  at  each  location  by  2  to 
12  days. 

Sampling  with  pots 

Crabs  were  sampled  with  commercial  crab  pots  (0.91  m 
in  diameter,  0.36  m  tall,  with  5-cm  wire  mesh).  Escape 
rings  were  sealed  with  webbing  on  each  pot  to  retain 
smaller  crabs.  Pots  were  baited  with  hanging  bait  com- 
prising salmon,  cod,  or  halibut  (depending  on  availabil- 
il  nd  bait  jars  that  were  filled  with  chopped  herring 
and  squid.  We  found  that  cod  was  predictably  available; 
therefore  from  1996  on,  we  consistently  used  cod  for 
hanging  bait.  Pots  were  soaked  for  24  hours. 


Within  each  study  site,  we  set  25  pots  in  shallow 
water  (0-9  m)  and  25  pots  in  deep  water  (10-25  m). 
Each  day  we  set  50  pots  in  one  of  the  study  sites  and 
retrieved  the  50  pots  that  had  been  set  the  previous  day 
at  one  of  the  other  study  sites.  The  pots  were  set  along 
strings  parallel  to  shore  at  intervals  of  approximately 
100  m.  Within  each  study  area,  the  strings  of  pots  were 
located  in  prime  Dungeness  crab  habitat  determined 
by  a  local  fisherman.  We  placed  the  pots  at  the  same 
locations  during  subsequent  sampling  events  by  using 
a  GPS  (Rockwell  PLGR+)  with  an  accuracy  of  ±3  m.  We 
estimate  that  the  pots  were  set  within  20  meters  from 
the  original  waypoints.  Water  depth  (standardized  to 
mean  lower  low  water),  set  and  retrieval  time,  and  GPS 
location  were  recorded  for  each  pot.  Water  temperature 
and  salinity  profiles  were  measured  at  each  study  site 
during  each  sampling  period  with  a  SEABIRD  SBE-19 
Profiler. 

As  the  pots  were  retrieved,  we  counted  and  identified 
all  organisms.  For  all  Dungeness  crabs  we  recorded  the 
sex,  carapace  width,  shell  condition,  and  damage  to 
appendages.  For  female  crabs  we  also  recorded  repro- 
ductive status.  Carapace  width  was  measured  to  the 
nearest  millimeter  immediately  anterior  to  the  10th 
anterolateral  spine  with  vernier  calipers  (Shirley  and 
Shirley,  1988;  Shirley  et  al..  1996).  All  organisms  were 
returned  to  the  water  at  the  location  where  they  were 
caught.  A  potential  problem  with  returning  the  crabs 
to  the  water  near  the  site  of  capture  is  the  possibility 
that  crabs  could  be  resampled  in  subsequent  pots,  which 
would  bias  the  catch  per  unit  of  effort.  Beginning  in 
April,  1995,  all  crabs  collected  in  the  South  Beardslee 
Islands  and  Berg  Bay  were  tagged  with  a  sequentially 
numbered,  double-T  Floy  tag  (Floy  Tag  and  Manufactur- 
ing Company,  Seattle,  WA)  inserted  along  the  postero- 
lateral margin  of  the  epimeral  suture.  Tags  placed  in 
this  location  are  retained  through  ecdysis  (Smith  and 
Jamieson,  1989).  Of  the  5226  crabs  tagged,  only  a  single 


Taggart  et  al.:  Estimating  abundance  of  Cancer  magister 


491 


crab  was  recovered  during  the  same  sampling  event. 
Thus,  the  probability  of  resampling  crabs  by  returning 
them  to  the  water  was  very  low. 

Sampling  by  divers 

Divers  using  scuba  equipment  censused  crabs  on  15 
to  20,  2x100  m  belt  transects  within  each  study  site. 
Approximately  one  day  of  sampling  was  required  at 
each  study  site.  The  dive  transects  were  conducted  per- 
pendicular to  the  shoreline  and  they  extended  from  the 
shallow  subtidal  (0  m,  mean  lower  low  water)  to  18  m 
depth  or  to  the  end  of  the  100  m  transect,  whichever 
came  first.  Divers  did  not  go  below  18  m  depth  in  an 
effort  to  reduce  nitrogen  accumulation  in  divers'  blood 
and  to  reduce  the  surface  intervals  required  between 
transects.  From  1992  to  1997,  transect  locations  were 
randomly  selected  in  the  same  areas  as  the  crab-pot 
sampling.  The  random  locations  selected  in  1997  were 
resampled  during  the  following  years  of  the  study. 

Divers  counted  all  Dungeness  crabs  located  within 
1  m  of  each  side  of  the  transect.  An  effort  was  made  to 
locate  buried  crabs  by  swimming  close  to  the  bottom 
and  looking  for  irregularities  in  the  bottom  or  protrud- 
ing crab  eyestalks.  Each  crab  was  examined  and  the 
following  were  recorded:  legal  males  >165  mm  carapace 
width),  sublegal  males  (<165  mm  carapace  width),  ovig- 
enous females,  and  nonovigerous  females. 

Data  analysis 

For  each  year,  we  calculated  the  average  pot  CPUE  for 
each  site  by  reproductive  class  (males,  nonovigerous 
females,  and  ovigerous  females).  The  number  of  pots 
sometimes  deviated  from  50  when  a  pot  was  lost  or  when 
the  degradable  cotton  string  securing  the  pot  lid  broke 
(range:  44-50  pots).  The  number  of  crabs  counted  on 
dive  transects  was  averaged  for  each  reproductive  class 
by  site  for  each  year.  All  dive  transects  were  conducted 
perpendicular  to  shore;  thus  the  transects  crossed  the 
shallow  habitat  where  the  shallow  string  of  pots  was 
set  and  terminated  at  18  m  which  was  the  center  of 
the  depth  we  targeted  for  the  deep  pot  set.  Because  the 
deep  pot  set  was  at  or  slightly  beyond  the  deep  end  of 
the  transect,  we  may  have  sampled  more  crabs  from 
deepwater  habitats  than  from  the  shallower  transects. 
However,  we  did  not  think  this  was  a  significant  bias 
because  we  sampled  crabs  from  a  relatively  large  area. 
We,  therefore,  pooled  the  pots  from  both  depth  strata 
for  analysis. 

We  tested  for  differences  between  April  and  Septem- 
ber for  the  pot  CPUE  data  and  the  dive  density  data 
with  paired  f-tests.  CPUE  and  density  data  were  not 
normally  distributed;  therefore  we  transformed  the 
data  with  a  square-root  transformation  [Y=JiY  +  3/8)] 
for  statistical  analyses  (Zar,  1996).  These  analyses  were 
conducted  for  males,  nonovigerous  females,  and  oviger- 
ous females.  Because  seasonal  increases  in  water  tem- 
perature could  drive  differences  in  CPUE  between  April 
and  September,  we  calculated  mean  water  temperatures 


by  averaging  the  water  temperatures  at  the  5  m  and 
15  m  depths  at  each  site  and  year.  This  analysis  was 
limited  to  years  and  sites  where  we  collected  samples 
in  both  April  and  September  (1992-97,  from  five  sites: 
North  Beardslee  Islands,  South  Beardslee  Islands,  Berg 
Bay,  Bartlett  Cove,  and  Gustavus  Flats).  We  assessed 
how  CPUE  was  influenced  by  two  independent  vari- 
ables, water  temperature  and  season,  with  stepwise 
regression.  Because  CPUE  declined  from  1992  to  1997 
(Taggart  et  al.,  in  press),  we  controlled  for  year  so  that 
it  would  not  confound  our  analysis. 

In  order  to  assess  sampling  bias  between  pots  and  dive 
transects,  the  percentages  of  females  (females/all  crabs), 
nonovigerous  females  (nonovigerous  females/all  crabs), 
and  ovigerous  females  (ovigerous  females/all  crabs)  were 
calculated  for  each  site  and  sampling  time.  We  also  com- 
pared the  percentage  of  the  male  population  that  was 
legal  size  (legal-size  male  crabs/all  male  crabs)  from  the 
pots  and  from  the  dives.  The  percentage  estimates  from 
the  pot  data  were  compared  to  estimates  from  the  dive 
transects  with  a  paired  sign  test  (Zar,  1996).  If  percent- 
age estimates  for  pot  data  were  unbiased  when  compared 
to  estimates  from  dive  data,  the  pot  percentage  esti- 
mates would  have  an  equal  chance  of  being  higher  or 
lower  than  the  percentage  estimates  for  the  dive  data. 
Because  small  sample  sizes  exaggerate  percentage  com- 
parisons, we  excluded  samples  where  the  total  number 
of  crabs  collected  was  less  than  25  crabs/site. 

The  power  of  pots  and  dive  transects  to  detect  trends 
in  populations  was  compared  with  Monitor,  a  power 
analysis  program  (Gibbs  and  Melvin,  1997;  Gibbs, 
1998).  For  our  analyses,  we  varied  the  number  of  tran- 
sects and  pots,  compared  males  and  nonovigerous  fe- 
males, and  varied  the  duration  of  the  study.  For  all 
analyses  the  following  input  parameters  of  the  model 
were  held  constant:  "survey  occasions"  =  annual,  "type" 
=  linear,  "significance  level"  =  0.05,  "number  of  tails"  = 
2,  "constant  added"  =  1,  "trend  variation"  =  0,  "round- 
ing" =  decimal,  "trend  coverage"  =  complete,  and  "rep- 
lications" =  10,000. 

To  estimate  power,  the  model  requires  "count"  and 
"variance"  for  each  plot  across  years  for  at  least  three 
years.  Pot  and  transect  data  collected  from  1992  to 
1998  from  five  sites  (North  Beardslee  Islands,  South 
Beardslee  Islands,  Berg  Bay,  Bartlett  Cove,  and  Gusta- 
vus Flats)  were  used  for  these  analyses.  The  data  were 
limited  to  September  to  avoid  seasonal  bias.  The  aver- 
age across  years  was  calculated  for  each  transect  and 
each  pot.  These  averages  were  input  into  the  model's 
variable  called  "plot  count."  For  each  pot  and  transect 
a  linear  regression  was  calculated  among  years  (CPUE 
vs.  year  for  pots;  density  vs.  year  for  dive  transects) 
and  the  residual  mean  square  was  the  "plot  variance" 
variable  (Thomas  and  Krebs,  1997). 

To  estimate  the  effect  of  sample  size  on  power  we  set 
the  "number  [surveys]  conducted"  to  four  and  limited 
the  analysis  to  males.  We  varied  the  number  of  "plots" 
(pots  and  transects).  For  pots,  we  randomly  selected 
subsamples  of  the  250  pots  and  ran  simulations  from 
25  pots  to  250  pots  in  25-pot  increments.  The  number 


492 


Fishery  Bulletin  102(3) 


of  dive  transects  for  which  data  were  collected  for  mul- 
tiple years  was  75.  For  simulations  with  a  sample  size 
less  than  75,  we  randomly  subsampled  the  data  in  the 
same  manner  as  we  did  with  pots.  For  simulations  with 
sample  sizes  greater  than  75,  we  amplified  the  samples 
with  simple  bootstrapping  to  obtain  samples  from  100  to 
250  transects  in  25-transect  increments  (Wonnacott  and 
Wonnacott,  1990).  For  each  sample  size,  we  modeled 
three  annual  rates  of  change  (0.02,  0.03,  and  0.05). 

To  evaluate  how  study  duration  affects  power,  we  lim- 
ited the  analysis  to  males,  varied  study  duration  ("num- 
ber [surveys]  conducted")  from  two  years  to  12  years  in 
two-year  increments,  and  compared  three  annual  rates 
of  change  (0.02,  0.03,  and  0.05)  for  both  pots  and  tran- 
sects. To  hold  effort  constant  between  the  two  sampling 
techniques,  we  set  the  pot  and  transect  sample  size  to 
the  number  we  could  accomplish  in  a  five-day  research 
cruise  (250  pots  and  75  transects). 

To  explore  the  relationship  between  annual  trend  in 
population  and  power,  we  held  effort  constant  (250  pots 
and  75  transects)  and  varied  the  annual  trend  (from 
-0.10  to  +0.10  in  0.01  increments)  for  both  males  and 
nonovigerous  females.  It  was  not  possible  to  conduct  a 
power  analysis  for  ovigerous  females  because  a  large 
proportion  of  the  pots  and  transects  had  no  ovigerous 
female  crabs. 


Results 

The  pot  CPUE  estimates  for  males,  nonovigerous  females, 
and  ovigerous  females  was  significantly  different  in 
April  than  in  the  following  September  (Fig.  2,  A,  C,  and 
E).  Male  and  nonovigerous  female  CPUE  was  higher  in 
September  (Fig.  2,  A  and  C)  and  ovigerous  female  CPUE 
was  lower  in  September  (Fig.  2E).  In  contrast,  April  den- 
sity estimates  from  dive  transects  were  not  significantly 
different  from  the  following  September  density  estimates 
for  males  (Fig.  2B).  Dive  density  estimates  for  nonovig- 
erous females  were  higher  in  September  than  in  April 
(Fig.  2D);  density  estimates  for  ovigerous  females  were 
lower  in  September  than  in  April  (Fig.  2F). 

When  we  tested  the  influence  of  temperature  and 
season  on  male  CPUE  with  stepwise  regression,  season 
was  selected  first;  temperature  was  not  selected  because 
it  did  not  have  a  significant  additional  effect  (Table  2). 
Because  no  significant  difference  was  found  between 
the  April  and  September  density  estimates  from  dive 
transects  (Fig.  2B),  we  did  not  conduct  a  stepwise  re- 
gression for  the  dive  data. 

Percentage  estimates  of  females  from  sampling  with 
pots  were  lower  than  percentage  estimates  from  dive 
transects  for  a  significant  number  of  samples  for  both 
April  and  September  (Fig.  3Ai;  therefore  pots  were  bi- 
ased against  sampling  females.  When  females  were 
split  by  reproductive  status,  no  bias  was  detected  for 
sampling  nonovigerous  females  with  pots  (Fig.  3B). 
In  contrast,  the  percentage  estimates  for  ovigerous  fe- 
males remained  biased  and  the  magnitude  of  the  bias 
increased  (Fig.  3C).  To  test  potential  sampling  bias 


co    2.5  • 


3.25 


2.75  ■ 


2.25 


O     1.5  • 


1.75 


2.75 


2.25 


P=0.40 


P=0.04 


0.75 


1.75 
3.75 

2.75  \ 


2.75 


P=0.04 


April 


September 


April 


September 


Figure  2 

Within-year  paired  comparisons  by  site  of  catch  in 
pots  (left  column)  and  density  on  dive  transects  (right 
column)  for:  (A  and  B)  male  Dungeness  crabs  {Cancer 
magister);  (C  and  D)  nonovigerous  female  crabs;  and 
(E  and  F)  ovigerous  female  crabs.  Catch  and  density 
data  were  transformed  with  a  square-root  transforma- 
tion. P-values  indicate  results  from  paired  /-tests  and 
significant  results  show  differences  between  April  and 
September.  Lines  on  the  graphs  are  parallel  if  measure- 
ments at  sites  were  consistently  higher  or  lower  in  April 
and  September. 


related  to  crab  size,  we  compared  the  proportion  of 
the  male  population  that  was  legal  size  sampled  with 
pots  and  dives  (Fig.  4).  There  was  no  significant  bias 
when  pots  and  transects  were  compared  with  a  sign 
test  (April.  P>0.999;  September,  P=0.06). 

CPUE  estimates  from  pots  had  a  higher  power  than 
density  estimates  from  dive  transects  for  the  same 
sample  size  (Fig.  5).  Because  more  time  is  required 
to  conduct  a  dive  transect  than  to  set  and  pull  a  crab 
pot,  the  power  of  transects  compared  to  pots  was  even 
lower  when  effort  was  incorporated  into  the  analysis 
(Fig  6).  The  power  can  be  increased  for  both  pots  and 


Taggart  et  al.:  Estimating  abundance  of  Cancer  magister 


493 


Table  2 

Stepwise  regression  results  of  CPUE  (male  crabs/pot) 
versus  three  independent  variables  (year,  season,  and 
temperature). 

Step 

Model 
parameters                r2 

P-value 
( parameter  1 

1 
2 
3 

Year                            0.1493 
Year  and  month        0.5589 
Year,  month.             0.5589 
and  temperature 

0.001  (year) 

0.04  (month I 

0.98  (temperature) 

transects  by  increasing  the  study  duration  or  increasing 
the  amount  of  change  in  the  population  that  the  study 
is  attempting  to  detect  (Fig.  6).  Although  pots  had  more 
power  than  dive  transects,  there  was  only  slightly  more 
power  to  detect  change  in  abundance  of  male  crabs 
versus  nonovigerous  females  (Fig.  7). 


Discussion 

For  male  Dungeness  crabs,  the  density  estimates  from 
the  dive  transects  showed  no  difference  between  April 
and  September  (Fig.  2B).  The  male  CPUE  estimates 
from  pots,  however,  were  consistently  lower  in  April  than 
in  the  following  September  (Fig.  2A).  Because  feeding 
rates  of  Dungeness  crabs  are  correlated  with  tempera- 
ture (Kondzela  and  Shirley,  1993),  we  thought  that  tem- 
perature was  likely  to  explain  the  differences  in  CPUE 
between  April  and  September.  We  found,  however,  that 
season  had  a  larger  effect  than  temperature  (Table  2). 
This  result  suggests  that  seasonal  factors  other  than 
temperature  influence  catchability.  Stone  and  O'Clair 
(2001)  followed  the  seasonal  movements  of  Dungeness 
crabs  in  a  glacial  estuary  in  southeastern  Alaska  and 
reported  that  mean  movement  of  male  crabs  was  lower 
during  the  spring  than  in  the  late  summer  and  fall.  It 
is  possible  that  our  spring  sampling  schedule  coincided 
with  low  male  activity  and  male  crabs  were  less  likely  to 
encounter  a  bait  plume  and  be  attracted  to  a  pot.  These 
results  indicate  that  if  pots  are  used  for  sampling,  late 
summer  and  early  fall  is  the  time  of  year  to  conduct 
population  assessment  surveys  of  male  crabs.  Similar 
seasonal  differences  in  CPUE  have  also  been  described 
for  edible  crabs  (Cancer  pagurus)  and  American  lobsters 
(Homarus  americanus)  (Bennett,  1974).  These  data  dem- 
onstrate the  importance  of  controlling  for  season  when 
comparing  CPUE  among  years  or  sites. 

The  proportion  of  large  crabs  caught  in  pots  increased 
with  longer  soak  time  for  Dungeness  crabs  in  British 
Columbia  (Smith  and  Jamieson,  1989)  and  red  king 
crabs  in  Britstol  Bay,  Alaska  (Pengilly  and  Tracy, 
1998).  We  found  no  bias  when  we  measured  the  legal- 
size  proportion  of  the  male  population  caught  in  pots 
and  compared  it  to  the  proportion  sampled  on  dives 


1 

0.8 -I 

0.6 

0.4 -I 

0.2 

0 
1 

0.8 


i    A 

April  P<0.0001           / 

Sept.  P<0.0001     / 

/        @ 

O     '         A 

/    °         O 

»/i*AA 

/O   n    A     ~ 

/  a4o        o 

A 

<3$»# 

/                                   u 

' 

1 1 1           i 

0.4- 
0.2- 


B 


April  P=0.17 
Sept.  P=0.49 


'  A 


0.6  9     o 


o 


i>4. 


O       "    Oft    /A 
*«0«'     *A 


0°A^ 

a 


*    A 

A     AA 


o 


O  April 

A  September 


0- 

i-i  c 


0.8 
0.6 
0.4 
0.2 


0* 


April  P<0.0001 
Sept.  P<0.0001 


O 


o 


cr 
o 


0         0.2       0.4      0.6        0.8         1 
Percent  crabs  on  transects 

Figure  3 

The  percentage  of  (A)  female  Dungeness  crabs 
(C.  magister).  (B)  nonovigerous  female  crabs, 
and  (Ci  ovigerous  female  crabs  estimated  from 
pots  and  from  dive  transects.  The  dashed  line 
in  each  graph  has  a  slope  of  1:  thus  half  of  the 
data  points  should  be  above  and  half  should  be 
below  the  dashed  line  if  percentage  estimates 
for  dives  and  pots  are  unbiased.  Pot  and  dive 
transect  data  for  each  sex  class  and  season 
were  compared  with  a  paired  sign  test  and 
P-values  are  reported. 


(Fig.  4).  We  expect,  however,  that  the  bias  observed  in 
British  Columbia  and  Bristol  Bay  would  occur  for  our 
study  sites  if  the  soak  time  of  pots  were  increased. 


494 


Fishery  Bulletin  102(3) 


O  April 

A  September 


(P>0.999) 
(P=0.06) 


1.0 
0.8 
0.6 


CD     CD 

Q-   E 


•     0A\f 


i, 
02OA 


O 


A. 


°^° 


O 

o 

A 


O 


A 


'    C*°J 


0D 


0.2 


0.4       0.6       0.1 


1.0 


Percent  legal-size 
male  crabs  on  transects 

Figure  4 

The  percentage  of  male  crabs  of  the 
Dungeness  crab  (C.  magister)  popu- 
lation that  were  legal  size  (>165  mm) 
estimated  from  pots  compared  to  the 
percentage  of  male  crabs  estimated 
from  dive  transects.  Data  from  pots 
and  dive  transects  were  compared  for 
each  season  with  a  paired  sign  test 
and  P-values  are  reported. 


0.8 


06 


g 
o 


0.2 


Pots    Transects 
-o-  •*■  0  05  change/year 


003  change/year 


50 


100  150 

Sample  size  (n) 


250 


Figure  5 

Relationship  between  power  and  sample  size  (/!)  in  comparing 
catch  from  pots  and  density  on  dive  transects  for  male  Dungeness 
crabs  (C.  magister)  at  three  levels  of  population  change. 


In  both  April  and  September,  pot  sampling  was  signif- 
icantly biased  against  females  (Fig.  3A).  When  females 
were  categorized  as  ovigerous  and  nonovigerous,  it  was 
clear  that  ovigerous  females  accounted  for  the  major- 
ity of  the  bias  because  pots  were  not  biased  against 
nonovigerous  females  (Fig.  3B).  Similar  results  have 
been  found  for  a  closely  related  species,  Cancer  pagurus; 
female  C.  pagurus  readily  enter  pots  when  they  are  in 
a  nonovigerous  reproductive  state  but  are  rarely  cap- 
tured when  they  are  ovigerous  (Bennett,  1995).  Move- 
ment studies  of  Dungeness  crabs  tagged  with  sonic 
transmitters  have  demonstrated  that  ovigerous  females 
move  less  frequently  and  move  slower  than  males  or 
nonovigerous  females  (O'Clair  et  al.,  1990).  Thus,  one 
explanation  for  the  bias  against  ovigerous  female  crabs 
is  that  their  restricted  movements  make  it  less  likely 
they  will  be  able  to  locate  and  become  entrapped  in 
pots.  In  addition  to  being  less  mobile,  ovigerous  females 
may  be  less  attracted  to  bait  than  nonovigerous  crabs. 
In  controlled  feeding  experiments,  ovigerous  females 
had  lower  feeding  rates  than  nonovigerous  females,  and 
ovigerous  females  took  longer  to  begin  feeding  (Schultz 
et  al.,  1996;  Schultz  and  Shirley,  1997  i.  Therefore,  ovig- 
erous females  may  be  less  responsive  to  the  bait  plume 
from  a  pot. 

The  estimate  of  nonovigerous  females  from  both  pot 
CPUE  and  dive  transect  density  increased  from  April 
to  September  (Fig.  2,  C  and  D).  As  with  males,  the  in- 
crease in  CPUE  for  nonovigerous  females  may  be  partly 
due  to  an  increase  in  catchability  in  September.  How- 


ever, the  fact  that  the  density  estimates  from  dives  also 
increased  suggests  that  the  number  of  nonovigerous 
females  actually  increased  between  April  and  Septem- 
ber. This  explanation  is  supported  by  the  decrease  in 
ovigerous  crabs  from  April  to  September  for  both  CPUE 
(Fig.  2E)  and  density  estimates  (Fig.  2F i. 

The  low  catchability  of  ovigerous  females  makes  it 
problematic  to  monitor  relative  abundance  of  females 
or  changes  in  sex  ratio  through  time.  However,  be- 
cause pots  were  not  biased  against  nonovigerous  fe- 
males (Fig.  3),  the  solution  may  be  to  estimate  the  rela- 
tive abundance  of  females  by  sampling  after  females 
hatch  their  eggs  and  before  they  extrude  a  new  clutch 
of  eggs  in  the  fall.  In  southeastern  Alaska,  most  females 
are  nonovigerous  in  late  July  and  early  August  (Stone 
and  O'Clair,  2001;  Swiney  et  al.,  2003);  therefore  this 
would  be  the  optimal  time  of  year  to  sample  females 
or  to  measure  sex  ratio  of  Dungeness  crab  populations. 
Unfortunately,  this  timing  coincides  with  the  summer 
commercial  fishing  season,  which  could  bias  sampling 
if  there  was  "competition"  between  survey  pots  and 
commercial  pots. 

For  both  males  and  females,  the  power  analyses  of 
the  pot  and  dive  data  indicated  that  for  most  population 
assessment  applications  it  would  be  extremely  difficult 
to  conduct  enough  dive  transects  to  obtain  sufficient 
statistical  power.  Even  if  it  were  possible  to  conduct 
as  many  dive  transects  as  pot  samples,  the  power  of 
a  dive  transect  was  still  lower  than  that  of  a  pot;  the 
higher  power  of  the  pots  was  due  to  lower  variance 
among  pots.  Pots  work  by  attracting  crabs  with  a  bait 
plume;  thus  the  area  and  number  of  crabs  sampled  is 


Taggart  et  al.:  Estimating  abundance  of  Cancer  magister 


495 


0  6- 


04- 


0.2 


Pots  Transects 

(n=250)  (n=75) 

-O-  -•-       0  05  change/year 

-O  ♦      0  03  change/year 

•V-  -T-       0  02  change/year 


0  2  4  6  8  10 

Study  duration  (years) 


12 


Figure  6 

Relationship  between  power  and  study  duration  in  comparing  catch  from 
crab  pots  and  density  on  dive  transects  for  male  Dungeness  crabs  (C. 
magister)  at  three  levels  of  population  change.  To  hold  effort  constant, 
we  set  the  sample  size  (n )  to  the  number  of  pots  and  dives  that  could 
be  accomplished  in  five  days. 


1 1 

0.8- 

0.6- 

53 

s 

o 
Q_ 

04- 

\        \\            II       ff        ^ots    Transects 

0.2- 

\      \\           //     J         ("=250)    (n=75) 

>.  »    3  y         ~D~    "*"  Males 

w^\   A?f                    -^-       •+.    Nonovigerous  females 

-.C 

12  -0.1 

-0  08-0.06-0  04-0  02      0      0  02   0.04    0.06   0.08     0  1     0.12 

Trand  (change/year) 

Figure  7 

Relationship 

between  power  and  trend  in  population  in  comparing  catch 

in  crab  pots 

and  density  on  dive  transects  for  male  and  nonovigerous 

female  Dung 

eness  crabs  iC.  magister). 

larger  with  pots  than  with  transects  and  the  variance 
with  pots  is  lower. 

Despite  their  low  power,  the  independent  measures 
of  abundance  provided  by  dives  helped  us  identify  bias 
in  our  Dungeness  crab  survey  method.  Our  analysis  of 
these  two  techniques  demonstrates  that  it  is  possible 


to  avoid  most  biases  with  pots  if  sampling  is  conducted 
at  optimal  times  of  year.  Similar  comparisons  could  be 
conducted  in  other  areas  to  identify  sampling  biases  so 
that  they  could  be  minimized  and  important  param- 
eters, such  as  abundance,  size,  and  sex  ratio,  could  be 
monitored  effectively. 


496 


Fishery  Bulletin  102(3) 


Acknowledgments 

This  long-term  study  was  made  possible  by  the  support  of 
a  large  number  of  people.  J.  de  La  Bruere  made  the  field 
work  efficient  and  enjoyable  through  his  expert  ability 
to  operate  the  RV  Alaskan  Gyre.  We  thank  A.  Andrews 
for  large  efforts  during  the  field  work,  data  manage- 
ment, and  analysis.  G.  Bishop,  C.  Dezan,  E.  Hooge,  P. 
Hooge,  E.  Leder,  J.  Luthy,  J.  Nielsen,  C.  Schroth,  D. 
Schultz,  L.  Solomon,  and  K.  Swiney  each  participated 
in  the  project  for  several  years.  The  manuscript  was 
improved  by  comments  from  E.  Mathews,  E.  Knudsen, 
and  three  anonymous  reviewers.  We  thank  M.  Jensen, 
J.  Brady,  T.  Lee,  M.  Moss,  and  S.  Rice  for  their  contin- 
ued support.  We  especially  thank  the  large  number  of 
unnamed  graduate  students,  faculty,  state  and  federal 
agency  researchers — over  70  people  total — who  gener- 
ously donated  their  time  and  efforts  to  this  long-term 
project.  This  project  was  funded  by  the  United  States 
Geological  Survey  and  the  National  Park  Service. 


Literature  cited 

Abbe,  G.  R„  and  C.  Stagg. 

1996.     Trends  in  blue  crab  {Callinectes  sapidus  Rathbun) 
catches  near  Calvert  Cliffs.  Maryland,  from  1968  to 
1995  and  their  relationship  to  the  Maryland  commercial 
fishery.     J.  Shellfish  Res.  15:751-758. 
Bennett,  D.  B. 

1974  The  effects  of  pot  immersion  time  on  catches  of  crabs. 
Cancer  pagurus  L.  and  lobster,  Homarus  gammarus  (L.). 
J.  Cons.  Int.  Explor.  Mer  35:332-336. 

1995.  Factors  in  the  life  history  of  the  edible  crab  [Cancer 
pagurus  L.)  that  influence  modelling  and  manage- 
ment.    ICES  Mar.  Sci.  Symp.  199:89-98. 

Caddy.  J.  F. 

1979.  Some  considerations  underlying  definitions  of  catch- 
ability  and  fishing  effort  in  shellfish  fisheries,  and  their 
relevance  for  stock  assessment  purposes.  Manuscript 
Report,  1489,  1-18  p.  Department  of  Fisheries  and 
Oceans  Canada,  Halifax,  Nova  Scotia,  Canada. 

Cowan.  E.  A.,  R.  D.  Powell,  and  N.  D.  Smith. 

1988.  Rainstorm-induced  event  sedimentation  at  the 
tidewater  front  of  a  temperate  glacier.  Geology  16: 
409-412. 

Dawe,  E.  G.,  D.  M.  Taylor,  and  J.  M.  Hoenig. 

1996.  Evaluating  an  index  of  snow  crab  (Chionocetes 
opilio)  biomass  from  trapping  surveys.  In  Proceedings 
of  the  international  symposium  on  biology,  management, 
and  economics  of  crabs  from  high  latitude  habitats. 
Anchorage,  Alaska,  October  11-13,  1995,  vol.  96-02, 
p.  301-314.  Alaska  Sea  Grant  College  Program,  Univ. 
Alaska,  Anchorage,  AK. 

Gibbs,  J.P. 

1998.     Monitoring  populations  of  plants  and  animals.     Bio- 
Science  48:935-940. 
Gibbs,  J.  P.,  and  S.  M.  Melvin. 

1997.  Power  to  detect  trends  in  water  bird  abundance 
with  call-response  surveys.  J.  Wildl.  Manag.  61(4): 
1262-1267. 

Gotshall,  D.  W. 

1978.     Catch-per-unit-of-effort  studies  of  northern  Cali- 


fornia Dungeness  crabs.  Cancer  magister.     Calif.  Fish 
Game  64(31:189-199. 

Higgins,  K.,  A.  Hastings,  and  L.  W.  Botsford. 

1997a.  Density  dependence  and  age  structure:  non- 
linear dynamics  and  population  behavior.  Am.  Nat. 
149(21:247-269. 

Higgins,  K.,  A.  Hastings,  J.  N.  Sarvela,  and  L.  W.  Botsford. 

1997b.  Stochastic  dynamics  and  deterministic  skele- 
tons: population  behavior  of  Dungeness  crab.  Science 
276(53171:1431-1435. 

High,  W.  L. 

1976.  Escape  of  Dungeness  crabs  from  pots.  Mar.  Fish. 
Rev.  38(41:19-23. 

Kondzela,  C.  M.,  and  T  C.  Shirley. 

1993.  Survival,  feeding,  and  growth  of  juvenile  Dunge- 
ness crabs  from  southeastern  Alaska  reared  at  different 
temperatures.     J.  Crustacean  Biol.  13(11:25-35. 

Kruse,  G.  H. 

1993.  Biological  perspectives  on  crab  management  in 
Alaska.  In  Proceedings  of  the  international  symposium 
on  management  strategies  for  exploited  fish  populations 
(G.  Kruse,  D.  M.  Eggers.  R.  J.  Marasco,  C.  Pautzke,  and 
T  J.  Quinn  II,  eds.  i,  355-384  p.  Lowell  Wakefield  Fish- 
eries Symposium.  Alaska  Sea  Grant  College  Program 
Report  93-02,  Univ.  Alaska,  Fairbanks,  AK. 

Miller,  R.  J. 

1974.  Saturation  of  crab  traps:  reduced  entry  and  es- 
capement.    J.  Cons.  Int.  Explor.  Mer  38(31:338-345. 

O'Clair.  C.  E.,  R.  P.  Stone,  and  J.  L.  Freese. 

1990.  Movements  and  habitat  use  of  Dungeness  crabs 
and  the  Glacier  Bay  fishery.  In  Second  Glacier  Bay 
science  symposium.  Glacier  Bay  National  Park  &  Pre- 
serve, AK,  Sept.  19-22,  1988  (A.  M.  Milner  and  J.  D. 
Wood  Jr.,  eds.  I,  74-77  p.  U.S.  National  Park  Service, 
Anchorage,  AK. 

Orensanz,  J.  M.,  J.  Armstrong,  D.  Armstrong,  and  R.  Hilborn. 
1998.     Crustacean  resources  are  vulnerable  to  serial 
depletion — the  multifaceted  decline  of  crab  and  shrimp 
fisheries  in  the  Greater  Gulf  of  Alaska.     Rev.  Fish  Biol. 
Fish.  8:117-176. 

Pengilly.  D„  and  D.  Tracy. 

1998.  Experimental  effects  of  soak  time  on  catch  of  legal- 
sized  and  nonlegal  red  king  crab  by  commercial  king 
crab  pots.     Alaska  Fish.  Res.  Bull.  5(2):81-87. 

Rumble.  J.,  and  G.  Bishop. 

2002.  Report  to  the  Board  of  Fisheries,  Southeast  Alaska 
Dungeness  Crab  Fishery.  Alaska  Department  of  Fish 
and  Game,  Regional  Information  Report  1J02-45.  p. 
2.2-2.16.     Alaska  Dep.  Fish  and  Game.  Juneau,  AK. 

Scheding,  K.,  T  C.  Shirley,  C.  E.  O'Clair.  and  S.  J.  Taggart. 

2001.  Critical  habitat  for  ovigerous  Dungeness  crabs.  In 
Spatial  processes  and  management  offish  populations 
October  27-30  (G.  H.  Kruse,  N.  Bez,  A.  Booth.  M  W. 
Dorn,  S.  Hills,  R.  N.  Lipcius,  D.  Pelletier,  C.  Roy,  S. 
J.  Smith,  and  D.  Withercll.  eds.),  431-446  p.  Alaska 
Sea  Grant,  report  AK-SG-01-02.  Univ.  Alaska,  Fair- 
banks, AK. 

Schultz.  D.  A.,  and  T.  C.  Shirlej 

1997.  Feeding,  foraging  and  starvation  capability  of  ovig- 
erous Dungeness  crabs  in  laboratory  conditions.  J. 
Crustacean  Res.  26:26-37. 

Schultz,  D.  A.,  T.  C.  Shirley,  C.  E.  O'Clair,  and  S.  J.  Taggart. 
1996.     Activity  and  feeding  of  ovigerous  Dungeness  crabs 
in  Glacier  Bay,  Alaska.     In  High  latitude  crabs:  biology. 
management,  and  economics.  October  11-13.  1995.  vol. 


Taggart  et  al.:  Estimating  abundance  of  Cancer  magister 


497 


96-02,  p.  411-424.     Alaska  Sea  Grant  College  Program, 
AK-SG-96-02,  Univ.  Alaska,  Anchorage,  AK. 
Shirley,  S.  M.,  and  T.  C.  Shirley. 

1988.  Appendage  injury  in  Dungeness  crabs,  Cancer 
magister,  in  southeastern  Alaska.  Fish.  Bull.  86:156- 
160. 

Shirley,  T.  C,  G.  Bishop,  C.  E.  O'Clair,  S.  J.  Taggart,  and 
J.  L.  Bodkin. 

1996.  Sea  otter  predation  on  Dungeness  crabs  in  Glacier 
Bay,  Alaska.  //;  High  latitude  crabs:  biology,  manage- 
ment, and  economics,  563-576  p.  Lowell  Wakefield  Fish- 
eries Symposium.  Alaska  Sea  Grant  College  Program 
Report  96-02,  Univ.  Alaska,  Fairbanks,  AK. 
Shirley,  S.  M.,  T.  C.  Shirley,  and  S.  D.  Rice. 

1987.     Latitudinal  variation  in  the  Dungeness  crab.  Cancer 
magister:  zoeal  morphology  explained  by  incubation 
temperature.     Mar.  Biol.  95(31:371-376. 
Smith,  B.  D.,  and  G.  S.  Jamieson. 

1989.  A  model  for  standardizing  Dungeness  crab  (Cancer 
magister\  catch  rates  among  traps  which  experienced 
different  soak  times.  Can.  J.  Fish.  Aquat.  Sci. 
46:1600-1608. 

Stone,  R.  P.,  and  C.  E.  O'Clair. 

2001.     Seasonal  movements  and  distribution  of  Dungeness 
crabs,  Cancer  magister,  in  a  glacial  southeastern  Alaska 
estuary.     Mar.  Ecol.  Prog.  Ser.  214:167-176. 
Swiney,  K.  M.,  T.  C.  Shirley.  S.  J.  Taggart.  and  C.  E.  O'Clair. 
2003.     Dungeness  crab,  Cancer  magister,  do  not  extrude 
eggs  annually  in  southeastern  Alaska:  An  in  situ 
study.     J.  Crustacean  Biol.  23(21:280-288. 
Taggart,  S.  J.,  P.  N.  Hooge,  J.  Mondragon,  E.  R.  Hooge,  and 
A.  G.  Andrews. 

2003.  Living  on  the  edge:  the  distribution  of  Dunge- 
ness crab,  Cancer  magister,  in  a  recently  deglaciated 
fjord.     Mar.  Ecol.  Prog.  Ser.  246:241-252. 


Taggart,  S.  J.,  T.  C.  Shirley.  C.  E.  O'Clair,  and  J.  Mondragon. 
In  press.  Dramatic  increase  in  the  relative  abundance  of 
large  male  Dungeness  crabs,  Cancer  magister,  following 
closure  of  commercial  fishing  in  Glacier  Bay.  Alaska.  In 
Aquatic  protected  areas  as  fisheries  management  tools 
(J.  B.  Shipley,  ed.).     Am.  Fish.  Soc.  Bethesda,  MD. 

Thomas,  L.,  and  C.  J.  Krebs. 

1997.  A  review  of  statistical  power  analysis  software. 
Bull.  Ecol.  Soc.  A.  78(21:128-139. 

Wonnacott,  T.  H.,  and  R.  J.  Wonnacott. 

1990.  Introductory  statistics  for  business  and  economics, 
815  p.     John  Wiley  &  Sons,  New  York.  NY. 

Wyngaard,  J.  G.,  and  M.  I.  Iorio. 

1996.  Status  of  the  southern  king  crab  (Lithodes  san- 
tolla)  fishery  of  the  Beagle  Channel.  Argentina.  In 
Proceedings  of  the  international  symposium  on  biology, 
management,  and  economics  of  crabs  from  high  latitude 
habitats.  Anchorage,  Alaska,  October  11-13,  1995,  vol. 
96-02,  25-39  p.  Alaska  Sea  Grant  College  Program, 
Univ.  Alaska.  Anchorage,  AK. 

Zar,  J.  H. 

1996.  Biostatistical  analysis,  663  p.  Prentice-Hall,  Inc.. 
Upper  Saddle  River,  NJ. 

Zheng,  J..  T.  J.  Quinn  II.  T.  J.,  and  G.  H.  Kruse. 

1993.  Comparison  and  evaluation  of  threshold  estimation 
methods  for  exploited  fish  populations.  In  Proceedings 
of  the  international  symposium  on  mnagement  strate- 
gies for  exploited  fish  populations,  Anchorage,  Alaska, 
October  21-24,  1992  iG.  H.  Kruse.  D.  M.  Eggers,  R. 
J.  Marasco,  C.  Pautzke.  and  T.  J.  Quinn  II.  eds.i.  vol. 
93-02,  267-289  p.  Alaska  Sea  Grant  College  Program, 
Anchorage,  AK. 


498 


Abstract— The  lengths  of  otoliths  and 
other  skeletal  structures  recovered 
from  the  scats  of  pinnipeds,  such  as 
Steller  sea  lions  iEumetopias  juba- 
tus),  correlate  with  body  size  and 
can  be  used  to  estimate  the  length 
of  prey  consumed.  Unfortunately, 
otoliths  are  often  found  in  too  few 
scats  or  are  too  digested  to  usefully 
estimate  prey  size.  Alternative  diag- 
nostic bones  are  frequently  recovered, 
but  few  bone-size  to  prey-size  cor- 
relations exist  and  bones  are  also 
reduced  in  size  by  various  degrees 
owing  to  digestion.  To  prevent  under- 
estimates in  prey  sizes  consumed 
techniques  are  required  to  account  for 
the  degree  of  digestion  of  alternative 
bones  prior  to  estimating  prey  size. 
We  developed  a  method  (using  defined 
criteria  and  photo-reference  material) 
to  assign  the  degree  of  digestion  for 
key  cranial  structures  of  two  prey 
species:  walleye  pollock  (Theragra 
chalcogramma)  and  Atka  mackerel 
(Pleurogrammus  monopterygius).  The 
method  grades  each  structure  into  one 
of  three  condition  categories;  good, 
fair  or  poor.  We  also  conducted  feeding 
trials  with  captive  Steller  sea  lions, 
feeding  both  fish  species  to  determine 
the  extent  of  erosion  of  each  structure 
and  to  derive  condition-specific  diges- 
tion correction  factors  to  reconstruct 
the  original  sizes  of  the  structures 
consumed.  In  general,  larger  struc- 
tures were  relatively  more  digested 
than  smaller  ones.  Mean  size  reduc- 
tion varied  between  different  types 
of  structures  (3.3-26.3%),  but  was 
not  influenced  by  the  size  of  the  prey 
consumed.  Results  from  the  observa- 
tions and  experiments  were  combined 
to  be  able  to  reconstruct  the  size  of 
prey  consumed  by  sea  lions  and  other 
pinnipeds.  The  proposed  method  has 
four  steps:  1)  measure  the  recovered 
structures  and  grade  the  extent  of 
digestion  by  using  defined  criteria 
and  photo-reference  collection;  2) 
exclude  structures  graded  in  poor  con- 
dition; 3)  multiply  measurements  of 
structures  in  good  and  fair  condition 
by  their  appropriate  digestion  correc- 
tion factors  to  derive  their  original 
size;  and  4)  calculate  the  size  of  prey 
from  allometric  regressions  relating 
corrected  structure  measurements  to 
body  lengths.  This  technique  can  be 
readily  applied  to  piscivore  dietary 
studies  that  use  hard  remains  of 
fish. 


Manuscript  submitted  28  April  2003 
I"  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
25  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:498-508(2004). 


A  method  to  improve  size  estimates  of 
walleye  pollock  (Theragra  chalcogramma)  and 
Atka  mackerel  (Pleurogrammus  monopterygius) 
consumed  by  pinnipeds:  digestion  correction 
factors  applied  to  bones  and  otoliths 
recovered  in  scats 

Dominic  J.  Tollit ' 
Susan  G.  Heaslip1 
Tonya  K.  Zeppelin2 
Ruth  Joy' 
Katherine  A.  Call2 
Andrew  W.  Trites1 

1  Marine  Mammal  Research  Unit,  Fisheries  Centre 
University  of  British  Columbia,  Room  18,  Hut  B-3 
6248  Biological  Sciences  Road 

Vancouver,  British  Columbia,  Canada,  V6T  1Z4 
E-mail  address  (for  D.  J  Tollit):  tollit  5zoology.ubc.ca 

2  National  Marine  Mammal  Laboratory 
Alaska  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
7600  Sand  Point  Way  NE 

Seattle,  Washington  98115 


Prey  skeletal  remnants  from  stom- 
ach samples  and  more  recently  from 
fecal  (scat)  samples  are  widely  used 
to  determine  what  pinnipeds  eat 
(Pitcher,  1981;  Olesiuk  et  al.,  1990; 
Tollit  and  Thompson,  1996;  Browne 
et  al.,  2002).  Prey  can  usually  be 
identified  from  taxon-specific  hard 
remains,  the  sizes  of  which  often  cor- 
relate with  the  length  and  mass  of 
the  prey  (Harkonen,  1986;  Desse  and 
Desse-Berset,  1996).  In  the  past,  sag- 
ittal otoliths  were  commonly  used  to 
estimate  prey  size  (Frost  and  Lowry, 
1981)  but  were  recognized  to  erode  or 
become  completely  digested  (Prime 
and  Hammond,  1987;  Harvey,  1989). 
Thus,  otolith  measurements  likely 
underestimated  sizes  and  numbers 
of  fish  ingested  (Jobling  and  Breiby, 
1986),  thereby  preventing  a  reliable 
assessment  of  overlap  of  prey  con- 
sumed with  catch  taken  by  commercial 
fisheries  (Beverton,  1985).  Accurate 
estimates  of  size  of  prey  consumed  by 
pinnipeds  are  also  important  in  order 
to  understand  foraging  behavior  and 
to  explain  spatial  and  temporal  vari- 
ability in  diet  composition. 


There  are  at  least  three  potential 
ways  to  deal  with  the  effect  of  diges- 
tion on  estimates  of  prey  size.  One  is 
to  measure  only  relatively  uneroded 
otoliths  and  assume  that  eroded  oto- 
liths are  from  the  same  size  fish  as 
uneroded  otoliths  (Frost  and  Lowry, 
1986;  Bowen  and  Harrison,  1994). 
Another  is  to  apply  a  single  species- 
specific  digestion  coefficient  or  correc- 
tion factor  (DCF),  derived  from  feed- 
ing experiments  with  captive  seals 
fed  fish  of  known  sizes  and  using 
measurements  of  all  the  eroded  oto- 
liths recovered  in  the  scats  produced 
(Prime  and  Hammond,  1987;  Harvey, 
1989).  The  third  is  to  estimate  and 
correct  for  the  degree  of  digestion 
(based  on  defined  losses  of  morpho- 
logical features)  of  each  recovered 
otolith  by  using  estimates  from  ref- 
erence material  (Sinclair  et  al..  1994; 
Antonelis  et  al.,  1997)  or  by  applying 
condition-specific  DCFs  derived  from 
fish  fed  in  captive  seal  feeding  studies 
(Tollit  et  al„  1997). 

Of  the  three  approaches  to  correctly 
estimate  prey  size  from  skeletal  re- 
mains, there  is  the  assumption  with 


To  Hit  et  al.:  A  method  to  improve  size  estimates  of  Theragra  chalcogramma  and  Pleurogrammus  monopterygius 


499 


the  use  of  only  uneroded  otoliths  that  recovery  and  the 
degree  of  digestion  is  independent  of  otolith  size,  result- 
ing in  a  potentially  biased  fraction.  For  certain  species 
it  can  also  result  in  a  notable  reduction  in  sample  size 
because  relatively  few  otoliths  pass  through  the  gut 
in  good  condition.  The  second  approach  of  applying 
mean  species-specific  DCFs  is  an  improvement  to  not 
accounting  for  size  reduction  (Laake  et  al.,  2002);  how- 
ever, there  is  the  assumption  with  this  approach  that 
all  structures  are  reduced  in  size  by  the  same  amount. 
Consequently,  mean  fish  mass  may  be  overestimated  if 
such  correction  factors  are  applied  to  relatively  undi- 
gested otoliths,  or  they  may  be  underestimated  if  ap- 
plied to  very  digested  otoliths  (Hammond  et  al.,  1994; 
Tollit  et  al.,  1997).  The  third  method  accounts  for  the 
intraspecific  variation  in  size  reduction  caused  by  di- 
gestion, reduces  systematic  error  (see  Hammond  and 
Rothery,  19961.  yields  estimates  of  mass  that  compare 
favorably  to  those  fed  to  captive  animals  (Tollit  et  al., 
1997).  and  hence  may  well  be  the  most  promising  ap- 
proach to  reconstructing  prey  size. 

The  dramatic  decline  of  the  western  population  of 
Steller  sea  lions  (Eumetopias  jubatus)  in  the  1980s 
(Loughlin  et  al.,  1992;  Trites  and  Larkin,  1996)  has 
prompted  a  number  of  studies  to  determine  what  they 
eat  and  the  extent  of  dietary  overlap  (prey  consumed) 
with  catch  taken  by  commercial  fisheries.  Stomach  con- 
tents analysis  was  used  to  determine  diet  until  the  late 
1980s  when  scat  analysis  became  the  preferred  method 
(e.g.,  Pitcher,  1981;  Frost  and  Lowry,  1986;  Sinclair  and 
Zeppelin,  2002).  However,  unlike  in  stomachs,  there  is 
an  overall  sparsity  of  otoliths  in  Steller  sea  lion  scats 
(Sinclair  and  Zeppelin,  2002)  and,  therefore  there  is  a 
need  to  also  use  other  skeletal  structures  to  describe 
the  size  of  prey  consumed. 

The  following  outlines  a  method  (using  defined  crite- 
ria and  photo-reference  material)  to  assign  the  degree 
of  digestion  for  otoliths  and  alternative  key  skeletal 
structures  of  walleye  pollock  (Theragra  chalcogramma) 
and  Atka  mackerel  (Pleurogrammus  monopterygius) 
recovered  from  scats.  We  also  present  the  results  of 
a  feeding  study  with  captive  Steller  sea  lions  used  to 
determine  the  extent  of  erosion  and  to  derive  condition- 
specific  digestion  correction  factors  to  reconstruct  the 
original  sizes  of  the  pollock  and  Atka  mackerel  struc- 
tures consumed.  Finally,  we  combine  these  DCFs  with 
newly  developed  regression  formulae  that  estimate  fish 
length  to  derive  a  more  accurate  method  of  estimating 
size  of  pollock  and  Atka  mackerel  consumed  by  Steller 
sea  lions  and  other  pinnipeds  (see  Zeppelin  et  al.,  2004, 
this  issue;  Tollit  et  al.,  2004,  this  issue). 


Materials  and  methods 

Experimentally  derived  digestion  correction  factors 

Feeding  experiments  were  conducted  with  two  3-year- 
old  female  Steller  sea  lions:  Steller  sea  lion  1  (SSL11 
[ID  no.  F97HA],  mean  mass  129  kg;  steller  sea  lion 


2  [SSL2]  [ID  no.  F97SI],  mean  mass  150  kg)  between 
October  2000  and  April  2002  at  the  Vancouver  Aquarium 
Marine  Science  Centre.  Over  the  experimental  period, 
the  sea  lions  were  fed  pollock  for  52  days  in  16  separate 
feeding  experiments,  and  Atka  mackerel  for  31  days 
in  5  separate  feeding  experiments,  at  between  -4-8% 
of  body  mass  per  day.  Fork  length  (FL)  and  weight  of 
all  fish  were  measured  to  ±0.1  cm  and  ±1  g.  Sea  lions 
were  fed  meals  of  pollock  of  three  size  categories  (small, 
28.5-32.5  cm  FL;  medium.  33.5-38.7  cm  FL;  large, 
40-45  cm  FLi  and  meals  of  Atka  mackerel  of  one  size 
category  (30-36  cm  FL).  Fish  of  one  particular  size  cat- 
egory were  fed  either  as  a  single  meal  or  as  a  seven-day 
block  of  meals.  Full  details  of  a  typical  experimental 
protocol  can  be  found  in  Tollit  et  al.  (2003).  Size  ranges 
for  any  category  offish  fed  within  separate  experiments 
were  usually  <3  cm.  Fecal  material  was  collected  until  no 
other  remains  of  experimental  meals  were  found  (7  days 
after  feeding),  and  was  washed  through  a  0.5-mm  sieve 
to  remove  hard  parts.  Each  animal  was  maintained  on 
whole  Pacific  herring  (Clupea  pallasi)  between  experi- 
ments at  ~6'7(  body  mass  per  day. 

The  strong  relationship  between  fish  size  and  otolith 
size  also  exists  for  other  skeletal  structures  (Desse  and 
Desse-Berset,  1996).  Thus,  we  quantified  the  types  and 
numbers  of  the  prey  structures  recovered  in  the  scats 
of  free-ranging  Steller  sea  lions  (from  the  collections 
of  Trites  et  al.1  and  Sinclair  and  Zeppelin,  2000)  and 
selected  seven  of  the  most  commonly  occurring  struc- 
tures for  pollock  and  Atka  mackerel.  These  were  the 
sagittal  otolith  (OTO),  as  well  as  the  interhyal  (INTEl, 
hypobranchial  3  (HYPO),  pharyngobranchial  2  (PHAR), 
angular  (ANGU),  quadrate  (QUAD),  and  the  dentary 
(DENT).  The  structures  selected  also  had  particular 
morphological  features  that  seemed  to  be  relatively 
resistant  to  digestion  and  could  effectively  be  used  to 
estimate  fish  size  (Figs.  1  and  2,  Table  1). 

Concurrent  with  our  feeding  study,  we  measured 
selected  structures  (Figs.  1  and  2)  from  randomly 
subsampled  fresh  fish  and  combined  these  data  with 
unpublished  NMFS  data  to  generate  allometric  regres- 
sion formulae  relating  structural  measurements  to  fish 
length  (see  Zeppelin  et  al.,  this  issue).  Fork  lengths 
(±0.1  cm)  and  weights  (±1  gl  of  an  extended  subsample 
of  pollock  (8.3-47.7  cm  FL)  were  measured  to  generate 
an  appropriate  regression  formula  for  estimating  fish 
mass  from  fork  length  estimates.  All  selected  structures 
are  located  in  the  cranium  as  illustrated  in  Zeppelin  et 
al.  (2004,  this  issue).  Naming  offish  structures  follows 
Rojo  (1991). 

Initial  inspection  of  selected  structures  found  in  scats 
from  the  wild  revealed  high  intraspecific  variation  in 
the  degree  of  digestion,  ranging  from  no  apparent  size 
reduction  to  about  a  60%  size  reduction  (heavily  digest- 
ed material).  Consequently,  we  extended  the  condition- 


Trites,  A.  W.,  D.  G.  Calkins,  and  A.  J.  Winship.  2003.  Un- 
publ.  data.  Marine  Mammal  Research  Unit,  Fisheries 
Centre,  University  of  British  Columbia,  Hut  B-3,  6248  Bio- 
logical Sciences  Road,  Vancouver,  B.C.,  Canada,  V6T  1Z4. 


500 


Fishery  Bulletin  102(3) 


Figure  1 

Photographs  showing  the  changes  in  morphological  features  in  seven  cranial  structures  of  walleye  pol- 
lock iTheragra  chalcogramma)  resulting  from  digestion.  Within  each  section  of  the  figure  three  condi- 
tion categories  (good,  fair,  and  poor)  are  represented  from  left  to  right  for  (A)  interhyal  (INTEi.  iBi 
hypobranchial  3  (HYPOl,  (Ci  pharyngobranchial  2  (PHARl,  (D)  angular  (ANGU),  lE  i  quadrate  (QUAD), 
(F)  dentary  (DENTi  and  (G)  sagittal  otolith  (OTO).  Key  features  used  in  classification  are  labeled  (see 
Table  1  for  details),  and  the  measurements  taken  to  calculate  fish  length  (solid  line  between  dashed  lines i. 


specific  DCF  technique  described  by  Tollit  et  al.  (1997). 
We  began  by  examining  the  external  morphological 
features  and  surface  topography  of  selected  structures 


from  undigested  fish  (<12  cm  to  >53  cm)  and  compared 
these  with  the  topography  of  the  same  structures  recov- 
ered from  scats  collected  from  wild  and  captive  animals 


"To Hit  et  al.:  A  method  to  improve  size  estimates  of  Theragra  chakogramma  and  Pleurogrammus  monopterygius 


501 


Figure  2 

Photographs  showing  the  changes  in  morphological  features  in  seven  cranial  structures  of  Atka  mack- 
erel (Pleurogrammus  monopterygius)  resulting  from  digestion.  Within  each  section  of  the  figure  three 
condition  categories  (good,  fair,  and  poor)  are  represented  from  left  to  right  for  (A)  interhyal  (INTE), 
(B)  hypobranchial  3  (HYPO),  (C)  angular  (ANGU),  (D)  quadrate  (QUAD),  (E)  dentary  (DENT)  and 
(F)  sagittal  otolith  (OTO).  Key  features  used  in  classification  are  labeled  (see  Table  1  for  details)  and 
measurements  taken  to  calculate  fish  length  (solid  line  between  dashed  lines). 


(Figs.  1  and  2).  The  morphological  features  used  to  as- 
sess level  of  digestion  showed  no  differences  in  relative 
shape,  structure,  or  in  proportion  across  the  size  range 


of  fresh  fish  examined.  We  then  devised  a  criteria-based 
method  to  assign  a  condition  category  to  each  structure 
depending  on  the  degree  of  digestion.  These  criteria 


502 


Fishery  Bulletin  102(3) 


take  into  account  only  the  loss  of  size  to  the  relevant 
feature  being  measured  to  estimate  fish  length  (Figs. 
1  and  2). 

The  grading  criteria  for  otoliths  (OTO)  were  based 
on  the  condition  categories  developed  by  Sinclair  et 
al.  (1996)  to  investigate  prey  selection  by  northern  fur 
seals  (Callorhinus  ursinus).  As  seen  in  Sinclair  et  al. 
(1996)  and  other  studies  (Frost  and  Lowry,  1986;  Tollit 
et  al.,  1997),  external  features  such  as  lobation  and  the 
general  shape  and  definition  of  the  sulcus  were  found  in 
our  study  to  be  good  indicators  of  the  degree  of  otolith 
digestion.  For  the  remaining  cranial  bones,  digestion 
indicators  included  the  loss  of  definition  or  breakage 
of  defined  structural  features  such  as  the  horns  and 
ridge  (QUAD),  hammerhead  and  stock  (DENT),  swan 
neck,  notch  and  ridge  (INTE),  honeycomb  and  crown 
iPHAR),  cap,  neck,  and  head  (ANGU)  and  tube  and 
cone  (HYPO).  We  used  changes  in  the  described  condi- 
tion-category criteria  (see  Table  1  for  full  details)  in 
tandem  with  photo-reference  material  (Figs.  1  and  2)  to 
classify  all  structures  into  one  of  three  digestion  grades 
or  condition  categories:  "good",  "fair,"  or  "poor." 

Hard  parts  recovered  from  feeding  experiments  were 
sorted,  and  all  selected  cranial  structures  were  as- 
signed a  condition  category  and  measured  with  cali- 
pers to  within  ±0.01  mm.  Because  otoliths  were  often 
chipped  or  partly  broken  lengthwise,  both  length  and 
width  were  measured.  To  test  our  grading  technique,  an 
independent  observer  (T.Z.)  reassigned  a  random  sub- 
sample  of  each  condition  category  of  pollock  structures 
(n  =  158)  in  a  blind  test. 

On  initial  investigation,  high  intraspecific  variation 
was  observed  within  the  selected  structures  assigned  in 
poor  condition  in  our  feeding  study  with  captive  Steller 
sea  lions.  Consequently,  structures  in  poor  condition 
were  not  used  to  calculate  DCFs  for  this  category.  Our 
basis  for  exclusion  was  supported  by  the  work  of  Sin- 
clair et  al.  (1994)  and  Tollit  et  al.  (1997).  Captive  sea 
lions  in  our  study  occasionally  regurgitated  prey  in  the 
swim  tank.  Recovered  structures  that  we  considered  to 
have  been  regurgitated  were  excluded  from  DCF  calcu- 
lations (i.e.,  vertebrae  still  articulated,  bones  that  had 
flesh  attached  or  that  were  of  a  size  to  exclude  passage 
through  the  pyloric  sphincter). 

Mean  reduction  (MR)  in  the  metric  of  each  structure 
(s)  recovered  from  our  feeding  experiment  was  estimated 
for  each  remaining  condition  category  (c)  according  to 


MR 


T 


xlOO, 


where  the  mean  size  of  egested  structures  (E)  of  each 
condition  category  was  calculated  from  measurements 
of  those  recovered  from  the  captive  feeding  experiments. 
and  the  mean  size  of  each  ingested  structure  (/.)  was 
estimated  from  the  fork  length  of  fish  fed  by  using 
inverse  predictions  of  the  regression  formulae  derived 
from  fresh  material  (Zeppelin  et  al.,  2004,  this  issue). 
Mean  ingested  size  was  estimated  by  using  bootstrap 


simulations  (1000  runs)  that  randomly  sampled  with 
replacement  and  selected  the  median  (500th  value  I  from 
the  sorted  bootstrapped  values  (Reynolds  and  Aebischer, 
1991). 

For  pollock,  mean  reduction  for  each  condition  cat- 
egory was  compared  across  size  ranges  by  using  a  Krus- 
kal-Wallis  analysis  of  variance.  A  significance  level  of 
P<0.0056  was  set  based  on  the  Bonferroni  adjusted 
probability  for  nine  multiple  comparisons  (Siegel  and 
Castellan,  1988).  Failing  to  find  any  significant  differ- 
ences resulted  in  pooling  the  data  from  each  size  range 
to  calculate  specific  condition  category  MR  values.  Con- 
dition category  DCFs  were  calculated  for  each  selected 
structure  as  /  JESC  except  for  PHAR  structures  of  Atka 
mackerel  because  too  few  elements  were  recovered  from 
the  scats  of  captive  animals. 

Estimating  confidence  limits  around  digestion 
correction  factors 

We  used  a  bootstrap  simulation  to  estimate  upper  and 
lower  bounds  of  the  95%  confidence  interval  (CI)  given 
that  the  DCF  is  a  ratio  of  two  means  (Reynolds  and 
Aebischer.  1991 1.  This  technique  allows  different  sources 
of  error  to  be  combined  or  partitioned.  There  were  two 
major  sources  of  error  associated  with  calculating  DCFs 
(Tollit  et  al.,  1997).  The  first  were  those  associated  with 
the  regression  formulae  used  to  calculate  the  mean  size 
of  structure  ingested  from  the  original  fish  fed.  and  the 
second  were  those  associated  with  the  errors  around  the 
mean  size  of  egested  structure  (i.e.,  resampling  errors). 

We  assessed  errors  associated  with  the  regression 
formulae  using  a  parametric  bootstrapping  procedure 
(Manly,  1997)  that  involved  regressing  structure  size 
against  fork  length.  This  was  repeated  1000  times  and 
95%  confidence  intervals  were  taken  as  the  25th  and 
975th  values  of  the  sorted  bootstrapped  regression  coef- 
ficient values.  Results  were  compared  to  those  computed 
analytically  by  using  the  resultant  standard  error  (Eq. 
17.23  in  Zar,  1984)  and  were  found  to  be  consistent  (see 
Zeppelin  et  al.,  2004,  this  issue). 

We  estimated  resampling  errors  related  to  the  vari- 
ability in  digestion  of  egested  structures  by  repeatedly 
selecting  n  structures,  at  random,  with  replacement  from 
the  original  sample  set  of  n  egested  structures.  Mean 
egested  size  was  recalculated  in  this  way  1000  times, 
as  were  a  mean  DCF  and  95%  CI  as  described  above. 
Both  regression  and  resampling  errors  were  combined  in 
sequence  to  derive  overall  95%  CIs  around  DCFs. 

Our  recommended  procedure  for  applying  our  DCFs  to 
cranial  structures  recovered  from  scats  collected  in  the 
wild  has  four  steps:  1)  measure  the  recovered  structures 
and  grade  the  extent  of  digestion  using  defined  criteria 
and  photo-reference  collection;  2)  exclude  structures 
graded  in  poor  condition;  3)  multiply  measurements  of 
structures  in  good  and  fair  condition  by  their  appropriate 
digestion  correction  factors  to  derive  their  original  size; 
and  4)  calculate  the  size  of  prey  from  allometric  regres- 
sions relating  corrected  structure  measurements  to  fish 
fork  lengths  (see  also  Tollit  et  al..  2004,  this  issue). 


"To Hit  et  al.:  A  method  to  improve  size  estimates  of  Theragra  cha/cogramma  and  Pleurogrammus  monopterygius 


503 


Results 

A  relatively  objective  method  to  estimate  the  degree  of 
digestion  of  dominant  structures  of  pollock  and  Atka 
mackerel  was  derived  by  using  defined  criteria  (Table  1) 
and  photo-reference  material  (Figs.  1  and  2).  Condition- 
specific  digestion  correction  factors  (and  derived  confi- 
dence intervals)  calculated  for  each  structure  augmented 
our  method  of  estimating  size  of  prey  from  bones  and 
otoliths  recovered  in  scats,  as  well  as  potentially  from 
bones  and  otoliths  taken  from  stomach  contents. 

Mean  reduction  (MR)  in  the  size  of  pollock  DENT 
and  QUAD  in  good  condition  and  ANGU,  HYPO,  IN- 
TE,  OTO,  and  PHAR  in  fair  condition  were  between 
12.2-18.5%,  and  larger  values  were  found  for  QUAD 
(22.8%)  and  DENT  (24.7%)  in  fair  condition  (Table  2). 
Our  overall  95%  confidence  intervals  were  generally 
symmetrical  and  converted  to  a  mean  range  of  ±2.2% 
(±0.5,  SD)  around  MR  values.  Mean  DCFs  ranged  be- 
tween 1.14  and  1.33,  and  lower  bounds  of  95%  CIs  ex- 
ceeded 1.11  in  all  instances,  confirming  that  egested 
structures  of  these  condition  categories  were  significant- 
ly smaller  than  the  size  at  which  they  were  ingested 
(Table  2).  Partitioning  errors  showed  that  resampling  of 
egested  structures  was  the  major  source  of  error  (>73% 
across  structures)  within  the  overall  total.  Our  overall 
95%  CIs  resulted  in  a  maximum  total  error  of  ±1.7  cm 
around  an  estimated  mean  of  40  cm  for  pollock. 

Mean  reduction  in  the  size  of  Atka  mackerel  struc- 
tures varied  more  widely  (3.3-26.3%),  leading  to  DCFs 
ranging  between  1.03  and  1.36.  QUAD  in  good  condition 
provided  the  smallest  DCF,  and  DENT  in  fair  condition 
the  largest.  Overall,  our  95%  CIs  converted  to  a  mean 
range  of  ±2.4%  (±0.6)  around  MR  values,  and  all  lower 
95%  CI  bounds  exceeded  1.0  (Table  2).  As  seen  when 
errors  were  partitioned  for  pollock,  errors  owing  to  resa- 
mpling of  egested  structures  were  the  major  source  of 
error  (>83'7t  across  structures)  within  the  overall  total 
for  Atka  mackerel.  Our  overall  95%  CIs  resulted  in  a 
maximum  total  error  of  ±1.2  cm  around  an  estimated 
mean  of  40  cm  for  Atka  mackerel. 

With  the  exception  of  the  two  largest  skeletal  struc- 
tures (DENT  and  QUAD,  Table  2),  some  selected  struc- 
tures (INTE,  HYPO,  PHAR,  ANGU,  and  OTO)  occurred 
in  scats  with  no  clear  loss  in  size  or  loss  of  morphologi- 
cal features  related  to  digestion.  For  these  five  struc- 
tures, we  ascribed  the  condition  category  good  and  as- 
signed a  DCF  of  1.0  (i.e.,  no  correction  for  partial  size 
reduction  due  to  digestion  required). 

Of  the  158  structures  in  our  blind  test,  141  (89.2%) 
were  assigned  identical  condition  categories.  Of  the 
remaining  17  structures,  11  (65%)  were  noted  as  be- 
ing borderline  between  categories.  Angulars  (ANGU) 
accounted  for  the  majority  (-60%)  of  all  differences, 
with  all  but  one  re-assigned  in  good  condition  as  op- 
posed to  fair  condition.  On  review,  differences  in  as- 
signing angulars  were  mainly  the  result  of  differences 
in  opinion  on  what  constituted  a  well-defined  and  sharp 
point  (Fig.  1,  Table  1).  Clarification  through  the  addi- 
tional use  of  reference  material  (including  both  pristine 


structures  and  examples  of  each  condition  category)  is 
advised,  particularly  for  angulars.  Comparison  of  the 
same  158  bones  between  two  observers  (D.T.  and  S.H.) 
using  the  same  structure  reference  collection  resulted 
in  assigning  more  than  93%  (147/158)  of  structures  to 
an  identical  category. 

The  regression  formula  for  estimating  pollock  mass 
(M)  from  fork  length  (FL)  estimates  was  best  described 
by  using  an  exponential  equation  (M=0.0051  xFL3n, 
n  =  981.  r2  =  0.987). 


Discussion 

The  size  of  prey  consumed  by  pinnipeds  can  usually  be 
reliably  estimated  from  otoliths  recovered  in  scats  if 
partial  digestion  is  accounted  for  (Tollit  et  al.,  1997). 
However,  otoliths  from  Steller  sea  lion  scats  are  often 
found  in  too  few  numbers,  or  are  too  digested  or  broken 
to  be  useful  (Sinclair  and  Zeppelin,  2002;  Tollit  et  al., 
2004,  this  issue).  It  was,  therefore,  necessary  to  use 
alternative  skeletal  structures  to  estimate  the  size  of 
prey  selected  by  Steller  sea  lions.  Zeppelin  et  al.  (2004, 
this  issue)  documented  good  relationships  (r2=0.78-0.99) 
between  the  size  of  selected  alternative  structures  and 
fork  length  for  pollock  and  Atka  mackerel.  However, 
all  skeletal  structures  are  susceptible  to  digestion  in 
the  stomach  (our  study,  and  Murie  and  Lavigne,  1986). 
Thus,  techniques  are  required  to  account  for  the  degree 
of  digestion  of  alternative  structures  prior  to  estimating 
prey  size. 

Reductions  in  the  size  of  otoliths  during  passage 
through  the  digestive  tract  of  pinnipeds  have  been 
widely  reported  (e.g.,  da  Silva  and  Neilson,  1985;  Prime 
and  Hammond.  1987;  Harvey,  1989;  Tollit  et  al.,  1997). 
Similarly,  we  found  significant  reduction  in  the  sizes  of 
all  selected  cranial  structures  from  pollock  and  Atka 
mackerel.  Size  reduction  also  showed  great  variability. 
Relatively  small  structures  were  found  with  no  obvious 
loss  in  size  due  to  digestion,  but  were  also  frequently 
heavily  eroded. 

The  degree  of  digestion  on  different  otoliths  and 
bones  may  be  related  to  species,  size  of  fish  (Bowen, 
2000).  or  even  its  shape,  but  seems  to  be  random  in  any 
one  meal  (Murie  and  Lavigne,  1986).  Degree  of  diges- 
tion likely  depends  on  a  range  of  factors  such  as  meal 
size,  meal  frequency,  meal  composition,  and  method  of 
consumption.  In  the  face  of  these  multiple  factors  we 
feel  our  method  for  classifying  the  degree  of  digestion 
into  one  of  three  condition  categories  is  practical  and 
relatively  objective.  However,  our  technique  does  not 
consider  potential  biases  of  enumeration  associated 
with  smaller  prey  being  more  susceptible  to  complete 
digestion  than  relatively  larger  prey,  or  of  individual 
fish  being  counted  more  than  once  if  all  multiple  struc- 
tures are  used.  Nevertheless,  resolution  to  these  biases 
have  been  advocated  (see  Tollit  et  al.,  1997;  Laake  et 
al.,  2002;  Tollit  et  al.,  2003;  Tollit  et  al.,  2004,  this  is- 
sue). The  category  selections  chosen  with  our  criteria 
showed  good  agreement  among  independent  observers. 


504 


Fishery  Bulletin  102(3) 


Table  1 

Distinctive  external  morphological  features  for  defining  the  degree  of  digestion  (condition  category)  as  good(G),  fair  (F),  and  poor 
(Pi  for  selected  cranial  structures  of  walleye  pollock  and  Atka  mackerel.  Features  are  given  in  order  of  importance.  See  Table  2 
for  definition  of  structure  codes  and  Figure  1  and  2  for  illustrations.  WP  =  walleye  pollock. 

Species  and 

structure  code         Category  Distinctive  external  morphological  features 


Walleye  pollock 
INTE 


HYPO 


PHAR 


ANGU 


QUAD 


DENT 


OTO 


G  1)  Retains  characteristic  shape,  notably  the  ridge  and  swan  neck.  2)  Both  ends  show  no  damage 
(except  for  the  loss  of  the  point  and  minor  nicks)  and  do  not  affect  length  measurement. 

F  li  Ridge  and  swan  neck  clearly  defined.  2)  One  end  can  show  limited  damage  with  <159c  reduction. 
Minor  nicks  on  opposite  end  acceptable,  if  there  is  no  further  loss  in  length  measurement. 

P  1)  Loss  of  characteristic  shape,  with  ridge  or  swan  neck  (or  both)  ill  defined.  Body  of  structure 
contains  holes.  2)  Both  ends  show  clear  damage. 

G  1)  Retains  characteristic  shape,  with  cone  ~2x  the  length  of  the  tube.  2)  Tube  end  and  area  1  show  no 
damage  (except  for  minor  nicks)  and  do  no  affect  the  total  length  measurement.  3)  Cone  end  angled 
when  viewed  from  the  front  elevation  (back  elevation  shown  in  Fig.  1). 

F  1)  Tube  end  or  area  1  shows  limited  damage  (cone  end  no  longer  angled)  clearly  preventing  an 
accurate  length  measurement. 

P        1)  Both  tube  end  and  area  1  show  damage,  and  a  general  loss  of  characteristic  shape  evident. 

G  1)  Retains  characteristic  shape,  notably  a  raised  spout,  honeycomb,  and  crown.  2)  Crown  clearly 
projects  above  honeycomb  (front  elevation)  and  is  intact  at  area  2.  3)  Clear  projection  of  honeycomb 
(back  elevation — see  area  3).  4)  No  affect  on  measurement. 

F  1)  No  clear  projection  of  honeycomb  at  area  3  or  crown  shows  damage  at  area  2  (preventing  an 
accurate  width  measurement).  2)  Crown  or  spout  (or  both)  can  show  minor  damage. 

P  li  Characteristic  shape  lost,  often  only  honeycomb  present.  2)  Honeycomb  smooth,  crown  heavily 
eroded  with  areas  2  and  3  eroded  or  damaged.  3)  Both  ends  show  clear  damage. 

G  1)  Point  sharp  and  well  defined  with  no  impact  to  measurement.  2)  Area  4  in  good  condition  and 
angled  curve  complete.  3)  Neck  present,  but  with  minor  damage.  4)  Material  of  cap  continues  to  point 
tip. 

F  1 )  Point  no  longer  extensive  or  sharp  or  area  4  damaged  and  poorly  defined.  2 )  Neck  usually  present, 
but  with  wear. 

P  1 )  Characteristic  shape  lost  with  neck  often  absent.  2 )  Point  heavily  eroded.  3 )  Area  4  shows  damage 
or  no  definition. 

G  1)  Groove  defined  from  all  angles  and  observable  with  the  naked  eye.  2)  Horns  rounded.  3)  Angle  of 
area  5  is  clearly  curvilinear.  4)  Evidence  of  ridge  and  spike  often  observable. 

F  1)  Groove  unclear,  forming  only  an  indisctinct  notch.  2)  Horns  have  lost  rounded  definition  and  may- 
be pointed  or  worn  on  one  side.  3)  Ridge  and  spike  often  only  residual. 

P  1)  Horns  pointed,  notch  absent.  2)  Ridge  and  spike  often  absent.  3)  Angle  of  area  5  flattened. 
4 1  Unable  to  determine  side  with  assurance. 

G  1)  Hammerhead  retains  rounded  end  elevation  features  (note:  both  sides  are  not  exactly  symmetrical), 
allowing  full  width  measurement.  2)  Material  in  addition  to  the  stock  may  be  present.  3)  Stock 
clearly  curved  from  side  elevation.  4)  Width  and  breadth  of  "rounded"  stock  similar. 

F  1)  Hammerhead  shows  erosion  on  one  side,  affecting  full  width  measurement.  2)  Breadth  of  stock 
reduced,  but  not  flattened. 

P  li  Hammerhead  eroded  and  flattened  with  both  sides  showing  erosion.  2)  Breadth  of  stock  flattened, 
stock  less  rounded  and  less  robust.  3)  Unable  to  determine  side  with  assurance. 

P  1)  Sulcus  and  scalloping  (on  most  margins)  well  defined,  and  no  obvious  reduction  in  size  due  to 
digestion.  2)  Able  to  determine  side.  3)  Inside  strongly  convex,  retains  characteristic  shape. 

F  1)  Sulcus  worn  but  shows  definition.  2)  Able  to  determine  side.  3)  Scalloping  worn  but  shows  no 
reverse  scalloping. 

P  1)  Unable  to  determine  side.  2)  Scalloping  worn  completely  smooth  and  reverse  scalloping  present. 
3)  Clearly  broken,  worn,  flattened,  and  unable  to  obtain  an  accurate  measurement. 

continued 


"To Hit  et  al.:  A  method  to  improve  size  estimates  of  Theragra  chalcogramma  and  P/eurogrammus  monopterygius 


505 


Table  1  (continued) 


Species  and 
structure  code 


Category 


Distinctive  external  morphological  features 


Atka  mackerel 
INTE 


HYPO 


ANGU 


QUAD 


DENT 


OTO 


G       Like  that  of  walleye  pollock  (WP)  (except  no  point),  with  ridge,  neck  and  notch  clearly  defined. 

F  li  Ridge  present,  but  shows  signs  of  wear.  2)  Swan  neck  shows  wear  resulting  in  a  "horseshoe"  shape. 
3 1  Notch  shows  only  minor  wear  or  chipping  and  does  not  prevent  accurate  measurement. 

P  1)  Loss  of  characteristic  ridge  and  neck  with  body  worn  (may  contain  holes).  2)  Both  neck  and  notch 
show  clear  damage. 

G  1)  Cone  rounded  and  complete,  tube  complete,  retains  characteristic  shape.  2)  Minor  nicks  on  cone 
and  tube  may  be  present  but  do  not  impact  total  length  measurement. 

F  Cone  worn,  loss  of  rounded  shape,  and  area  1  shows  minor  chipping  or  damage  or  tip  of  tube  is  broken 
or  clearly  chipped. 

P  1)  Cone  body  and  area  1  show  major  wear,  chips,  and  breaks.  2)  Tube  broken  or  absent  entirely, 
unable  to  measure  length. 

G       Like  WP.  additionally  cap  rounded  and  head  shows  only  minor  wear. 

F  Like  WP,  additionally  1)  cap  worn  with  loss  of  shape.  2)  Head  worn,  chipped,  and  often  has  holes, 
3)  Ridge  on  dorsal  side  above  neck  worn  smooth. 

P  Like  WP.  additionally  1 )  head  shows  major  damage,  wear,  breaks,  and  holes,  2  )  Difficult  to  determine 
side  with  confidence. 

G  It  Horns  rounded  and  in  good  condition,  with  angle  between  horns  clearly  curvilinear.  (Note:  Horns 
are  of  unequal  size  and  shape  and  one  side  is  more  robust,  rounded,  and  sloped.)  2 1  Evidence  of  ridge 
and  spike  observable.  3)  Definition  of  left  and  right  sides  is  easily  achievable. 

F  1)  Horns  have  lost  rounded  definition  and  may  be  pointed  or  worn  on  one  side,  making  distinction 
between  sides  difficult.  2)  Ridge  and  spike  often  only  residual. 

P        Like  WP,  additionally  no  distinction  between  horns  easily  achievable. 

G  Like  WP  (except  no  hammerhead),  additionally  1)  head  retains  characteristic  features,  tooth  sockets 
present,  2)  Ventral  side  of  head  a  defined  point. 

F  Like  WP,  additionally  1)  head  eroded  or  chipped  with  tooth  sockets  noticeably  worn,  2)  Point  on 
ventral  side  of  head  eroded  or  chipped. 

P  Like  WP,  additionally  head  eroded  or  flattened  with  point  often  heavily  eroded  or  badly  chipped, 
accurate  measurement  unattainable. 

G  1 1  Rostrum  not  chipped  or  broken.  2 )  Sulcus  clearly  defined,  as  are  anterior  and  posterior  colliculums. 
3 )  Scalloping  on  antirostrum  and  posterior  end  clearly  distinguishable.  4 )  No  obvious  wear  or  chipping 
with  no  obvious  reduction  in  length  (width).  5)  Cristae  of  antirostrum  forms  a  well-defined  ridge. 

F  1)  Rostrum  shows  some  wear  but  remains  unbroken  and  retains  characteristic  shape.  2 1  Sulcus 
still  has  definition  despite  wear,  shown  as  a  uniform  channel,  anterior  and  posterior  colliculums 
indistinct.  3)  Cristae  and  scalloping  on  antirostrum  and  posterior  end  worn  smooth. 

P  1 )  Rostrum  or  posterior  end  broken  or  worn  to  such  a  degree  that  accurate  measurement  cannot  be 
obtained.  2)  Sulcus  difficult  to  distinguish  or  worn  smooth.  3)  Cristae  and  scalloping  on  antirostrum 
and  posterior  end  worn  completely  smooth.  4)  Side  cannot  be  easily  obtained. 


Nevertheless,  we  recommend  that  a  hands-on  reference 
collection  be  used. 

The  procedure  we  recommend  to  estimate  fish  length 
after  classification  involves  excluding  structures 
considered  heavily  digested  (condition  category  poor) 
and  applying  specific  condition-category  DCFs  (Ta- 
ble 2)  to  the  remaining  structure  prior  to  calculating 
fish  length  from  allometric  regressions  (see  Tollit  et 
al.,  2004,  this  issue).  The  exclusion  of  structures  in 
poor  condition  was  necessary  because  of  the  large  and 


variable  size  reduction  observed  in  this  category.  Our 
technique  uses  changes  noted  in  the  morphological 
features  of  the  structures  themselves  and  is  therefore 
not  specific  to  Steller  sea  lions.  Because  structures 
are  likely  to  erode  in  a  predictable  manner  whatever 
the  species  of  the  stomach  they  are  held  within,  it 
seems  probable  that  they  can  also  be  classified  into 
a  particular  condition  category  for  use  with  DCFs. 
Consequently,  our  technique  may  be  appropriate  to 
marine  piscivore  dietary  studies  where  prey  size  needs 


506 


Fishery  Bulletin  102(3) 


Table  2 

Condition-specific  digestion 

correction  factors  (DCFs)  for  selected  cranial  structures  of  walleye  pollock  and  Atka  mackerel  with 

associated  condition  categories  good  (G)  and  fair  (F 

).  Lower  and  upper 

bounds  of  the  95^ 

confidence  intervals  (CIs)  were  calcu- 

lated  by  using  bootstrap  ret 

ampling  procedures. 

Structure 

CI 

Species  and  structure 

code 

Grade 

n 

DCF 

SD 

Lower 

Upper 

Walleye  pollock 

Interhyal 

INTE 

F 

54 

1.1423 

0.054 

1.1168 

1.1714 

Hypobranchial  3 

HYPO 

F 

22 

1.1658 

0.063 

1.1343 

1.1970 

Pharyngobranchial  2 

PHAR 

F 

39 

1.2109 

0.067 

1.1717 

1.2566 

Angular 

ANGU 

F 

85 

1.2065 

0.103 

1.1670 

1.2462 

Quadrate 

QUAD 

G 

20 

1.2272 

0.039 

1.2025 

1.2512 

F 

27 

1.2958 

0.074 

1.2623 

1.3280 

Dentary 

DENT 

G 

17 

1.1950 

0.074 

1.1546 

1.2337 

F 

31 

1.3285 

0.071 

1.2941 

1.3649 

Otolith  (length) 

OTOL 

F 

37 

1.1593 

0.059 

1.1400 

1.1788 

Otolith  (width) 

OTOW 

F 

49 

1.2107 

0.089 

1.1901 

1.2419 

Atka  mackerel 

Interhyal 

INTE 

F 

37 

1.0729 

0.089 

1.0374 

1.1085 

Hypobranchial  3 

HYPO 

F 

23 

1.1361 

0.040 

1.1160 

1.1568 

Angular 

ANGU 

F 

40 

1.1361 

0.097 

1.1053 

1.1700 

Quadrate 

QUAD 

G 

23 

1.0343 

0.053 

1.0070 

1.0597 

F 

23 

1.0886 

0.078 

1.0551 

1.1213 

Dentary 

DENT 

G 

34 

1.2068 

0.098 

1.1666 

1.2466 

F 

37 

1.3563 

0.143 

1.3063 

1.4119 

Otolith  (length) 

OTOL 

F 

109 

1.1691 

0.109 

1.1459 

1.1921 

Otolith  (width) 

OTOW 

F 

115 

1.2062 

0.104 

1.1837 

1.2277 

to  be  determined  from  partially  digested  prey  hard 
remains. 

Experimentally  derived  pollock  DCFs  were  deter- 
mined from  three  distinct  size  ranges  of  fish  (28.5-45 
cm  FL),  but  the  degree  of  erosion  for  each  structure 
within  each  condition  category  did  not  show  any  sig- 
nificant differences  across  this  range.  We  also  found  the 
relative  shape,  structure,  and  proportion  of  the  morpho- 
logical features  used  to  estimate  erosion  were  consistent 
for  both  smaller  and  larger  fish.  We  therefore  believe 
DCFs  can  be  used  for  fish  outside  of  the  experimental 
size  range  of  this  study.  Average  size  reduction  varied 
between  different  pollock  structures  (12.2-24.7%)  and 
also  between  condition  categories,  as  they  did  for  Atka 
mackerel  (Table  2).  We  determined  that  pollock  otoliths 
in  fair  condition  were  reduced  by  149J  in  length,  close  to 
the  20%  value  estimated  from  reference  material  (Sin- 
clair et  al.,  1994).  Our  criteria  for  defining  a  condition 
category  of  fair  for  pollock  otoliths  equates  to  a  grade 
between  low  amounts  and  medium  amounts  of  diges- 
tion as  defined  by  Tollit  et  al.  (1997)  for  Atlantic  cod 
(which  has  a  similar  looking  otolith).  Our  value  of  14% 
lies  midway  between  those  determined  for  cod  otoliths 
graded  low  and  medium. 

Jaw  bones  (DENT)  were  by  far  the  largest  structure 
used  in  our  study  but  do  not  appear  to  pass  through 


the  pyloric  sphincter  without  some  level  of  digestion. 
Usually  only  the  hammerhead  and  stock  (representing 
less  than  a  third  of  the  whole  structure)  are  recov- 
ered in  scats.  The  large  size  accounts  for  the  relatively 
greater  percent  mean  reduction  and  hence  higher  DCF 
of  DENT  structures  graded  either  in  good  or  fair  con- 
dition (Table  2).  Although  quadrates  (QUAD)  are  also 
relatively  large  structures  with  a  projecting  ridge  that 
is  often  much  reduced  when  found  in  scats,  we  found 
QUAD  structures  of  Atka  mackerel  recovered  in  scats 
from  field  studies  and  captive  sea  lion  studies  in  rela- 
tively better  condition  than  those  of  pollock,  leading 
to  differences  in  grading  criteria  and  resulting  DCFs 
(Tables  1  and  2).  Part  of  the  reason  may  be  that  the 
horns  on  a  pollock  QUAD  project  widthwise  more  than 
those  of  Atka  mackerel,  presenting  a  greater  surface 
area  for  digestive  erosion  of  the  structural  feature  that 
is  measured  to  estimate  size  (Fig.  1). 

Our  overall  95^f  confidence  intervals  around  DCFs 
were  generally  narrow  (Table  2),  highlighting  the  tight 
fit  of  the  regression  formulae  used  and  the  benefits  of 
partitioning  the  data  into  specific  categories.  Our  boot- 
strap analysis  suggests  that  resampling  errors  were 
the  major  source  of  error  in  calculating  DCFs.  Future 
research  should  concentrate  on  improving  sample  sizes 
for  data  on  percentage  size  reduction  of  bones  for  each 


Tollit  et  al.:  A  method  to  improve  size  estimates  of  Theragra  chalcogramma  and  Pleurogrammus  monopterygius 


507 


category,  rather  than  on  improving  regression  formulae. 
For  both  prey  species,  QUAD  in  good  condition  and 
OTO  in  fair  condition,  in  addition  to  pollock  INTE  in 
fair  condition  and  Atka  mackerel  HYPO  in  fair  condi- 
tion, provided  the  most  reliable  estimates  of  prey  size 
(Table  2).  DENT  in  fair  condition,  particularly  for  Atka 
mackerel,  provided  the  least  reliable  estimate  of  prey 
size  (Table  2).  Measurement  error  was  relatively  insig- 
nificant, but  attention  should  be  taken  when  measuring 
ANGU  and  HYPO  (Tollit  et  al.,  2004,  this  issue). 

Companion  studies  by  Tollit  et  al.  (2004,  this  issue) 
and  Zeppelin  et  al.  (2004,  this  issue)  demonstrate  the 
feasibility  of  applying  DCFs  to  structures  other  than 
otoliths  and  the  need  to  consider  the  degree  of  diges- 
tion to  correctly  estimate  the  length  of  prey  eaten  by 
pinnipeds  and  other  piscivores.  Applying  appropriate 
digestion  correction  factors  will  lead  to  more  refined 
estimates  of  consumption  (mass  of  prey)  by  marine 
mammals,  as  well  as  the  extent  of  potential  overlap 
(length  of  prey)  with  the  length  of  fish  caught  by  com- 
mercial fisheries. 


Acknowledgments 

Funding  was  provided  to  the  North  Pacific  Universities 
Marine  Mammal  Research  Consortium  by  the  National 
Oceanographic  Atmospheric  Administration  and  the 
North  Pacific  Marine  Science  Foundation.  We  would 
like  to  thank  the  marine  mammal  trainers  and  staff  of 
the  Vancouver  Aquarium  Marine  Science  Centre,  the 
contribution  of  personnel  of  the  UBC  Marine  Mammal 
Research  Unit  of  the  UBC  EM  facility,  J.  L.  Laake  for 
statistical  advice,  T.  J.  Orchard,  C.  J.  Gudmundson,  S.  J. 
Crockford,  M.  Wong,  E.  H.  Sinclair,  and  two  anonymous 
reviewers.  We  would  also  like  to  express  gratitude  to  the 
organizations  and  companies  that  have  donated  fish  to 
the  project.  Work  was  undertaken  in  accordance  with 
UBC  Animal  Care  Committee  guidelines. 


Literature  cited 

Antonelis,  G.  A.,  E.  H.  Sinclair.  R.  R.  Ream,  and  B.  W.  Robson. 
1997.     Inter-island  variation  in  the  diet  of  female  northern 
fur  seals  tCallorhinus  ursinus)  in  the  Bering  Sea.     J. 
Zool.  lLond.l  242(31:435-51. 
Beverton,  R.  J.  H. 

1985.     Analysis  of  marine  mammal-fisheries  interaction. 
In  Marine  mammals  and  fisheries  (J.  R.  Beddington,  R. 
J.  H.  Beverton,  and  D.  M.  Lavigne  eds.l,  p.  3-33.     George 
Allen  &  Unwin,  Boston,  MA. 
Bowen,  W.  D. 

2000.     Reconstruction  of  pinniped  diets:  Accounting  for 
complete  digestion  of  otoliths  and  cephalopod  beaks. 
Can.  J.  Fish.  Aquat.  Sci.  57:898-905. 
Bowen,  W.  D.,  and  G.  D.  Harrison. 

1994.  Offshore  diet  of  grey  seals  Halichoerus  grypus 
near  Sable  Island,  Canada.  Mar.  Ecol.  Prog.  Ser.  112 
(1-2):1-11. 


Browne,  R,  J.  L.  Laake.  and  R.  L.  DeLong. 

2002.     Improving    pinniped    diet    analyses    through 

identification  of  multiple  skeletal  structures  in  fecal 

samples.     Fish.  Bull.  100:423-433. 
da  Silva,  J.,  and  J.  D.  Neilson. 

1985.  Limitations  of  using  otoliths  recovered  in  scats 
to  estimate  prey  consumption  in  seals.  Can.  J.  Fish. 
Aquat.  Sci.  42:1439-1442. 

Desse,  J.,  and  N.  Desse-Berset. 

1996.     On  the  boundaries   of  osteometry   applied  to 
fish.     Archaeofauna  5:171-179. 
Frost,  K.  J.,  and  L.  F.  Lowry. 

1981.  Trophic  importance  of  some  marine  gadids  in 
northern  Alaska  and  their  body-otolith  size  relation- 
ships.    Fish.  Bull.  79:187-192. 

1986.  Sizes  of  walleye  pollock,  Theragra  chalcogramma, 
consumed  by  marine  mammals  in  the  Bering  Sea.  Fish. 
Bull.  84:192-197. 

Hammond,  P.  S„  A.  J.  Hall,  and  J.  H.  Prime. 

1994.     The  diet  of  grey  seals  around  Orkney  and  other 
island  and  mainland  sites  in  North-eastern  Scotland.     J. 
Appl.  Ecol.  31:340-350. 
Hammond,  P.  S.,  and  P.  Rothery. 

1996.  Application  of  computer  sampling  in  the  estimation 
of  seal  diet.     J.  Appl.  Stat.  23:525-533. 

Harkonen,  T. 

1986.  Guide  to  the  otoliths  of  the  bony  fishes  of  the  North- 
east Atlantic,  256  p.  Danbiu  ApS.  Biological  Consul- 
tants, Hellerup,  Denmark. 

Harvey,  J.  T. 

1989.  Assessment  of  errors  associated  with  harbor 
seal  (Phoca  vitulina)  fecal  sampling.  J.  Zool.  (Lond.) 
219:101-111. 

Jobling,  M.,  and  A.  Breiby. 

1986.     The  use  and  abuse  offish  otoliths  in  studies  of  feed- 
ing habits  of  marine  piscivores.     Sarsia  71:265-274. 
Laake.  J.  L.,  P.  Browne,  R.  L.  DeLong.  and  H.  R.  Huber. 

2002.     Pinniped  diet  composition:  a  comparison  of  esti- 
mation models.     Fish.  Bull.  100:434-477. 
Loughlin,  T.  R.,  A.  S.  Perlov,  and  V.  A.  Vladimirov. 

1992.     Range-wide  survey  and  estimation  of  total  number 
of  Steller  sea  lions  in  1989.     Mar.  Mamm.  Sci.  8:220- 
239. 
Manly,  B.  F.  J. 

1997.  Randomization,  bootstrap  and  Monte  Carlo  meth- 
ods in  biology,  2nd  ed.,  399  p.  Chapman  and  Hall,  New 
York.  NY. 

Murie,  D.  J.,  and  D.  M.  Lavigne. 

1986.  Interpretation  of  otoliths  in  stomach  content  analy- 
ses of  phocid  seals  quantifying  fish  consumption.  Can. 
J.  Zool.  64:1152-1157. 

Olesiuk,  P.  F.,  M.  A.  Bigg,  G.  M.  Ellis,  S.  J.  Crockford,  and 
R.  J.  Wigen. 

1990.  An  assessment  of  the  feeding  habits  of  harbour 
seals  [Phoca  vitulina)  in  the  Strait  of  Georgia,  British 
Columbia,  based  on  scat  analysis.  Can.  Tech.  Rep. 
Fish.  Aquat.  Sci.  1730.  135  p. 

Pitcher.  K.  W. 

1981.     Prey  of  the  Steller  sea  lion.  Eumetopias  jubatus, 
in  the  Gulf  of  Alaska.     Fish.  Bull.  79:467-472. 
Prime,  J.  H.,  and  P.  S.  Hammond. 

1987.  Quantitative  assessment  of  gray  seal  diet  from 
fecal  analysis.  In  Approaches  to  marine  mammal  ener- 
getics (A.  C.  Huntley,  D.  P.  Costa.  G.  A.  J.  Worthy, 
and  M.  A.  Castellini.  eds.),  p.  165-181.  Allen  Press, 
Lawrence,  KS. 


508 


Fishery  Bulletin  102(3) 


Reynolds,  J.  O,  and  N.  J.  Aebischer. 

1991.     Comparison  and  quantification  of  carnivore  diet 
by  fecal  analysis — a  critique,  with  recommendations, 
based  on  a  study  of  the  fox  Vulpes  vulpes.     Mamm.  Rev. 
21:97-122. 
Rojo,  A.  L. 

1991.     Dictionary  of  evolutionary  fish  osteology,  273  p. 
CRC  Press,  Boca  Raton. 
Siegel,  S.,  and  N.  J.  Castellan. 

1988.     Nonparametric  statistics  for  the  behavioral  sci- 
ences, 2nd  ed.,  399  p.     McGraw-Hill,  New  York,  NY. 
Sinclair,  E.  H.,  G.  A.  Antonelis.  B.  W.  Robson,  R.  R.  Ream,  and 
T.  R.  Loughlin. 

1996.  Northern  fur  seal,  Callorhinus  ursinus,  predation 
on  juvenile  walleye  pollock,  Theragra  chalcogramma. 
In  Ecology  of  juvenile  walleye  pollock,  Theragra  chalco- 
gramma. Papers  from  the  workshop  "The  importance  of 
prerecruit  walleye  pollock  to  the  Bering  Sea  and  North 
Pacific  ecosystems"  Seattle,  WA,  October  28-30,  1993 
iR.  D.  Brodeur,  P.  A.  Livingston,  T.  R.  Loughlin,  and 
A.  B.  Hollowed,  eds.),  p. 167-178.  NOAA  Tech.  Rep. 
NMFS  126.  [NTIS  No.  PB97-155188.] 
Sinclair,  E.,  T.  Loughlin,  and  W.  Pearcy. 

1994.     Prey  selection  by  northern  fur  seals  {Callorhi- 
nus ursinus)  in  the  eastern  Bering  Sea.     Fish.  Bull. 
92:144-156. 
Sinclair,  E.  H..  and  T.  K.  Zeppelin. 

2002.     Seasonal  and  spatial  differences  in  diet  in  the  west- 
ern stock  of  Steller  sea  lions  (Eumetopias  jubatus).     J. 
Mammal.  83(4):973-990. 
Tollit,  D.  J.,  S.  G.  Heaslip,  and  A.  W.  Trites. 

2004.     Sizes  of  walleye  pollock  (Theragra  chalcogramma) 


consumed  by  the  eastern  stock  of  Steller  sea  lions 
(Eumetopias  jubatus)  in  Southeast  Alaska  from  1994 
to  1999.     Fish.  Bull.  102:522-532. 
Tollit,  D.  J.,  M.  J.  Steward,  P.  M.  Thompson,  G.  J.  Pierce, 
M.  B.  Santos,  and  S.  Hughes. 

1997.     Species  and  size  differences  in  the  digestion  of 
otoliths  and  beaks:  implications  for  estimates  of  pin- 
niped diet  composition.     Can.  J.  Fish.  Aquat.  Sci. 
54:105-119. 
Tollit,  D.  J.,  and  P.  M.  Thompson. 

1996.     Seasonal  and  between-year  variations  in  the  diet 
of  harbour  seals  in  the  Moray  Firth,  Scotland.     Can. 
J.  Zool.  74:1110-1121. 
Tollit,  D.  J.,  M.  Wong.  A.  J.  Winship,  D.  A.  S.  Rosen,  and 
A.  W.  Trites. 

2003.  Quantifying  errors  associated  with  using  prey 
skeletal  structures  from  fecal  samples  to  determine  the 
diet  of  the  Steller's  sea  lion  (Eumetopias  jubatus).  Mar. 
Mamm.  Sci.  19l4):724-744. 

Trites,  A.  W.,  and  P.  A.  Larkin. 

1996.     Changes  in  the  abundance  of  Steller  sea  lions 
(Eumetopias  jubatus)  in  Alaska  from  1956  to  1992:  how 
many  were  there?     Aquat.  Mamm.  22:153-166. 
Zar.  J.  H. 

1984.     Biostatistical  analysis,  2nd  ed.,  718  p.     Prentice- 
Hall,  Englewood  Cliffs,  N.J. 
Zeppelin,  T.  K.,  D.  J.  Tollit.  K.  A.  Call.  T.  J.  Orchard,  and 
C.  J.  Gudmundson. 

2004.  Sizes  of  walleye  pollock  (Theragra  chalcogramma  i 
and  Atka  mackerel  (Pleurogrammus  monopterygius)  con- 
sumed by  the  western  stock  of  Steller  sea  lions  (Eume- 
topias jubatus)  in  Alaska  from  1998  to  2000.  Fish. 
Bull.  102:509-521. 


509 


Abstract— Prey-size  selectivity  by 
Steller  sea  lions  lEumetopias  juba- 
tus)  is  relevant  for  understanding 
the  foraging  behavior  of  this  declin- 
ing predator,  but  studies  have  been 
problematic  because  of  the  absence 
and  erosion  of  otoliths  usually  used 
to  estimate  fish  length.  Therefore, 
we  developed  regression  formulae  to 
estimate  fish  length  from  seven  diag- 
nostic cranial  structures  of  walleye 
pollock  (Theragra  chalcogramma) 
and  Atka  mackerel  (Pleurogrammus 
monopterygius).  For  both  species, 
all  structure  measurements  were 
related  with  fork  length  of  prey  (r2 
range:  0.78-0.99).  Fork  length  (FL) 
of  walleye  pollock  and  Atka  mackerel 
consumed  by  Steller  sea  lions  was 
estimated  by  applying  these  regres- 
sion models  to  cranial  structures 
recovered  from  scats  (feces)  collected 
between  1998  and  2000  across  the 
range  of  the  Alaskan  western  stock 
of  Steller  sea  lions.  Experimentally 
derived  digestion  correction  factors 
were  applied  to  take  into  account  loss 
of  size  due  to  digestion.  Fork  lengths 
of  walleye  pollock  consumed  by  Steller 
sea  lions  ranged  from  3.7  to  70.8  cm 
(mean=39.3  cm,  SD  =  14.3  cm,  n  =  666l 
and  Atka  mackerel  ranged  from  15.3 
to  49.6  cm  (mean  =  32.3  cm,  SD  = 
5.9  cm,  rc  =  1685).  Although  sample 
sizes  were  limited,  a  greater  propor- 
tion of  juvenile  (<20  cm)  walleye  pol- 
lock were  found  in  samples  collected 
during  the  summer  (June-September) 
on  haul-out  sites  (64^  juveniles,  ;;=11 
scats)  than  on  summer  rookeries  (9% 
juveniles,  n  =  132  scats)  or  winter 
l  February-March)  haul-out  sites 
(3%  juveniles,  n  =  69  scats).  Annual 
changes  in  the  size  of  Atka  mackerel 
consumed  by  Steller  sea  lions  cor- 
responded to  changes  in  the  length 
distribution  of  Atka  mackerel  result- 
ing from  exceptionally  strong  year 
classes.  Considerable  overlap  (>51%) 
in  the  size  of  walleye  pollock  and  Atka 
mackerel  taken  by  Steller  sea  lions 
and  the  sizes  of  these  species  caught 
by  the  commercial  trawl  fishery  were 
demonstrated. 


Sizes  of  walleye  pollock 

(Theragra  chalcogramma)  and  Atka  mackerel 
(Pleurogrammus  monopterygius)  consumed  by 
the  western  stock  of  Steller  sea  lions 
(Eumetopias  jubatus)  in  Alaska  from  1998  to  2000 

Tonya  K.  Zeppelin' 
Dominic  J.  Tollit2 
Katherine  A.  Call1 
Trevor  J.  Orchard3 
Carolyn  J.  Gudmundson' 

E-mail  address:  Tonya  Zeppelin ifflnoaa  gov 

1  National  Marine  Mammal  Laboratory 
Alaska  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
7600  Sand  Point  Way  NE 

Seattle,  Washington  98115 

2  Marine  Mammal  Research  Unit 
Fisheries  Center,  Room  18,  Hut  B-3 
University  ot  British  Columbia 
6248  Biological  Sciences  Road 
Vancouver,  British  Columbia,  Canada  V6T  1Z4 

3  Department  of  Anthropology 
University  of  Toronto 

100  St.  George  Street 

Toronto,  Ontario,  Canada  M5S  3G3 


Manuscript  submitted  28  April  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
25  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:509-521  (2004). 


The  western  stock  of  Steller  sea  lions 
(Eumetopias  jubatus)  in  the  Gulf  of 
Alaska  and  the  Bering  Sea  has  experi- 
enced dramatic  and  continued  declines 
since  the  mid-1970s  (Loughlin  et  al„ 
1992;  Loughlin  and  York,  2000).  It  is 
likely  that  changes  in  prey  availabil- 
ity linked  to  commercial  fisheries  and 
large-scale  oceanographic  changes  are 
among  the  reasons  for  the  continued 
decline  (Loughlin  and  Merrick,  1989; 
NRC,  1996).  The  diet  of  the  western 
stock  of  Steller  sea  lions  has  been 
recently  assessed  (Sinclair  and  Zep- 
pelin, 2002),  but  discrete  selection  of 
prey  by  size  has  not  been  described. 
The  size  of  prey  is  relevant  for  under- 
standing the  foraging  behavior  of  the 
predator  as  well  as  the  ecological  role 
of  the  prey  (e.g.,  mortality  at  a  given 
life  history  stage).  In  the  case  of  the 
Steller  sea  lion,  prey-size  selectivity  is 
particularly  important  for  understand- 
ing spatial  and  temporal  changes  in 


diet  and  is  needed  for  making  fishery 
management  decisions. 

Size  of  fish  prey  consumed  by  ma- 
rine mammals  has  been  estimated 
by  using  sagittal  otoliths  recovered 
from  stomach  and  more  recently  scat 
samples  (Pitcher,  1981;  Frost  and 
Lowry,  1986;  Browne  et  al.,  2002). 
Significant  relationships  have  been 
demonstrated  between  fish  fork  length 
(FL)  and  otolith  length  (Templeman 
and  Squires,  1956;  Frost  and  Low- 
ry, 1981;  Harvey  et  al.,  2000).  The 
use  of  otoliths  to  describe  the  size  of 
prey  taken  by  Steller  sea  lions  has 
proved  useful  in  data  collected  from 
stomach  samples  (e.g.,  Pitcher,  1981; 
Calkins  and  Goodwin1).  However,  few 

1  Calkins,  D.  G.,  and  E.  Goodwin.  1988. 
Unpubl.  report.  Investigation  of  the 
declining  sea  lion  population  in  the  Gulf 
of  Alaska,  76  p.  Alaska  Department  of 
Fish  and  Game,  333  Raspberry  Road, 
Anchorage,  Alaska,  99518-1599. 


510 


Fishery  Bulletin  102(3) 


otoliths  are  recovered  from  Steller  sea  lion  scat,  and 
measurements  of  otoliths  recovered  from  scats  likely 
underestimate  prey  size  because  of  partial  erosion 
from  digestion  (Prime  and  Hammond,  1987;  Del- 
linger  and  Trillmich,  1988;  Harvey,  1989).  Because 
of  the  impracticality  of  collecting  stomachs  and  the 
low  number  and  poor  quality  of  otoliths  found  in 
scats,  alternative  methods  are  needed  to  accurately 
describe  the  size  of  prey  consumed  by  Steller  sea 
lions. 

Archaeological  studies  routinely  use  skeletal  struc- 
tures other  than  otoliths  to  estimate  either  fish 
length  or  mass  (Keys,  1928;  Casteel,  1976;  Owen 
and  Merrick,  1994;  Desse  and  Desse-Berset,  1996). 
Wise  11980)  used  a  regression  offish  length  on  ver- 
tebrae length  to  estimate  prey  size  from  scat  samples 
of  otters  (Lutra  lutra)  and  mink  tMustela  vison). 
The  regression  approach  relies  on  the  assumption 
that  the  overall  size  of  a  given  fish  and  the  size  of 
skeletal  structures  are  highly  correlated.  This  as- 
sumption has  been  substantiated  for  cranial  and 
skeletal  structures  other  than  otoliths  in  various 
North  Pacific  fish  species  (Orchard,  2001).  Thus,  the 
use  of  cranial  structures  appear  to  be  a  viable  alter- 
native to  the  use  of  otoliths  for  studying  prey  size  of 
Steller  sea  lions. 

Walleye  pollock  (Theragra  chalcogramma)  and  At- 
ka  mackerel  (Pleurogrammus  monopterygius)  rank 
among  the  top  prey  items  of  Steller  sea  lions  (Sin- 
clair and  Zeppelin,  2002)  as  well  as  being  valuable 
in  the  U.S.  commercial  fishery  (NMFS,  2003).  We 
estimated  fork  length  for  these  two  primary  prey 
species  from  scats  collected  between  1998  and  2000 
across  the  range  of  the  Alaskan  western  stock  of  sea 
lions.  Fish  length  was  estimated  by  using  regres- 
sion formulae  relating  bone  or  otolith  measurement 
to  fork  length  for  seven  cranial  structures  found  in 
sufficient  quantities  and  in  good  and  fair  condition  in 
scat  samples.  Experimentally  derived  digestion  cor- 
rection factors  (Tollit  et  al.,  2004b,  this  issue)  were 
applied  to  bone  and  otolith  measurements  to  account 
for  loss  of  size  due  to  erosion.  The  methods  developed 
here  proved  to  be  an  effective  tool  to  estimate  size  of 
prey  selected  by  Steller  sea  lions  and  are  applicable 
for  other  marine  mammal  diet  studies  particularly 
where  otoliths  are  highly  eroded. 


Materials  and  methods 

Development  of  regression  formulae 

Fork-length  to  bone  and  otolith-length  regression 
equations  were  developed  for  seven  cranial  struc- 
tures from  walleye  pollock  and  Atka  mackerel.  Bones 
and  otoliths  were  selected  according  to  species-specific 
features,  predictability  in  condition,  and  prevalence 
in  scats.  Bones  included  the  angular  (ANG),  quadrate 
(QUAD),  interhyal  (INTE),  dentary  (DENT),  pharyn- 
gobranchial  2  (PHAR),  and  hypobranchial  3  (HYPO) 


A    Walleye  pollock 


Pharyngobranchial  #2 


Quadrate 


Interhyal 


B     Atka  mackerel 


Otolith 


I 

Hypobranchial  #3 


Pharyngobranchial  #2 


Interhyal 


Otolith 


Figure  1 

Illustrations  of  the  various  planes  for  bone  and  otolith 
measurements  used  to  solve  the  bone-length  to  fish-length 
regression  equations  for  (A)  walleye  pollock  and  (Bl  Atka 
mackerel.  The  structures  from  the  right  side  of  the  body 
are  shown  for  all  structures  except  for  quadrates. 


(Fig.  1).  Fork  length  regressions  were  developed  for  sagit- 
tal otolith  length  (OTOL),  as  well  as  for  width  (OTOW) 
measurements.  All  selected  cranial  structures  were 
paired  (having  a  left  and  right  side)  which  allowed  for 
enumeration  of  prey  species.  Only  right-sided  bones  and 
otoliths  were  used  to  develop  the  regression  equations. 


Zeppelin  et  al.:   Sizes  of  walleye  pollock  and  Atka  mackerel  consumed  by  Eumetopias  jubatus 


511 


In  symmetrical  fishes  such  as  walleye  pollock  and  Atka 
mackerel  the  left  and  right  otoliths  are  mirror  images 
of  each  other  (Harkonen,  1986).  We  compared  the  left 
and  right-sided  measurements  for  all  seven  structures 
using  a  subsample  of  the  structures  used  to  develop  the 
regression  equations.  There  was  no  significant  difference 
for  either  walleye  pollock  (paired  Ntest,  P<0.05,  /?  =  13  for 
HYPO,  15  for  QUAD,  and  14  for  all  other  structures)  or 
Atka  mackerel  (paired  t-test,  P<0.05,  ra=14  for  OTOS 
and  17  for  all  other  structures). 

Fish  specimens  used  for  regressions  were  collected 
from  the  Gulf  of  Alaska  and  Bering  Sea.  Standard 
length  (SLi  was  converted  to  fork  length  for  walleye 
pollock  (when  fork  length  was  not  available  for  a  small 
number  of  otoliths  included  in  the  regressions)  by  using 
the  following  equation:  FL  =  0.40+1. 07(SL)  (Wilson2). 
We  chose  to  use  FL  over  SL  for  the  regressions  because 
all  fish  were  in  good  condition,  thus  allowing  for  ac- 
curate measurements.  Additionally,  FL  is  the  standard 
used  for  commercial  fishery  and  survey  data  by  the 
National  Marine  Fisheries  Service  for  direct  compari- 
sons. A  partial  analysis  of  these  data  was  previously 
reported  in  Orchard  (2001).  We  expanded  the  data  set 
reported  in  Orchard  (2001)  to  reflect  the  size  range  of 
bones  found  in  Steller  sea  lion  scats  and  included  only 
fish  specimens  collected  within  our  study  area. 

Linear  regression  models  were  fitted  for  most  cranial 
structures  by  using  the  following  equation: 

Y=  a+  PX, 

where   Y  =  the  fork  length  of  the  fish; 

X  =  the  measurement  of  the  cranial  structure; 
and 
a  and  P   are  constants  that  define  the  regression 
formula. 

However,  some  cranial  structures  provided  a  better  fit 
with  the  following  quadratic  regression  equation: 

Y  =  a  +  PX  +  pX2. 

The  strength  of  the  relationship  of  the  regression  models 
was  assessed  by  using  a  coefficient  of  determination 
(r2). 

Erosion  is  a  potential  source  of  bias  when  estimating 
prey  body  size  from  digested  otoliths  (Prime  and  Ham- 
mond, 1987;  Dellinger  and  Trillmich,  1988;  Harvey, 
1989).  We  used  condition-specific  digestion  correction 
factors  (DCFs)  developed  by  Tollit  et  al.  (2004b,  this 
issue)  to  correct  for  the  high  degree  of  variation  in  the 
erosion  of  cranial  structures.  DCFs  were  obtained  from 
feeding  experiments  on  captive  juvenile  Steller  sea  lions 
by  using  a  subsample  of  fish  collected  for  the  regres- 
sion analysis  (Tollit  et  al.,  2004b,  this  issue).  Selected 
cranial  structures  from  three  size  groups  of  pollock 


2  Wilson,  M.     2003.     Persona]  commun.     Alaska  Fisheries  Sci- 
ence Center,  Natl.  Mar.  Fish.  Serv.,  NOAA.  Seattle,  WA. 


(28.5-45.0  cm  FL)  and  one  size  group  of  Atka  mackerel 
(30-36  cm  FL)  were  used  to  develop  the  DCFs. 

Estimation  of  size  of  walleye  pollock  and  Atka  mackerel 
consumed  by  Steller  sea  lions  in  the  Bering  Sea  and 
Gulf  of  Alaska 

Steller  sea  lion  scats  were  collected  from  1998  to  2000 
along  most  of  the  U.S.  range  of  the  Alaskan  western  stock. 
Scats  were  collected  from  rookery  (breeding)  and  haul- 
out  (nonbreeding)  sites  in  summer  (June-September) 
and  haul-out  sites  in  winter  (February-March).  We 
assumed  that  scats  collected  on  summer  rookery  sites 
primarily  represent  the  diet  of  adult  females  because 
adult  males  present  on  rookeries  usually  fast  during 
this  time.  Juveniles  of  both  sexes  come  ashore  on  rook- 
eries during  summer  and  undoubtedly  are  represented 
in  the  data,  but  to  a  lesser  degree  than  adult  females. 
Scats  from  juvenile  Steller  sea  lions  are  more  likely  to 
be  sampled  on  haul-out  sites  during  summer,  where 
juveniles  make  up  the  greatest  proportion  of  individuals. 
Scats  collected  on  summer  haul-out  sites  or  any  winter 
site  presumably  represent  a  greater  cross-section  of 
ages  and  sexes  than  collections  from  rookeries  during 
summer. 

Scats  were  rinsed  through  nested  sieves  of  4.8-,  1.4-, 
0.7-,  and  0.5-mm  mesh.  Bones  and  otoliths  were  iden- 
tified to  the  lowest  possible  taxon  by  using  reference 
collection  specimens.  All  recovered  otoliths  and  selected 
bones  identified  as  either  walleye  pollock  or  Atka  mack- 
erel were  given  a  condition  grade  based  on  the  degree  of 
erosion  (Tollit  et  al.,  2004b,  this  issue).  In  general,  cra- 
nial structures  considered  in  "good"  condition  had  little 
or  no  erosion,  "fair"  were  moderately  eroded  (generally 
up  to  about  20%),  and  "poor"  were  heavily  digested 
(Tollit  et  al.,  2004b,  this  issue).  All  structures  that  were 
given  a  condition  grade  of  "good"  or  "fair"  were  identi- 
fied as  being  from  the  left  or  right  side  and  measured 
to  the  nearest  0.01  mm  with  digital  calipers.  Cranial 
structures  graded  as  "poor"  were  not  measured  and  ex- 
cluded from  further  analyses  because  of  high  observed 
intraspecific  variation  (Tollit  et  al.,  1997;  Tollit  et  al., 
2004b,  this  issue). 

Fork-length  estimates  with  and  without  DCFs  applied 
were  calculated  for  each  cranial  structure  and  for  all 
structures  combined.  Otoliths  were  treated  separate- 
ly because  most  diet  studies  currently  rely  on  otolith 
length  to  estimate  fish  fork  length.  Ninety-five  percent 
confidence  intervals  around  all  mean  size  estimates 
were  calculated  by  using  parametric  bootstrapping  pro- 
cedures (Manly,  1997)  in  which  error  associated  with 
the  regression  equation  and  resampling  error  resulting 
from  variability  within  correction  factors,  and  variabil- 
ity in  scats  were  taken  into  account.  Full  details  of  the 
bootstrapping  procedure  are  presented  in  Tollit  et  al. 
(2004b,  this  issue). 

The  same  fish  may  be  represented  by  multiple  cranial 
structures  within  a  scat;  therefore,  in  order  to  avoid 
pseudoreplication.  we  selected  a  minimum  number  of 
individuals  (MNI;  Ringrose,  1993)  for  each  scat  sample. 


512 


Fishery  Bulletin  102(3) 


Table  1 

Relationship  between  bone  measui 

•ement  and  fish  fork  length  (FL)  in 

millimeters.  For 

each  equation 

the  number  of  bones  mea- 

sured  In),  coefficient  of  determination  (r2),  standard  error  of  the  regression  coefficient  (SE  and  SE2for 

quadratic  regression  coef- 

ficients),  range  offish  lengths  and 

mean  of  fork  lengths  are  g 

ven.  All  measurements  are  given  in  millimeters. 

Species 

Structure  code 

Regression 

r2 

ii 

SE, SE2 

Range  of  FL 

Mean  FL 

Walleye  pollock 

INTE 

FL  =  49.78* +  5.12 

0.98 

49 

1.12 

83-477 

201.61 

HYPO 

FL  =  43.14*  +  14.12 

0.99 

49 

0.78 

83-477 

231.58 

PHAR 

FL  =  80.19a-  +  19.43 

0.95 

51 

2.58 

83-477 

204.37 

ANGU 

FL  =  59.25*  +15.27 

0.96 

44 

1.82 

83-477 

208.75 

QUAD 

FL  =  89.47*  +  6.77 

0.99 

59 

1.32 

83-477 

203.92 

DENT 

FL  =  108.46x- 1.52 

0.99 

60 

1.75 

83-477 

206.61 

OTOL 

FL  =  0.50*2  +  15.74*  + 

13.3 

0.99 

504 

0.68,  0.34 

49-530 

187.35 

OTOW 

FL  =  2.32*2  +  44.74*  + 

3.73 

0.99 

508 

1.54,0.19 

49-530 

188.66 

Atka  mackerel 

INTE 

FL  =  57.38*  +  95.57 

0.86 

106 

2.26 

185-500 

355.37 

HYPO 

FL  =  38.58*  80.64 

0.95 

105 

0.85 

185-500 

355.62 

PHAR 

FL  =  81.32*  +  70.40 

0.91 

107 

2.48 

185-500 

354.90 

ANGU 

FL  =  58.38*  +  73.86 

0.91 

105 

1.85 

185-500 

355.34 

QUAD 

FL  =  -8.90*2  +  129.38* 

+  9.16 

0.96 

108 

7.07,0.96 

185-500 

354.69 

DENT 

FL  = -7.10*2  +  115.83* 

-21.68 

0.94 

108 

7.08,  0.73 

185-500 

354.69 

OTOL 

FL  =  62.54*  +24.24 

0.83 

165 

2.19 

185-500 

349.82 

OTOW 

FL=  188.19* -77.71 

0.78 

170 

7.71 

185-500 

350.09 

Minimum  number  of  individuals  for  each  species  in 
each  scat  was  estimated  by  counting  species-specific 
sided  elements  and  choosing  the  greatest  number  of  left 
or  right  elements.  If  more  than  one  structure  had  the 
same  number,  the  structure  with  the  highest  r2  value 
in  its  regression  on  fork  length  (Table  1)  was  selected 
as  a  representative  length  estimate  for  that  fish.  If  an 
equal  number  of  left  and  right  bones  were  present,  right 
bones  were  selected. 

Temporal  variation  in  size  of  walleye  pollock  and 
Atka  mackerel  consumed  by  Steller  sea  lions 

Temporal  differences  were  assessed  by  grouping  fish 
into  stage-class  categories.  Stage-class  categories  were 
defined  for  pollock  as  follows:  juvenile  or  1-year-old 
fish  (<20  cm  FL),  adolescent  (20.1-34  cm  FL),  subadult 
(34.1-45  cm  FL),  and  adult  (>45.1  cm  FL;  Dorn  et  al., 
2001;  Smith,  1981;  Walline,  1983).  Walleye  pollock  sub- 
adults  are  likely  3-4  years  old,  of  which  -50%  have 
matured  and  recruited  into  the  fishery,  whereas  adults 
are  sexually  mature  fish,  likely  5  years  or  older.  Stage- 
class  categories  for  Atka  mackeral  were  defined  as  fol- 
lows: juvenile  up  to  2-year-old  fish  (<30  cm),  adolescent 
or  3-year-old  fish  (30.1-35.2  cm),  subadult  or  4-year-old 
fish  (35.3-45  cm),  and  adults  (>45.1  cm;  Lowe  et  al., 
2001;  McDermott  and  Lowe,  1997).  Atka  mackerel  ado- 
lescents are  -50%  sexually  mature  and  adult-size  fish 
are  fully  mature. 

A  chi-squared  contingency  test  was  used  to  test  for 
differences  in  the  proportion  offish  stage-classes  occur- 


ring in  scats  among  rookeries  and  haul-out  sites,  years, 
and  seasons  by  using  corrected  fork-length  estimates 
from  all  cranial  structures  (S-PLUS  2000,  Mathsoft, 
Inc.,  Cambridge,  MA).  To  avoid  pseudoreplication,  we 
used  presence  or  absence  of  cranial  elements  of  a  stage 
class  in  a  scat  particilarily  because  multiple  elements 
from  the  same  stage-class  within  a  sample  may  not  be 
independent  (Hunt  et  al.,  1996).  By  using  presence- 
absence  data  we  also  avoided  the  problems  associated 
with  the  variability  in  passage  and  recovery  rates  of 
different  size  structures  (Tollit  et  al.,  1997).  Because 
sample  sizes  were  small,  juvenile  and  adolescent  wall- 
eye pollock  stage  classes  and  recruiting  adult  and  adult 
Atka  mackerel  stage  classes  were  combined  for  seasonal 
comparisons  among  years.  Fisher's  exact  test  was  used 
for  comparisons  when  samples  sizes  for  any  stage  class 
were  less  than  5  (S-PLUS  2000,  Mathsoft,  Inc.,  Cam- 
bridge, MA). 

We  obtained  size  composition  data  from  commercial 
bottom  trawls  of  walleye  pollock  and  Atka  mackerel 
from  the  NMFS  North  Pacific  Groundfish  Observer  Pro- 
gram. Data  were  divided  into  winter  (January-April  I 
and  fall  (August-November)  seasons  and  compared 
with  our  seasonal  scat  data  (February-March  and 
June-September).  The  percentage  of  overlap  in  sizes 
of  fish  caught  by  the  commercial  groundfish  fishery 
with  sizes  of  fish  consumed  by  Steller  sea  lions  was 
calculated  by  comparing  size-frequency  distributions. 
Two-cm  size  bins  were  used  for  the  overlap  calculation 
and  Steller  sea  lion  prey-size  data  were  rounded  to  the 
nearest  integer  to  be  consistent  with  the  fishery  data. 


Zeppelin  et  al.:   Sizes  of  walleye  pollock  and  Atka  mackerel  consumed  by  Eumetopias  lubatus 


513 


Results 

Regression  formulae 

A  total  of  517  pollock  and  191  Atka  mackerel  samples 
were  used  to  develop  the  regression  equations  of  bone 
and  otolith  measurement  to  fork  length.  The  sample  size 
and  range  of  fish  lengths  used  for  the  regressions  varied 
between  species  and  cranial  structures  (Table  1).  No 
clear  indications  of  sample  size  required  for  regression 
analysis  are  currently  provided  in  the  literature;  how- 
ever, Owen  and  Merrick  (1994)  recommend  a  minimum 
sample  size  of  30-40.  Sample  sizes  used  to  develop  equa- 
tions presented  here  ranged  from  44  to  508. 

In  general,  linear  models  were  used  for  regression 
equations;  however,  several  cranial  structures  were 
best  fitted  with  a  quadratic  model.  For  both  species, 
all  structures  were  strongly  related  to  fork  length  (r2 
range:  0.78-0.99;  Table  1).  The  regressions  encompassed 
the  majority  of  sizes  of  bones  and  otoliths  found  in  Stell- 
er  sea  lion  scat  samples  for  this  study.  However,  a  small 
proportion  of  walleye  pollock  bones  from  scats  were 
larger  than  those  used  to  develop  the  regressions. 

Frost  and  Lowry  (1981)  developed  otolith  linear  re- 
gression equations  for  walleye  pollock  from  the  Bering 
Sea  using  a  double-regression  approach  that  produced 
an  inflection  point  at  10  mm.  We  examined  the  double 
regression  approach  but  found  a  higher  degree  of  corre- 
lation using  a  quadratic  regression  model.  We  compared 
the  results  of  our  model  with  Frost  and  Lowry 's  (1981) 
model  and  found  that  estimated  fork  lengths  of  walleye 
pollock  differed  less  than  2  cm  across  the  length  range 
in  our  samples. 

Estimation  of  size  of  walleye  pollock  and  Atka  mackerel 
consumed  by  Steller  sea  lions  in  the  Bering  Sea 
and  Gulf  of  Alaska 

A  total  of  714  scats  from  39  sites  contained  3646  selected 
cranial  elements  from  either  walleye  pollock  or  Atka 
mackerel.  Of  those,  212  scats  contained  666  walleye 
pollock  cranial  elements  with  a  condition  grade  of  either 
"good"  (ft  =236)  or  "fair"  (n  =  430).  The  minimum  number 
of  individual  pollock  per  scat  ranged  from  1  to  18  with 
a  mean  of  1.6  (SD  =  1.7).  For  Atka  mackerel,  379  scats 
contained  1685  skeletal  elements  with  condition  grade 
of  either  "good"  (;?=755)  or  "fair"  (rc=930).  The  minimum 
number  of  individual  Atka  mackerel  per  scat  ranged 
from  1  to  14  with  a  mean  of  1.9  (SD  =  1.6). 

The  mean  fork  length  of  walleye  pollock  consumed  by 
Steller  sea  lions  in  the  Bering  Sea  and  Gulf  of  Alaska  es- 
timated from  uncorrected  otoliths  found  in  scats  was  23.7 
cm  (SD=12.0;  «  =  88).  Application  of  the  DCF  increased 
the  mean  estimate  to  28.4  cm  (SD=14.75;  rc=88).  The  size 
distribution  estimated  from  corrected  otoliths  had  three 
modes:  a  major  mode  around  32  cm  and  minor  modes 
around  5  cm  and  13  cm  (Fig.  2A).  Confidence  intervals  for 
all  grade-corrected  estimates  can  be  found  in  Table  1. 

The  mean  fork  length  of  walleye  pollock  estimated 
from  all  seven  structures  was  39.8%  greater  than  the 


mean  estimated  from  otoliths  alone.  The  uncorrected 
mean  was  33.1  cm.  Applying  the  DCF  increased  the 
mean  length  of  walleye  pollock  by  18.7%  to  39.3  cm 
(paired  t  test,  ?665=37.9,  P<0.001).  Mean  grade-corrected 
size  estimates  for  cranial  structures  other  than  otoliths 
ranged  from  34.5  cm  (PHAR)  to  47.2  cm  (HYPO)  and 
95%  confidence  intervals  ranged  from  25.2  to  50.6  cm 
(Table  2).  The  condition-specific  DCFs  increased  length 
estimates  between  6.8%  (HYPO)  and  28.3%  (DENT). 
The  size  distribution  estimated  from  all  grade-corrected 
structures  had  three  modes:  a  major  mode  around  44  cm 
and  minor  modes  around  5  cm  and  15  cm  (Fig.  2A). 

The  mean  fork  length  of  Atka  mackerel  consumed  by 
Steller  sea  lions  in  the  Bering  Sea  and  Gulf  of  Alas- 
ka estimated  from  uncorrected  otoliths  was  30.3  cm 
(SD=4.0;  n=117).  Application  of  the  DCF  increased  the 
mean  estimate  to  34.7  (SD  =  4.8;  n=U7). 

The  mean  fork  length  of  Atka  mackerel  estimated 
from  all  structures  (30.7  cm;  SD  =  5.9  cm,  corrected 
32.3  cm;  SD  =  5.9  cm,  rc=1685,  paired  t  test,  f1684=39.1, 
P<0.001)  was  similar  to  the  mean  estimated  from  oto- 
liths (6.9%  less  without  a  DCF  and  1.3%  less  with  a 
DCF;  Fig.  2B).  Mean  grade-corrected  size  estimates 
for  structures  other  than  otoliths  ranged  from  26.6  cm 
(QUAD)  to  34.2  cm  (INTE)  and  95%  confidence  inter- 
vals ranged  from  24.0  cm  (DENT)  to  35.0  cm  (INTE; 
Table  2).  Use  of  the  condition-specific  DCFs  increased 
length  estimates  between  2.1%  (INTE)  and  24.0% 
(DENT).  Fork  length  estimates  for  all  structures  did 
not  include  PHAR  because  too  few  were  recovered  in 
scats  in  the  feeding  studies  of  captive  Steller  sea  lions 
to  develop  a  correction  factor. 

When  mean  prey  size  was  calculated  by  using  MNI, 
the  mean  corrected  and  uncorrected  size  estimate  of 
both  walleye  pollock  and  Atka  mackerel  differed  by  less 
than  0.2  cm  from  estimates  derived  by  using  all  struc- 
tures. There  was  little  difference  in  the  standard  devia- 
tions or  distributions  when  MNI  estimates  were  used 
compared  with  all  structures  (Table  2).  Unsurprisingly, 
the  use  of  MNI  estimates  did  substantially  reduce  the 
sample  size  (336/666  for  walleye  pollock  and  722/1685 
for  Atka  mackerel). 

Spatial  and  temporal  variation  in  size  of  pollock  and 
Atka  mackerel  consumed  by  Steller  sea  lions 

No  statistical  difference  was  found  in  the  proportion  of 
pollock  stage  classes  among  years  on  summer  rookery 
sites  (P=0.29,  ^2  =  4.9,  df=3)  or  winter  haul-out  sites 
(P=0.10;  Fisher's  exact  test).  Scats  were  collected  only 
on  summer  haul-out  sites  during  2000.  Although  sample 
sizes  were  limited,  we  found  significant  differences  in 
the  proportion  of  pollock  stage  classes  between  summer 
rookery  and  haul-out  scats  (P=0.02;  Fisher's  exact  test) 
and  between  summer  and  winter  haul-out  sites  (P=0.018; 
Fisher's  exact  test)  for  year  2000.  A  greater  proportion 
of  juvenile  pollock  were  found  on  summer  haul-outs 
(64%' juveniles,  n=ll  scats)  than  on  summer  rookeries 
(9%  juveniles,  n  =  132  scats)  or  winter  haul-out  sites  (3% 
juveniles,  n  =  69  scats,  Fig.  3).  No  statistical  difference 


514 


Fishery  Bulletin  102(3) 


Walleye  pollock 
All  structures 


MNI 


8 

6  1 


n  =  666 


No  DCFs 


DCFs 

applied 


— i — i — i — i — i — i — i — 

0     10    20    30    40    50    60    70    80  0     10    20    30    40    50    60    70    80 
Estimated  fork  length  (cm) 


B     Atka  mackerel 

All  structures 


1NI 


No  DCFs 


DCFs 
applied 


0     10    20    30    40    50    60    70    80  0 


i — i — r 

10    20    30   40    50    60    70    80 


Estimated  fork  length  (cm) 

Figure  2 

Relative  frequency  histograms  of  the  estimated  fork  length  of  <A>  walleye  pol- 
lock and  iB)  Atka  mackerel  consumed  by  Steller  sea  lions.  Fork  lengths  were 
predicted  from  cranial  structures  in  good  and  fair  condition.  Comparisons 
were  made  on  the  application  of  correction  factors  (DCFs)  which  account  for 
digestion  and  for  using  minimum  number  I  MNI  I  estimates  as  a  selection  tech- 
nique versus  using  all  structures.  Otoliths  (black  bars)  are  stacked  beneath 
all  other  structures  (gray  bars). 


was  found  in  the  proportion  of  stage  classes  between 
summer  rookery  (9.09%  juvenile;  20.45%  adolescent: 
53.03$   subadult;  65.15'^   adult)  and  winter  haul-out 


(2.90%  juvenile;  21.74r;  adolescent;  57.97%  subadult; 
46.38%  adult)  sites  for  all  years  combined  (P=0.32, 
*2=2.3,  df=2). 


Zeppelin  et  al.:    Sizes  of  walleye  pollock  and  Atka  mackerel  consumed  by  Eumetopias  jubatus  515 


Table  2 

Estimated  mean  fork  length  of  walleye  pollock  and  Atka  mackerel  consumed  by  Steller  sea  lions  based  on  selected  structures 
with  or  without  application  of  condition-specific  digestion  correction  factors  (DCFsl.  Data  sets  exclude  all  structures  graded  as 
heavily  digested.  Remaining  total  sample  sizes  of  elements  in1)  are  given  along  with  proportion  of  grade  "good"  structures  (n«). 
For  data  sets  where  DCFs  were  applied,  95'r  confidence  intervals  (95%  CI)  were  estimated  by  using  bootstrap  resampling  pro- 
cedures (Tollit  et  al.,  2004b,  this  issue). 


Species 


Structure  code 


DCF 


/!' 


Mean  FLlcml 


SDlcm)        Range  (cm) 


95%  CI 


Walleye  pollock 


Atka  mackerel 


INTE 

HYPO 

PHAR 

ANGU 

QUAD 

DENT 

OTOL 

All 

INTE 

HYPO 

ANGU 

QUAD 

DENT 

OTOL 

All 


No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 

No 

Yes 


60 

60 

38 

38 

23 

23 

136 

136 

134 

134 

187 

187 

88 

88 

666 

666 

601 

601 

238 

238 

488 

488 

161 

161 

80 

80 

117 

117 

1685 

1685 


0.45 
0.45 
0.55 
0.55 
0.61 
0.61 
0.40 
0.40 
0.34 
0.34 
0.37 
0.37 
0.03 
0.03 
0.35 
0.35 

0.58 
0.58 
0.42 
0.42 
0.45 
0.45 
0.37 
0.37 
0.28 
0.28 
0.06 
0.06 
0.45 
0.45 


43.7 
47.0 
44.2 
47.2 
32.2 
34.5 
36.1 
40.2 
35.1 
44.5 
28.6 
36.7 
23.7 
28.4 
33.1 
39.3 

33.5 
34.2 
31.1 
32.9 
30.2 
31.8 
25.3 
26.6 
22.5 
27.9 
30.3 
34.7 
30.7 
32.3 


8.0 

8.5 

7.2 

7.8 

14.3 

14.8 

8.4 

9.0 

12.0 

15.3 

11.8 

15.1 

12.0 

14.8 

12.4 

14.3 

5.0 
5.1 
4.8 
5.5 
4.7 
5.1 
5.4 
5.6 
7.7 
8.0 
4.0 
4.8 
5.9 
5.9 


16.7- 

16.7- 

30.5- 

34.9- 

9.7- 

10.9- 

10.6- 

10.6- 

9.4- 

11.9- 

3.1- 

3.7- 

4.6- 

4.6- 

3.1- 

3.7- 

19.5- 
19.5- 
18.8- 
19.3- 
17.3- 
17.3- 
14.8- 
15.3- 
13.0- 
17.7- 
21.2- 
21.2 
13.0- 
15.3- 


59.4 
65.9 
60.4 
62.7 
53.1 
5.3.1 
55.3 
60.6 
■57.8 
■70.8 
-57.2 
■70.2 
46.8 
-57.1 
60.4 
-70.8 

46.8 

49.6 
46.2 
48.3 
43.0 
■46.1 
40.6 
-41.4 
-38.7 
■44.1 
-40.6 
-47.0 
■46.9 
-49.6 


44.9-49.8 
44.5-50.6 
25.2-44.5 
38.5-42.4 
38.8-49.6 
30.3-42.4 
17.0-32.4 
35.9-42.4 
33.4-35.0 
32.4-34.6 
31.7-33.3 
25.1-28.4 
24.0-33.0 
33.5-35.8 
31.7-33.4 


Significant  differences  were  found  in  the  proportion 
of  Atka  mackerel  stage  classes  between  1998  and  1999 
on  summer  rookery  sites  (P=0.05,  x2  =6.0,  df=2  )  and 
winter  haul-out  sites  (P=0.01,  ^  =  9.9,  df=2)  and  be- 
tween 1998  and  2000  winter  haul-out  sites  (P=<0.01, 
Fisher's  exact  test).  Significant  seasonal  differences 
were  found  only  in  1998  (P=0.03,  r  =  7.1,  df=2)  which 
may  be  the  result  of  the  small  sample  size  in  winter 
2000.  In  summer  and  winter,  annual  differences  in 
size  of  Atka  mackerel  consumed  by  Steller  sea  lions 
corresponded  to  changes  in  the  length  distribution  of 
Atka  mackerel  resulting  from  exceptionally  strong  year 
classes  in  1995  and  1998  (Lowe  et  al.,  2001).  The  1995 
year  class  is  represented  as  a  mode  around  30  cm  in 
1998  (3-year-old  fish),  35  cm  in  1999  and  >40  cm  in 


2000  (Fig.  4).  The  1998  year  class  is  represented  most 
clearly  as  2  year  olds  (mode  20-25  cm)  in  summer  2000 
(Fig.  4).  Strong  annual  modes  found  in  our  data  match 
those  recorded  in  surveys  of  Atka  mackerel  in  the  Ber- 
ing Sea  and  Gulf  of  Alaska  (Lowe  et  al.,  2001). 

For  walleye  pollock  and  Atka  mackerel  there  was  no 
difference  in  the  mean  size  of  fish  caught  by  the  com- 
mercial fishery  among  years  (P>0.4,  one-way  ANOVA). 
There  was  a  significant  difference  (P<0.05,  one-way 
ANOVA)  in  the  size  of  fish  caught  between  seasons. 
This  difference  is  likely  due  to  aggregations  of  spawning 
adult  fish  caught  during  the  roe  fishery.  In  the  winter 
there  is  a  56%  overlap  between  the  size  of  fish  caught  in 
the  commercial  pollock  fishery  and  those  taken  by  sea 
lions  and  a  54%  overlap  in  the  size  taken  by  the  Atka 


516 


Fishery  Bulletin  102(3) 


25  - 
20 
15 
10  -\ 

5 

0 


Winter  1998 
ns  =  5 
n  =  8 


25 

20 
15 
10 

5 

0 


0  102030405060  7080 


25  - 
20  - 
15   - 
10 
5 
0 


Summer  1998  (rookey) 
n,  =  81 
n„  =  226 


0  1020304050607080 


Winter  1999 
ns  =  24 
n  =46 


Winter  2000 
ns  =  40 
n  =96 


0  1020304050607080 


0  1020304050607080 

Summer  2000  (rookery) 
ns  =  24 
a,  =  81 


Summer  2000  (haul-out) 
ns=  1 1 
n„  =  77 

l 

j 

I       l!                                  I.       . 

0  1020304050607080 


0  1020304050607080 


0  1020304050607080 


Estimated  fork  length  (cm) 


Figure  3 

Relative  frequency  histograms  of  the  estimated  fork  length  of  walleye  pollock  consumed  by  Steller  sea  lions 
across  seasons  and  years  for  rookeries  and  haul-outs.  Fork  lengths  are  predicted  from  corrected  cranial  struc- 
tures in  good  and  fair  condition.  Sample  sizes  for  cranial  elements  inel  and  scats  (ra  )  are  provided.  All  winter 
sites  are  considered  haul-out  sites. 


Winter  1999 
ns  =  48 
n  =218 


Winter  2000 

ns  =  13 

n„  =  32 

- 

1 

"         1 

0    10  20  30  40  50 

Summer  1998 
_  ns  =  201 
ne  =  965 


0    10  20  30  40  50 

Summer  1999 
ns  =  44 
n  =193 


0    10  20  30  40  50 

Summer  2000 
n,  =  35 
n  .  =  142  J 


0    10  20  30  40  50 


0    10  20  30  40  50 
Estimated  fork  length  (cm) 


0    10  20  30  40  50 


Figure  4 

Relative  frequency  histograms  of  the  estimated  fork  length  of  Atka  mackerel  of 
consumed  by  Steller  sea  lions  by  season  and  year.  Fork  lengths  are  predicted  from 
corrected  cranial  structures  in  good  and  fair  condition.  Sample  sizes  for  cranial 
elements  (ne)  and  scats  (ns)  are  provided.  All  summer  sites  are  rookeries  and  winter 
sites  are  haul-out  sites. 


Zeppelin  et  al.:   Sizes  of  walleye  pollock  and  Atka  mackerel  consumed  by  Eumetopias  jubatus 


517 


12-1 

A   Walleye  pollock 

Overlap  68% 

10- 

SSL,  n=666 

■  trawl.  n=92133 

8- 

6  - 

4  - 

ll 

2- 

I    III         I 

ll.. 

n  - 

10 


20 


30 


40 


-      25 


20 


B   Atka  mackerel 


50  60  70 


Overlap  53°o 


80 


15 


10 


SSL.  n=1685 

■  trawl.  n=92877 

ill 

.III 

1. 

10  20  30 

Estimated  fork  length  (cm) 


40 


50 


Figure  5 

Relative  frequency  histograms  of  the  estimated  fork  length  of  wall- 
eye pollock  and  Atka  mackerel  consumed  by  Steller  sea  lions  (SSL) 
compared  with  relative  frequency  histograms  offish  caught  by  the 
walleye  pollock  and  Atka  mackerel  commercial  trawl  fishery. 


mackerel  fishery.  In  the  summer  the  overlap  in  size  of 
fish  consumed  by  sea  lions  and  the  size  of  fish  caught 
in  the  pollock  fishery  is  67%  and  there  is  a  51%  overlap 
in  the  size  of  fish  caught  in  the  Atka  mackerel  fishery. 
When  seasonal  data  were  pooled,  overlap  between  the 
size  of  fish  caught  in  the  commercial  fishery  and  the 
size  of  fish  consumed  by  sea  lions  was  68%  for  walleye 
pollock  (Fig.  5A)  and  53%  for  Atka  mackerel  (Fig.  5B). 


Discussion 

Regression  formulae 

Regressions  of  cranial  structure  measurement  on  fish 
fork  length  with  the  use  of  multiple  structures  was 
an  effective  tool  for  estimating  size  of  fish  consumed 
by  Steller  sea  lions.  Sample  sizes  of  measurable  prey 


remains  from  scats  were  enhanced  by  using  a  number 
of  cranial  structures  in  addition  to  otoliths.  Body  size 
estimates  of  only  13.2%  of  the  pollock  and  6.9%  of  the 
Atka  mackerel  prey  in  this  study  were  based  on  otoliths 
alone.  Fork-length  estimates  can  be  considered  accurate 
regardless  of  which  structure  was  used  in  the  estimate 
because  all  r2  values  were  high  (range:  0.78-0.99).  Like- 
wise errors  associated  with  the  application  of  DCFs 
were  consistent  across  structures  (Tollit  et  al.,  2004b, 
this  issue).  Confidence  intervals  around  size  estimates 
generally  overlapped  across  structures;  however,  it  was 
not  surprising  that  different  structures  yielded  slightly 
different  mean  sizes  because  different  bones  can  origi- 
nate from  different  scats. 

The  use  of  multiple  cranial  structures  may  also  re- 
duce bias  resulting  from  variability  in  recovery  and 
passage  rates  of  structures  from  different  species  or 
sizes  of  fish  (Pierce  and  Boyle,  1991;  Browne  et  al., 


518 


Fishery  Bulletin  102(3) 


2002;  Tollit  et  al.,  2003).  Even  after  applying  a  DCF, 
the  estimated  mean  size  of  walleye  pollock  based  on 
otoliths  was  10.9  cm  smaller  than  the  mean  size  esti- 
mated by  using  all  cranial  structures.  Because  walleye 
pollock  otoliths  are  relatively  large  and  have  a  different 
composition  than  other  cranial  structures,  the  larger 
otoliths  may  be  regurgitated,  fully  digested,  or  crushed 
by  rocks  in  the  stomach  and  not  pass  through  in  scat  as 
readily  as  smaller  otoliths  or  other  cranial  structures, 
thereby  reducing  their  occurrence  in  scat  and  use  in 
generating  prey-size  estimates.  Atka  mackerel  otoliths 
are  much  smaller  at  older  ages  in  relation  to  walleye 
pollock,  which  may  explain  why  the  size  of  prey  esti- 
mated from  otoliths  was  similar  to  the  size  estimated 
from  other  cranial  structures. 

The  use  of  DCFs  for  all  structures,  including  otoliths, 
to  account  for  erosion  increased  mean  size  estimates 
for  both  walleye  pollock  (33.1  vs.  39.3  cm  FL)  and  Atka 
mackerel  (30.7  vs.  32.3  cm  FLi.  The  relatively  small 
increase  in  the  corrected  size  of  Atka  mackerel  re- 
flects that  the  structures  from  this  species  were  found 
in  better  condition  than  those  from  pollock  (Table  2), 
as  well  as  that  correction  factors  were  found  to  be 
species-,  structure-,  and  condition-specific  (Tollit  et 
al.,  2004b,  this  issue).  Overall,  our  results  emphasize 
the  importance  of  using  appropriate  condition-specific 
DCFs.  Other  studies  with  captive  sea  lions  have  also 
demonstrated  that  grade-specific  DCFs  can  reduce  sys- 
tematic error  and  increase  precision  of  body  mass  es- 
timates (Tollit  et  al.  1997).  For  walleye  pollock,  there 
was  no  significant  difference  in  the  degree  of  erosion 
across  the  three  size  ranges  for  each  structure  within 
each  condition  category  (Tollit  et  al.,  2004b,  this  issue). 
We  assume  the  DCFs  can  be  used  for  fish  outside  of 
this  size  range  because  the  relative  shape,  structure, 
and  proportion  of  the  morphological  features  are  con- 
sistent for  both  smaller  and  larger  fish  (Tollit  et  al., 
2004b,  this  issue).  Further  research  is  necessary  to 
test  whether  there  are  differences  across  the  size  range 
for  Atka  mackerel. 

Size  of  walleye  pollock  and  Atka  mackerel  consumed  by 
Steller  sea  lions  in  the  Bering  Sea  and  Gulf  of  Alaska 

In  general,  Steller  sea  lions  on  summer  rookery  and 
winter  haul-out  sites  consumed  primarily  subadult  and 
adult-size  walleye  pollock  and  Atka  mackerel  year-round 
in  1998-2000.  Steller  sea  lions  typically  forage  near 
shore,  in  shallow  water  (<50  m)  and  at  night  (Raura- 
Suryan  et  al.,  2002;  Loughlin  et  al.,  2003).  Likewise, 
adult  walleye  pollock  migrate  vertically  to  shallower 
depths  during  the  night  (Smith,  1981).  Adult-size  Atka 
mackerel  also  are  commonly  found  in  nearshore  coastal 
areas  during  their  spawning  season  (Zolotov,  1993). 

Juvenile  walleye  pollock  were  found  in  relatively  high 
numbers  only  in  scats  collected  on  summer  haul-out 
sites.  Scats  collected  from  summer  haul-out  sites  likely 
represent  a  larger  proportion  of  juvenile  Steller  sea  li- 
ons than  those  collected  on  summer  rookery  or  winter 
haul-out  sites.  Previous  studies  indicate  t hat  juvenile 


sea  lions  (<4  years  old)  consume  smaller  walleye  pollock 
than  adult  sea  lions  (Pitcher,  1981;  Frost  and  Lowry, 
1986;  Merrick  and  Calkins,  1996).  Juvenile  walleye  pol- 
lock are  often  found  at  shallow  depths  in  bays  and  near 
shore  habitat  (Smith,  1981).  Likewise,  Loughlin  et  al. 
(2003)  reported  that  juvenile  Steller  sea  lions  are  typi- 
cally shallow  divers  and  frequently  make  short  range 
foraging  trips  (<15  km).  Additional  scat  collections  on 
summer  haul-out  sites  are  necessary  to  determine  more 
conclusively  prey-size  selectivity  for  juvenile  Steller  sea 
lions. 

Annual  changes  in  the  size-frequency  distribution  of 
Atka  mackerel  consumed  by  Steller  sea  lions  followed 
changes  in  the  size  distribution  of  Atka  mackerel  re- 
sulting from  exceptionally  strong  year  classes.  Merrick 
and  Calkins  (1996)  also  showed  that  the  size  of  prey 
consumed  by  Steller  sea  lions  can  reflect  the  size  dis- 
tribution of  the  fish  population.  From  the  mid-1990s  on, 
only  1999  was  a  strong  recruitment  year  for  walleye 
pollock  in  the  Gulf  of  Alaska  (Dorn  et  al.,  2001),  but  we 
did  not  find  a  significantly  greater  proportion  of  juvenile 
fish  eaten  by  Steller  sea  lions  in  2000  than  in  1999  or 
1998  perhaps  because  sufficient  numbers  of  larger  size 
fish  were  available  in  regions  where  walleye  pollock 
were  consumed. 

Historical  studies  of  Steller  sea  lion  prey  size  have 
primarily  been  based  on  measurements  of  walleye  pol- 
lock otoliths  found  in  stomach  samples  but  often  with- 
out application  of  correction  factors  for  erosion  (Pitcher, 
1981;  Merrick  and  Calkins,  1996;  Calkins.  1998).  Prey- 
size  estimates  based  on  stomach  contents  will  likely 
differ  from  estimates  derived  from  scats  because  of 
differences  in  digestion  rates  and  breakage  ( Jobling  and 
Breiby,  1986  i.  However,  results  of  studies  examining 
the  variability  in  prey  size  with  sample  type  are  vari- 
able. Sinclair  et  al.  (1996)  suggested  that  in  northern 
fur  seals  (Callorhinus  ursinus),  another  otariid,  small 
otoliths  tend  to  flush  through  the  digestive  system  more 
quickly  than  larger  ones,  resulting  in  a  possible  bias  in 
scats  towards  smaller  otoliths.  In  contrast,  experiments 
with  captive  sea  lions  have  shown  that  smaller  otoliths 
are  recovered  in  lower  relative  frequencies  than  are 
larger  ones  (Tollit  et  al..  1997).  Frost  and  Lowry  (1980) 
found  no  significant  difference  between  the  size  of  oto- 
liths obtained  from  stomach  and  intestines  of  ribbon 
seals.  Overall,  we  believe  useful  comparisons  of  prey 
size  consumed  by  Steller  sea  lions  can  be  made  between 
our  study  and  earlier  studies. 

Steller  sea  lions  have  been  reported  to  consume  a 
wide  size  range  of  walleye  pollock.  However,  in  most 
prior  studies  a  larger  proportion  of  juvenile  fish  were 
found  than  what  we  estimated  from  scats.  Otoliths  from 
stomach  samples  collected  from  1975  to  1978  in  the 
Gulf  of  Alaska  contained  primarily  juvenile  age  pollock 
(mean  FL=29.8cm;  SD=  11.6;  Pitcher,  1981).  Undigested 
otoliths  from  stomach  samples  collected  between  1975 
and  1981  in  the  Bering  Sea  also  contained  mostly  juve- 
nile fish  (mean  FL=29.3  cm)  but  had  a  distinct  mode  of 
adult-size  pollock  (48  cm  FL:  Frost  and  Lowry.  1986). 
Likewise,  43  stomach  samples  collected  between  1985 


Zeppelin  et  al.:   Sizes  of  walleye  pollock  and  Atka  mackerel  consumed  by  Eumetopias  jubatus 


519 


and  1986  in  the  central  Gulf  of  Alaska  contained  pri- 
marily juveniles  (mean  FL=25.4  cm;  SD  =  12.4)  and  had 
a  weak  mode  of  adult-size  fish  (39-43  cm;  Merrick  and 
Calkins,  1996;  Calkins  and  Goodwin1).  Mostly  adult-size 
fish  were  found  in  stomachs  recovered  from  Steller  sea 
lions  caught  in  trawl  nets  in  the  central  Gulf  of  Alaska 
11983-84;  Loughlin  and  Nelson,  1986)  and  in  stomach 
samples  collected  from  1994  to  1995  in  Japanese  waters 
(Goto  and  Shimazaki,  1998).  However,  in  both  these 
studies  the  samples  of  prey  size  may  have  been  biased 
by  the  selectivity  of  the  fishing  gear  for  larger  fish. 

Using  identical  methods  to  those  of  our  study,  Tollit 
et  al.  (2004a,  this  issue)  estimated  the  size  of  wall- 
eye pollock  consumed  by  the  eastern  stock  of  Steller 
sea  lions  between  1994  and  1999.  The  average  size  of 
walleye  pollock  consumed,  estimated  from  all  grade- 
corrected  structures  (mean=42.4  cm;  SD  =  11.6),  was 
similar  to  the  average  size  found  in  our  study  of  the 
western  stock  of  Steller  sea  lions.  Furthermore,  Tollit 
et  al.  (2004a,  this  issue)  also  found  a  greater  occur- 
rence of  adult  pollock  in  scats  collected  on  rookery  sites 
than  from  scats  collected  on  haul-out  sites.  However, 
Steller  sea  lions  from  the  western  stock  consumed  a 
greater  proportion  of  juvenile  and  adolescent  fish  and 
less  adult  fish  than  those  from  the  eastern  stock  dur- 
ing summer  (June-July)  and  similar-size  fish  were 
consumed  on  haul-out  sites  in  winter  (March)  in  both 
regions.  Neither  study  indicated  the  high  occurrence 
of  juvenile  walleye  pollock  reported  in  the  1970s  and 
1980s.  The  greater  occurrence  of  juvenile  walleye  pol- 
lock in  historical  studies  may  be  a  result  of  prey  avail- 
ability or  differences  in  sampling  methods. 

By  examining  the  relative  size-frequency  distributions 
of  prey  selected  by  Steller  sea  lions  and  those  taken 
in  the  commercial  trawl  fishery,  we  found  considerable 
overlap  (689r  walleye  pollock  and  53%  Atka  mackerel). 
Likewise,  high  levels  of  potential  overlap  in  size  were 
found  between  walleye  pollock  selected  by  Steller  sea 
lions  from  the  eastern  stock  and  caught  by  the  small 
commercial  fishery  bordering  Southeast  Alaska  (Tollit 
et  al.,  2004a,  this  issue).  The  extent  of  overlap  through- 
out the  range  of  Steller  sea  lions  between  the  size  of 
prey  consumed  by  sea  lions  and  the  size  of  fish  targeted 
and  taken  by  the  pollock  and  Atka  mackerel  trawl  fish- 
eries could  result  in  competition  between  fisheries  and 
foraging  sea  lions  if  resources  are  limited. 


Acknowledgments 

Fish  specimens  for  the  regression  equations  were  pro- 
vided by  the  National  Marine  Fisheries  Service,  the 
University  of  Victoria,  and  the  University  of  British 
Columbia  bone  reference  collections.  Fish  remains  were 
identified  by  Pacific  Identifications,  Victoria,  BC.  We 
thank  J.  Laake,  A.  York,  and  R.  Joy  for  statistical  advice, 
K.  Chumbley,  E.  Sinclair,  and  S.  Crockford  for  help  in 
initiating  the  study,  A.  Browne  and  M.  Wilson  for  wall- 
eye pollock  otolith  data,  and  S.  Heaslip  for  graphics  and 
laboratory  assistance.  Reviews  by  E.  Sinclair,  S.  Melin, 


W.  Walker,  B.  Robson,  and  three  anonymous  reviewers 
greatly  improved  this  manuscript. 


Literature  cited 

Browne,  P.,  J.  L.  Laake,  and  R.  L.  DeLong. 

2002.     Improving  pinniped  diet  analyses  through  iden- 
tification of  multiple  skeletal  elements  in  fecal  sam- 
ples.    Fish.  Bull.  100:423-433. 
Calkins,  D.  G. 

1998.     Prey  of  Steller  sea  lions  in  the  Bering  Sea.     Biosph. 
Cons.  1(11:33-44. 
Casteel,  R.  W. 

1976.     Fish  remains  in  archaeology  and  paleo-environ- 
mental  studies,  180  p.     Acad.  Press.  London. 
Desse,  J.,  and  N.  Desse-Berset, 

1996.     On  the  boundaries   of  osteometry  applied   to 
fish.     Archaeofauna  5:171-179. 
Dellinger,  T.,  and  F.  Trillmich. 

1988.  Estimating  diet  composition  from  scat  analysis  in 
otariid:  is  it  reliable?     Can.  J.  Zool.  66:1865-1870. 

Dorn,  M.,  A.  Hollowed.  E.  Brown,  B.  Megrey,  C.  Wilson,  and 
J.  Blackburn. 

2001.  Assessment  of  the  walleye  pollock  stock  in  the  Gulf 
of  Alaska.  In  Stock  assessment  and  fishery  evaluation 
report  for  the  groundfish  resources  of  the  Gulf  of  Alaska, 
90  p.  Prepared  by  the  Gulf  of  Alaska  Groundfish  Plan 
Team,  North  Pacific  Fishery  Management  Council,  W 
4th  Avenue,  Suite  306,  Anchorage,  Alaska  99501. 
Frost,  K.  J.,  and  L.  F.  Lowry. 

1980.  Feeding  of  ribbon  seals  (Phoca  fasciata)  in  the 
Bering  Sea  in  spring.     Can.  J.  Zool.  58:1601-1607. 

1981.  Trophic  importance  of  some  marine  gadids  in 
Northern  Alaska  and  their  body-otolith  size  relation- 
ships.    Fish.  Bull.  84:192-197. 

1986.     Sizes  of  walleye  pollock,  Theragra  chalcogramma, 
consumed  by  marine  mammals  in  the  Bering  Sea.     Fish 
Bull.  79:187-192. 
Goto,  Y.,  and  K.  Shimazaki. 

1998.     Diet  of  Steller  sea  lions  off  the  Coast  of  Rausu, 
Hokkaido,  Japan.     Biosph.  Cons.  1(  2  1:141-148. 
Harkonen,  T 

1986.     Guide  to  the  otoliths  of  the  bony  fishes  of  the 
northeast  Atlantic,  256  p.     Danbiu  Aps.   Hellerup. 
Denmark. 
Harvey,  J.  T 

1989.  Assesment  of  errors  associated  with  harbor  seal 
[Phoca  vitulina)  fecal  sampling.  J.  Zool.  (London)  219: 
101-111. 

Harvey,  J.  T,  Loughlin.  T.  R..  Perez,  M.  A.,  and  D.  S.  Oxman. 
2000.     Relationship  between  fish  size  and  otolith  length 
for  63  species  of  fishes  from  the  eastern  north  Pacific 
Ocean,  36  p.     NOAA  Tech.  Rep.  NMFS  150. 

Hunt.,  G.  L.,  Jr.,  A.  S.  Kitaysky,  M.  B.  Decker.  D.  E.  Dragoo, 
and  A.  M.  Springer. 

1996.  Changes  in  the  distribution  and  size  of  juvenile 
walleye  pollock.  Theragra  chalcogramma.  as  indicated 
by  seabird  diets  at  the  Pribilof  Islands  and  by  bottom 
trawl  surveys  in  the  Eastern  Bering  Sea,  1975  to 
1993.  //;  Ecology  of  juvenile  walleye  pollock,  Theragra 
chalcogramma.  Papers  from  the  workshop  "The  impor- 
tance of  prerecruit  walleye  pollock  to  the  Bering  Sea  and 
North  Pacific  ecosystems"  Seattle,  WA,  October  28-30, 
1993  (R.  D.  Brodeur,  P.  A.  Livingston,  T.  R.  Loughlin, 


520 


Fishery  Bulletin  102(3) 


and  A.  B.  Hollowed,  eds.),  p.  125-139.     NOAA  Tech. 
Rep.  NMFS  126.     [NTIS  no.  PB97-155188.] 
Jobling,  M.,  and  A.  Breiby. 

1986.     The  use  and  abuse  offish  otoliths  in  studies  of  feed- 
ing habits  of  marine  piscivores.     Sarsia  71:265-274. 
Keys,  A.  B. 

1928.     The  weight-length  relation  in  fishes.     Proceedings 
of  the  National  Academy  of  Sciences  of  the  USA  14 
(121:922-925. 
Loughlin,  T  R.,  and  R.  L.  Merrick. 

1989.  Comparison  of  commercial  harvest  of  walleye  pol- 
lock and  northern  sea  lion  abundance  in  the  Bering  Sea 
and  Gulf  of  Alaska,  p.  679  -700.  In  Proceedings  of  the 
international  symposium  on  the  biology  and  manage- 
ment of  walleye  pollock,  Nov.  14-16,  1988,  Anchorage, 
Alaska.  Alaska  Sea  Grant  Report  89-1,  Anchorage, 
AK. 
Loughlin,  T.  R.,  and  R.  Nelson  Jr. 

1986.     Incidental  mortality  of  northern  sea  lions  in  She- 
likof  Strait,  Alaska.     Mar.  Mamm.  Sci.  2:14-33. 
Loughlin  T.  R.,  A.  S.  Perlov,  and  V.  A.  Vladimirov. 

1992.     Range-wide  survey  and  estimation  of  total  num- 
ber of  Steller  sea  lions  in  1989.     Mar.  Mamm.  Sci.  83(3): 
220-239. 
Loughlin,  T.  R.,  J.  T.  Sterling,  R.  L.  Merrick,  J.  L.  Sease,  and 
A.  E.  York. 

2003.     Diving  behavior  of  immature  Steller  sea  lions 
(Eumetopias  jubatus).     Fish.  Bull.  101:566-582. 
Loughlin,  T.  R.,  and  A.  E.  York. 

2000.  An  accounting  of  the  source  of  Steller  sea  lion 
mortality.     Mar.  Fish.  Rev.  62(41:40-45. 

Lowe,  S.  A.,  R.  F.  Reuter,  and  H.  Zenger. 

2001.  Assessment  of  Bering  Sea  /Aleutian  Island  Atka 
mackerel.  In  Stock  assessment  and  fishery  evalua- 
tion report  for  the  groundfish  resources  of  the  Bering 
Sea  /Aleutian  Islands  region,  44  p.  Prepared  by  the 
Bering  Sea  and  Aleutian  Islands  Groundfish  Plan 
Team,  North  Pacific  Fishery  Management  Council,  W 
4th  Avenue,  Suite  306,  Anchorage,  Alaska  99501. 

Manly,  B.  F  J. 

1997.     Randomization,  bootstrap  and  Monte  Carlo  methods 
in  biology,  2nd  ed.,  399  p.     Chapman  and  Hall,  London: 
New  York,  NY. 
McDermott,  S.  F,  and  S.  A.  Lowe. 

1997.     The  reproductive  cycle  and  sexual  maturity  of  Atka 
mackerel,  Pleurogrammus  monopterygius,  in  Alaska 
waters.     Fish.  Bull.  95:231-333. 
Merrick,  R.  L.,  and  D.  G.  Calkins. 

1996.  Importance  of  juvenile  walleye  pollock  in  the  diet  of 
Gulf  of  Alaska  Steller  sea  lions.  In  Ecology  of  juvenile 
walleye  pollock,  Theragra  chalcogramma  (R.  D.  Brodeur, 
P.  A.  Livingston,  T.  R.  Loughlin,  and  A.  B.  Hollowed, 
eds.  I,  p.  153-166.  NOAA  Tech.  Rep.  NMFS  126. 
NMFS  (National  Marine  Fisheries  Service). 

2003.     Fisheries  of  the  United  States,  2002.     Current 
Fishery  Statistics  No.  2002,  126  p.     Fisheries  Statistics 
and  Economics  Division,  Natl.  Mar.  Fish.  Serv.,  Silver 
Spring,  MD. 
NRC  (National  Research  Council). 

1996.     The  Bering  Sea  ecosystem,  307  p.     National  Acad- 
emy Press.  Washington,  D.C. 
Orchard,  T.  J. 

2001.     The  role  of  selected  fish  species  in  Aleut  paleo- 
diet.     M.A.  thesis,  228  p.     Univ.  Victoria,  Victoria.  BC. 
Owen,  J.  F,  and  J.  R.  Merrick. 

1994.     Analysis  of  coastal  middens  in  south-eastern  Aus- 


tralia: Sizing  offish  remains  in  holocene  deposits.     J. 
of  Archaeol.  Sci.  21:3-10. 
Pierce,  G  J.,  and  P.  R.  Boyle. 

1991.     A  review  of  methods  for  diet  analysis  in  piscivo- 
rous marine  mammals.     Oceanogr.  Mar.  Biol.  Annu. 
Rev.  29:409-486. 
Pitcher,  K.  W. 

1981.     Prey  of  the  Steller  sea  lion,  Eumetopias  jubatus, 
in  the  Gulf  of  Alaska.     Fish.  Bull.  79:467-472. 
Prime  J.  H.,  and  P.  S.  Hammond. 

1987.  Quantitative  assessment  of  gray  seal  diet  from 
fecal  analysis.  In  Approaches  to  marine  mammal 
energetics  (A.  C.  Huntley,  D.  P.  Costa,  G.  A.  J.  Worthy 
and  M.  A.  Castellini,  eds.  I  p.  165-181.  Allen  Press, 
Lawrence,  KS. 
Raum-Suryan,  K.  L..  K.  W.  Pitcher.  D.  G.  Calkins,  J.  L.  Sease, 
and  T  R.  Loughlin. 

2002.     Dispersal,  rookery  fidelity,  and  metapopulation 
structure  of  Steller  sea  lions  (Eumetopias  jubatus)  in  an 
increasing  and  a  decreasing  population  in  Alaska.     J. 
Mammal.  18(3):746-764. 
Ringrose,  T.  J. 

1993.     Bone  counts  and  statistics:  a  critique.     J.  Archaeol. 
Sci.  20:121-157. 
Sinclair,  E.  H.,  and  T  K.  Zeppelin. 

2002.     Seasonal  and  spatial  differences  in  diet  in  the  west- 
ern stock  of  Steller  sea  lions  (Eumetopias  jubatus).     J. 
Mammal.  83(4):973-990. 
Sinclair,  E.  H.,  G.  A.  Antonelis,  B.  W.  Robson,  R.  R.  Ream,  and 
T.  R.  Loughlin. 

1996.  Northern  fur  seal,  Callorhinus  ursinus,  predation 
on  juvenile  walleye  pollock,  Theragra  chalcogramma. 
In  Ecology  of  walleye  pollock,  Theragra  chalcogramma 
(R.  D.  Brodeur,  P.  A.  Livingston,  T.  R.  Loughlin,  and 
A.  B.  Hollowed,  eds.)  p.  167-178.  NOAA  Tech.  Rep.. 
NMFS-126. 

Smith,  G.  B. 

1981.     The  biology  of  walleye  pollock.     In  The  eastern 
Bering  Sea  shelf:  oceanography  and  resources,  vol.  1 
(D.  W.  Hood  and  J.  A.  Calder,  eds.)  p.  527-552.     Univ. 
Wash.  Press,  Seattle,  WA. 
Templemann,  W.,  and  H.  J.  Squires. 

1956.     Relationship  of  otolith  lengths  and  weights  in  the 
haddock,  Melanogrammus  aeglefinus  (L.)  to  the  rate 
of  growth  of  the  fish.     J.  Fish.  Res.  Board  Can.   13: 
467-487. 
Tollit,  D.  J..  S.  G.  Heaslip.  and  A.  W.  Trites. 

2004a.     Sizes  of  walleye  pollock  ( Theragra  chalcogramma) 
consumed  by  the  eastern  stock  of  Steller  sea  lions 
(Eumetopias  jubatus)  in  Southeast  Alaska  from  1994 
to  1999.     Fish.  Bull.  102:522-532. 
Tollit,  D.  J..  S.  G.  Heaslip,  T  K.  Zeppelin,  R.  Joy,  K.  A.  Call, 
and  A.  W.  Trites. 
2004b.     A  method  to  improve  size  estimates  of  walleye 
pollock  (Theragra  chalcogramma)  and  Atka  mackerel 
(.Pleurogrammus  monopterygius)  consumed  by  pinnipeds: 
digestion  correction  factors  applied  to  bones  and  otoliths 
recovered  in  scats.     Fish.  Bull.  102:498-508. 
Tollit  D.  J.,  M.  J.  Steward,  P.  M.  Thompson.  G  J.  Pierce, 
M.  B.  Santos,  and  S.  Hughes. 

1997.  Species  and  size  differences  in  the  digestion  of 
otoliths  and  beaks:  implications  for  estimates  of  pin- 
niped diet  composition.  Can.  J.  Fish  Aquat.  Sci. 
54:105-119. 


Zeppelin  et  al.:   Sizes  of  walleye  pollock  and  Atka  mackerel  consumed  by  Eumetopias  jubatus 


521 


Tollit,  D.  J.,  M.  Wong,  A.  J.  Winship,  D.  A.  S.  Rosen,  and 
A.  W.  Trites. 

2003.     Quantifying  errors  associated  with  using  prey 
skeletal  structures  from  fecal  samples  to  determine  the 
diet  of  the  Steller  sea  lion  (Eumetopias  jubatus).     Mar. 
Mamm.  Sci.  19(4):724-744. 
Walline,  P.  D. 

1983.  Growth  of  larval  and  juvenile  walleye  pollock  re- 
lated to  year-class  strength.  Ph.D.  diss.,  144  p.  Univ. 
Washington,  Seattle,  WA. 


Wise,  M.  H. 

1980.     The  use  offish  vertebrae  in  scats  for  estimating  prey 
size  of  otters  and  mink.     J.  Zool.,  Lond.  192:25-31. 
Zolotov,  O.  G. 

1993.  Notes  on  the  reproductive  biology  of  Pleurogram- 
mus  monopterygius  in  Kamchatkan  waters.  J.  Ichthyol. 
33(41:25-37. 


522 


Abstract— Lengths  of  walleye  pollock 
i  Theragra  chalcogramma  I  consumed 
by  Steller  sea  lions  (Eumetopias 
jubatus)  were  estimated  by  using 
allometric  regressions  applied  to 
seven  diagnostic  cranial  structures 
recovered  from  531  scats  collected 
in  Southeast  Alaska  between  1994 
and  1999.  Only  elements  in  good  and 
fair  condition  were  selected.  Selected 
structural  measurements  were  cor- 
rected for  loss  of  size  due  to  erosion 
by  using  experimentally  derived 
condition-specific  digestion  correc- 
tion factors.  Correcting  for  digestion 
increased  the  estimated  length  of 
fish  consumed  by  23% ,  and  the  aver- 
age mass  offish  consumed  by  88%. 
Mean  corrected  fork  length  (FLl  of 
pollock  consumed  was  42.4  ±11.6  cm 
(range  =  10. 0-78.1  cm,  n=909).  Adult 
pollock  (FL>45.0  cm)  occurred  more 
frequently  in  scats  collected  from 
rookeries  along  the  open  ocean  coast- 
line of  Southeast  Alaska  during  June 
and  July  (74%  adults,  mean  FL  =  48.4 
cm l  than  they  did  in  scats  from  haul- 
outs  located  in  inside  waters  between 
October  and  May  (51%  adults,  mean 
FL  =  38.4  cm).  Overall,  the  contribu- 
tion of  juvenile  pollock  (<20  cm)  to 
the  sea  lion  diet  was  insignificant; 
whereas  adults  contributed  44%  to 
the  diet  by  number  and  74%  by  mass. 
On  average,  larger  pollock  were  eaten 
in  summer  at  rookeries  throughout 
Southeast  Alaska  than  at  rookeries 
in  the  Gulf  of  Alaska  and  the  Bering 
Sea.  Overall  it  appears  that  Steller 
sea  lions  are  capable  of  consuming 
a  wide  size  range  of  pollock,  and  the 
bulk  offish  fall  between  20  and  60  cm. 
The  use  of  cranial  hard  parts  other 
than  otoliths  and  the  application  of 
digestion  correction  factors  are  fun- 
damental to  correctly  estimating  the 
sizes  of  prey  consumed  by  sea  lions 
and  determining  the  extent  that  these 
sizes  overlap  with  the  sizes  of  pollock 
caught  by  commercial  fisheries. 


Sizes  of  walleye  pollock  (Theragra  chalcogramma) 
consumed  by  the  eastern  stock  of  Steller  sea  lions 
(Eumetopias  jubatus)  in  Southeast  Alaska  from 
1994  to  1999 


Dominic  J.  Tollit 

Susan  G.  Heaslip 

Andrew  W.  Trites 

Marine  Mammal  Research  Unit,  Fisheries  Centre 
University  ol  British  Columbia 
Room  18,  Hut  B-3,  6248  Biological  Sciences  Road 
Vancouver,  British  Columbia,  Canada,  V6T  1Z4 
E-mail  (for  D  J  Tollit).  tollitia'zoology  ubc  ca 


Manuscript  submitted  28  April  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
25  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:522-532(2004). 


The  dramatic  decline  of  the  western 
population  of  Steller  sea  lions  (Eume- 
topias jubatus)  in  the  1980s  (Loughlin 
et  al.,  1992;  Trites  and  Larkin,  1996) 
prompted  a  number  of  studies  to  deter- 
mine what  they  eat  and  the  extent  of 
overlap  of  the  fish  consumed  by  Steller 
sea  lions  and  fish  caught  by  commer- 
cial fisheries.  The  eastern  population 
of  sea  lions  (east  of  longitude  144°) 
located  mainly  in  Southeast  Alaska 
and  British  Columbia  gradually 
increased  as  the  western  population 
declined  (e.g..  Calkins  et  al.,  1999), 
permitting  insightful  comparative 
studies  to  be  undertaken  (e.g.,  Mer- 
rick et  al.,  1995;  Milette  and  Trites, 
2003).  Possible  explanations  for  the 
different  population  trends  include 
ocean  climate,  competition  with  fish- 
eries, predation,  and  the  amount  or 
the  sizes  of  pollock  in  the  diets  of  sea 
lions  in  the  two  regions  (Loughlin 
and  York,  2000;  Benson  and  Trites, 
2002;  NRC,  2003;  Trites  and  Don- 
nelly, 2003;  Calkins  and  Goodwin1; 
Loughlin  and  Merrick-). 

Reliable  estimates  of  prey  size  are 
important  not  only  to  investigate  prey 
selectivity  and  the  extent  of  overlap 
in  size  of  prey  with  size  of  the  same 
species  caught  by  commercial  fisher- 
ies and  by  other  marine  piscivores 
but  are  also  vital  for  accurately  as- 
sessing prey  numbers,  biomass,  and 
total  consumption  (Beverton,  1985; 
Ringrose,  1993;  Laake  et  al.,  2002). 
One  means  of  estimating  prey  size  is 
to  measure  hard  parts  recovered  from 
fecal  remains  and  to  apply  allometric 


regressions  relating  fork  length  to 
the  size  of  otoliths  (Frost  and  Lowry, 
1981)  and  other  bones  (Zeppelin  et  al., 
2004,  this  issue).  However,  the  extent 
of  digestion  incurred  by  both  otoliths 
and  bones  as  they  pass  through  the 
digestive  tract  must  be  accounted  for 
to  ensure  that  prey  size  is  not  un- 
derestimated (Tollit  et  al.,  2004,  this 
issue).  Application  of  these  two  steps 
is  integral  to  correctly  estimate  the 
size  of  prey  consumed  by  Steller  sea 
lions  and  other  pinnipeds. 

The  goal  of  our  study  was  to  esti- 
mate the  size  of  walleye  pollock  (Ther- 
agra chalcogramma)  consumed  by 
Steller  sea  lions  in  Southeast  Alaska 
between  1994  and  1999  by  using 
new  methods  outlined  by  Tollit  et  al. 
(2004,  this  issue)  and  Zeppelin  et  al. 
(2004,  this  issue).  Previous  size  esti- 
mates for  this  region  of  Alaska  are 
based  on  the  analysis  of  only  eight 
stomachs  collected  in   1986  i Calkins 


1  Calkins,  D.  G.,  and  E.  Goodwin.  1988. 
Unpubl.  report.  Investigation  of  the 
declining  sea  lion  population  in  the  Gulf 
of  Alaska,  76  p.  Alaska  Department 
of  Fish  and  Game,  333  Raspberry  Rd, 
Anchorage,  AK  99518. 

2  Loughlin,  T.  R.,  and  R.L.  Merrick. 
1989.  Comparison  of  commercial  har- 
vest of  walleye  pollock  and  northern  sea 
lion  abundance  in  the  Bering  Sea  and 
Gulf  of  Alaska.  In  Proceedings  of  the 
international  symposium  on  the  biology 
and  management  of  w-alleye  pollock. 
Nov.  14-16,  1988,  Anchorage,  AK,  p. 
679-700.  Alaska  Sua  Grant  KVp  Sil- 
01.  LIniv.  Alaska  Fairbanks,  Fairbanks. 
AK 


To  Hit  et  al.:  Sizes  of  walleye  pollock  consumed  by  Eumetopias  /ubatus 


523 


58    N 


56 


A 


50 


Petersburg 


"      14- 

15 

fh.        r 


50 


Kilometers 


Pollock  trawl 
19 '       Dixon  ,lsherV 

Entrance 


138   W 


134 


130 


Figure  t 

Location  of  Steller  sea  lion  (Eumetopias  jubatus)  haul-outs  and  rookeries  visited  during 
1994-99  to  collect  scats  containing  pollock  hard  remains.  Symbols:  haul-outs  in  inside 
waters  (•),  haul-outs  in  outside  waters  (©),  haul-outs  where  scats  were  not  collected  or 
sites  at  which  no  pollock  hard  remains  were  found  (O),  rookeries    ■     and  cities  (*i. 


and  Goodwin1).  We  sought  to  compare  the  sizes  of  pol- 
lock consumed  in  the  1990s  with  these  earlier  samples, 
as  well  as  with  the  sizes  consumed  by  the  declining 
population  of  sea  lions  in  the  Gulf  of  Alaska  and  Ber- 
ing Sea  during  the  1970s  and  1980s  (e.g.,  Pitcher,  1981; 
Merrick  and  Calkins,  1996)  and  between  1998  and  2000 
(Zeppelin  et  al.,  2004,  this  issue).  We  also  wanted  to 
evaluate  the  use  of  digestion  correction  factors  (DCFs) 
and  skeletal  structures  other  than  otoliths  to  estimate 
prey  size,  and  to  compare  the  different  size  estimates 
for  fish  consumed  by  sea  lions  in  Southeast  Alaska 
with  sizes  of  fish  caught  by  a  nearby  commercial  trawl 
fishery. 


Materials  and  methods 

Estimating  sizes  of  pollock  consumed 

Scats  that  contained  pollock  hard  remains  were  collected 
from  four  rookeries  and  16  haul-outs  from  both  inside 
and  outside  waters  of  Southeast  Alaska  between  1994 


and  1999  (Fig.  1  and  Table  1).  Scats  from  three  haul- 
outs  and  four  rookeries  in  outside  waters  were  collected 
from  May  through  October  1994-99,  but  most  were 
collected  from  June  and  July.  Scats  from  inside  waters 
were  collected  at  13  haul-outs  located  in  the  straits 
and  sounds  between  Juneau  and  Petersburg,  Alaska 
(56.8-58.6°N,  132. 8-134. 9°W)  (Fig.  1).  The  majority  of 
these  "inside"  scats  were  collected  from  Frederick  Sound 
(Fig.  1)  between  October  1995  and  February  1997.  Most 
were  collected  in  the  winter  and  spring,  but  some  were 
collected  in  the  summer  of  1999  (Trites  et  al.3).  In  gen- 
eral, the  haul-out  sites  visited  to  collect  scats  were  those 
with  relatively  high  numbers  of  animals  across  South- 
east Alaska  (Calkins  et  al.,  1999;  Sease  et  al.,  2001). 

Scats  were  washed  and  sieved  (0.5  mm)  and  hard 
remains  were  identified  by  Pacific  IDentifications  Inc. 
(Univ.  of  Victoria,  Victoria,  B.C.).  Seven  commonly 


3  Trites,  A.  W.,  D.  G.  Calkins,  and  A.  J.  Winship.  2003.  Unpubl. 
data.  Marine  Mammal  Research  Unit,  Fisheries  Centre, 
University  of  British  Columbia,  Room  18,  Hut  B-3.  6248  Bio- 
logical Sciences  Rd.,  Vancouver,  B.C.,  Canada,  V6T  1Z4. 


524 


Fishery  Bulletin  102(3) 


Table  1 

Steller  sea  lion  scat  collection  sites  in  Southeast  Alaska,  as  illustrated  in  Figure  1,  giving  details 

of  the  type  (HO=haul-outl.  fish 

element  sample  size 

irir),  and  the  estimated  corrected  mean  fork  length  Imean  FL,  cm) 

of  walleye  pollock 

based  on  seven  cranial 

structures  found  in 

scats  at  each  site. 

Region 

Site  no. 

Site  name                               Type 

"/ 

Mean  FL 

SD 

Inside  waters 

1 

Benjamin  Island                           HO 

11 

39.7 

13.9 

2 

Dorothy  Island                              HO 

3 

38.6 

13.6 

3 

Circle  Point                                    HO 

31 

45.1 

13.7 

4 

Point  League                               HO 

37 

42.9 

9.9 

5 

Sunset  Point                                  HO 

196 

37.4 

10.3 

6 

Sail  Island                                     HO 

36 

35.5 

10.0 

7a 

W  Brother  Island                          HO 

8 

40.9 

14.7 

7b 

SW  Brother  Island                      HO 

152 

37.1 

9.9 

8 

Turnabout  Island                         HO 

34 

47.8 

11.0 

9 

Yasha  Island                                HO 

19 

44.8 

5.3 

10 

Sukoi  Islets                                    HO 

14 

26.9 

7.8 

11 

Horn  Cliffs                                     HO 

19 

44.3 

12.7 

12 

Liesnoi  Island                                HO 

7 

31.8 

10.2 

Outside  waters 

13 

Cape  Cross                                     HO 

7 

45.3 

3.0 

14 

Timbered  Island                           HO 

5 

39.3 

8.7 

15 

Point  Addington                           HO 

1 

53.5 

— 

16 

Graves  Rock                                Rookery 

49 

42.9 

7.7 

17 

White  Sisters                              Rookery 

33 

43.4 

9.6 

18 

Hazy  Islands                               Rookery 

54 

45.4 

8.7 

19 

Forrester  Islands                          Rookery 

193 

51.4 

10.0 

occurring,  robust,  and  diagnostic  pollock  structures 
were  removed  from  all  scats  containing  pollock  (see 
Tollit  et  al.,  2004,  this  issue).  All  were  from  the  cra- 
nium region  (see  Zeppelin  et  al.,  2004,  this  issue)  and 
included  the  sagittal  otolith  (OTO),  as  well  as  the  inter- 
hyal  (INTE),  hypobranchial  3  (HYPO),  pharyngobran- 
chial  2  (PHAR),  angular  (ANGU),  quadrate  (QUAD), 
and  the  dentary  (DENT).  Each  individual  fish  element 
was  assigned  one  of  three  condition  categories  (good, 
fair,  or  poor)  and  was  measured  three  times  (±0.01  mm) 
at  a  specific  location  to  calculate  a  mean  estimate  (see 
Tollit  et  al.,  2004,  this  issue). 

Fork  lengths  of  pollock  eaten  by  Steller  sea  lions  in 
Southeast  Alaska  were  first  estimated  by  applying  al- 
lometric  regressions  (Zeppelin  et  al.,  2004,  this  issue) 
to  otolith  lengths  (OTOL)  without  correcting  for  par- 
tial digestion  (see  Pitcher,  1981;  Merrick  and  Calkins, 
1996).  We  also  measured  and  substituted  otolith  width 
(OTOW)  when  otoliths  were  broken  lengthwise.  We  then 
applied  appropriate  DCFs  and  regression  formulae  to 
otoliths  assigned  in  good  and  fair  condition  (Tollit  et 
al.,  2004,  this  issue).  Finally,  we  applied  allometric 
regressions  (Zeppelin  et  al.,  2004,  this  issue)  to  all  ele- 
ments of  the  remaining  six  cranial  structures  (bones) 
assigned  to  good  or  fair  condition  categories  to  provide 
estimates  of  fish  size  across  structures  both  with  and 
without  applying  the  appropriate  DCFs  (Tollit  et  al., 
2004,  this  issue).  Structures  in  poor  condition  were 


excluded  because  of  large  intraspecific  size  variation 
noted  from  feeding  experiments  with  captive  sea  lions 
(see  also  Sinclair  et  al..  1994;  Tollit  et  al.,  1997;  Tollit 
et  al.,  2004,  this  issue). 

To  incorporate  the  major  sources  of  error  in  our 
method,  we  calculated  confidence  intervals  (95^r)  for 
fork-length  estimates.  First,  we  applied  a  random 
bootstrapped  regression  equation,  followed  by  a  boot- 
strapped correction  factor  applicable  to  each  selected 
structure  (see  Tollit  et  al.,  2004,  this  issue).  For  the 
five  structures  (INTE,  HYPO.  PHAR,  ANGU,  and  OTO) 
in  good  condition  for  which  Tollit  et  al.  (this  issue) 
recommended  a  DCF  of  1.0  (no  correction),  we  drew 
bootstrapped  values  from  a  discrete  declining  triangu- 
lar probability  distribution  (hj  ranging  from  1.0  to  1.05 
(to  simulate  a  limited  degree  of  digestion).  Finally,  we 
bootstrapped  individual  scats  at  random,  by  selecting 
n  scats  with  replacement  from  the  original  sample  size 
n  (to  account  for  resampling  variability  across  scats) 
and  included  only  selected  elements  within  those  ran- 
domly bootstrapped  scats.  Bootstrap  randomizations  for 
these  steps  were  done  1000  times  and  959!  confidence 
intervals  were  taken  as  the  25th  and  975,h  values  of  the 
sorted  bootstrapped  values. 

Finally,  consideration  was  also  given  to  whether 
an  individual  fish  might  be  represented  by  different 
structures  within  a  single  scat.  We  therefore  compared 
length  estimates  using  all  structures  with  those  esti- 


To  Hit  et  al.:  Sizes  of  walleye  pollock  consumed  by  Eumetopias  jubatus 


525 


mated  with  the  minimum  number  of  individuals  (MNI) 
technique  (Ringrose,  1993;  Browne  et  al.,  2002).  This 
technique  is  used  to  select  structures  within  each  scat 
that  preclude  pseudoreplication  or  double  counting  of 
fish.  Within  each  scat,  the  structure  with  the  great- 
est MNI  was  selected,  and  right-sided  structures  were 
selected  over  left-sided  structures  if  both  sides  were 
found  in  equal  number  because  right-sided  structures 
are  used  in  regression  formulae.  If  two  structures  had 
the  same  MNI  estimate,  then  selection  was  made  on 
the  structure  with  the  larger  regression  determination 
coefficient,  r2  (OTO-W>OTO-L>QUAD>DENT>HYPO> 
INTER>ANGU>PHAR ). 

Geographical  and  temporal  variation  in  sizes  of 
prey  consumed 

All  elements  from  the  seven  cranial  structures  in  good  or 
fair  condition  were  used  to  compare  size  of  pollock  con- 
sumed by  Steller  sea  lions  in  Southeast  Alaska  between 
regions  (inside  haul-outs  versus  outside  rookeries),  across 
years  and  across  rookeries  (with  rookery  data  collected 
in  June  and  July),  and  across  months  (with  data  col- 
lected from  inside  haul-outs).  Biologically  meaningful 
differences  in  FL  of  pollock  were  assessed  by  grouping 
corrected  lengths  into  stage-class  categories  (juvenile 
or  1-year-old  fish  FL<20  cm;  adolescent  20<FL<34  cm; 
subadult  34<FL<45  cm;  and  adult  FL>45  cm)  (Smith, 
1981;  Walline,  1983;  Dorn  et  al.,  2001).  Adults  were 
considered  to  be  mature  fish  >5  years  old  and  targeted 
by  fisheries  (Smith,  1981).  Subadults  were  likely  3  or 
4  years  old,  of  which  only  a  proportion  had  matured  or 
were  targeted  by  the  fishery.  To  avoid  the  possibility  of 
pseudoreplication  in  our  chi-squared  comparisons,  we 
used  only  the  presence  or  absence  of  structures  of  each 
stage  class  in  a  scat  because  individual  fish  eaten  by  a 
sea  lion  may  have  come  from  an  age-specific  school  and 
were  therefore  not  independent  (Hunt  et  al.,  1996).  Pres- 
ence-absence data  was  chosen  over  MNI  data  because 
the  former  greatly  reduces  potential  concerns  regarding 
size-dependent  recovery  of  cranial  structures  (Tollit  et 
al.,  1997).  With  the  exception  of  our  regional  comparison, 
data  from  juvenile  and  adolescent  stage-classes  were 
pooled  because  of  the  low  sample  sizes  of  juvenile  fish.  A 
Fisher's  exact  test  was  used  as  an  alternative  test  to  chi- 
square  comparisons  when  counts  for  a  stage-class  group- 
ing were  <5  (S-PLUS  2000,  Mathsoft  Inc.,  Seattle,  WA). 

Overlap  of  prey  size  with  size  of  fish  caught  by  fisheries 

To  assess  the  impact  of  using  the  new  methods  described 
and  to  compare  the  size  of  pollock  consumed  by  sea  lions 
with  the  size  of  pollock  typically  caught  by  fisheries,  we 
obtained  randomly  subsampled  size-frequency  landing 
data  from  the  Canadian  commercial  pollock  fishery  in 
Dixon  Entrance  (1993-1999)  (Saunders4).  This  area  is 


4  Saunders,  M.  2002.  Unpubl.  data.  Fisheries  and  Oceans 
Canada,  3190  Hammond  Bay  Road,  Nanaimo,  B.C.,  Canada, 
V9T  6N7. 


115-135  km  SE  of  the  Forrester  Island  rookery  on  the 
southern  border  of  Southeast  Alaska  (Fig.  1). 


Results 

Sizes  of  pollock  consumed 

The  traditional  method  of  estimating  prey  size  from 
otoliths  alone  was  not  satisfactory  because  most  otoliths 
were  in  poor  condition  (86%,  ra=247)  or  were  broken 
lengthwise  (>89%)  (or  were  both  broken  and  in  poor 
condition).  Cranial  bones,  on  the  other  hand,  occurred 
in  higher  numbers  than  otoliths  and  were  therefore  more 
useful  for  estimating  prey  size  (Table  2). 

Sixty-one  percent  of  scats  (1215  of  1987)  collected 
from  Southeast  Alaska  (1994-99)  contained  pollock 
remains,  with  an  average  MNI  of  1.57  ±1.66  individual 
pollock  per  scat  (range:  1-37  individuals).  Many  scats 
contained  hard  parts  that  were  not  useful  for  estimat- 
ing prey  size  (e.g.,  gill  rakers),  leaving  531  scats  (26%) 
with  measurable  selected  structures.  Of  these,  303  scats 
contained  1746  elements  in  good  (n  =  225),  fair  (n  =  684). 
and  poor  condition  («  =  837). 

Applying  digestion  correction  factors  had  a  consider- 
able effect  on  the  estimated  length  and  mass  of  fish 
consumed,  and  on  the  proportion  that  were  deemed 
to  be  adults  (Fig.  2).  The  estimated  lengths  of  pollock 
calculated  from  all  structures  graded  in  good  or  fair 
condition  (without  accounting  for  digestion)  was  34.4 
±9.7  cm  (ra=909,  modal  range:  32-40)  (Table  2,  Fig.  2). 
Lengths  increased  by  23%  on  average  when  appropri- 
ate DCFs  were  applied  to  each  structure  to  account  for 
the  observed  degree  of  digestion  (mean  FL  =  42.4  ±11.6 
cm,  modal  range:  44-52,  95%  CI  =  41. 0-43.9)  (paired 
t-test,  i908=67.1,  P<0.001).  A  DCF  of  1.0  (no  correction 
required  to  account  for  digestion)  was  applied  to  62  ele- 
ments in  good  condition,  resulting  in  a  mean  fork  length 
of  39.6  ±11.9  cm  estimated  from  those  bones. 

The  size-frequency  distribution  of  pollock  consumed  by 
sea  lions  also  varied  significantly  following  the  applica- 
tion of  DCFs  (Kolmogorov-Smirnov,  KS  =  176.2,  P<0.001) 
and  led  to  an  increase  in  the  proportion  of  fish  thought 
to  have  been  adult  (>45  cm  FL)  from  16%  to  44%.  This 
result  in  turn  reduced  the  proportion  of  fish  thought 
to  have  been  subadults  (29%),  adolescents  (25%),  and 
juveniles  (<2%,  <20  cm  FL)  (Fig.  2).  The  size  range  of 
pollock  eaten  ranged  widely  regardless  of  whether  DCFs 
were  applied  (10-78  cm)  or  not  (10-64  cm).  When  we 
calculated  fork  lengths  using  only  elements  selected 
according  to  MNI  criteria,  the  means  increased  by  just 
0.5  cm  for  corrected  and  by  just  0.3  cm  for  uncorrected 
lengths,  with  near  identical  standard  deviations  and 
distributions  (Fig.  2)  (Kolmogorov-Smirnov,  uncorrected 
KS  =  0.33,  P=0.89,  corrected  KS  =  0.032,  P=0.91). 

The  use  of  all  otoliths  regardless  of  digestion  state 
resulted  in  a  mean  fork  length  that  was  only  about 
half  of  that  derived  by  using  all  structures  corrected 
for  digestion  (Table  2).  Excluding  otoliths  in  poor  con- 
dition significantly  reduced  sample  size  (Table  2)  but 


526 


Fishery  Bulletin  102(3) 


Fable  2 

Estimated  mean 

fork  length 

mean 

FL,  cmi  of 

walleye  pollock  consumed  by  Stellei 

sea  lions. 

Values  were  determined  by  using 

selected 

cranial 

structui 

es  w 

ith  or 

without  the  application 

of  condition-specific  di 

gestion 

correction  factors 

DCFs 

.  Data  sets 

exclude 

all  struc 

ures  gr 

aded 

in  poor  condition 

(with  the  exception  of  data  sets  marked  w 

th 

an  asterisk l.  Fi 

ih  element  sample 

sizes  (n,- 

are  given  along 

with 

propo 

'tion  of  elements  assign* 

d  condition  category  good  ing). 

When  DCFs  were  applied 

95ri  confi- 

dence  intervals  (95<~t  CI  i 

were 

estimated  by  using  bootstrap 

resampling  I  see  "Materials  and  methods"). 

Structure  code 

DCF 

n 

ng 

Mean  FL 

SD 

Range 

95%  CI 

INTE 

No 

37 

0.35 

44.0 

8.0 

28.0-54.5 



Yes 

37 

0.35 

48.0 

9.3 

31.9-62.2 

45.0-52.2 

HYPO 

No 

47 

0.19 

35.3 

8.9 

19.0-52.0 

— 

Yes 

47 

0.19 

39.8 

10.1 

19.0-60.4 

36.7-43.6 

PHAR 

No 

20 

0.25 

38.1 

8.5 

20.4-50.3 

— 

Yes 

20 

0.25 

43.7 

9.5 

20.4-56.1 

39.9-48.4 

ANGU 

No 

207 

0.16 

34.0 

10.2 

10.0-62.8 

— 

Yes 

207 

0.16 

39.4 

11.4 

10.0-63.2 

37.4-41.5 

QUAD 

No 

238 

0.36 

33.1 

10.4 

14.0-63.8 

— 

Yes 

238 

0.36 

41.9 

13.2 

17.3-78.1 

39.5-44.7 

DENT 

No 

326 

0.24 

34.9 

8.1 

11.0-63.0 

— 

Yes 

326 

0.24 

45.1 

9.8 

14.7-75.3 

43.3-46.8 

OTOL 

No 

10 

0.10 

30.6 

13.8 

14.2-54.8 

— 

Yes 

10 

0.10 

36.6 

17.6 

16.7-67.2 

27.0-51.1 

OTOL  or  OTOW 

No 

34 

0.03 

27.2 

16.1 

10.8-54.8 

— 

Yes 

34 

0.03 

33.7 

12.8 

13.3-67.2 

29.5-39.5 

All  structures 

No 

909 

0.25 

34.4 

9.7 

9.8-63.8 

— 

Yes 

909 

0.25 

42.4 

11.6 

10.0-78.1 

41.0-43.9 

OTOL* 

No 

27 

0.04 

23.3 

11.9 

7.9-54.8 

— 

OTOL  or  OTOW" 

No 

247 

<0.01 

20.2 

9.7 

5.0-58.0 

— 

increased  our  estimate  of  fork  length  by  approximately 
.'!.'!'<  Applying  grade-specific  DCFs  increased  these  es- 
timates by  another  19%  (to  36.6  cm)  for  otolith  length 
and  by  24%  (to  33.7  cm)  for  a  combination  of  otolith 
length  and  width  (Table  2).  All  six  remaining  struc- 
tures in  good  or  fair  condition  provided  larger  corrected 
mean  length  estimates  than  did  otoliths  alone,  but  95% 
confidence  intervals  derived  from  otoliths  did  overlap 
with  other  structures  (Table  2).  The  smaller  estimate 
provided  by  otoliths  may  reflect  that  >83%  of  grade 
corrected  otoliths  (n  =  34)  came  from  the  inside  haul- 
out  sites,  where  animals  seem  to  eat  smaller  fish  (see 
following  section).  On  the  other  hand,  large  pollock 
otoliths  were  observed  to  have  been  regurgitated  in 
feeding  studies  on  captive  sea  lions  and  also  may  be 
more  easily  crushed  by  rocks  often  found  in  the  stom- 
achs of  Steller  sea  lions  (Tollit  et  ah,  2003).  Regres- 
sion formulae  used  in  our  study  to  predict  pollock  FL 
from  otolith  length  were  similar  to  those  of  Frost  and 
Lowry  (1981)  for  juvenile  fish  (<10  mm  otolith  length) 
but  led  to  smaller  fish  size  estimates  (—1—1.7  cm)  over 
the  range  of  30-50  cm. 

The  lengths  of  the  biggest  fish  (corrected  mean 
FL  =  48.0  cm)  were  estimated  from  measurements  of 
INTE  (Table  2),  the  structure  with  the  lowest  DCF. 
Dentary  bones  (the  most  abundant  structure  recov- 


ered) predicted  mean  (Table  2)  and  modes  (FL  44- 
50  cm)  similar  to  those  predicted  from  all  structures. 
Applying  DCFs  increased  our  length  estimates  by  be- 
tween 9%  (INTE)  and  29%  (DENT)  (paired  r- tests,  all 
P<0.001).  Overall,  corrected  fork  length  estimates  from 
elements  in  good  condition  were  similar  to  those  from 
elements  in  fair  condition  (Mann-Whitney  U,  P=0.47), 
but  multiple  comparisons  indicated  a  significant  differ- 
ence (P<0.05>  between  condition  categories  for  INTE 
and  DENT  (lengths  were  estimated  to  be  longer  from 
elements  graded  in  fair  condition). 

Repeat  measurements  of  individual  elements  were 
all  within  0.04  mm  of  the  mean,  and  88.9-99.5%  were 
within  0.02  mm.  The  highest  variability  was  associated 
with  ANGU,  HYPO,  and  PHAR  with  88.9%,  91.7%,  and 
94.9%  of  their  respective  measurements  falling  within 
0.02  mm.  A  0.02-mm  measurement  difference  corre- 
sponded to  only  a  0.1-0.2  cm  difference  in  fish  length, 
depending  on  the  structure  used. 

Small  differences  in  estimates  of  fork  lengths  can 
have  large  effects  on  estimated  body  mass  (given  the 
exponential  mass-length  relationship,  see  Tollit  et  al., 
2004,  this  issue)  and  can  increase  the  mean  mass  of 
fish  by  more  than  sixfold  depending  on  which  method 
is  used  to  estimate  body  length  (all  otoliths  and  no  cor- 
rection versus  condition-corrected  structures).  The  ap- 


"To Hit  et  al.:  Sizes  of  walleye  pollock  consumed  by  Eumetopias  jubatus 


527 


All  structures                                        MNI 

8  ■ 

■  Otoliths        n 
d  Other 
bones         r    r 

n  =  909 

I 

n  =  478 

6  " 

n. 

lU 

4  ■ 

ill 

J 

No  DCFs 

Percent  relative  frequency 
^         a>         03        o         i\3 

_A 

..  .     Irun 

f 

^n 

Ii 

k 

n-  f 

Ik 

DCFs 
applied 

2  " 

o  ■ 

__jLi... 

kn>,„ 

r,      rlR.       L     II 

II  k„ 

h — r~   i      i      i      i      i      i      i    i      i      i      

0      10    20    30    40    50    60    70    80  0      10    20    30    40    50    60    70    80 

Estimated  fork  length  (cm) 

Figure  2 

Relative  frequency  histograms  of  the  estimated  fork  length  of  walleye  pollock  (Theragra 
ehalcogramma)  predicted  from  seven  cranial  structures  in  good  and  fair  condition. 
Otoliths  (black  bars)  are  stacked  beneath  all  other  selected  structures  (gray  barsl. 
Comparisons  were  made  with  the  application  of  digestion  correction  factors  (DCFs) 
to  take  account  of  partial  digestion  and  with  the  use  of  a  selection  technique  (MNI) 

versus  using  all  structures  present  in  scats. 

plication  of  our  DCFs  to  all  structures  in  good  and  fair 
condition  increased  the  estimated  mean  mass  of  pollock 
consumed  by  88%  (from  a  mean  of  388  g  to  731  g). 
Thus,  although  we  estimate  that  44%  of  the  number  of 
pollock  eaten  by  Steller  sea  lions  were  adults,  based  on 
length,  their  contribution,  based  on  weight,  increased 
to  74%.  In  contrast,  the  contribution  of  juvenile  fish  (in 
terms  of  mass)  dropped  to  <0.1%  (compared  to  <2%  by 
number). 

Geographical  and  temporal  variation  in  sizes  of 
pollock  consumed 

More  pollock  elements  in  good  and  fair  condition  were 
recovered  from  inside  haul-outs  (??  =  567)  than  from  sites 
on  the  outside  coastline  (n  =  342)  (Table  1).  Upon  investi- 
gation, we  found  the  size  of  pollock  consumed  by  Steller 
sea  lions  varied  over  time  and  across  regions  (Fig.  3). 
In  particular,  the  frequency  of  occurrence  of  pollock 
stage  classes  differed  significantly  in  the  scats  of  sea 
lions  resting  at  rookeries  on  the  outside  coastline  of 
Southeast  Alaska  in  summer  (mean  FL=48.4  cm,  n  =  328, 
modal  range:  44-50  cm,  95%  CI  =  46.5-50.2,  ns=126 
scats)  compared  to  those  collected  between  October  and 
May  at  haul-outs  in  the  waters  of  the  inside  passages 
(mean  FL  =  38.4  cm,  /;  =  499,  modal  range;  30-34  cm, 


95%  CI=36.9-40.3,  ns=168  scats)  (^=45. 2,  P<0.001). 
Scats  from  these  inside  haul-outs  contained  a  greater 
diversity  of  stage  classes,  and  there  was  an  equal  prob- 
ability of  any  given  scat  containing  adults  (51.2%),  sub- 
adults  (47.6%).  and  adolescents  (53.0%),  but  a  far  less 
probability  of  containing  juveniles  (6.5%).  In  contrast, 
the  pollock  found  in  scats  from  the  outside  rookeries 
contained  mostly  adults  (73.8%)  and  fewer  occurrences 
of  the  remaining  three  stage  classes  (38.1%,  9.5%,  and 
3.2%.  respectively).  Notably,  the  stage-class  compar- 
ison of  summer  1999  with  scats  from  inside  versus 
outside  waters  was  not  significant  (Fishers  exact  test, 
P=0.11). 

Similar  proportions  of  each  pollock  stage  class  were 
found  in  scats  collected  between  years  (Forrester, 
Fisher's  exact  test,  P=0.54;  Hazy,  Fisher's  exact  test, 
P=0.16),  and  between  rookeries  (1994  only,  Fisher's  ex- 
act test,  P=0.57;  all  years,  Fisher's  exact  test,  P=0.22). 
Scats  from  inside  haul-outs  collected  in  spring  1996 
contained  comparatively  smaller  fish  than  in  other 
months  and  years  examined  (Fig.  3).  However,  there 
were  no  significant  monthly  differences  in  the  pro- 
portions of  age  classes  from  October  1995  to  Febru- 
ary 1997  (*2  =  16.52,  P=0.28)  or  when  all  monthly  data 
(June  and  July  1999)  were  included  from  inside  haul- 
outs  (/=23.4,  P=0.10). 


528 


Fishery  Bulletin  102(3) 


Rookeries  (outside)                         Haul-outs  (inside) 

80- 

F                                           H                    W     G 
94    95    96    97    98    99    94    98    99    94    94 
o 

0 

95             96                              97     99 
Oct   Dec  Mar    Apr  May  Nov   Feb    Jun    Jul 

60- 
E 

o> 

c 
0 

£    40- 

o 

"O 
CD 

I 

( 

>    8 
1 

V 

-- 

0    I 

o         o  | 

o  oo  oo    h 

o 

Adult 
Subadult 

1 

1 

o 

Correc 

|\J 

o 
i            i 

1                 o                       °                 1 

■.  ,8 

r 

8    . 

Adolescent 

§                                       o 

0 

o 

Juvenile 

ty=42    29    35    59    9     18    17    18    19    33   49 

n,=108    56     76     62     69     75     50     35     33 

0- 

ns=19    13    10   26    3     10    9     10    9      6     11 

ns=33     22     20     19     28     22     23     10     14 

Figure  3 

Box  plots  of  corrected  fork  length  of  walleye  pollock  iTheragra  chalcogramma)  across 

months  and  years  for  haul-outs  and  rookeries  in  Southeast  Alaska  (Forrester  Island.  F; 

Hazy,  H;  White  Sisters,  W;  Graves  Rock.  G).  Box  widths  correspond  to  relative  sample 

sizes  for  numbers  offish  elements  in,)  and  numbers  of  scats  (ns)  are  also  provided,  and 

stage-class  categories  are  illustrated.  Gray  areas  represent  95%  confidence  intervals. 

the  whiskers  bound  1.5x  the  interquartile  range  I  boxes),  the  circles  denote  outliers,  and 

the  stars  (*)  denote  extremes. 

Overlap  of  size  of  pollock  consumed  by  Steller  sea  lions 
with  size  of  pollock  caught  by  the  fisheries 

The  Canadian  commercial  pollock  trawl  fishery  in  Dixon 
Entrance  between  1993  and  1999  landed  mostly  (93%) 
adult  fish  (mean  FL  =  52.2  ±5.9  cm,  «=2103,  modal  range: 
48-54  cm).  The  majority  (79%)  of  scats  containing  pol- 
lock from  the  Forrester  Island  rookery  in  June  and  July 
also  contained  remains  of  adult  pollock  (mean  corrected 
FL=51.4  ±10  cm,  a?  =  192,  modal  range:  46-52  cm,  ns=81 
scats).  Percentage  overlap  based  on  a  comparison  of 
size-frequency  distributions  totaled  75.1%  for  those  fish 
eaten  around  Forrester  Island  and  52.1%  for  all  fish 
eaten.  However,  the  estimated  overlap  would  have  been 
assumed  incorrectly  to  be  half  these  values  if  DCFs  had 
not  been  applied  to  the  selected  digested  otoliths  and 
bones  (i.e.,  36.7%  overlap  at  Forrester  and  24.1%  for  all 
areas  combined).  Clearly  overlap  levels  would  have  been 
further  underestimated  if  structures  in  poor  condition 
had  been  included  in  our  analyses. 


Discussion 

Only  57%  of  the  scats  (303  of  531)  that  contained  suit- 
able pollock  remains  had  structures  that  were  in  good 


enough  condition  to  be  measured  reliably.  Numbers  of 
elements  in  good  or  fair  condition  (rc  =  909)  averaged 
three  per  scat,  and  a  very  small  fraction  of  these  con- 
sisted of  otoliths  (<4%).  The  most  numerous  structures 
were  DENT,  QUAD,  and  ANGU  (Table  1).  This  finding 
is  inconsistent  with  feeding  trials  with  captive  Steller 
sea  lions  where  otoliths  were  found  to  be  the  most  com- 
monly occurring  structure  (Cottrell  and  Trites,  2002; 
Tollit  et  al.,  2003). 

Different  structures  yielded  somewhat  different  mean 
sizes  of  pollock,  although  95%  confidence  intervals  gen- 
erally overlapped,  ranging  between  37  and  52  cm  for 
bones  (Table  2).  Such  discrepancy  is  not  surprising 
given  that  different  bones  originate  from  different  scats 
and  possibly  different  fish  (even  within  a  single  scat). 
Our  comparison  of  estimates  with  all  structures  versus 
MNI  selections  indicates  that  the  potential  effect  of 
double  counting  (and  measuring)  fish  within  a  single 
scat  is  likely  negligible  with  large  sample  sizes  (Fig.  2). 
Although  the  use  of  all-structure  data  to  estimate  fish 
length  results  in  a  greatly  increased  sample  size,  there 
remains  an  underlying  assumption  that  all  structures 
are  affected  equally  by  digestion.  Tollit  et  al.  (2004, 
this  issue)  found  no  significant  difference  in  the  degree 
of  erosion  across  the  three  size  ranges  (28.5-45.0  cm 
FL)  for  each  structure  within  each  condition  category. 


Tollit  et  al.:  Sizes  of  walleye  pollock  consumed  by  Eumetopias  /ubatus 


529 


They  also  found  that  the  relative  shape,  structure,  and 
proportion  of  the  morphological  features  used  to  es- 
timate erosion  were  consistent  for  both  smaller  and 
larger  fish.  We  therefore  assumed  that  the  DCFs  in  that 
study  could  be  used  reliably  for  the  fish  in  our  study 
outside  of  the  experimental  size  range  in  which  they 
were  considered. 

Applying  DCFs  increased  mean  fork  length  estimates 
by  23%  (from  34.4  to  42.4  cm)  on  average  and  resulted 
in  adult  fish  contributing  44%  to  the  sea  lion  diet  by 
number  and  74%  by  mass.  The  contribution  of  juvenile 
fish  was  insignificant.  Applying  valid  correction  factors 
clearly  provides  better  insights  into  prey-size  selection 
and  consequently  niche  overlap.  It  should  also  lead  to 
more  precise  estimates  of  mass  of  prey  consumed  and 
the  number  of  prey  within  a  scat  (Ringrose,  1993;  Tollit 
et  al.,  1997;  Laake  et  al.,  2002). 

Over  61  species  of  prey  were  identified  in  the  diet  of 
Steller  sea  lions  in  Southeast  Alaska  from  1993  to  1999 
(Trites  et  al.3).  The  most  common  prey  were  walleye  pol- 
lock, Pacific  herring  (Clupea  pallasi).  Pacific  sand  lance 
(Ammodytes  hexapterus),  salmon  iOncorhynchus  spp.), 
arrowtooth  flounder  (Remhardtius  stomias),  rockfish  (Se- 
bastes  spp.),  skates  {Raja  spp.),  and  cephalopods.  During 
summer,  gadids  (most  of  which  were  pollock)  made  up 
27%  of  the  diet,  and  increased  to  49-62%  of  the  diet  at 
other  times  of  the  year  (Trites  et  al.3),  confirming  that 
pollock  are  a  significant  component  of  the  diet. 

Steller  sea  lions  consumed  a  wide  size  range  of  pollock 
in  Southeast  Alaska;  the  bulk  of  fish  fell  between  20  and 
60  cm  and  peaked  between  44  and  52  cm  (Fig.  2).  The 
contribution  of  juvenile  fish  (<20  cm)  was  insignificant. 
The  only  historical  data  to  compare  with  these  results 
are  those  from  the  stomach  samples  of  eight  Steller  sea 
lions  collected  from  Southeast  Alaska  in  1986  (Calkins 
and  Goodwin1).  Pollock  lengths  backcalculated  from  all 
otoliths  found  in  the  stomachs  were  generally  shorter 
(mean  FL  =  25.5  ±10.4  cm,  range;  4. 8-55. 7cm,  n  =  80) 
than  our  estimates  from  multiple  structures  found  in 
scats  collected  during  the  1990s  (mean  FL  =  42.4  ±11.6 
cm,  range:  10.0-78.1  cm,  n  =  909).  It  should  be  noted 
that  we  derived  our  estimates  after  removing  heavily 
eroded  structures  and  applying  DCFs,  whereas  Calkins 
and  Goodwin1  did  not  account  for  partial  digestion. 
However  our  estimates  of  pollock  length  would  have 
been  similar  to  those  of  Calkins  and  Goodwin1  if  we 
had  used  only  otoliths  and  had  not  corrected  for  diges- 
tion (Table  2).  Although  Frost  and  Lowry  (1980)  found 
no  significant  difference  between  the  size  of  otoliths 
obtained  from  stomachs  and  intestines  of  ribbon  seals, 
underestimates  of  fish  size  determined  from  otoliths 
from  stomach  samples  will  depend  on  the  time  since 
ingestion  (i.e.,  on  the  extent  of  digestion). 

One  possible  explanation  for  the  virtual  absence  of  ju- 
venile pollock  in  the  scats  we  examined  is  that  the  rela- 
tively smaller  structures  of  smaller  fish  were  more  likely 
to  be  completely  digested,  and  were  therefore  underrep- 
resented  in  the  scats  (Tollit  et  al.,  1997;  Bowen,  2000). 
However,  juvenile  pollock  otoliths  and  bones  were  found 
in  large  numbers  in  a  number  of  scats  collected  from 


the  western  stock  (Zeppelin  et  al.,  2004,  this  issue). 
Clearly,  the  potential  for  underestimating  smaller  fish 
depends  heavily  on  the  balance  between  relative  re- 
covery rates  and  the  number  of  different  size  fish  con- 
sumed in  a  meal.  For  example,  if  an  animal  needs  to 
eat  5  kg  a  day,  then  it  would  have  to  consume  195  15.5- 
cm  pollock,  but  less  than  ten  41-cm  pollock.  Given  that 
large  pollock  bones  are  at  least  three  times  more  likely 
than  small  bones  to  pass  through  the  digestive  tract 
(Tollit  et  al.,  2003;  D.  J.  Tollit,  unpubl.  data),  the  sheer 
numbers  of  small  pollock  in  this  example  would  lead 
to  a  conclusion  that  smaller  fish  were  more  important 
numerically,  when  in  fact  they  were  equally  important. 
Conversely,  the  relative  proportion  of  large  fish  is  likely 
to  be  overestimated  if  ten  large  and  ten  small  pollock 
are  consumed  together.  The  generally  low  number  of 
structures  per  scat  provides  little  information  to  assess 
this  balance.  Hence  we  must  assume  that  our  results 
are  representative  and  unbiased. 

Steller  sea  lions  in  Southeast  Alaska  did  not  seem  to 
eat  fish  over  65  cm.  Whether  or  not  sea  lions  do  not  tar- 
get large  fish,  or  whether  large  fish  are  harder  to  catch 
and  handle,  or  are  encountered  at  a  lower  rate  is  not 
known.  However,  large  fish  could  be  under-represented 
in  scats  if  large  fish  cannot  be  swallowed  whole,  and 
head  skeletal  parts  are  lost  while  the  fish  is  torn  apart 
on  the  surface  (Olesiuk  et  al.,  1990;  Wazenbock,  1995) 
or  if  bone  regurgitation  is  size  specific. 

Regional,  geographical,  and  temporal  variation  in  sizes  of 
pollock  consumed 

Stomach  samples  collected  in  1975-78  and  1985-86  in 
the  Gulf  of  Alaska  contained  substantial  numbers  of 
juvenile  pollock,  as  well  as  larger  fish  (mode:  39-43  cm). 
In  1985,  the  distribution  of  sizes  consumed  by  sea  lions 
around  Kodiak  Island  appeared  to  mimic  that  of  the  pol- 
lock population  (Merrick  and  Calkins.  1996).  However, 
juvenile  sea  lions  ate  significantly  smaller  and  relatively 
more  juvenile  pollock  than  adult  sea  lions.  Stomachs 
from  the  Gulf  of  Alaska  contained  an  average  of  49  pol- 
lock (1975-78)  and  72  pollock  (1985)  compared  with  1.6 
pollock  per  scat  in  Southeast  Alaska.  In  the  Bering  Sea, 
90  stomachs  were  examined  between  1975  and  1981  by 
using  only  non-eroded  otoliths,  and  these  also  contained 
mainly  (76%)  juvenile  pollock  (mean  FL=29.3  cm),  but 
also  some  adult  fish  (Frost  and  Lowry,  1986). 

Between  1998  and  2000,  Steller  sea  lions  across  the 
range  of  the  western  population  in  Alaska  consumed 
pollock  averaging  39.3  ±14.3  cm  (range:  3.7-70.8  cm, 
Zeppelin  et  al.,  2004,  this  issue).  This  finding  suggests 
that  sea  lions  may  have  been  less  reliant  on  juvenile 
pollock  than  they  were  during  the  1970s  and  1980s. 
Apparent  differences  may  reflect  differences  in  pollock 
year-class  strength,  and  thus  differences  in  the  domi- 
nant size  classes  that  were  available  to  be  consumed. 
However,  Zeppelin  et  al.  (2004,  this  issue)  reported 
that  the  size  distribution  of  walleye  pollock  consumed 
by  Steller  sea  lions  between  1998  and  2000  did  not  ap- 
pear to  fluctuate  with  year-class  strength,  unlike  the 


530 


Fishery  Bulletin  102(3) 


sizes  of  Atka  mackerel  {Pleurogrammus  monopterygius) 
consumed  in  western  Alaska. 

Comparing  samples  collected  at  rookeries  from  the 
eastern  and  western  populations  reveals  that  sea  lions 
in  the  western  stock  ate  significantly  greater  numbers 
of  smaller  pollock  and  fewer  adults  in  summer  than  sea 
lions  in  Southeast  Alaska  (Zeppelin  et  al.,  2004,  this  is- 
sue; and  our  study).  However,  both  eastern  and  western 
stock  sea  lions  using  haul-outs  in  March  (winter)  ate 
similar  size  pollock.  Adult  pollock  occurred  more  fre- 
quently in  scats  collected  from  rookeries  along  the  open 
ocean  coastline  of  Southeast  Alaska  during  June  and 
July  (74%  adults)  than  they  did  in  scats  from  haul-outs 
located  in  inside  waters  between  October  and  May  (51% 
adults).  Scats  collected  at  rookeries  can  be  considered  to 
be  from  adult  female  sea  lions  and  to  a  lesser  extent  from 
adult  males,  whereas  those  collected  at  haul-outs  during 
other  times  of  the  year  contain  a  more  diverse  mix  of  age 
groups,  including  greater  numbers  of  younger  sea  lions. 
Thus  it  is  uncertain  whether  observed  size  differences 
in  pollock  between  these  two  groups  are  seasonal  or  due 
more  to  size  preferences  of  different  aged  animals.  Lim- 
ited support  for  the  former  comes  from  the  similar  size 
pollock  observed  in  the  scats  between  the  two  groups  in 
June  and  July  of  1999.  Overall,  however,  it  is  unknown 
whether  the  consumption  patterns  observed  are  a  result 
of  an  actual  size  selection  of  prey  or  if  they  result  from  co- 
incidental distributions  of  sea  lions  and  prey-size  classes. 
Some  pinnipeds  may  select  prey  of  particular  sizes  (Sin- 
clair et  al.,  1994)  and  may  encounter  difficulties  if  they 
cannot  switch  to  other  sizes  or  species  if  the  abundance 
of  preferred  prey  is  reduced.  Fine-scale  studies  are  now 
being  undertaken  to  address  such  uncertainties. 

There  are  few  assessments  of  pollock  stock  size  for 
the  1990s  in  Southeast  Alaska  (Martin,  1997).  However 
the  biomass  is  believed  to  have  been  low  compared  to 
other  regions  of  Alaska.  Juvenile  pollock  are  known  to 
congregate  in  the  shallow  inside  waters  of  Southeast 
Alaska  during  winter  (Sigler0)  but  are  also  known  to 
occur  in  significant  numbers  in  the  summer  in  waters 
shallower  than  200  meters  on  the  outer  coastline  (Mar- 
tin, 1997).  Recruitment  of  1-year-old  fish  was  found  to 
be  high  during  acoustic  studies  in  1994  and  1999  in  the 
Gulf  of  Alaska  (Guttormsen  et  al.,  2003). 

Steller  sea  lions  using  rookeries  in  Southeast  Alaska 
consumed  mainly  adult  pollock  between  1994  and  1999 
and  showed  no  evidence  of  tracking  any  abundant  age 
class  of  pollock.  However,  the  trend  in  increasing  length 
estimates  for  inside  haul-outs  after  1995  (Fig.  3 1  does 
suggest  that  sea  lions  might  be  tracking  a  particu- 
lar age  class  of  prey.  Certainly  a  greater  range  of  age 
classes  were  consumed  at  these  haul-outs  (Fig.  3). 

Scientific  trawls  in  1996  indicated  that  the  larger  pol- 
lock on  the  outside  coastline  occurred  generally  in  wa- 
ters 201-300  m  deep  during  daylight  hours  (Martin, 


B  Sigler,  M.  F.  2003.  Unpubl.  data.  Auke  Bay  Lab,  National 
Marine  Fisheries  Service.  1  L305  Glacier  Highway,  Juneau. 
AK  99801. 


1997)  and  that  smaller  pollock  were  present  in  shallower 
depths.  Larger  pollock  tend  to  disperse  and  move  to 
shallow  waters  to  feed  at  night  (Smith,  1981).  Thus,  the 
observed  crepuscular  and  nighttime  foraging  by  lactating 
Steller  sea  lions  (Higgins  et  al.,  1988;  Trites  and  Porter, 
2002)  would  be  a  logical  foraging  strategy  to  capture 
adult  pollock.  Other  important  factors,  in  addition  to 
depth,  that  likely  influence  size  selection  include  prey 
density  and  spatial  distribution  in  relation  to  rookeries 
and  haul-outs.  Given  both  the  greater  mass  and  energy 
content  of  adults  compared  with  juveniles  (Perez,  1994; 
Anthony  et  al.,  2000),  the  selection  of  adults  would  be  an 
energy  efficient  strategy — all  other  things  being  equal. 

Overlap  in  sizes  of  pollock  consumed  by  Steller  sea  lions 
and  sizes  of  pollock  caught  by  fisheries 

There  was  no  commercial  fishery  for  pollock  in  South- 
east Alaska  during  the  1990s.  However,  a  small  fishery 
occurred  in  nearby  Dixon  Entrance,  B.C.,  that  might 
indicate  sizes  that  could  have  been  caught  in  Southeast 
Alaska  if  a  fishery  had  occurred.  Overlap  in  sizes  of 
pollock  caught  by  the  B.C.  fishery  with  those  taken  by 
Steller  sea  lions  further  north  (our  study)  more  than 
doubled  after  applying  digestion  correction  factors  (from 
24%  to  52%).  Similarly,  high  levels  of  overlap  were  also 
found  between  the  sizes  of  pollock  consumed  by  the 
western  stock  (1998-2000)  and  those  caught  in  the 
same  region  by  fisheries  (after  our  DCFs  were  applied  to 
structures  recovered  from  scats — Zeppelin  et  al.,  2004. 
this  issue).  A  high  degree  of  overlap  in  size  highlights 
a  potential  conflict  between  fisheries  and  sea  lions,  but 
this  overlap  cannot  be  considered  indicative  of  competi- 
tion unless  the  resource  that  fisheries  and  sea  lions  seek 
is  limited  across  the  space  and  time  in  question  (Krebs 
and  Davies,  1991). 


Conclusions 

Our  study  provides  the  first  substantial  description  of 
the  size  of  pollock  eaten  by  Steller  sea  lions  in  South- 
east Alaska.  It  also  shows  the  benefits  of  using  bones 
other  than  otoliths  to  estimate  the  sizes  of  prey  eaten 
by  Steller  sea  lions,  and  the  importance  of  correcting 
for  degree  of  digestion.  Accurately  reconstructing  the 
sizes  of  bones  and  otoliths  recovered  from  scats  has  a 
significant  bearing,  in  turn,  on  accurately  determining 
the  mass  of  prey  consumed,  and  on  the  extent  of  overlap 
of  sizes  of  prey  consumed  and  sizes  of  the  same  resource 
caught  in  commercial  fisheries. 

We  found  that  Steller  sea  lions  in  Southeast  Alaska 
consumed  a  large  proportion  of  adult  pollock  and  few 
juveniles  between  1994  and  1999.  Although  greater 
proportions  of  juvenile  and  adolescent  pollock  were  con- 
sumed over  the  same  period,  during  the  summer  in  the 
Gulf  of  Alaska  and  Bering  Sea,  larger  size  fish  still  were 
the  most  abundant  prey  item  in  the  diet  of  sea  lions. 
A  comparison  of  these  estimates  with  the  lengths  of  pol- 
lock consumed  during  the  1970s  and  1980s  shows  that 


"To Hit  et  al.:  Sizes  of  walleye  pollock  consumed  by  Eumetopias  jubatus 


531 


Steller  sea  lions  can  consume  a  wide  range  of  different 
size  pollock  (4-78  cm).  Whether  or  not  these  differences 
in  sizes  of  pollock  consumed  between  regions  and  de- 
cades reflect  differences  in  availability,  size  preferences, 
or  year-class  strength  is  not  known  and  requires  further 
study  primarily  with  fine-scale  data  from  scientific  sur- 
veys and  concurrent  scat  collections. 


Acknowledgments 

Funding  was  provided  to  the  North  Pacific  Universities 
Marine  Mammal  Research  Consortium  by  the  National 
Oceanographic  Atmospheric  Administration  and  the 
North  Pacific  Marine  Science  Foundation.  We  would 
like  to  thank  the  contribution  of  personnel  of  the  UBC 
Marine  Mammal  Research  Unit,  ADF&G,  T.  K.  Zep- 
pelin, K.  A.  Call,  A.  J.  Winship,  E.  H.  Sinclair,  and 
two  anonymous  reviewers.  We  are  also  grateful  to  J.  L. 
Laake  and  R.  Jov  for  statistical  advice. 


Literature  cited 

Anthony,  J.  A..  D.  D.  Roby,  and  K.  R.  Turco. 

2000.     Lipid  content  and  energy  density  of  forage  fishes 
from  the  northern  Gulf  of  Alaska.     J.  Exp.  Mar.  Biol. 
Ecol.  248:53-78. 
Benson,  A.  J.,  and  A.  W.  Trites. 

2002.     A  review  of  the  effects  of  regime  shifts  on  the 
production   domains   in   the   eastern   North   Pacific 
Ocean.     Fish  and  Fisheries  3:95-113. 
Beverton,  R.  J.  H. 

1985.     Analysis  of  marine  mammal-fisheries  interaction. 
In  Marine  mammals  and  fisheries  (J.  R.  Beddington.  R. 
J.  H.  Beverton,  and  D.  M.  Lavigne  eds.),  p.  3-33.     George 
Allen  &  Unwin,  Boston,  MA. 
Bowen,  W.  D. 

2000.  Reconstruction  of  pinniped  diets:  accounting 
for  complete  digestion  of  otoliths  and  cephalopod 
beaks.     Can.  J.  Fish.  Aquat.  Sci.  57:898-905. 

Browne,  R.  J.  L.  Laake,  and  R.  L.  DeLong. 

2002.     Improving  pinniped  diet  analyses  through  iden- 
tification  of  multiple   skeletal   structures   in   fecai 
samples.     Fish.  Bull.  100:423-433. 
Calkins,  D.  G„  D.  C.  McAllister,  K.  W.  Pitcher,  and 
G.  W.  Pendleton. 

1999.     Steller  sea  lion  status  and  trend  in  Southeast 
Alaska:  1979-1977.     Mar.  Mamm.  Sci.  15(21:462-477. 
Cottrell,  P.  E.,  and  A.  W.  Trites. 

2002.     Classifying  prey  hard  part  structures  recovered 
from  fecal  remains  of  captive  Steller  sea  lions  (Eume- 
topias jubatus).     Mar.  Mamm.  Sci.  18:525-539. 
Dorn,  M.,  A.  Hollowed,  E.  Brown,  B.  Megrey,  C.  Wilson,  and 
J.  Blackburn. 

2001.  Assessment  of  the  walleye  pollock  stock  in  the  Gulf 
of  Alaska.  In  Stock  assessment  and  fishery  evaluation 
report  for  the  groundfish  resources  of  the  Gulf  of  Alaska, 
90  p.  Prepared  by  the  Gulf  of  Alaska  Groundfish  Plan 
Team,  North  Pacific  Fishery  Management  Council,  W 
4th  Avenue,  Suite  306,  Anchorage,  AK  99501. 

Frost,  K.  J.,  and  L.  F.  Lowry. 

1980.  Feeding  of  ribbon  seals  fPhoca  fasciata)  in  the 
Bering  Sea  in  spring.     Can.  J.  Zool.  58:1601-1607. 


1981.  Trophic  importance  of  some  marine  gadids  in 
northern  Alaska  and  their  body-otolith  size  relation- 
ships. Fish.  Bull.  79:187-192. 
1986.  Sizes  of  walleye  pollock,  Theragra  chalcogramma, 
consumed  by  marine  mammals  in  the  Bering  Sea.  Fish. 
Bull.  84:192-197. 
Guttormsen,  M.  A.,  C.  D.  Wilson,  and  S.  Stienessen. 

2003.     Results  of  the  February  and  March  2003  echo 
integration-trawl  surveys  of  walleye  pollock  (Theragra 
chalcogramma  i  conducted  in  the  gulf  of  Alaska,  cruises 
MF2003-01  and  MF2003-05.     In  Stock  assessment  and 
fishery  evaluation  report  for  the  groundfish  resources 
of  the  Gulf  of  Alaska,  August  2003,  45  p.     Prepared 
by  the  Gulf  of  Alaska  Groundfish  Plan  Team,  North 
Pacific  Fishery  Management  Council.  P.O.  Box  103136, 
Anchorage,  AK  99510. 
Higgins,  L.  V..  D.  P.  Costa,  A.  C.  Huntley,  and  B.  J.  Le  Boeuf. 
1988.     Behavioral  and  physiological  measurements  of 
maternal  investment  in  the  Steller  sea  lion,  Eumetopias 
jubatus.     Mar.  Mamm.  Sci.  4:44-58. 
Hunt,  G.  L.,  Jr.,  A.  S.  Kitaysky,  M.  B.  Decker.  D.  E.  Dragoo,  and 
A.  M.  Springer. 

1996.  Changes  in  the  distribution  and  size  of  juvenile 
walleye  pollock.  Theragra  chalcogramma.  as  indicated  by 
seabird  diets  at  the  Pribilof  Islands  and  by  bottom  trawl 
surveys  in  the  Eastern  Bering  Sea,  1975  to  1993.  In 
Ecology  of  juvenile  walleye  pollock,  Theragra  chalco- 
gramma. Papers  from  the  workshop  "The  importance  of 
prerecruit  walleye  pollock  to  the  Bering  Sea  and  North 
Pacific  ecosystems"  Seattle,  WA,  October  28-30,  1993 
(R.  D.  Brodeur.  P.  A.  Livingston,  T.  R.  Loughlin,  and 
A.  B.  Hollowed,  eds.),  p.  125-139.  NOAA  Tech.  Rep. 
NMFS  126.     [NTIS  no.  PB97-155188.1 

Krebs,  J.  R„  and  N.  B.  Davies. 

1991.  Behavioral  ecology:  an  evolutionary  approach,  3rd 
ed.,  483  p.     Blackwell  Scientific,  Boston,  MA. 

Laake,  J.  L..  P.  Browne.  R.  L.  DeLong.  and  H.  R.  Huber. 

2002.     Pinniped  diet  composition:  a  comparison  of  esti- 
mation models.     Fish.  Bull.  100:434-477. 
Loughlin,  T.  R.,  A.  S.  Perlov,  and  V.  A.  Vladimirov. 

1992.  Range-wide  survey  and  estimation  of  total 
number  of  Steller  sea  lions  in  1989.  Mar.  Mamm.  Sci. 
8:220-239. 

Loughlin.  T.  R„  and  A.  E.  York. 

2000.     An  accounting  of  the  sources  of  Steller  sea  lion 
mortality.     Mar.  Fish.  Rev.  62(41:40-45. 
Martin,  M.  H. 

1997.  Data  report:  1996  Gulf  of  Alaska  bottom  trawl 
survey.  NOAA  Tech.  Memo.  NMFS-AFSC-82,  235  p. 
[NTIS  no.  PB98-103930.I 

Merrick,  R.  L.,  R.  Brown.  D.  G.  Calkins,  and  T  R.  Loughlin. 

1995.  A  comparison  of  Steller  sea  lion,  Eumetopias  juba- 
tus, pup  masses  between  rookeries  with  increasing  and 
decreasing  populations.     Fish.  Bull.  93(4):753-758. 

Merrick,  R.  L.,  and  D.  G.  Calkins. 

1996.  Importance  of  juvenile  walleye  pollock,  Theragra 
chalcogramma,  in  the  diet  of  Gulf  of  Alaska  Steller 
sea  lions,  Eumetopias  jubatus.  In  Ecology  of  juvenile 
walleye  pollock,  Theragra  chalcogramma.  Papers  from 
the  workshop  "The  importance  of  prerecruit  walleye 
pollock  to  the  Bering  Sea  and  North  Pacific  ecosystems" 
Seattle,  WA,  October  28-30,  1993  (R.  D.  Brodeur,  P.  A. 
Livingston,  T.  R.  Loughlin,  and  A.  B.  Hollowed,  eds.), 
p.  153-166.  NOAA  Tech.  Rep.  NMFS  126.  [NTIS  no. 
PB97-155188.] 


532 


Fishery  Bulletin  102(3) 


Milette,  L.  L.,  and  A.W.  Trites. 

2003.  Maternal  attendance  patterns  of  Steller  sea  lions 
(Eumetopias  jubatus)  from  stable  and  declining  popula- 
tions in  Alaska.     Can.  J.  Zool.  81:  340-348. 

NRC  (National  Research  Council) 

2003.  Decline  of  the  Steller  sea  lion  in  Alaskan  waters: 
untangling  food  webs  and  fishing  nets,  204  p.  Prepared 
by  the  Committee  on  the  Alaska  Groundfish  Fishery 
and  Steller  sea  lions.  National  Academies  Press.  Wash- 
ington, DC. 

Olesiuk,  P.  F.,  M.  A.  Bigg,  G.  M.  Ellis,  S.  J.  Crockford,  and 
R.  J.  Wigen. 

1990.     An  assessment  of  the  feeding  habits  of  harbour 
seals  [Phoca  vitulina)  in  the  Strait  of  Georgia,  British 
Columbia  based  on  scat  analysis.     Can.  Tech.  Rep.  Fish. 
Aquat.  Sci.  1730,  135  p. 
Perez,  M.  A. 

1994.     Calorimetry  measurements  of  energy  value  of  some 
Alaskan  fishes  and  squids.     NOAA  Tech.  Memo.  NMFS- 
AFSC-32,  32  p.     [NTIS  no.  PB94-152907.] 
Pitcher,  K.  W. 

1981.     Prey  of  the  Steller  sea  lion,  Eumetopias  jubatus, 
in  the  Gulf  of  Alaska.     Fish.  Bull.  79:467-472. 
Ringrose,  T.  J. 

1993.  Bone  counts  and  statistics:  a  critique.  J.  Archaeol. 
Sci.  20:121-157. 

Sease,  J.  L.,  W.  P.  Taylor,  T  R.  Loughlin,  and  K.  W.  Pitcher. 

2001.  Aerial  and  land-based  surveys  of  Steller  sea  lions 
(Eumetopias  jubatus)  in  Alaska,  June  and  July  1999  and 
2000.     NOAA  Tech.  Mem.  NMFS-ASFC-122,  52  p. 

Sinclair,  E.  H.,  T.  Loughlin,  and  W.  Pearcy. 

1994.  Prey  selection  by  northern  fur  seals  (Callorhinus 
ursinus)  in  the  eastern  Bering  Sea.  Fish.  Bull.  92: 
144-156. 

Smith,  G.  B. 

1981.     The  biology  of  walleye  pollock.     In  The  eastern 
Bering  Sea  Shelf:  oceanography  and  resources  (D.  W. 
Hood  and  J.  A.  Calder  eds.),  vol.  1,  p.  527-551.     Univ. 
Washington  Press,  Seattle,  WA. 
Tollit,  D.  J.,  S.  G.  Heaslip,  T.  K.  Zeppelin,  R.  Joy,  K.  A.  Call, 
and  A.  W.  Trites. 

2004.  A  method  to  improve  size  estimates  of  walleye 
pollock  (Theragra  ehaleogramma)  and  Atka  mackerel 


(Pleurogrammus  monopterygius)  consumed  by  pinnipeds: 
digestion  correction  factors  applied  to  bones  and  otoliths 
recovered  in  scats.     Fish.  Bull.  102:498-508. 
Tollit,  D.  J.,  M.  J.  Steward,  P.  M.  Thompson,  G.  J.  Pierce, 
M.  B.  Santos,  and  S.  Hughes. 

1997.     Species  and  size  differences  in  the  digestion  of 
otoliths  and  beaks:  implications  for  estimates  of  pin- 
niped diet  composition.     Can.  J.  Fish.  Aquat.   Sci. 
54:105-119. 
Tollit,  D.  J.,  M.  Wong,  A.  J.  Winship,  D.  A.  S.  Rosen,  and 
A.  W.  Trites. 

2003.     Quantifying  errors  associated  with  using  prey 
skeletal  structures  from  fecal  samples  to  determine  the 
diet  of  the  Steller  sea  lion  (Eumetopias  jubatus).     Mar. 
Mamm.  Sci.  19(41:724-744. 
Trites,  A.  W.,  and  L.  P.  Donnelly. 

2003.  The  decline  of  Steller  sea  lions  in  Alaska:  a  re- 
view of  the  nutritional  stress  hypothesis.  Mamm.  Rev. 
33:3-28. 

Trites,  A.  W.,  and  P.  A.  Larkin. 

1996.     Changes  in  the  abundance  of  Steller  sea  lions 
{Eumetopias  jubatus)  in  Alaska  from  1956  to  1992:  how 
many  were  there?     Aquat.  Mamm.  22:153-166. 
Trites,  A.  W.,  and  B.  T.  Porter. 

2002.     Attendance  patterns  of  Steller  sea  lions  (Eumeto- 
pias jubatus)  and  their  young  during  winter.     J.  Zool. 
(Lond.l  256:547-556. 
Walline,  P.  D. 

1983.     Growth  of  larval  and  juvenile  walleye  pollock 
related  to  year-class  strength.     Ph.D.  diss.,  144  p.     Univ. 
Washington,  Seattle,  WA. 
Wazenbock,  J. 

1995.     Changing  handling  times  during  feeding  and  con- 
sequences for  prey  size  selection  of  0+  zooplanktivorous 
fish.     Oecologia  (Heidelb.l  104:372-586. 
Zeppelin,  T.  K.,  D.  J.  Tollit,  K.  A.  Call.  T.  J.  Orchard,  and 
C.  J.  Gudmundson. 

2004.  Sizes  of  walleye  pollock  (Theragra  ehaleogramma) 
and  Atka  mackerel  (Pleurogrammus  monopterygius)  con- 
sumed by  the  western  stock  of  Steller  sea  lions  (Eume- 
topias jubatus)  in  Alaska  from  1998  to  2000.  Fish. 
Bull.  102:509-521. 


533 


Abstract— We  examined  movement 
patterns  of  sportfish  that  were  tagged 
in  the  northern  Indian  River  Lagoon, 
Florida,  between  1990  and  1999  to 
assess  the  degree  of  fish  exchange 
between  an  estuarine  no-take  zone 
iNTZ)  and  surrounding  waters.  The 
tagged  fish  were  from  seven  spe- 
cies: red  drum  (Sciaenops  ocella- 
tus);  black  drum  (Pogonias  cromis); 
sheepshead  (Archosargus  probato- 
cephalus);  common  snook  iCentropo- 
mus  undecimalis);  spotted  seatrout 
(Cynoscion  nebulosus);  bull  shark 
{Carcharhinus  leucas);  and  crevalle 
jack  (Caranx  hippos).  A  total  of  403 
tagged  fish  were  recaptured  during 
the  study  period,  including  65  indi- 
viduals that  emigrated  from  the  NTZ 
and  16  individuals  that  immigrated 
into  the  NTZ  from  surrounding  waters 
of  the  lagoon.  Migration  distances 
between  the  original  tagging  location 
and  the  sites  where  emigrating  fish 
were  recaptured  were  from  0  to  150 
km,  and  these  migration  distances 
appeared  to  be  influenced  by  the  prox- 
imity of  the  NTZ  to  spawning  areas 
or  other  habitats  that  are  important 
to  specific  life-history  stages  of  indi- 
vidual species.  Fish  that  immigrated 
into  the  NTZ  moved  distances  rang- 
ing from  approximately  10  to  75  km. 
Recapture  rates  for  sportfish  species 
that  migrated  across  the  NTZ  bound- 
ary suggested  that  more  individuals 
may  move  into  the  protected  habitats 
than  move  out.  These  data  demon- 
strated that  although  this  estuarine 
no-take  reserve  can  protect  species 
from  fishing,  it  may  also  serve  to 
extract  exploitable  individuals  from 
surrounding  fisheries;  therefore,  if 
the  no-take  reserve  does  function 
to  replenish  surrounding  fisheries, 
then  increased  egg  production  and 
larval  export  may  be  more  important 
mechanisms  of  replenishment  than 
the  spillover  of  excess  adults  from  the 
reserve  into  fishable  areas. 


Multidirectional  movements  of 
sportfish  species  between  an  estuarine 
no-take  zone  and  surrounding  waters 
of  the  Indian  River  Lagoon,  Florida 

Derek  M.  Tremain 
Christopher  W.  Harnden 
Douglas  H.  Adams 

Florida  Fish  and  Wildlife  Conservation  Commission 

Florida  Marine  Research  Institute 

1220  Prospect  Avenue,  Suite  285 

Melbourne,  Florida  32901 

Email   Derek  Tremainia  fwcstate-fl. us 


Manuscript  submitted  12  May  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 

20  January  2004  by  the  Scientific  Editor. 

Fish.  Bull  102:533-544  (2004). 


Fishery  reserves  or  no-take  sanctu- 
aries, defined  as  areas  where  all 
fishing  activities  are  prohibited,  are 
increasingly  proposed  as  an  addi- 
tional measure  to  traditional  fishery 
management  practices  for  protecting 
fish  populations  from  overexploita- 
tion  (PDT,  1990;  Bohnsack  and  Ault, 
1996).  The  American  Fisheries  Soci- 
ety recently  issued  a  policy  statement 
on  the  protection  of  marine  fish  stocks 
at  risk  of  extinction  and  supported  the 
development  of  large  marine  reserves 
to  protect  and  rebuild  vulnerable  popu- 
lations (Musick  et  al.,  2000).  Although 
reserves  have  been  established  pri- 
marily in  reef  or  coastal  marine  habi- 
tats, the  potential  to  apply  similar 
management  strategies  in  estuarine 
systems  may  also  be  possible  (Johnson 
et  al.,  1999;  Roberts  et  al.,  2001). 

Reserves  in  estuarine  areas  may 
help  protect  exploitable  fishery  spe- 
cies. Increases  in  species'  sizes  and 
densities  within  these  reserves  may 
also  enhance  adjacent  fisheries  by  two 
separate  mechanisms.  Johnson  et  al. 
(1999)  found  that  an  existing  estua- 
rine no-take  sanctuary  on  Florida's 
central  east  coast  protected  popula- 
tions of  larger,  spawning-age  sport- 
fish species.  As  a  result,  they  sug- 
gested that  protection  of  populations 
in  no-take  sanctuaries  could  also  lead 
to  the  replenishment  of  surrounding 
fisheries  through  increased  egg  pro- 
duction, larval  export,  and  juvenile 
recruitment.  Additionally,  mark-re- 
capture data  have  demonstrated  that 
large  juvenile  and  adult  fishes  emi- 


grate from  estuarine  protected  ar- 
eas to  surrounding  waters  (Bryant 
et  al.,  1989;  Funicelli  et  al.,  1989; 
Johnson  et  al.,  1999;  Roberts  et  al., 
2001;  Stevens  and  Sulak,  2001)  and 
these  data  have  been  used  to  suggest 
that  spillover  of  excess  adult  fish 
from  estuarine  reserve  areas  can  di- 
rectly supplement  nearby  fisheries. 
Roberts  et  al.  (2001)  concluded  that 
the  abundance  of  International  Game 
Fish  Association  based  on  line-class- 
record  catches  in  the  vicinity  of  the 
estuarine  no-take  sanctuary  on  Flor- 
ida's east  coast  resulted  indirectly 
from  protection  and  spillover  of  large 
adults  to  outlying  waters. 

It  has  also  been  suggested  that  re- 
serves protect  areas  of  undisturbed 
habitat  (PDT,  1990),  either  by  design 
or  through  cessation  of  destructive 
practices,  and  reserves  are  common- 
ly established  in  areas  of  pristine, 
productive,  or  otherwise  important 
habitats  required  by  the  species  be- 
ing protected  (e.g.,  Russ.  1985).  Fur- 
thermore, studies  have  shown  that 
protecting  fishery  species  can  indi- 
rectly change  the  overall  community 
structure  (Cole  and  Keuskamp,  1998) 
and,  under  certain  circumstances, 
can  increase  primary  and  secondary 
productivity  (Sala  and  Zabala.  1996; 
Babcock  et  al.,  1999).  The  influence 
of  habitat  quality  on  fish  movements 
in  relation  to  protected  areas  has  not 
been  investigated;  however,  reserve 
habitats  that  offer  potential  advan- 
tages in  the  form  of  improved  habitat 
quality  (Chapman  and  Kramer,  1999) 


534 


Fishery  Bulletin  102(3) 


or  increased  food  and  habitat  availability  could  be  ex- 
pected  to  attract,  or  at  least  retain,  individuals  that 
immigrate  to  the  reserves  from  surrounding  unpro- 
tected habitats.  Reserve  areas  that  attract  and  retain 
exploitable  individuals  from  surrounding  habitats  at 
higher  rates  than  they  replenish  the  surrounding  habi- 
tats could  be  considered  to  be  sinks  in  terms  of  their 
ability  to  directly  supplement  adjacent  fisheries  through 
spillover  of  exploitable-size  individuals.  Fish  emigration 
from  reserve  habitats  and  the  replenishment  of  nearby 
fisheries  is  a  commonly  predicted  benefit  of  harvest  re- 
serves (see  reviews  in  Roberts  and  Polunin,  1991.  and 
Rowley,  1994).  However,  there  are  currently  no  studies 
that  simultaneously  examine  emigration  and  immigra- 
tion in  relation  to  estuarine  reserves  or  that  document 
the  extent  to  which  reserve  areas  may  also  function  to 
withdraw  individuals  from  surrounding  fisheries.  With- 
out an  assessment  of  net  exchange,  the  interpretation  of 
reserve  benefits  with  respect  to  replenishment  cannot 
be  properly  evaluated. 

The  National  Aeronautics  and  Space  Administration 
(NASA)  closed  a  portion  of  the  Indian  River  Lagoon  at 
the  Merritt  Island  National  Wildlife  Refuge  (MINWRi 
on  Florida's  east  coast  for  security  purposes  in  1962.  A 
direct  result  of  this  closure  was  the  effective  creation  of 
an  estuarine  no-take  zone  that  remains  to  the  present 
time.  The  proximity  of  this  no-take  zone  to  productive 
estuarine  fisheries  provided  an  opportunity  to  examine 
sportfish  movements  in  the  area  with  mark-recapture 
methods.  Johnson  et  al.  (1999)  first  documented  sport- 
fish  migrations  out  of  this  no-take  sanctuary,  and  in  a 
related  study,  Stevens  and  Sulak  (2001)  provided  more 
complete  descriptions  of  movement  patterns  of  indi- 
vidual species;  each  of  these  studies  provided  evidence 
that  the  restricted  habitats  protected  fish  populations 
and  that  adult  sportfish  egressed  into  surrounding  wa- 
ters open  to  fishing.  However,  because  all  tagged  fish 
originated  from  within  restricted  habitats,  in  neither  of 
these  studies  was  it  possible  to  consider  the  potential 
for  the  movements  of  fish  into  protected  areas  from 
surrounding  waters.  Therefore,  we  (sponsored  by  tin- 
Florida  Fish  and  Wildlife  Conservation  Commission- 
Florida  Marine  Research  Institute  [hereafter  referred  to 
as  FMRII  Fisheries-Independent  Monitoring  Program) 
tagged  fish  species  throughout  the  northern  Indian 
River  Lagoon  system,  including  both  the  MINWR  no- 
take  zone  and  the  surrounding  lagoon  waters,  from 
1990  to  1999.  We  investigated  the  relationship  between 
sportfish  egress  and  ingress  in  relation  to  the  MINWR 
no-take  zone  and  offer  a  quantitative  foundation  for 
the  discussion  of  net  fish  movements  into  or  away  from 
protected  estuarine  habitats. 


central  east  coast  of  Florida  between  Ponce  de  Leon 
Inlet  in  Volusia  County  and  Jupiter  Inlet  in  Palm  Beach 
County.  The  lagoon  is  composed  of  three  relatively  iso- 
lated basins:  Mosquito  Lagoon,  the  Indian  River  proper, 
and  the  Banana  River  i  Fig.  1).  These  three  basins  main- 
tain hydrological  connections  with  each  other  through 
narrow  man-made  channels  at  Haulover  Canal  and 
the  Merritt  Island  Barge  Canal  (shown  on  Fig.  2)  and 
through  a  natural  channel  at  the  southern  end  of  the 
Banana  River.  Hydrodynamie  exchange  and  fish  passage 
between  the  lagoon  and  the  Atlantic  Ocean  occur  pri- 
marily through  five  inlets,  which  are  concentrated  in  the 
southern  half  of  the  system.  The  hydraulic  lock  system 
located  at  Port  Canaveral  provides  only  an  intermittent 
opportunity  for  exchange  between  the  IRL  and  Atlantic 
Ocean.  Gilmore  et  al.  (1981)  and  Mulligan  and  Snelson 
(1983)  have  provided  detailed  descriptions  of  the  lagoon 
and  its  habitats. 

The  no-take  zone  (NTZ)  created  by  NASA  and  MIN- 
WR is  located  at  the  northern  terminus  of  the  Banana 
River  basin  of  the  lagoon.  An  earthen  causeway  defines 
the  southern  boundary  of  this  no-access  security  area 
and  contains  only  two  openings  that  permit  fish  to  mi- 
grate to  and  from  adjacent  waters.  Much  of  the  natural 
shoreline  and  saltmarsh  habitats  in  the  lagoon  have 
been  altered  for  mosquito  control  purposes.  However, 
actual  shoreline  habitats  surrounding  MINWR — in- 
cluding the  NTZ,  the  northern  Banana  River  basin, 
the  northern  Indian  River  basin,  and  Mosquito  La- 
goon— remain  relatively  undeveloped  in  comparison  to 
the  urban  shoreline  development  in  the  southern  IRL. 
Detailed  descriptions  of  the  habitat  composition  within 
the  NTZ  and  surrounding  study  area  were  provided  by 
Johnson  et  al.  (1999). 

Data  collection 

Fish  were  tagged  as  part  of  several  related  FMRI  proj- 
ects (stratified-random,  fixed-station,  and  directed  sam- 
pling designs)  in  the  northern  IRL  between  1990  and 
1999  (FMRI1).  In  most  cases,  tagging  was  conducted 
opportunistically  on  healthy  fish  following  capture  in 
multipanel  monofilament  gill  nets,  nylon  trammel  nets, 
nylon  haul  seines,  or  on  hook  and  line.  In  other  cases, 
projects  were  designed  specifically  to  assess  tag-recap- 
ture information  (Murphy  et  al.,  1998).  Because  of  the 
focus  of  our  sampling  programs  in  this  area,  the  major- 
ity of  our  tagging  efforts  occurred  north  of  Sebastian 
Inlet  within  the  Indian  and  Banana  River  basins  of 
the  lagoon.  A  small  percentage  of  tags  were  placed  in 
fish  captured  south  of  Sebastian  Inlet  or  in  Mosquito 
Lagoon.  Overall,  our  sampling  collections  in  the  NTZ 


Materials  and  methods 

Study  area 

The  Indian  River  Lagoon  (IRL)  is  a  shallow  barrier 
island  estuarine  system  spanning  25.3  km  along  the 


1  FMRI  (Florida  Fish  and  Wildlife  Conservation  Commis- 
sion). 1999.  Florida  Marine  Research  Institute.  Fisheries- 
independent  monitoring  program,  1999  annual  data  summary 
report  In-house  Report.  Florida  Fish  and  Wildlife  Conser- 
vation Commission.  Florida  Marine  Research  Institute.  LOO 
Eighth  Ave.  S.E.,  St.  Petersburg,  Florida,  33701. 


Tremain  et  ai.:  Sportfish  species  movements  in  relation  to  an  estuarine  no-take  zone 


535 


\  80°30'W 

.  Ponce  Jt  Leon  Inlet 


mi  00'  \\ 


\1I\WR  No-take  Zone  (NTZ) 

upper  Banana  River  I  BR  i 


Atlantic 
Ocean 


GulfoJ 

l/l'Wl  0 


29° 00'  N- 


28   JO'  N- 


28  OO'N- 


Sebasritjn  Inlet 


Ft.  Pierce  Inlet 


27   in  \ 


Florida 


Atlantu 
Ocean 


Si  Lucielnlet 


27"00'N 


Palm  Bead 
County 


Figure  1 

Map  of  Florida  and  the  Indian  River  Lagoon  study  area. 


accounted  for  approximately  20%  of  our  total  sampling 
efforts  and  averaged  approximately  1-2  days/month  over 
the  study  period. 

Fish  were  tagged  by  inserting  50-mm,  70-mm.  or 
100-mm  Hallprint  dart  tags  (Halprint  Ltd.,  Victor  Har- 
bor, South  Australia)  into  the  dorsal  musculature;  the 
plastic  dart  was  lodged  beneath  the  pterygiophores 
of  the  dorsal  fin.  Each  tag  contained  a  visible  exter- 
nal streamer  with  a  unique  alphanumeric  code  and 
instructions  for  anglers  to  contact  us  with  recapture 
information  in  order  to  collect  a  reward  (five  dollars 
or  equivalent).  Information  recorded  at  the  time  of  ini- 
tial tagging  included  the  tag  number,  species  tagged, 
date,  location  (latitude  and  longitude i,  and  fish  length 
(standard,  fork,  and  total  lengths  as  appropriate  for 
the  species).  Recapture  information  on  tagged  fish  was 
collected  through  August  2000  from  angler  reports 
and  from  fish  recaptured  during  FMRI  sampling  ac- 


tivities. Because  of  public-access  prohibitions,  recap- 
ture information  from  inside  the  MINWR  NTZ  was 
gathered  exclusively  through  FMRI  sampling  efforts. 
Data  requested  for  recaptured  fish  included  the  same 
information  as  that  recorded  at  initial  tagging;  however, 
in  several  cases,  length  or  precise  location  information 
returned  from  anglers  was  considered  to  be  unreliable, 
which  prevented  accurate  statistical  comparisons  of 
relationships  involving  recapture  lengths  or  distances 
traveled.  Therefore,  reported  length  data  are  limited  to 
initial  tagging  information  only  (total  length;  TL).  To 
prevent  problems  with  pseudoreplication  for  individuals 
recaptured  on  multiple  occasions,  we  included  only  the 
initial  tag  recovery  data  in  our  calculations  of  recapture 
percentages. 

Overall  patterns  offish  migrations,  including  general 
recapture  locations  and  direction  of  movements  into 
or  away  from  the  NTZ,  were  described  by  using  data 


536 


Fishery  Bulletin  102(3) 


A 

B 

!';;:;""  rb                          Egress 

Catul 

s                   Ingress 

\ 

"   R 

B 

TR 

Merritt  Island 

B*                                      ^-^. 

National  Wildlife 

f                  \ 

1    No-take  ] 

Reflige            /^^        ^\ 

|     No-take  j 

V   zone      / 
Kr     R                   V            J 

R 

y  zone    J 

b  Rft; 

V y 

r              R     :r    Cape 

R  R    Cape 

Canaveral 

B      Canaveral 

Cocoa                     Sr   i} 

• 
Cocoa 

R 

8B           Por, 

Barge  Cona/                Canaveral 

H 

H 

R 

2R        R 

Co 
D 

3 

P6 

R      2R 

• 

• 

Melbourne 

Melbourne 

R 

% 

also  one  "S"  and  one  "J" 

-. 

approximately  75  km  south 

in  the  St.  Lucie  River 

\s 

3RH 

/                                               /« 

Sebastian  Inlet 

Figure  2 

(A)  Recapture  locations  of  tagged  fish  that  migrated  out  of  the  Merritt  Island 
National  Wildlife  Refuge  no-take  zone.  (B)  Original  tagging  locations  offish  that 
migrated  into  the  Merritt  Island  National  Wildlife  Refuge  no-take  zone.  R  =  red 
drum,  B  =  black  drum,  S  =  common  snook,  H  =  sheepshead,  T  =  spotted  seatrout, 
J  =  crevalle  jack,  K  =  bull  shark.  Numbers  before  species  codes  (letters)  indicate 
the  number  of  individuals  of  that  species  that  were  captured  at  that  location. 


from  all  available  recapture  sources.  In  contrast,  we 
calculated  migration  rates  exclusively  from  the  recap- 
ture data  collected  during  FMRI  sampling  activities. 
Although  this  procedure  excluded  tag-return  data  from 
recreational  anglers,  it  permitted  a  quantitative  assess- 
ment of  recapture  rates  based  on  standardized  FMRI 
col  lection  gear,  comparable  sampling  effort,  and  lOO'* 
tag  reporting  rates.  We  resolved  potential  problems 
related  to  differences  in  habitat  characteristics  and 
sampling  intensity  by  including  only  data  from  the  NTZ 
and  a  fishable  area  of  a  similar  size  and  habitat  type 
in  the  adjacent  Banana  River  (BR,  Fig.  1).  This  BR 


zone  corresponded  precisely  to  the  sampling  zone  used 
for  population  comparisons  in  Johnson  et  al.  (1999), 
denoted  as  "FBR"  (fished  Banana  River)  in  that  study. 
Species  that  did  not  contribute  any  FMRI  recapture 
information  in  either  of  these  two  areas  were  excluded 
from  our  analyses.  Tag  recovery  and  migration  rates 
were  calculated  separately  for  the  NTZ  and  BR.  For  our 
purposes,  "migration"  was  defined  as  a  directional  fish 
movement  across  the  NTZ  boundary  from  the  original 
tagging  location,  and  we  made  the  assumption  that 
the  migration  patterns  of  recaptured  fish  represented 
the  migration  patterns  of  the  overall  population.  Rela- 


Tremain  et  al.:  Sportfish  species  movements  in  relation  to  an  estuarine  no-take  zone 


537 


Table  1 

Summary  of  tagging  and  recapture  data  for  seven  of  the  most  common  sportfish  species  tagged  by  FMRI  scientists 
ern  Indian  River  Lagoon  study  area.  Locations  where  tag  and  recapture  data  were  collected  are  separated  into  the 
(NTZ)  and  the  surrounding  waters  of  the  Indian  River  Lagoon  (IRL). 

in  the  north- 
no-take  zone 

Species 

No-take  zone 

Indian  River  Lagoon 

Total  no. 
offish 
tagged 

Total  no. 
recaptured 
and  percent 
recaptured 

Tagged 
inside 
NTZ 

Recapture 
location 

Tagged 

outside 

NTZ 

Recapture 
location 

NTZ 

IRL 

NTZ 

IRL 

Bull  shark  (Carcharhinus  leucas) 

1 

1 

24 

1 

25 

2(8.0) 

Common  snook  (Centropomus  undecimalis) 

104 

1 

9 

406 

32 

510 

42(8.2) 

Crevalle  jack  ( Caranx  hippos  I 

55 

1 

59 

1 

114 

2(1.8) 

Sheepshead  tArchosargus probatocephalus) 

597 

6 

520 

26 

1117 

32(2.9) 

Spotted  seatrout  (Cynoseion  nebulosus) 

193 

2 

171 

1 

3 

364 

6(1.6) 

Black  drum  (Pogonias  cromis) 

637 

4 

8 

831 

9 

32 

1468 

53(3.6) 

Red  drum  (Seiaenops  ocellatus) 

720 

30 

40 

1344 

6 

190 

2064 

266(12.9) 

Total 

2307 

37 

65 

3355 

16 

285 

5662 

403)7.1) 

tive  migration  rates  were  calculated  as  the  percentage 
of  recaptured  fish  that  migrated  from  their  original 
tagging  location.  These  migration  rates  and  their  re- 
ciprocal (retention  rates)  were  compared  between  the 
NTZ  and  the  BR  to  determine  the  relative  potential  for 
sportfish  movements  into  or  away  from  protected  habi- 
tats. Chi-square  contingency  tests  for  frequency  data 
(with  Yates's  correction  for  small  sample  sizes)  were 
used  to  test  the  hypothesis  that  recapture  location  was 
independent  of  the  tagging  location. 


Results 

A  total  of  5951  fish  of  27  species  were  tagged  during 
FMRI  sampling  within  the  IRL  between  September  1990 
and  December  1999.  However,  because  95%  of  these  fish 
were  represented  by  only  seven  species  (Table  1),  which 
included  all  fish  that  migrated  across  the  reserve  bound- 
aries, only  these  seven  species  were  considered  further 
in  our  analyses.  Red  drum  (Seiaenops  ocellatus)  was 
the  most  commonly  tagged  species  (n=2064),  followed 
by  black  drum  (Pogonias  cromis,  n  =  1468),  sheepshead 
(Archosargus  probatocephalus,  /?  =1117 ),  common  snook, 
(Centropomus  undecimalis,  n  =  510),  spotted  seatrout 
(Cynoscion  nebulosus,  /;  =  364),  crevalle  jack  (Caranx 
hippos,  n=114),  and  bull  shark  (Carcharhinus  leucas, 
n=25).  Approximately  41%  (n=2307)  of  these  fish  were 
tagged  inside  the  boundaries  of  the  NTZ.  The  remain- 
der («  =  3355>  were  tagged  in  the  surrounding  lagoon. 
Through  August  2000,  403  tagged  fish  (7.1%  of  total) 
were  recaptured  and  reported  either  by  FMRI  staff 
sampling  in  the  lagoon  or  by  the  public.  Overall  recap- 
ture rates  were  highest  for  red  drum  (12.9%),  followed 
by  those  for  common  snook  (8.2%),  bull  shark,  (8.0%), 
black  drum  (3.6%),  and  sheepshead  (3.0%). 


Tagged  fish  were  generally  representative  of  the  larg- 
er mobile  members  of  the  species  and  encompassed  the 
legally  exploitable  size  ranges  for  species  with  man- 
agement restrictions  (Table  2).  For  species  except  the 
bull  shark  and  red  drum,  mean  lengths  of  fish  tagged 
inside  the  NTZ  exceeded  those  of  fish  tagged  outside 
the  NTZ. 

Approximately  25%  (n  =  W2)  of  the  403  total  recap- 
tured fish  were  fish  originally  tagged  inside  the  NTZ 
(Table  1).  Thirty-seven  of  these  fish  were  also  recovered 
inside  the  NTZ.  including  three  red  drum  that  were 
subsequently  recaptured  on  multiple  occasions  in  the 
protected  area.  The  remaining  65  recaptured  fish  were 
caught  after  emigrating  to  outlying  waters,  including 
one  red  drum  that  was  recaptured  a  second  time  outside 
the  NTZ.  Species  that  migrated  out  of  the  NTZ  were 
red  drum  (n  =  40,  mean  TL  =  643  mm,  SD  =  135  mm), 
common  snook  (;?  =  9,  mean  TL  =  570  mm,  SD  =  97  mm), 
black  drum  (n  =  8,  mean  TL  =  845  mm,  SD  =  88  mm), 
sheepshead  (n  =  6,  mean  TL  =  398  mm,  SD  =  38  mm), 
bull  shark  (n  =  l,  TL=789  mm),  and  crevalle  jack  (n  =  l, 
TL  =  628  mm).  Recapture  distances  ranged  from  0  km 
immediately  outside  the  NTZ  to  approximately  150  km 
south  in  the  St.  Lucie  River  estuary,  but  recaptured 
fish  were  more  abundant  closer  to  the  NTZ  (Fig.  2A). 
Most  of  the  recaptured  fish  were  concentrated  in  areas 
of  high  fishing  pressure,  such  as  causeways,  inlets,  and 
waters  near  the  boundary  of  the  NTZ.  Collectively,  fish 
that  emigrated  from  the  NTZ  did  not  appear  to  show  a 
bias  for  any  one  direction  of  movement:  recaptured  fish 
were  found  both  northward  in  the  Indian  River  and 
southward  throughout  both  the  Indian  River  and  Ba- 
nana River  basins  of  the  lagoon.  For  individual  species, 
red  drum  that  emigrated  were  distributed  throughout 
the  lagoon  system  and  coastal  habitats,  whereas  black 
drum  were  predominantly  recaptured  in  the  northern 


538 


Fishery  Bulletin  102(3) 


Table  2 

Total  length  ITL) 
and  the  outlying 

size  ranges  (in  mm)  and  le 
ndian  River  Lagoon  study 

gal  size 
area. 

limits  ( 

as  of  August  2000)  foi 

tagged  sportfish  species 

from  the  no-take  zone 

Species 

No-take  zone 

Indian 

River  Lagoon 

Legal  size  limits 

Mean(SD) 

min 

max 

MeanlSDl 

min 

max 

(mm  TL) 

Bull  shark 

789(  —  i 

789 

789 

974(135) 

684 

1180 

None 

Common  snook 

570(106) 

330 

844 

506(138) 

227 

944 

660-8641+  lover) 

Crevalle  jack 

486(140) 

305 

720 

443(113] 

264 

720 

None 

Sheepshead 

398(68) 

235 

614 

365(76) 

171 

594 

305  minimum 

Spotted  seatrout 

415(129) 

185 

754 

335(111) 

212 

678 

381-508  (+  1  over) 

Black  drum 

786(129) 

249 

1156 

742(240) 

225 

1135 

356-610  (  +  1  over) 

Red  drum 

613(166) 

308 

1245 

624(229) 

203 

1210 

457-686 

estuarine  portion  of  the  study  area.  Sheepshead  and 
common  snook  were  recaptured  primarily  to  the  south 
at  inlets  or  in  the  adjacent  Atlantic  coastal  waters  out- 
side the  lagoon. 

The  remaining  75%  («  =  301)  of  the  total  recaptured 
fish  were  from  fish  originally  tagged  outside  the  NTZ 
(Table  1).  The  majority  of  these  (/!=285)  were  also  re- 
covered in  outlying  waters,  including  16  red  drum  and 
1  sheepshead  that  were  subsequently  recaptured  on 
multiple  occasions.  Sixteen  fish  were  recaptured  after 
they  had  immigrated  into  the  reserve.  These  recaptured 
fish  were  from  three  sciaenid  species:  predominantly 
black  drum  in  =  9,  mean  TL  =  907  mm,  SD  =  66  mm)  and 
red  drum  ln  =  6,  mean  TL  =  656  mm,  SD  =  170  mm),  but 
also  one  spotted  seatrout  (TL=420  mmMFig.  2B).  The 
longest  migration  distances  into  the  NTZ  were  up  to  75 
km  for  red  drum  and  spotted  seatrout  tagged  in  south- 
ern Mosquito  Lagoon  and  the  northern  Indian  River 
basins.  All  black  drum  that  immigrated  into  the  NTZ 
were  tagged  in  the  adjacent  Banana  River  basin. 

A  relatively  large  number  of  red  drum,  common  snook, 
and  sheepshead  that  were  tagged  inside  the  NTZ  or  in 
the  outlying  waters  were  recaptured  in  close  proximity 
(0  to  2.75  km  distance)  to  inlet  habitats.  Recaptured 
red  drum  from  inlet  habitats  (n  =  45,  mean  TL=647  mm, 
SD  =  135  mm)  peaked  during  September  through  No- 
vember. Recaptured  common  snook  from  inlet  habitats 
(n=13,  mean  TL  =  598  mm,  SD  =  111  mm)  were  distrib- 
uted throughout  much  of  the  year  but  peaked  in  late 
fall.  Few  common  snook  were  recaptured  from  inlet 
spawning  habitats  during  the  peak  summer  spawning 
months  (June-August)  when  their  fishery  was  closed. 
Recaptured  sheepshead  from  inlet  habitats  («=8,  mean 
TL  =  373  mm,  SD  =  53  mm)  were  concentrated  in  the 
winter  and  early  spring. 

Estimated  migration  rates  were  calculated  by  using 
only  those  fish  that  were  tagged  and  recovered  from 
FMRI  sampling  in  the  NTZ  and  the  immediately  ad- 
jacent upper  Banana  River  (BR).  The  number  of  fish 
tagged  in  the  NTZ  (n=1654)  was  approximately  1.7 
times  the  number  tagged  in  the  BR  (/(=965)  (Table  3); 


however,  the  overall  recapture  rates  of  fish  that  were 
originally  tagged  in  each  of  these  two  areas  were  equal 
(2.4%).  Black  drum  and  red  drum  made  up  the  majority 
of  tagged  and  recaptured  fish  in  both  areas  and  were 
the  only  species  recaptured  that  had  migrated  both  into 
and  away  from  the  NTZ  in  this  comparison.  For  total 
sportfish  (all  species  pooled),  there  was  a  significant 
relationship  between  the  tagging  location  and  the  direc- 
tion offish  movements  (fh  005=13.8,  P=0.0002).  A  total 
of  40  fish  originating  from  the  NTZ  were  recaptured, 
but  that  number  included  only  2  fish  (one  red  drum 
and  one  black  drum)  that  emigrated  to  the  BR  (5% 
overall  migration  rate).  In  contrast,  23  fish  originat- 
ing in  the  BR  were  recaptured  overall,  including  12 
that  immigrated  into  the  NTZ  (52%  overall  migration 
rate).  Species-specific  migration  rates  were  highest  for 
black  drum,  and  relative  immigration  rates  (90%)  were 
higher  than  emigration  rates  (25"7<).  For  this  species, 
the  frequency  of  immigration  and  emigration  were  sta- 
tistically independent  of  tagging  location  (xZi  005=0.01. 
P=0.9039),  which  is  probably  due  to  the  low  number  of 
recaptures  offish  tagged  inside  the  NTZ  (Table  3).  For 
red  drum,  relative  immigration  rates  (27%)  were  also 
higher  than  emigration  rates  (3%),  but  in  this  case, 
there  was  a  significant  relationship  between  fish  move- 
ments and  tagging  location  (^  ,,  (ia=20.58,  P<0.0001). 
Common  snook,  spotted  seatrout.  and  sheepshead  were 
also  recaptured  by  FMRI  scientists  in  these  compari- 
sons, but  none  of  these  recaptured  fish  represented 
evidence  of  migrations  across  the  NTZ  boundary  from 
their  original  tagging  location. 


Discussion 

This  study  demonstrated  both  the  emigration  and  immi- 
gration of  sportfish  species  across  the  boundaries  of  an 
estuarine  no-take  zone  (NTZ).  Legal-size  large  juveniles 
and  adults  of  six  of  the  recreationally  valuable  species 
tagged  within  NTZ  boundaries — red  drum,  black  drum, 
common  snook,  sheepshead,  bull  shark,  and  crevalle 


Tremain  et  al.:  Sportfish  species  movements  in  relation  to  an  estuarine  no-take  zone 


539 


Table  3 

Summary  of  tag  and  recapture 
and  the  adjacent  fished  waters 
included  in  calculations  of  tota 

data  from  only  the  Florida  Marine  Research  Institute  sampl 
of  the  Banana  River  (BR).  Species  that  did  not  contribute 
s  or  of  migration  percentages. 

ing  efforts  in  the  no-take  zone  ( NTZ ) 
any  recapture  information  were  not 

No-take  zone 

Total 

Percent  that 
migrated 

Banana  River 

Total 

Percent  that 
migrated 

No.  of 
fish  tagge 

No.  fish 
recaptured 

No.  of 
fish  tagged 

No.  fish 
recaptured 

i      NTZ 

BR 

NTZ 

BR 

Red  drum 

720 

32 

1 

33 

3.3 

176 

3 

8 

11 

27.3 

Black  drum 

637 

3 

1 

4 

25.0 

495 

9 

1 

10 

90.0 

Common  snook 

104 

1 

0 

1 

0 

62 

0 

1 

1 

0.0 

Spotted  seatrout 

193 

2 

0 

2 

0 

121 

0 

0 

0 

— 

Sheepshead 

597 

0 

0 

0 

— 

232 

0 

1 

1 

0.0 

Totals 

1654 

38 

2 

40 

5.0 

965 

12 

11 

23 

52.2 

Tag  recovery  (% ) 

2.4 

2.4 

jack — were  documented  to  migrate  out  of  the  protected 
area.  Johnson  et  al.  (1999)  and  Stevens  and  Sulak  (2001) 
also  observed  many  of  these  same  species  emigrat- 
ing from  no-take  zones  within  the  same  refuge  system 
during  the  late  1980s,  although  the  species  with  the 
highest  recapture  rates  in  their  studies  (common  snook) 
differed  from  the  current  study  (red  drum).  This  differ- 
ence may  reflect  an  increase  in  the  popularity  of  the  red 
drum  fishery  on  Florida's  east  coast  during  the  current 
study  period.  Since  1989,  when  the  recreational  red 
drum  fishery  reopened  under  strict  management  regula- 
tions, there  has  been  a  significant  increase  in  both  the 
total  red  drum  landings  on  the  Atlantic  coast  and  in  the 
estimated  number  of  fishing  trips  made  by  anglers  seek- 
ing or  catching  red  drum  each  year  (Murphy2).  Tagging 
studies  in  estuarine  areas  of  the  Everglades  National 
Park  have  previously  documented  emigrations  of  striped 
mullet  (Mugil  cephalus),  gray  snapper  (Lutjanus  gri- 
seus),  and  spotted  seatrout  away  from  protected  habitats 
(Bryant  et  al.,  1989;  Funicelli  et  al.,  1989).  Recent  stud- 
ies suggest  that  fish  moving  out  of  protected  areas  in 
the  IRL  may  help  to  replenish  nearby  fisheries  and  may 
contribute  to  trophy  fisheries  in  the  surrounding  system 
(Johnson  et  al.,  1999;  Roberts  et  al.,  2001). 

In  our  study,  overall  emigration  rates  were  low,  but 
many  of  the  fish  that  emigrated  from  the  estuarine  NTZ 
moved  comparatively  large  distances.  The  egress  pat- 
terns of  exploitable  species  may  affect  both  the  species' 
potential  for  protection  and  the  degree  to  which  fisheries 
located  adjacent  to  protected  reserves  will  be  enhanced 
(DeMartini,  1993).  In  coastal  marine  and  tropical  reef 
systems,  where  the  large  majority  of  reserves  have  been 
established,  long-distance  movements  greater  than  a 


2  Murphy,  M.  D.  2002.  A  stock  assessment  of  red  drum. 
Seiaenops  ocellatus,  in  Florida:  status  of  the  stocks  through 
2000,  32  p.  Florida  Fish  and  Wildlife  Conservation  Com- 
mission Report,  Melbourne,  FL. 


few  kilometers  by  demersal  fishery  species  are  limited 
to  a  very  small  percentage  of  individuals  (Beaumar- 
iage,  1969;  PDT,  1990  and  references  therein;  Rowley, 
1994),  and  the  direct  supplementation  of  nearby  fisher- 
ies by  exploitable  species  appears  to  be  highly  localized 
(Buxton  and  Allen,  1989;  Russ  and  Alcala,  1996).  The 
majority  of  fish  that  emigrated  from  the  NTZ  were 
recaptured  between  10  and  75  km  from  the  boundary, 
but  fish  were  also  recovered  as  far  as  150  km  from  the 
NTZ  boundary.  Our  observations  on  migration  distances 
and  recapture  locations  corresponded  well  with  those 
reported  from  previous  studies  of  fish  movements  out 
of  this  same  reserve  system  (Johnson  et  al.,  1999;  Ste- 
vens and  Sulak,  2001).  although  maximum  recapture 
distances  in  earlier  studies  were  even  greater. 

Many  of  the  fish  that  emigrated  from  the  NTZ — such 
as  red  drum,  common  snook,  and  sheepshead — were 
recaptured  at  inlet  locations  or  in  the  nearshore  coastal 
waters  at  sizes  that  were  large  enough  to  include  re- 
productively  mature  adults  (Murphy  and  Taylor,  1990; 
Render  and  Wilson,  1992;  Taylor  et  al.,  2000).  The 
seasonality  of  inlet-associated  recaptures  was  consistent 
with  the  seasonality  of  documented  spawning  and  move- 
ment patterns  for  these  species.  In  Florida,  red  drum 
typically  spawn  in  nearshore  coastal  waters  during 
the  fall  (Murphy  and  Taylor,  1990),  although  spawning 
within  the  IRL  has  also  been  documented  (Johnson 
and  Funicelli,  1991).  Spawning  by  common  snook  may 
occur  year-round  on  Florida's  east  coast  (Gilmore  et  al., 
1983),  but  most  spawning  takes  place  between  May  and 
October  in  or  near  major  inlets  to  the  Atlantic  Ocean 
(Taylor  et  al.,  1998).  The  limited  number  of  common 
snook  recaptured  from  inlet  spawning  habitats  dur- 
ing the  peak  summer  spawning  season  (June-August) 
was  likely  due  to  the  fishery  being  closed  during  those 
months.  Sheepshead  move  offshore  with  the  onset  of 
cool  weather  in  the  late  fall  (Gunter,  1945;  Kelly,  1965), 
and  spawning  likely  occurs  in  offshore  waters  during 


540 


Fishery  Bulletin  102(3) 


the  spring  (Springer  and  Woodburn,  1960;  Jennings, 
1985;  Tucker  and  Barbera,  1987).  In  the  northern  por- 
tion of  the  IRL,  where  the  NTZ  is  located,  the  closest 
access  to  the  coastal  environment  is  through  two  inlets 
located  approximately  75  km  (Sebastian  Inlet)  and  100 
km  (Ponce  de  Leon  Inlet)  swimming  distance  away  or 
through  an  intermittent  lock  opening  at  Port  Canaveral 
approximately  12  km  to  the  south.  In  order  to  reach 
nearshore  or  tidal-pass  spawning  habitats,  species  must 
first  migrate  to  these  locations.  The  coincidence  of  tag 
recoveries  from  these  areas  during  identified  spawning 
or  migration  periods  likely  indicated  that  the  relatively 
long  movement  distances  we  observed  resulted  from  a 
combination  of  geographical,  environmental,  and  bio- 
logical factors,  including  the  proximity  of  the  NTZ  to 
habitats  that  are  important  for  specific  life-history  re- 
quirements of  individual  species.  From  a  management 
viewpoint,  these  relationships  can  affect  the  spatial 
extent  of  species'  migrations  in  relation  to  protected 
habitats,  as  well  as  the  degree  of  protection  provided  to 
individuals  that  are  migratory,  and  should  be  consid- 
ered carefully  in  the  design  of  estuarine  reserves. 

This  study  documented  the  ingress  of  exploitable  es- 
tuarine sportfish  species  into  protected  habitats  and 
demonstrated  that  these  movements  can  also  cover  sub- 
stantial distances.  Species  moving  towards  the  NTZ 
traveled  distances  of  at  least  10-75  km.  The  original 
tagging  locations  of  these  fish  were  distributed  through- 
out the  northern  Indian  and  Banana  rivers  and  southern 
Mosquito  Lagoon,  which  paralleled  the  primary  region 
of  our  tagging  efforts.  Whether  or  not  fish  from  more 
southerly  locations  in  the  IRL  system  would  migrate 
into  the  NTZ  is  largely  unknown  because  of  the  lack 
of  tagging  effort  in  those  areas.  However,  for  tropical 
species  such  as  the  common  snook,  permit  (Trachinotus 
falcatus),  gray  snapper,  and  others  whose  abundances 
increase  seasonally  in  the  northern  lagoon  habitats  dur- 
ing the  warmer  months  (Tremain  and  Adams,  1995),  it 
seems  probable  that  seasonal  movements  could  bring 
them  into  contact  with  the  protected  habitats.  In  such 
cases,  these  species  would  benefit  only  temporarily  from 
fishing  protection  until  their  return  migrations  made 
them  again  vulnerable  to  capture.  In  contrast,  species 
observed  migrating  into  the  NTZ  that  typically  have  a 
high  degree  of  site  fidelity  during  specific  life-history 
stages,  such  as  the  red  drum  (Beaumariage,  1969;  Ad- 
ams and  Tremain.  2000),  black  drum  (Murphy  et  al., 
1998),  and  spotted  seatrout  (Moffett,  1961),  should  de- 
rive greater  long-term  benefits  from  reserve  protection 
following  immigration  into  protected  areas. 

Tagging  studies  that  examine  the  transfer  of  fishery 
species  between  reserve  and  outlying  habitats  are  rare, 
and  we  have  found  only  one  recent  study  on  any  fishery 
species,  the  American  lobster  (Homarus  americanus\, 
that  investigated  the  effects  that  multidirectional  spe- 
cies migrations  may  have  upon  protective  reserve  func- 
tions (Rowe,  2001).  Studies  in  which  fish  movements 
have  been  examined,  in  both  estuarine  and  marine 
protected  areas,  have  focused  exclusively  on  fish  egress 
from  reserve  habitats  (Bryant  et  al.,  1989;  Buxton  and 


Allen,  1989;  Funicelli  et  al.,  1989;  Holland  et  al.,  1996; 
Zeller  and  Russ,  1998;  Johnson  et  al.,  1999.  Stevens 
and  Sulak,  2001)  or  on  home  ranges  of  species  associ- 
ated with  reserve  habitats  (Eristhee  and  Oxenford. 
2001;  Starr  et  al.,  2002).  In  the  present  study,  we  simul- 
taneously examined  both  egress  and  ingress  of  sportfish 
in  relation  to  a  no-take  reserve  and  the  surrounding 
unprotected  waters,  and  the  results  provide  a  starting 
point  to  quantitatively  discuss  the  relationship  between 
fish  emigration  and  immigration,  as  well  as  the  implica- 
tions of  such  movements  to  the  resulting  functions  of 
replenishment  to  or  withdrawal  from  nearby  estuarine 
fisheries.  When  all  recapture  sources  were  considered, 
the  ratio  of  migrating  to  nonmigrating  individuals  was 
much  higher  for  fish  tagged  inside  the  NTZ  (1.58)  than 
for  those  tagged  outside  the  NTZ  (0.05);  this  ratio  im- 
plies that  there  is  a  spillover  effect  from  the  reserve. 
However,  this  difference  is  less  apparent  when  measured 
against  the  large  disparity  between  recapture  effort 
from  inside  the  NTZ  (12-24  FMRI  sampling  days/year 
+  12-24  angler  days/year)  and  recapture  effort  from  the 
surrounding  lagoon  waters  of  Brevard  County  (50-100 
FMRI  sampling  days/year  +  114,000-181,000  angler 
days/year  [FMRI.  unpubl.  data]).  Furthermore,  this 
direct  comparison  assumes  that  recapture  potential  was 
the  same  in  protected  and  unprotected  areas,  which  is 
unlikely  given  the  differences  between  the  primary  re- 
capture gear  used  in  scientific  research  activities  inside 
the  reserve  (nets)  and  the  gear  used  in  recreational  an- 
gling outside  the  reserve  (hook  and  line).  There  were  no 
reliable  estimates  of  sportfish  species  landings  available 
for  the  limited  study  region  that  could  have  enabled 
us  to  intercalibrate  for  these  differences;  therefore. 
we  limited  further  comparisons  to  only  data  recovered 
through  FMRI  sampling  activities  in  the  northern  Ba- 
nana River  basin.  This  limitation  came  at  the  expense 
of  important  tag-recovery  data  collected  by  anglers  or 
collected  from  more  outlying  areas  of  the  lagoon  but 
permitted  a  more  quantitative  comparison  of  migra- 
tion potential  that  focused  comparisons  on  immediately 
adjacent  areas  where  the  effects  of  spillover  would  most 
likely  be  realized  (Buxton  and  Allen,  1989;  Russ  and 
Alcala,  1996).  In  these  comparisons,  a  disproportionate 
number  of  fish  were  tagged  inside  the  NTZ,  but  overall 
tag-recovery  rates  for  fish  originating  in  both  the  NTZ 
and  the  adjacent  Banana  River  were  equivalent.  This 
finding  indicated  that  tagged  individuals  from  both 
areas  were  equally  susceptible  to  recapture.  However, 
there  were  substantial  differences  in  the  migration 
patterns  of  fish  between  the  two  areas.  In  the  vicinity 
of  the  NTZ,  the  relative  potential  for  overall  sportfish 
migrations  (primarily  for  red  drum  and  black  drum, 
which  provided  the  greatest  quantity  of  tag  recovery 
data)  towards  the  NTZ  from  unprotected  habitats  (52%) 
was  greater  than  the  potential  for  migrations  out  of  the 
NTZ  (5%). 

Two  potential  limitations  must  be  considered  when 
comparing  these  migration  rates.  First,  it  is  possible 
that  recreational  fishing  in  the  upper  Banana  River 
could  have  reduced  the  number  of  tags  available  to  FM- 


Tremain  et  al.:  Sportfish  species  movements  in  relation  to  an  estuarine  no-take  zone 


541 


RI  sampling  activities  outside  the  NTZ,  leading  to  lower 
tag  recovery  rates  from  this  area.  However,  several  fish 
from  the  Banana  River  study  area  were  recaptured  on 
multiple  occasions — a  common  occurrence  in  this  region 
where  fish  are  caught  and  released  in  fishing  practices. 
Although  there  is  some  postrelease  cryptic  mortality 
associated  with  catch-and-release  practices,  these  re- 
leases likely  limited  the  effects  of  local  fishing  on  our 
analyses.  Second,  our  assumption  that  the  migration 
patterns  of  recaptured  fish  represented  the  migration 
patterns  of  the  overall  population  may  not  be  valid  if 
the  respective  length  frequencies  were  not  also  equally 
represented.  The  use  of  multiple  gear  types  and  sam- 
pling strategies  to  collect  fish  for  tagging  increased  the 
likelihood  that  the  length  frequencies  of  species  in  our 
collections  represented  the  available  population.  Report- 
ed recapture  length  frequencies  closely  approximated 
the  population  length  frequencies  in  our  collections  for 
red  drum,  black  drum,  and  sheepshead  but  over-repre- 
sented the  frequency  of  larger  individuals  for  common 
snook  and  spotted  seatrout.  Because  red  drum  and 
black  drum  were  the  principal  species  that  displayed 
multidirectional  migration  patterns,  we  considered  the 
potential  for  size  bias  to  be  minimal  in  our  comparisons 
of  estimated  ingress  and  egress  rates. 

Ultimately,  a  determination  of  the  net  result  of  these 
migration  patterns,  in  terms  of  replenishment  to  or 
withdrawal  from  adjacent  fisheries,  would  require  ac- 
curate assessments  of  species  population  abundances 
that  were  beyond  the  scope  of  this  study.  If  there  are 
large  enough  differences  in  population  densities  across 
the  NTZ  boundary,  either  as  a  result  of  increased  pro- 
duction inside  the  reserve  or  high  fishing  mortality 
outside,  then  the  relatively  low  emigration  rates  that  we 
observed  could  still  result  in  a  net  export  of  exploitable 
individuals  to  fished  populations  in  surrounding  waters. 
In  trammel-net  collections  from  this  same  reserve  dur- 
ing the  late  1980's,  Johnson  et  al.  (1999)  estimated 
that  in  the  protected  habitats,  relative  abundances  of 
red  drum  populations  were  6.3  times  greater  and  of 
black  drum  were  12.8  times  greater  than  the  relative 
abundances  of  these  populations  in  adjacent  unpro- 
tected areas.  More  recent  shoreline  haul-seine  data 
from  1997-2000  show  that  these  abundances  were  only 
1.8  times  greater  for  red  drum  and  1.5  times  greater 
for  black  drum  (FMRI,  unpubl.  data).  To  what  extent 
the  difference  in  abundance  estimates  between  these 
two  temporally  separate  studies  is  related  to  fish  move- 
ments, to  stringent  changes  in  management  regulations 
that  have  occurred,  or  to  the  difference  in  sampling 
methods  used  is  undetermined.  However,  if  we  consider 
the  more  recent  population  level  differences  between  the 
NTZ  and  adjacent  waters,  then  the  emigration  and  im- 
migration rates  observed  in  the  present  study  indicate 
that  there  is  a  potential  for  more  substantial  move- 
ments by  these  species  towards  protected  habitats  than 
away  from  them. 

One  limitation  of  tag-recapture  data  is  that  such  data 
provide  only  a  snapshot  view  of  overall  fish  movements, 
and  the  whereabouts  of  tagged  individuals  between 


the  time  of  tagging  and  recapture  are  unknown.  It 
is  possible  that  the  movements  we  observed  for  red 
drum  and  black  drum  in  the  vicinity  of  the  NTZ  were 
simply  instantaneous  views  of  a  more  complex  series  of 
movements  between  the  NTZ  and  adjacent  waters.  One 
possibility  is  that  these  movements  could  be  related 
to  daily  or  seasonal  home  ranges  that  extend  across 
reserve  boundaries.  Studies  that  attempt  to  quantify 
home  ranges  for  these  species  at  any  temporal  scale 
are  limited.  Carr  and  Chaney  1 1976 1  followed  a  single 
red  drum,  which  was  fitted  with  an  ultrasonic  trans- 
mitter, for  up  to  two  days  after  releasing  it  into  the 
Intracoastal  Waterway  near  St.  Augustine,  Florida. 
During  that  time,  fish  movements  were  oriented  against 
the  direction  of  tidal  flow  but  remained  within  2  km 
of  the  release  point.  Adams  and  Tremain  (2000)  found 
that  large  juvenile  red  drum  repeatedly  used  or  were 
continually  associated  with  a  2-km  section  of  a  northern 
IRL  tidal  creek  for  periods  of  up  to  18  months.  Tag- 
ging studies  from  estuarine  waters  generally  indicate 
that  the  majority  of  red  drum  and  black  drum  do  not 
make  substantial  movements  from  their  release  sites, 
although  some  individuals  are  capable  of  migrating  up 
to  several  hundred  kilometers  (Beaumariage,  1969: 
Osburn  et  al.  1982;  Music  and  Pafford,  1984;  Murphy  et 
al.,  1998i.  During  the  present  study,  20  red  drum  were 
recaptured  on  multiple  occasions;  however,  none  of  these 
fish  exhibited  movements  that  could  provide  evidence 
for  home  ranges  that  overlapped  the  NTZ  boundar- 
ies. Another  possibility  for  the  movement  patterns  we 
observed  is  that  they  are  related  to  population  equilib- 
rium adjustments  that  occur  when  the  relative  attri- 
butes of  the  NTZ  and  surrounding  areas  change  with 
respect  to  each  other.  For  example,  beginning  in  1990 
and  coinciding  with  the  onset  of  the  present  study,  the 
Banana  River  adjacent  to  the  NTZ  (including  much  of 
our  BR  study  area)  was  closed  to  motorized  boat  traffic. 
Although  the  area  remained  open  to  fishing,  it  became 
considerably  more  difficult  to  access  by  fishermen.  If 
this  limitation  resulted  in  lower  fishing  pressure  (i.e., 
predation)  and  fewer  habitat  disturbances,  then  the 
relative  habitat  value  and  rates  of  migration  into  this 
area  may  have  increased  during  that  time.  There  are 
no  quantifiable  estimates  of  migration  rates  prior  to 
this  study  for  comparison,  but  our  results  do  not  dem- 
onstrate an  equilibrium  adjustment  toward  potentially 
higher  quality  BR  habitats  during  our  study  period.  If 
species  movements  are  not  equilibrium  adjustments, 
but  rather  are  driven  by  an  attraction  to  or  retention 
within  habitats  that  offer  protective  benefits,  then  ul- 
timately reserve  habitats  should  become  saturated. 
Predicted  equilibrium  population  sizes  for  queen  conch 
iStrombus  gigas)  and  spiny  lobster  (Panulirus  argus) 
were  achieved  in  just  three  years  after  the  effective 
creation  of  a  Caribbean  reef  harvest  refuge,  but  models 
suggested  that  relatively  minor  changes  in  refuge  area 
and  boundary  condition  (i.e.,  permeability)  could  result 
in  major  population-level  responses  by  exploited  species, 
depending  upon  dispersal  dynamics  and  habitat  avail- 
ability (Acosta,  2002).  The  estuarine  no-take  zone  at 


542 


Fishery  Bulletin  102(3) 


MINWR  has  been  in  effect  for  approximately  40  years, 
presumably  long  enough  for  fish  populations  to  reach 
equilibrium  levels,  yet  we  observed  a  net  movement  of 
fish  into  protected  habitats  over  the  past  decade. 

A  wide  range  of  factors  interact  to  determine  the 
distributions  of  large  mobile  fish  in  the  IRL,  where 
physical  environmental  conditions  (salinity,  inlet  dis- 
tance, temperature,  etc.)  have  a  primary  influence  on 
the  species'  distributions  over  a  lagoon-wide  scale,  and 
where  species  responses  to  biological  variables  (sea- 
grass  cover,  depth,  seasonality,  etc.)  act  secondarily  to 
influence  distributions  at  smaller  scales  (Kupschus  and 
Tremain,  2001).  The  specific  mechanisms  that  lead  to 
the  greater  ingress  rates  into  the  NTZ  for  red  drum  and 
black  drum  in  the  present  study  cannot  be  determined 
from  our  data.  Possibilities  include  a  behavioral  attrac- 
tion to  the  NTZ  due  to  the  interrelated  influences  of 
habitat  preference,  spawning,  and  social  structure,  or 
due  to  potentially  higher  retention  rates  after  migra- 
tion into  the  reserve.  Red  drum  and  black  drum  were 
routinely  observed  foraging  in  large  schools  within  both 
the  NTZ  and  surrounding  waters,  which  suggested  that 
food  resources  were  available  in  each  of  these  habitats; 
however,  there  are  few  studies  that  have  attempted  to 
quantify  differences  in  resource  availability  between 
these  areas.  Johnson  et  al.  (1999)  described  the  habitat 
characteristics  of  their  study  areas  within  the  same  re- 
serve system  but  found  that  protection  from  fishing,  and 
not  habitat  difference,  was  the  primary  factor  contribut- 
ing to  differences  in  the  abundance  of  sportfish  species 
between  fished  and  unfished  areas.  The  availability  of 
suitable  spawning  habitats  within  the  NTZ  may  also 
attract  red  drum  and  black  drum  to  the  reserve  habi- 
tats. We  observed  indications  of  reproductive  behavior 
by  both  of  these  species  inside  the  NTZ  that  is  common 
among  members  of  the  drum  family,  including  concen- 
trations of  drumming  fish  (Mok  and  Gilmore,  1983) 
and  repeated  side-to-side  contact  among  individual  fish 
(Tabb,  1966)  in  the  presence  of  ripe  and  running  males. 
Although  we  did  not  directly  observe  these  behaviors 
for  either  species  outside  of  the  NTZ,  black  drum  and 
red  drum  are  documented  to  spawn  elsewhere  within 
the  IRL  system  (Mok  and  Gilmore,  1983;  Johnson  and 
Funicelli.  1991)  and  we  cannot  automatically  presume 
that  suitable  spawning  habitats  do  not  also  occur  in  the 
surrounding  waters.  If  there  is  a  behavioral  attraction 
to  protected  habitats,  then  the  subsequent  retention  of 
individuals  that  have  immigrated  into  these  areas  may 
be  prolonged  by  the  limited  boundary  permeability  of 
this  reserve,  which  contains  only  two  potential  egress 
pathways  back  into  the  adjacent  waters.  In  order  to  ful- 
ly understand  the  protective  functions  of  this  estuarine 
reserve  and  others,  it  will  be  important  to  identify  the 
biological,  behavioral,  and  physical  mechanisms  that 
influence  species  movements  in  relation  to  the  reserve 
boundaries. 

The  opportunistic  nature  of  our  tagging  efforts  within 
the  design  of  a  larger  sampling  program  precluded  sta- 
tistically valid  sample  replication,  and  only  one  reserve 
and  adjacent  fished  area  were  examined;  therefore, 


the  results  of  this  study  should  not  be  generalized  to 
other  areas.  Still,  the  IRL  is  typical  of  other  bar-built 
estuaries  where  access  by  estuarine  fishes  to  coastal 
waters  through  passes  or  inlets  may  be  limited,  and 
it  is  reasonable  to  expect  that  the  geographical,  en- 
vironmental, and  biological  processes  that  influence 
species  movements  in  the  IRL  would  also  be  important 
in  other  estuaries  of  similar  structure.  Studies  show- 
that  no-take  areas  in  estuarine  systems  can  have  an 
effect  on  species'  abundances  and  size  distributions 
within  these  protected  areas  and  may  indicate  that 
these  areas  protect  species  from  the  effects  of  fishing 
pressure  (Johnson  et  al.,  1999;  FMRI  unpubl.  datai. 
Whether  or  not  these  areas  will  actually  increase  fish 
abundance  in  adjacent  waters  or  benefit  surrounding 
fisheries  through  direct  supplemental  replenishment  of 
exploitable  species  is  less  evident.  Certainly,  some  indi- 
viduals will  migrate  out  of  protected  areas  in  response 
to  environmental,  biological,  or  physiological  stimuli, 
and  these  individuals  may  contribute  to  trophy  fisheries 
in  surrounding  waters  (Roberts  et  al..  2001);  however, 
our  data  indicated  that  within  estuaries,  reciprocal 
movements  over  relatively  large  distances  into  protected 
areas  also  occur  and  have  the  potential  to  extract  ex- 
ploitable individuals  from  surrounding  fisheries.  The 
overall  impact  of  such  withdrawals  on  these  fisheries 
will  depend  on  the  degree  of  retention  following  migra- 
tions into  protected  areas.  If  retention  rates  are  high, 
then  increased  egg  production,  larval  export,  and  juve- 
nile recruitment  may  be  more  important  mechanisms 
for  replenishment  of  nearby  fisheries  than  spillover  of 
exploitable  species,  but  production  and  export  will  be 
limited  unless  reserves  encompass  spawning  or  nursery- 
habitats  (or  both)  that  will  support  long-term  protection 
and  population  growth.  For  estuarine-dependent  coastal 
species  that  support  estuarine  fisheries,  the  benefits 
obtained  within  protected  areas  will  be  determined, 
in  part,  by  their  specific  life-history  characteristics, 
movement  patterns,  and  the  reserve  design.  Although 
the  establishment  and  study  of  reserves  in  marine  or 
coastal  systems  has  increased  in  recent  years,  research 
on  the  effects  of  protected  no-take  reserves  in  estuarine 
habitats  is  still  in  its  infancy.  Information  on  the  daily, 
seasonal,  or  annual  movement  patterns  of  estuarine- 
resident  or  estuarine-dependent  coastal  species  is  neces- 
sary for  understanding  and  designing  effective  reserve 
areas  in  these  habitats. 


Acknowledgments 

We  wish  to  thank  the  crewmembers  and  volunteers  at 
FMRI's  Indian  River  Field  Laboratory  for  collecting  data 
and  assisting  in  this  study  and  the  many  fishermen  who 
willingly  provided  us  with  recapture  information.  We 
are  grateful  to  U.S.  Fish  and  Wildlife  Service  personnel 
for  providing  access  to  sampling  areas  within  restricted 
areas  of  the  Merritt  Island  National  Wildlife  Refuge. 
This  paper  benefitted  from  reviews  by  R.  Cody.  J.  Col- 
vocoresses,  L.  French,  J.  Leiby,  R.  Paperno,  J.  Quinn, 


Tremain  et  al.:  Sportfish  species  movements  in  relation  to  an  estuarine  no-take  zone 


543 


T.  Tuckey,  and  F.  Vose,  and  two  anonymous  reviewers. 
This  work  was  supported  in  part  by  funding  from  the 
Department  of  Interior,  U.  S.  Fish  and  Wildlife  Service, 
Federal  Aid  for  Sport  Fish  Restoration  Grant  Number 
F-43,  and  by  the  State  of  Florida  Recreational  Saltwater 
Fishing  License  monies. 


Literature  cited 

Acosta,  C.  A. 

2002.     Spatially  explicit  dispersal  dynamics  and  equilib- 
rium population  sizes  in  marine  harvest  refuges.     ICES 
J.  Mar.  Sci.  59:458-468. 
Adams.  D.  H„  and  D.  M.  Tremain. 

2000.  Association  of  large  juvenile  red  drum,  Sciaenops 
ocellatus,  with  an  estuarine  creek  on  the  Atlantic  coast 
of  Florida.     Environ.  Biol.  Fish.  58:183-194. 

Babcock,  R.  C,  S.  Kelly,  N.  T.  Shears,  J.  W.  Walker,  and 
T.  J.  Willis. 

1999.     Changes  in  community  structure  in  temperate 
marine  reserves.     Mar.  Ecol.  Prog.  Ser.  189:125-134. 
Beaumariage.  D.  S. 

1969.     Returns  of  the  1965  Schlitz  tagging  program  includ- 
ing a  cumulative  analysis  of  previous  results.     Fla. 
Board  Conserv.  Mar.  Res.  Lab.  Tech.  Ser.  59.  38  p. 
Bohnsack,  J.  A.  and  J.  S.  Ault. 

1996.     Management  strategies  to  conserve  marine  bio- 
diversity.    Oceanography  (91:73-82. 
Bryant,  H.  E.,  M.  R.  Dewey,  N.  A.  Funicelli,  G.  M.  Ludwig, 
D.  A.  Meineke,  and  L.  J.  Mengal. 

1989.     Movement  of  five  selected  sports  species  offish  in 
Everglades  National  Park.     Bull.  Mar.  Sci.  44:515. 
Buxton,  C.  D.,  and  J.  A.  Allen. 

1989.     Mark  and  recapture  studies  of  two  reef  sparids 
in  the  Tsitsikamma  Coastal  National  Park.     Koedoe 
32:39-45. 
Carr,  W.  E.  S.,  and  T.  Chaney. 

1976.     Harness  for  attachment  of  an  ultrasonic  transmit- 
ter to  the  red  drum,  Sciaenops  ocellatus.     Fish.  Bull. 
74:998-1000. 
Chapman,  M.  R.,  and  D.  L.  Kramer. 

1999.     Gradients  in  coral  reef  fish  density  and  size  across 
the  Barbados  Marine  Reserve  boundary:  effects  of 
reserve  protection  and  habitat  characteristics.     Mar. 
Ecol.  Prog.  Ser.  181:81-96. 
Cole,  R.  G.,  and  D.  Keuskamp. 

1998.     Indirect  effects  of  predation  from  exploitation: 
patterns  from  populations  of  Evechinus  chloroticus  (Echi- 
noideal  in  northwestern  New  Zealand.     Mar.  Ecol.  Prog. 
Ser.  173:215-226. 
DeMartini,  E.  E. 

1993.     Modeling  the  potential  of  fishery  reserves  for 
managing  Pacific  coral  reef  fishes.     Fish.  Bull.  91: 
414-427. 
Eristhee.  N..  and  H.  A.  Oxenford. 

2001.  Home  range  size  and  use  of  space  by  Bermuda 
chub  Kyphosis  seetatrix  (L.)  in  two  marine  reserves 
in  the  Soufriere  Marine  Management  Area,  St.  Lucia, 
West  Indies.     J.  Fish  Biol.  59:129-151. 

Funicelli,  N.  A.,  D.  A.  Meineke,  H.  E.  Bryant.  M.  R.  Dewey, 
G.  M.  Ludwig,  and  L.  S.  Mengel. 

1989.  Movements  of  striped  mullet,  Mugil  cephalus, 
tagged  in  Everglades  National  Park,  Florida.  Bull. 
Mar.  Sci.  44:171-178. 


Gilmore,  R.  G..  Jr..  C.  J.  Donohoe,  and  D.  W.  Cooke. 

1983.     Observations  on  the  distribution  and  biology  of  the 
common  snook,  Centropomus  undecimalis  i  Block).     Fla. 
Sci.  46:313-336. 
Gilmore,  R.  G,  Jr..  C.  J.  Donohoe.  D.  W.  Cooke,  and 
D.  J.  Herrema. 

1981.     Fishes  of  the  Indian  River  Lagoon  and  adjacent 
waters.     Harbor  Branch  Foundation  Tech.  Rep.  No.  41, 
64  p. 
Gunter,  G. 

1945.     Studies  on  marine  fishes  of  Texas.     Inst.  Mar. 
Sci.  Univ.  Tex.  1:1-190. 
Holland.  K.  N,  C.  G.  Lowe,  and  B.  M.  Wetherbee. 

1996.     Movements  and  dispersal  patterns  of  blue  trev- 
ally  (Caranx  melamgypus)  in  a  fisheries  conservation 
zone.     Fish.  Res.  25:279-292. 
Jennings,  C.  A.. 

1985.  Species  profiles:  life  histories  and  environmental 
requirements  of  coastal  fishes  and  invertebrates  (Gulf  of 
Mexico) — sheepshead.  U.S.  Fish  and  Wildlife  Service 
Biological  Report  82(11.29)  and  U.  S.  Army  Corps  of 
Engineers,  TR  EL-82-4.  10  p. 
Johnson.  D.  R..  and  N.  A.  Funicelli. 

1991.     Spawning  of  the  red  drum  in  Mosquito  Lagoon, 
east-central  Florida.     Estuaries  14:74-79. 
Johnson,  D.  R..  N.  A.  Funicelli,  and  J.  A.  Bohnsack. 

1999.     Effectiveness  of  an  existing  no-take  fish  sanctuary 
within  the  Kennedy  Space  Center,  Florida.     N.  Am.  J. 
Fish.  Manag.  19:436-453. 
Kelly,  JR. 

1965       A  taxonomic  survey  of  the  fishes  of  the  Delta 
National  Wildlife  Refuge  with  emphasis  upon  distri- 
bution and  abundance.     M.S.  thesis.  133  p.     Louisiana 
State  Univ.,  Baton  Rouge,  LA. 
Kupschus,  S..  and  D.  Tremain. 

2001.     Associations  between  fish  assemblages  and  envi- 
ronmental factors  in  nearshore  habitats  of  a  subtropical 
estuary.     J.  Fish  Biol..  58:1383-1403. 
Moffett,  A.  W. 

1961.     Movement  and  growth  of  spotted  seatrout.  Cynoscion 
nebulosus  (Cuvier),  in  west  Florida..  35  p.     Fla.  Board 
Conserv.  Mar.  Res.  Lab.  Tech.  Ser.  No.  36. 
Mok,  H.,  and  R.  G.  Gilmore. 

1983.     Analysis  of  sound  production  in  estuarine  aggre- 
gations of  Pogonias  cromis,  Bairdiella  chrysoura,  and 
Cynoscion  nebulosus  (Sciaenidae  I.     Bull.  Inst.  Zool, 
Academia  Sinica  (Taipei).  22:157-186. 
Mulligan,  T.  J.,  and  F.  F.  Snelson  Jr. 

1983.  Summer-season  populations  of  epibenthie  marine 
fishes  in  the  Indian  River  Lagoon  system,  Florida. 
Fla.  Sci.  46:252-275. 

Murphy,  M.  D.,  D.  H.  Adams.  D.  M.  Tremain.  and  B.  W.  Winner. 
1998.     Direct  validation  of  ages  determined  for  adult 

black  drum,  Pogonias  cromis,  in  east  central  Florida, 

with  notes  on  black  drum   migration.     Fish.   Bull. 

96:382-387. 
Murphy,  M.  D.,  and  R.  G.  Taylor. 

1990.     Reproduction,  growth,  and  mortality  of  red  drum 

Sciaenops  ocellatus  in  Florida  waters.     Fish.  Bull. 

88:531-542. 
Music,  J.  L.,  and  J.  M.  Pafford. 

1984.  Population  dynamics  and  life  history  aspects  of 
major  marine  sportfishes  in  Georgia's  coastal  waters.. 
382  p.  Georgia  Department  of  Natural  Resources, 
Contribution  Series  38. 


544 


Fishery  Bulletin  102(3) 


Musick,  J.  A.,  S.  A.  Berkeley,  G.  M.  Cailliet,  M.  Camhi, 
G.  Huntsman,  M.  Nammack,  and  M.  L.  Warren  Jr. 

2000.  Protection  of  fish  stocks  at  risk  of  extinction.  Fish- 
eries 25(31:6-8. 

Osburn,  H.  R.,  G.  C.  Matlock,  and  A.  W.  Green. 

1982.     Red  drum  tSciaenops  ocellatus)  movement  in  Texas 
bays.     Contr.  Mar.  Sci..  25:85-97. 
PDT  (Plan  Development  Team). 

1990.  The  potential  of  marine  fishery  reserves  for  reef 
fish  management  in  the  U.S.  Southern  Atlantic.  NOAA 
Tech.  Memo.  NMFS-SEFC-261,  40  p. 

Render,  J.  H.,  and  C.  A.  Wilson. 

1992.     Reproductive  biology  of  sheepshead  in  the  northern 
Gulf  of  Mexico.     Trans.  Am.  Fish.  Soc.  121:757-764. 
Roberts,  C.  M„  J.  A.  Bohnsack,  F.  Gell,  J.  P.  Hawkins, 
and  R.  Goodridge. 

2001.  Effects  of  marine  reserves  on  adjacent  fisheries. 
Science  294:1920-1923. 

Roberts,  C.  M„  and  N.  V.  C.  Polunin. 

1991.  Are  marine  reserves  effective  in  management  of 
reef  fisheries?     Rev.  Fish  Biol,  and  Fish.  1:65-91. 

Rowe,  S. 

2001.     Movement  and  harvesting  mortality  of  American 
lobsters  (Homarus  americanus)  tagged  inside  and  outside 
no-take  reserves  in  Bonavista  Bay,  Newfoundland.     Can. 
J.  Fish.  Aquat.  Sci.  58:1336-1346. 
Rowley,  R.  J. 

1994.     Marine  reserves  in  fisheries  management.     Aquat. 
Conserv.  Mar.  Freshwater  Ecosyst.  4:233-254. 
Russ,  G.  R. 

1985.     Effects  of  protective  management  on  coral  reef 
fishes  in  the  central  Philippines.     Proc.  of  the  5th  Inter- 
national Coral  Reef  Congress  4:219-224. 
Russ,  G.  R.,  and  A.  C.  Alcala. 

1996.     Do  marine  reserves  export  adult  fish  biomass? 
Evidence  from  Apo  Island,  central  Philippines.     Mar. 
Ecol.  Prog.  Ser.  132:1-9. 
Sala,  E.,  and  M.  Zabala. 

1996.     Fish  predation   and   the   structure  of  the  sea 


urchin  Paracentrotus  lividus  populations  in  the  NW 
Mediterranean.     Mar.  Ecol.  Prog.  Ser.  140:71-81. 
Springer,  V.  G.,  and  K.  D.  Woodburn. 

1960.     An  ecological  study  of  the  fishes  of  the  Tampa 
Bay  area.     Fla.  Board  Conserv.  Mar.  Res.  Lab.  Prof. 
Pap.  Ser.  1.  104  p. 
Starr,  R.  M..  J.  N.  Heine,  J.  M.  Felton,  and  G.  M.  Cailliet. 

2002.     Movements  of  bocaccio  iSebastes  paucispinis)  and 
greenspotted  (S.  chlorostictus)  rockfishes  in  a  Monterey- 
submarine  canyon:  implications  for  the  design  of  marine 
reserves.     Fish.  Bull.  100:324-337. 
Stevens,  P.  W.,  and  K.  J.  Sulak. 

2001.     Egress  of  adult  sport  fish  from  an  estuarine 
reserve  within  Merritt  Island  National  Wildlife  Refuge, 
Florida.     Gulf  Mex.  Sci.  2:77-89. 
Tabb,  D.  C. 

1966.     The  estuary  as  a  habitat  for  spotted  seatrout  Cyno- 
scion  nebulosus.     Am.  Fish.  Soc.  Special  Publication 
3:59-67. 
Taylor,  R.  G.,  H.  J.  Grier,  and  J.  A.  Whittington. 

1998.     Spawning  rhythms  of  common  snook  in  Florida.     J. 
Fish  Biol.  53:502-520. 
Taylor,  R.  G„  J.  A.  Whittington.  H.  J.  Grier.  and 
R.  E.  Crabtree. 

2000.     Age,  growth,  maturation,  and  protandric  sex 
reversal  in  common  snook,  Centropomus  undecimalis, 
from  the  east  and  west  coasts  of  South  Florida.     Fish. 
Bull.  98:612-624. 
Tremain,  D.  M.,  and  D.  H.  Adams. 

1995.     Seasonal  variations  in  species  diversity,  abun- 
dance, and  composition  of  fish  communities  in  the 
northern  Indian  River  Lagoon,  Florida.     Bull.  Mar. 
Sci.  57:171-192. 
Tucker,  J.  W.,  Jr..  and  P.  A.  Barbera. 

1987.     Laboratory  spawning  of  sheepshead.     Prog.  Fish- 
Cult.  49:229-230. 
Zeller,  D.  C,  and  G.  R.  Russ. 

1998.  Marine  reserves:  patterns  of  adult  movement  of  the 
coral  trout  (Plectropomus  leopardus)  (Serranidae).  Can. 
J.  Fish.  Aquat.  Sci.  55:917-924. 


545 


Abstract— Patterns  of  distribution 
and  growth  were  examined  for  young- 
of-the-year  (YOY)  greater  amberjack 
{Seriola  dumerili)  associated  with 
pelagic  Sargassum  in  the  NW  Gulf 
of  Mexico.  Seriola  dumerili  were  col- 
lected off  Galveston,  Texas,  from  May 
to  July  over  a  two-year  period  (2000 
and  20011  in  both  inshore  (<15  nauti- 
cal miles  [nmijl  and  offshore  zones 
(15-70  nmi).  Relative  abundance  of 
YOY  S.  dumerili  (32-210  mm  stan- 
dard length)  from  purse-seine  col- 
lections peaked  in  May  and  June, 
and  abundance  was  highest  in  the 
offshore  zone.  Ages  of  S.  dumerili 
ranged  from  39  to  150  days  and  hatch- 
ing-date analysis  indicated  that  the 
majority  of  spawning  events  occurred 
from  February  to  April.  Average  daily 
growth  rates  of  YOY  S.  dumerili  for 
2000  and  2001  were  1.65  mm/d  and 
2.00  mm/d.  respectively.  Intra-annual 
differences  in  growth  were  observed; 
the  late-season  I  April  I  cohort  expe- 
rienced the  fastest  growth  in  both 
years.  In  addition,  growth  was  signifi- 
cantly higher  for  S.  dumerili  collected 
from  the  offshore  zone.  Mortality  was 
approximated  by  using  catch-curve 
analysis,  and  the  predicted  instan- 
taneous mortality  rate  (Z)  of  YOY  S. 
dumerili  was  0.0045  (0.45%/d). 


Distribution,  age,  and  growth  of  young-of-the-year 
greater  amberjack  {Seriola  dumerili) 
associated  with  pelagic  Sargassum 


R.  J.  David  Wells 
Jay  R.  Rooker 

Texas  A&M  University 

Department  ol  Marine  Biology 

5007  Avenue  U 

Galveston,  Texas  77551 

Present  address  (lor  R.  J,  D  Wells):  Coastal  Fisheries  Institute 

Louisiana  State  University 
Baton  Rouge,  Louisiana  70803 

E-mail  address  (for  R  J  D  Wells)  rwells4@lsu.edu 


Manuscript  submitted  9  December  2002 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
2  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:545-554  (2004). 


Recruitment  of  marine  fishes  is  highly 
variable  and  closely  linked  to  early 
life  events  (Houde,  1996;  Cole,  1999). 
Early  life  survival  is  dependent  upon 
several  biological  and  environmen- 
tal factors  including  spawning  time, 
prey  availability,  predation  pres- 
sure, growth,  and  physical  transport 
mechanisms  (Bricelj,  1993;  Schnack 
et  al.,  1998).  Recruitment  success 
is  commonly  assessed  by  examining 
patterns  of  relative  abundance  (Sano, 
1997),  whereas  estimates  of  growth 
and  mortality  are  commonly  used  to 
index  recruitment  potential  (Rilling 
and  Houde,  1999;  Rooker  et  al.,  1999). 
Early  life  growth  and  mortality  are 
linked  because  fishes  with  high  growth 
rates  often  exhibit  decreased  size-spe- 
cific predator  vulnerability  (Meekan 
and  Fortier,  1996).  As  a  result,  esti- 
mates of  juvenile  abundance,  growth, 
and  mortality  provide  insight  into 
patterns  of  nursery  habitat  quality 
and  thus  may  be  used  to  delineate 
essential  fish  habitat  (EFH)  (Pihl  et 
al.,  2000;  Sullivan  et  al,  2000). 

Greater  amberjack  ( Seriola  dumer- 
ili) is  a  reef-associated  species  with 
a  circumglobal  distribution  in  sub- 
tropical and  temperate  waters  (Ma- 
nooch  and  Potts,  1997a).  In  the  Gulf 
of  Mexico,  S.  dumerili  is  the  largest 
carangid  and  supports  important 
recreational  and  commercial  fisher- 
ies (Thompson  et  al.,  1999).  Owing 
to  increased  fishing  effort  and  land- 
ings, S.  dumerili  in  the  Gulf  are  cur- 
rently assessed  as  overfished  (NOAA, 


2000).  Consequently,  detailed  life  his- 
tory information  is  needed  to  effec- 
tively guide  fishery  management  of 
this  valuable  resource.  To  date,  avail- 
able life  history  data  on  S.  dumerili 
have  almost  entirely  been  based  on 
assessments  of  subadults  and  adults 
(Manooch  and  Potts,  1997a,  1997b; 
Thompson  et  al.1).  Despite  the  impor- 
tance of  early  life  processes,  data  on 
juvenile  or  young-of-the-year  (YOY) 
S.  dumerili  are  limited  to  qualitative 
surveys  of  pelagic  Sar-gassum  (Bor- 
tone  et  al.,  1977;  Settle,  1993). 

The  National  Marine  Fisheries  Ser- 
vice has  recently  designated  Sargas- 
sum as  essential  fish  habitat  (EFH) 
of  several  coastal  migratory  species 
including  S.  dumerili  (NOAA,  1996). 
In  response,  the  goal  of  this  study 
was  to  examine  the  distribution  and 
growth  of  S.  dumerili  associated  with 
pelagic  Sargassum  mats  in  the  NW 
Gulf  of  Mexico.  Specifically,  objectives 
of  this  research  were  to  quantify  spa- 
tial and  temporal  patterns  of  habitat 
use  by  S.  dumerili  and  to  determine 
age,  hatching-date,  growth,  and  mor- 
tality of  S.  dumerili  by  using  otolith- 
based  techniques. 


Thompson,  B.  A.,  C.  A.  Wilson,  J.  H. 
Render,  M.  Beasley,  and  C.  Cauthron. 
1992.  Age,  growth,  and  reproductive 
biology  of  greater  amberjack  and  cobia 
from  Louisiana  waters.  Final  report 
NA90AA-H-MF722,  77  p.  Marine  Fish- 
eries Initiative  (MARFIN)  program. 
National  Marine  Fisheries  Service, 
NOAA,  St.  Petersburg,  FL. 


546 


Fishery  Bulletin  102(3) 


96    00' W       95 

i                        i 

50'  W 

94 :  40'W 

i 

94 

00'W 

93-20'W 

i 

30 

00'  N" 

/    ^STUDY       * 
\                              I 
"7        SITE            ;v 

\ 

Louisiana 

29 

20'  N" 

Galveston.  TX 

Inshore 

zone       y 

Offshore 

,'T---- 

zone 

/" 

28c 

40'  NJ 

^)r^                              >"■-- 

_ ..  ~  t~ 

';  /' 

28 

00'  N- 

20  m  depth  contour 

t 

N 

25  km 

Figure  1 

Map  of  sampling  locations  along  the  Texas  Gulf  coast  for 
S.  dumerili.  Inshore  (<15  nautical  miles)  and  offshore  (>15 
nautical  miles  I  zones  off  Galveston,  TX,  are  shown. 


Materials  and  methods 


Field  collections 


Seriola  dumerili  associated  with  pelagic  Sar-gassum  mats 
were  collected  off  Galveston,  Texas,  from  May  to  July 
over  a  two-year  period  (2000  and  2001)  (Fig.  1).  Inshore 
(<15  nautical  miles  |nmi])  and  offshore  (15-70  nmil 
zones  were  sampled  to  evaluate  the  potential  importance 
of  physiochemical  conditions  because  inshore  waters  off 
the  coast  of  Texas  are  heavily  influenced  by  estuarine 
processes  (Smith,  1980;  Sahl  et  al.,  1993).  Replicate 
samples  (3-5  per  trip)  in  both  the  inshore  and  offshore 
zones  were  collected  monthly  by  using  a  larval  purse 
seine  (20  m  long x 3.3  m  deep,  1000-pm  mesh).  The  purse 
seine  was  deployed  into  the  water  as  the  boat  encircled  a 
randomly  chosen  mat.  The  seine  was  pursed,  the  Sargas- 
sum  was  discarded,  and  fishes  were  tunneled  into  the 
codend,  collected,  and  frozen  on  dry  ice.  Distribution  and 
abundance  were  expressed  as  relative  abundance,  and 
catch  per  unit  of  effort  (CPUE)  represented  the  number 
of  fishes  per  purse-seine  collection.  In  addition,  a  small 
number  of  YOY  S.  dumerili  were  collected  with  hook- 
and-line  for  age  and  growth  information  only.  Standard 
lengths  (SL)  were  measured  to  the  nearest  0.1  mm,  and 
weights  to  the  nearest  0.1  g  before  otolith  extraction. 
GPS  locations  and  mat  volume  (lengthx widthxdepthi 
were  recorded  at  each  sample  location.  Environmental 
parameters  measured  included  sea  surface  temperature, 
salinity,  and  dissolved  oxygen.  Daily  sea  surface  tem- 


perature data  were  also  taken  from  NOAA  buoy  42035, 
22  nmi  offshore  of  Galveston,  TX. 

Otolith  procedures 

Sagittal  otoliths  were  extracted  from  S.  dumerili.  Oto- 
liths were  measured  to  the  nearest  0.001  mm  and 
weighed  to  the  nearest  0.0001  g.  Left  or  right  sagittae 
were  randomly  selected  and  mounted  in  epoxy  resin 
(Spurr,  1969).  Once  mounted,  a  Buehler  isomet  low- 
speed  saw  equipped  with  a  diamond  wafering  blade 
was  used  to  transversely  cut  embedded  otoliths.  Otolith 
sections  were  then  attached  to  petrographic  slides  with 
Crystalbond  thermoplastic  cement.  Type  A  alumina 
powder  1 0.3  fim)  and  400-  and  600-grit  sandpaper  were 
used  to  grind  both  sides  of  the  otolith,  and  a  polishing 
cloth  was  used  for  final  preparations. 

Age  was  determined  by  counting  growth  increments 
along  the  sulcus  from  the  core  to  the  outer  margin  by 
using  a  Nikon  Labophot-2  light  microscope  and  Opti- 
mas  6.2  image  analysis  software  (Media  Cybernetics, 
Silver  Spring,  MD).  Because  of  the  difficulty  of  enu- 
merating some  inner  increments  near  the  otolith  core, 
a  relationship  between  age  and  otolith  radius  of  several 
clear  specimens  was  used  to  predict  the  number  of 
increments  within  the  unclear  region.  Age  was  deter- 
mined by  adding  the  correction  factor  to  the  increment 
count  from  the  first  identifiable  increment  to  the  otolith 
margin  (Rooker  and  Holt,  1997).  Correction  factors 
consisting  of  mure  than  five  days  were  applied  to  499r 


Wells  and  Rooker:  Distribution,  age,  and  growth  of  young-of-the-year  Senola  dumerili 


547 


of  the  fishes  and  the  average  correction  accounted  for 
9.5%  of  the  actual  age  estimate.  Otolith  readings  with 
correction  factors  accounting  for  more  than  20%  of  the 
predicted  age  were  not  used  for  estimates  of  growth. 
The  following  correction  factor  was  used 

Age  (d)  =  2.88  x  otolith  radius  (jUm)  -  0.096 

(r2  =  0.88,  n=20). 

Additionally,  all  otolith  counts  were  repeated  twice 
to  ensure  adequate  precision.  Differences  in  readings 
of  more  than  20%  were  not  incorporated  into  growth 
estimates. 

Daily  deposition  of  growth  increments  on  sagit- 
tal otoliths  was  validated  by  using  wild  S.  dumerili 
(re=14,  136-193  mm  SL).  Fishes  caught  in  the  wild 
were  brought  into  the  laboratory  and  placed  in  a  cir- 
cular holding  tank  (1.71  m  diameterx0.75  m  depth) 
for  48  hours.  Fishes  were  then  placed  in  a  separate 
tank  containing  80  liters  of  seawater  with  100  mg/L  of 
alizarin  complexone  for  two  hours  (Thomas  et  al.,  1995) 
and  returned  to  the  circular  holding  tank.  Individuals 
were  fed  approximately  10%  of  their  body  weight  daily. 
Fishes  marked  with  alizarin  were  removed  from  the 
tank  after  5  (;?  =  5),  10  (n=5),  and  15  (n=4)  days.  The 
number  of  otolith  increments  between  the  alizarin  mark 
and  outer  edge  were  then  counted  for  daily  increment 
verification.  Otolith  slides  were  coded  so  that  all  read- 
ings were  blind. 

Hatching  dates  were  determined  for  all  individuals 
by  subtracting  daily  age  from  date  of  capture.  An  age- 
specific  mortality  adjustment  was  made  for  individuals 
because  larger  S.  dumerili  have  spent  more  time  in 
the  early  life  stages  and  hence  individuals  from  these 
cohorts  have  experienced  greater  cumulative  mortality. 
Because  of  the  limited  number  of  individuals  in  2001, 
the  mortality  correction  was  calculated  only  for  year 
2000  collections  and  applied  to  hatching-date  distribu- 
tions in  2000  and  2001.  Age-specific  mortality  adjust- 
ments were  made  according  to  the  method  described  by 
Rooker  and  Holt  (1997). 

Growth  and  mortality  of  S.  dumerili  were  estimated 
by  using  otolith-derived  ages.  Daily  growth  rates  were 
estimated  by  using  the  linear  growth  equation 

SL  -  slope  (age)  +  y-intercept 

and  were  reported  as  mm/d.  Length-at-age  data  were 
also  fitted  with  curvilinear  growth  models  (von  Ber- 
talanffy,  Laird-Gompertz).  Percent  variation  in  length 
explained  by  age  for  both  curvilinear  models  was  slightly 
better  at  times  than  the  percent  variation  in  length 
explained  by  age  for  the  linear  model;  however,  certain 
model  parameters  (i.e.  LJ  were  biologically  unrealistic 
and  thus  the  linear  model  was  deemed  more  appropri- 
ate. Moreover,  when  possible,  L_  values  were  used  to 
model  length-at-age  data  and  the  nonlinear  models  were 
essentially  linear  over  the  limited  size  range  examined. 
Mortality  estimates  for  year  2000  S.  dumerili  were 
determined  by  using  a  regression  on  the  decline  in  log(>- 


transformed  abundance  on  age.  A  regression  coefficient 
(slope)  was  used  to  predict  the  instantaneous  mortality 
rate: 

\r\N,  =  ln7V0  -  Zt, 

where  Nt   =  abundance  at  age  t  (expressed  in  days); 

N0  =  an  estimate  of  abundance  at  hatching; 
and 
Z  (slope)  =  the  instantaneous  mortality  coefficient. 

Mortality  estimates  were  based  upon  10-day  cohort 
groupings.  Individuals  <40  days  old  were  not  included 
in  the  mortality  regression  because  of  an  ascending 
catch  curve  and  because  there  were  too  few  individuals 
>139  days  old  in  our  sample — probably  owing  to  gear 
avoidance  or  emigration  (or  both).  Therefore,  only  S. 
dumerili  between  40  and  139  days  (45-192  mm)  were 
used  to  estimate  mortality. 

Data  analysis 

Effects  of  location  and  date  on  CPUE  and  size  estimates 
were  examined  by  using  a  two-way  analysis  of  vari- 
ance (ANOVA).  Levene's  test  and  residual  examination 
established  if  the  homogeneity  of  variance  assumption 
was  met.  Normality  was  evaluated  by  plotting  residuals 
versus  expected  values.  Abundance  data  were  log  (.v+1) 
transformed  when  necessary  to  normalize  data  and 
reduce  heteroscedasticity.  Tukey's  honestly  significant 
difference  ( HSD )  test  was  used  to  determine  a  posteriori 
differences  among  means.  Comparisons  of  spatial  and 
temporal  variation  in  growth  were  performed  by  using 
analysis  of  covariance  (ANCOVA).  Prior  to  ANCOVA 
testing,  the  homogeneity  of  slopes  assumption  was  exam- 
ined using  an  interaction  regression  (Ott,  1993).  If  no 
significant  interaction  was  detected,  ANCOVA  models 
were  used  to  test  for  differences  in  length-at-age  (y- 
intercepts)  (Ott,  1993).  Statistical  analysis  was  car- 
ried out  by  using  SYSTAT  8.0  (SYSTAT  Software  Inc., 
Richmond,  CA),  and  significance  was  set  at  the  alpha 
level  of  0.05. 


Results 

Environmental  conditions 

Average  temperatures  from  May  to  July  ranged  from 
27.9  to  30.1°C  in  2000  and  from  24.5  to  30.4°C  in  2001 
(Fig.  2).  Mean  temperatures  over  the  sampling  period 
were  29.2°C  and  27.9°C  for  2000  and  2001,  respectively. 
Zonal  differences  occurred:  the  inshore  zone  averaged 
28.7°C  (±0.3)  in  2000  and  28.1°C  (±0.9)  in  2001,  and 
the  offshore  zone  averaged  29.8°C  (±0.3)  in  2000  and 
27.6°C  (±0.9)  in  2001.  Similar  to  temperature  trends, 
mean  salinity  was  higher  in  2000  (34.6%< )  than  in  2001 
(31.9%o)  (Fig.  2).  Average  salinity  values  gradually 
increased  from  an  average  of  31.5%o  in  May  to  37.2%r  in 
July  of  2000.  A  large  drop  in  salinity  occurred  during 


548 


Fishery  Bulletin  102(3) 


mid-summer  of  2001,  from  37.6r/ic  in  May  to 
25.7'7<c  in  June  (owing  to  tropical  storm  Allison) 
and  rose  to  32. 3^  in  July.  Salinity  values  were 
lower  and  more  variable  within  the  inshore  zone, 
ranging  from  29f«  to  37^  (33.4%«  average)  in 
2000  and  from  15%c  to  37%o  (average  28.87« )  in 
2001.  In  contrast,  the  offshore  zone  exhibited 
higher  and  more  stable  salinity  values,  ranging 
between  337fc  and  38f;<  (36%<  average)  in  2000, 
and  between  287«  and  36%f  (34.97«  average) 
in  2001.  Temperature  and  salinity  values  are 
likely  to  be  influenced  by  variation  in  precipita- 
tion between  years.  Precipitation  from  January 
through  July  of  2000  (14.29  inches)  was  half  that 
of  2001  (29.92  inches)  and  well  below  the  30-year 
average  of  22.17  inches  (National  Weather  Ser- 
vice, Dickinson,  TX).  Dissolved  oxygen  content 
was  similar  between  years;  values  decreased 
throughout  the  summer  months  and  were  higher 
within  the  inshore  zone. 

Spatial  and  temporal  distribution 

A  total  of  181  YOY  S.  dumerili  was  collected 
from  42  purse  seines  over  the  two-year  study 
period.  CPUE  values  were  fourfold  higher  in 
2000  than  in  2001.  averaging  6.38  (±3.0)  and 
1.50  (±0.8)  per  seine,  respectively  (Fig.  3A). 
A  significant  year  effect  indicated  that  rela- 
tive abundance  was  higher  in  2000  (P=  0.019). 
Additionally,  CPUE  values  were  higher  in  the 
offshore  zone  in  both  years  (Fig.  3,  B  and  C). 
However,  no  significant  zonal  difference  existed  in  abun- 
dance between  the  inshore  and  offshore  zones  in  2000 
(P=0.063)  or  2001  (P=0.058).  Temporal  patterns  indi- 
cated S.  dumerili  was  highly  abundant  in  May  and  June, 
declining  in  July  in  both  years  (Fig.  3A).  A  significant 
seasonal  effect  occurred  for  2000  when  highest  relative 
abundance  occurred  in  June  with  a  CPUE  of  16.2  (±0.8) 
(Tukey  HSD,  P<0.05). 

Size  comparison 

Sizes  of  S.  dumerili  ranged  from  33  to  210  mm  SL 
(mean  125  mm  SL  ±3.8).  Juveniles  greater  than  100 
mm  accounted  for  68%  of  the  total  catch,  whereas  indi- 
viduals less  than  50  mm  accounted  for  only  15%.  Size 
differences  of  S.  dumerili  were  observed  between  2000 
(average  125.5  mm)  and  2001  (average  141.5  mm);  sig- 
nificantly larger  S.  dumerili  were  collected  from  the 
offshore  zone  in  2001  (P=0.001).  A  significant  interaction 
(yearxmonth)  occurred  that  indicated  that  the  magni- 
tude of  size  differences  was  variable  over  time.  Sizes 
were  also  significantly  different  between  zones  in  2000; 
larger  individuals  were  collected  within  the  offshore  zone 
(P=0.025).  No  zonal  comparison  was  performed  for  2001 
because  few  individuals  were  collected  from  the  inshore 
zone.  In  addition,  a  trend  existed  within  both  years: 
mean  sizes  significantly  increased  from  May  to  June, 
then  decreased  in  July  (Tukey  HSD,  P<0.05). 


40    - 

A      Temperature                                                    2^° 

-  40 

35    - 
30   - 

A      Buoy  temperature                                   e>~-~~^'^ 
O      Salinity                                     a — - —                ■ 

-  35 

-  30 

25   - 

^*<^*~~ 

-  25 

20   - 
15   - 

^^^^^ 

-  20 

£                 2T 

-   15 

10   - 

-   10 

o      5- 

-  5 

U                         i                   i                    i                    i                 i                      i                        \j          uj 

2?                 Jan        Feb        Mar        Apr       May       June       July                 = 

CO                                                                                                                                                                                            ^ 
CD 

if    40  - 

A      Temperature                             n                      ^UUl 

"40    i 

H      35   - 

A      Buoy  temperature                      ^v 

-  35 

30   - 

O      Salinity                                           \.          -^^L 

-  30 

25   - 

Jr-—***^^ 

-  25 

20   - 

^^~ 

-  20 

15   - 

tr-——*^^^ 

~  15 

10  - 

-  10 

5  - 

-  5 

0 

0 

■               i              i              i              i               i 

Jan        Feb        Mar       Apr        May       June       July 

Month 

Figure  2 

Environmental  conditions  from  January  to  July  of  2000  and 

2001.  Average  temperature  (°C)  and  salinity  ('«)  values.  Open 

triangles  represent  temperature  data  from  NOAA  buoy  42035, 

located  22  nautical  miles  offshore  of  Galveston,  TX. 

Hatching-date  distribution 

Hatching-date  distributions  for  S.  dumerili  were  pro- 
tracted in  both  2000  and  2001.  Fishes  collected  in  2000 
exhibited  hatching-dates  from  29  January  to  25  May 
(117  days),  whereas  those  collected  in  2001  hatched 
from  11  January  to  30  May  (139  days)  (Fig.  4).  In  2000, 
over  80%  of  the  fishes  appeared  to  result  from  spawning 
events  in  March  and  early  April.  The  adjusted  distri- 
butions from  the  age-specific  mortality  correction  for 
both  2000  and  2001  were  indistinguishable  from  those 
without  the  correction. 

Age  and  growth 

Results  of  the  age-validation  exercise  indicated  that 
juvenile  S.  dumerili  deposit  otolith  increments  on  a 
daily  basis  (Fig.  5).  Average  increment  counts  at  day 
5,  10,  and  15  were  4.8  (±0.2  SD),  9.2  (±0.4),  and  14.0 
(±0.7),  respectively.  A  relationship  between  the  observed 
versus  expected  increments  was  described  by  the  follow- 
ing equation: 

Observed  increments  -  0.92  (expected  increments)  +  0.14 

(r2=0.95) 

where  days  after  staining  represent  expected  increment 
count. 


Wells  and  Rooker:  Distribution,  age,  and  growth  of  young-of-the-year  Senola  dumenli 


549 


Validation  of  daily  growth  increments  has  been  ob- 
served in  a  similar  study  involving  juvenile  (0-60 
days)  Seriola  quinqueradiata  (Sakakura  and  Tsuka- 
moto,  1997). 

Age  of  S.  dumerili  was  similar  between  years;  es- 
timated ages  ranged  from  41  to  150  days  (35  to  210 
mm  SL)  in  2000  and  from  35  to  120  days  (33  to  198 
mm  SL)  in  2001  (Fig.  6).  Interannual  differences  in 
growth  were  observed:  2000  (1.65  mm/d).  2001  (2.00 
mm/d)  (ANCOVA,  slope,  P<0.001)  (Fig.  7).  A  signifi- 
cant cohort  effect  was  also  observed;  the  late-season 
(April)  cohort  experienced  the  fastest  growth  (ANCO- 
VA, slopes,  P<0.001)  (Fig.  8).  Average  cohort-specific 
growth  rates  of  S.  dumerili  spawned  in  February, 
March,  and  April  of  2000  were  0.85  mm/d,  1.15  mm/d, 
and  2.76  mm/d,  respectively.  In  addition,  a  signifi- 
cant difference  in  growth  was  observed  for  S.  du- 
merili collected  from  inshore  (1.55  mm/d)  and  offshore 
(1.65  mm/d)  zones  of  2000  (ANCOVA,  slope,  P<0.001) 
(Fig.  9).  Again,  the  lack  of  individuals  within  the  in- 
shore zone  in  2001  precluded  a  comparison  between 
zones  for  that  year. 

Mortality 

Owing  to  the  limited  number  of  S.  dumerili  collected 
in  2001,  a  single  catch  curve  was  developed  for  the 
2000  year  class,  and  the  mortality  coefficient  (Z)  was 
0.0045  (0.45%/d)  for  individuals  between  40  and  139 
days  (Fig.  10).  Cumulative  mortality  was  estimated 
for  the  100-day  period  (40-139  days),  resulting  in  an 
overall  mortality  of  36%. 


20 

15 

10 

5 

0 


1  A 


□  2000 
■  2001 


MAY 


JUNE 


JULY 


40  " 
UJ     30  - 

B 

□  Offshore 
■  Inshore 

z> 

CL 

O     20- 

T 

10  - 

o- 

1 

MAY 


JUNE 


JULY 


o  - 

8  - 
6  - 

c 

4  J 

2  " 

U 

□  Offshore 
■  Inshore 


r^i^ 


MAY 


JUNE 
Months 


JULY 


Figure  3 

Relative  abundance  (number  per  purse  seine)  (±1  SE) 
of  S.  dumerili  collected  in  association  with  Sargassum 
mats:  (A)  2000  and  2001:  (B)  2000  by  zones;  (C)  2001 
bv  zones. 


Discussion 

The  size  range  of  S.  dumerili  collected  in  association 
with  Sargassum  ranged  from  approximately  30  to  210 
mm  (SL),  and  these  sizes  are  similar  to  those  reported  in 
other  studies  investigating  fish  assemblages  associated 
with  pelagic  Sargassum.  Bortone  et  al.  (1977)  collected 
several  small  S.  dumerili  (12-72  mm  SL)  in  the  eastern 
Gulf,  whereas  individuals  collected  in  the  western  Atlan- 
tic by  Dooley  (1972)  ranged  from  13  to  108  mm  (SL). 
Cho  et  al.  (2001)  found  juvenile  S.  dumerili  (35-120  mm 
TL)  associated  with  drifting  Sargassum  in  the  western 
Pacific.  Additionally,  Sakakura  and  Tsukamoto  (1997) 
collected  over  200  juvenile  Japanese  amberjack  (S.  quin- 
queradiata) (18-114  mm  TL)  associated  with  pelagic 
Sargassum  in  the  East  China  Sea.  Results  of  the  present 
study  and  others  indicate  that  pelagic  Sargassum  mats 
in  the  NW  Gulf  of  Mexico  serve  as  nursery  habitat  for 
S.  dumerili. 

The  limited  size  range  of  S.  dumerili  associated  with 
pelagic  Sargassum  indicates  that  a  shift  in  habitat  use 
may  occur  at  approximately  5-6  months  of  age.  Indi- 
viduals greater  than  210  mm  (SL)  have  not  been  found 
in  association  with  pelagic  Sargassum,  and  larger  S. 
dumerili  (ca.  300  mm  TL)  are  relatively  common  in  the 
recreational  headboat  fishery  in  the  Gulf  of  Mexico  (Ma- 


nooch  and  Potts,  1997a).  As  a  consequence,  S.  dumerili 
may  transition  from  a  pelagic  to  a  demersal  existence 
at  the  late  juvenile  stage  (between  200  mm  SL  and 
300  mm  TL).  Pipitone  and  Andaloro  (1995)  found  a  shift 
in  the  diet  of  S.  dumerili,  from  a  diet  predominately 
consisting  of  crustaceans  toward  one  of  fish  >200  mm 
(SL),  further  supporting  this  hypothesis. 

Seriola  dumerili  abundance  was  greater  in  the  off- 
shore zone  than  the  inshore  zone  throughout  the  sam- 
pling period.  These  patterns  of  habitat  use  are  consis- 
tent with  earlier  information  that  indicates  S.  dumerili 
is  an  offshore  species  (Hildebrand  and  Cable,  1930). 
The  proximity  to  spawning  grounds  may  contribute  to 
the  observed  spatial  patterns  because  S.  dumerili  are 
known  to  spawn  in  offshore  areas  (Fahay,  1975).  Physi- 
ological preferences  may  also  contribute  to  the  domi- 
nance of  S.  dumerili  in  the  offshore  zone.  In  our  study, 
salinity  values  were  higher  in  the  offshore  zone  but 
more  variable  within  the  inshore  zone,  suggesting  that 
freshwater  inflow  influences  conditions  within  the  in- 
shore zone.  Chen  et  al.  (1997)  determined  that  optimum 
salinity  conditions  for  S.  dumerili  larvae  were  between 
32%<r  and  35%t,  and  larvae  remained  inactive  below  a 
salinity  of  30%<r.  Zonal  differences  in  temperature  and 
dissolved  oxygen  were  also  observed.  Tzeng  et  al.  (1997) 


550 


Fishery  Bulletin  102(3) 


6- 

5  " 
4 
3 
2 


7- 

6 

5  • 

4 

3 

2 

1 

0 


2000 


r-        t-       cy 


,-  -r-  CM 


CO  U"> 


l-         i-         CM 


2001 


T-  T-  C\J 


t-      r>      ■*     *- 

CO         Rl         <-        CM 


t-        *-       CN 


CO         CO         CO 


Hatching  date 


Figure  4 

Hatch-date  distributions  of  S.  dumerili  associated  with  Sargassum  mats 
in  2000  and  2001. 


16  i 

£ 

14  - 

/  & 

f*  / 

12  - 

/           0 

CD 

S/ 

E      10  - 

tft/ 

cd 

'/rV 

CJ 

''/ 

£        8  - 

'/          ° 

o 

CD          c 

a       °  - 

ty 

E 

&y 

3 

Z        4  - 

/                      Increments  (expected) 

2  - 

/                                   O    Increments  (observed) 

u               I           I           I           I           I           I           1           1 
0          2           4          6           8         10         12         14         16 

Days  after  staining 

Figure  5 

Linear  regression  of  age  verification  for  S. 

dumerili.  Circles  and  solid  line  represent  the 

number  of  daily  increments  observed  after 

staining,  and  dotted  line  represents  the  number 

of  increments  expected. 

attributed  the  distribution  of  fishes  from  nearshore  to 
offshore  stations  to  environmental  factors,  season,  and 
life  history  strategies.  Furthermore,  the  combination  of 


available  resources  (i.e.  food  and  habitat),  seasons,  and 
physiochemical  tolerances  may  account  for  the  observed 
spatial  patterns  of  habitat  use. 

Temporal  patterns  of  size-specific  habitat  use  showed 
similar  trends  between  years  and  appeared  to  be  relat- 
ed to  spawning  season.  Relative  abundance  of  small  S. 
dumerili  was  highest  early  in  the  season  (May),  declined 
in  June,  and  further  increased  late  into  the  season 
(July)  for  both  2000  and  2001.  Nevertheless,  small  juve- 
niles were  collected  during  the  entire  collection  period, 
which  suggests  that  S.  dumerili  spawning  in  the  NW 
Gulf  is  protracted.  Previous  studies  have  found  that 
S.  dumerili  spawn  throughout  the  spring  and  summer 
months  ( March- July)  (Marino  et  al.,  1995;  Cummings 
and  McClellan,  1996).  In  addition,  Fahay  (1975)  sug- 
gested, on  the  basis  of  larval  collections  in  the  western 
Atlantic,  that  spawning  occurs  in  the  winter.  Despite 
the  limited  duration  of  our  collection  efforts,  our  results 
are  consistent  with  these  findings  with  63%  of  year- 
2000  S.  dumerili  and  36f>  of  year-2001  fish  resulting 
from  spring  spawning  events.  The  remaining  individu- 
als were  spawned  January  through  early  March. 

Growth  estimates  indicated  that  S.  dumerili  have 
rapid  growth  throughout  early  life  stages.  Based  on 
linear  growth  models,  average  growth  of  S.  dumerili 
was  1.45  mm/d — an  estimate  similar  to  that  of  Manooch 
and  Potts's  (1997b)  study  in  the  Gulf  (average  growth  of 
1.17  mm/d  for  age-1  individuals).  However,  growth  com- 
parisons may  be  invalid  because  their  study  estimated 
growth  based  on  counts  of  annuli  and  no  temperature 


Wells  and  Rooker:  Distribution,  age,  and  growth  of  young-of-the-year  Senolo  dumerili 


551 


data  were  presented.  Because  of  the  lack  of  studies 
investigating  growth  of  YOY  S.  dumerili,  we  compared 
our  estimates  to  those  in  Sakakura  and  Tsukamoto's 
(1997)  study  of  YOY  S.  quinqueradiata  where  growth 
rates  were  estimated  at  1.3  mm/d.  Average  temperature 
in  their  study  was  21.2°C,  which  was  considerably  lower 
than  the  average  during  our  study  (28.6°C)  and  may 
account  for  their  slower  growth  rates. 

Variation  in  growth  of  S.  dumerili  was  observed  and 
rates  were  significantly  higher  in  the  offshore  zone 
and  greater  for  the  late  season  cohort.  Differences  in 
water  temperature  may  be  partly  responsible  for  ob- 
served differences  in  growth.  Planes  et  al.  (1999)  sug- 
gested that  spatial  differences  in  growth  of  juvenile 
sparid  fishes  were  a  result  of  water  temperature  and 
currents.  The  proximity  between  zones  in  this  study 
may  have  masked  differences  in  hydrography;  however, 
temperatures  were  higher  in  the  offshore  zone  (29.8°C, 
CV=0.03)  than  in  the  inshore  zone  (28.7°C,  CV=0.04), 
and  warmer  temperatures  were  likely  contributing  to 
faster  growth  rates  in  offshore  waters.  Intra-annual 
(cohort-specific)  growth  patterns  indicated  that  the  late- 
season  cohort  had  the  fastest  growth.  Similar  to  trends 
between  zones,  temperature  was  lowest  for  the  slowest 
growing  cohort  (early  season)  and  highest  for  the  fast- 
est growing  cohort  (late  season).  Although  temperature 
may  affect  early  life  growth  of  S.  dumerili,  differences 
in  growth  may  be  attributed  to  other  factors  such  as 
prey  availability  and  predator  activity  (Houde,  1987; 
Paperno  et  al.,  2000;  Plaganyi  et  al.,  2000).  Moreover, 
a  clear  distinction  exists  in  the  size  classes  of  YOY  S. 
dumerili  in  comparisons  of  growth  rates  and  these  dif- 
ferences likely  contribute  to  the  observed  results. 

The  mortality  rate  of  YOY  S.  dumerili  associated  with 
pelagic  Sargassum  was  estimated  at  0.45  %/d  for  fishes 


4-1 
2- 

0- 

8" 
6" 

2000 

4- 

2  r 

...M„. 

-  f " 

.-d-Lo-n 

2001 


mm 


^      w      n     n 


Age  class  (days) 

Figure  6 

Age-frequency  distribution  of  S.  dumerili  collected  in 
association  with  Sargassum  in  2000  and  2001. 


collected  in  2000.  These  findings  are  well  below  similar 
studies  investigating  mortality  of  YOY  individuals.  Nelson 
(1998)  calculated  a  mortality  estimate  of  2.1-2.3^/d  for 
pinfish  in  three  different  bay  areas  in  the  eastern  Gulf  of 
Mexico.  In  addition,  Deegan  (1990)  estimated  YOY  men- 
haden mortality  between  1.7  and  2.1'S/d  in  the  northern 


250" 

,'     a 

200" 

A 

t&y^u   D 

E 

A°£a$ 

$ffi^° 

~X      150  - 

a    Lt^PSjaKi- 

CT 

J^r 

0> 

■Vx       D 

"D 

/y 

ffl      100  - 

/? 

"D 

C 

ns 

CO 

*P  D 

50  - 

/Z& 

fSS                -B-  2000 

SL=  1.65Mge)-15.33 
r2=0.86 

-A-  2001 

Si.  =  2.00(/lc;e)-37.32 

r  =  0  95 

0 

1             ' 

'               '                ' 

0             20             40 

60             80            1 00 

Age  (days) 
Figure  7 

120           140            160 

Age-length  relationship 

of  S.  dumerili  determined  with  linear 

growth  curves  for  inters 

nnual  comparison. 

552 


Fishery  Bulletin  102(3) 


250  " 

Standard  length  (mm) 

O                   en                   O 
o                o                o 

s.*£          o 

a/       -&-  February        SU  0  85(4ge)+65.12 
MT°                                                      i2  =  0.34 

50  - 

Djfl 

Van        ~°~  March            SL  =  1  1 5(4ge)+40  69 
r  =  0.55 

-a—   April                SL=2  76(Age)-80  48 
^=0  95 

1                       ' 

I                                                  t                 i 

0             20            40 

60            80           100           120           140          160 

Age  (days) 

Figure  8 

Age-length  relationship 

of  S.  dumerili  determined  with  linear 

growth  curves  for  a  com 

parison  of  cohorts  of  different  hatching 

dates  in  2000. 

250  " 

200  - 

E 
P 

s' 

/ 

y 

/         a 

_□ 

0       n       &  B}flO 

Standard  length  ( 

en                   o                   oi 
o                o                o 

^JSsS        —*-  Inshore         SL  =  1  55{Age)-\2  11 

'rffi'                                                                       f  =  0.74 

-a-  Offshore        SL  =  1.65(/*ge)-15.33 

o  ■ 

^  =  086 

0             20            40            60             80            100          120           140          160 

Age  (days) 

Figure  9 

Age-length  relationship  of  S.  dumerili  determined  with  linear 

growth  curves  for  a  comparison  of  zones  in  2000. 

Gulf.  These  studies  included  estuarine-dependent  spe- 
cies and  consisted  of  smaller  individuals.  Because  our 
estimates  were  limited  to  age  40-139  d  individuals,  the 
lack  of  smaller  fishes  precluded  any  mortality  estimates 
of  younger  S.  dumerili.  These  estimates  provide  baseline 
information  on  mortality  of  YOY  S.  dumerili;  however, 
more  detailed  studies  will  be  needed  to  adequately  de- 
termine mortality  rates  of  YOY  S.  dumerili. 

Based  on  observed  patterns  of  distribution  and 
growth  in  the  NW  Gulf  of  Mexico,  early  life  survival  of 


S.  dumerili  may  depend  on  pelagic  Sargassum.  Results 
of  this  study  suggest  that  S.  dumerili  are  associated 
with  this  habitat  over  a  limited  size  range  and  exhibit 
rapid  growth  during  the  first  six  months.  Addition- 
ally, S.  dumerili  were  more  abundant  and  exhibited 
higher  growth  in  offshore  areas  where  potential  spawn- 
ing may  occur.  Thus,  Sargassum  appears  to  provide 
nursery  habitat  for  YOY  S.  dumerili,  and  may  influ- 
ence the  recruitment  potential  of  this  valuable  fishery- 
species. 


Wells  and  Rooker:  Distribution,  age,  and  growth  of  young-of-the-year  Senola  dumerili 


553 


Log.W  =  1  506-0  0045(4ge) 

x                        r  =0.315 

"i    i.o  - 

CD 
■D 

C 

CD 

*     0.5    - 

en 
O 

_l 

♦ 

♦ 

O         OO         OOOOOOOO         o         O         OOO 

i-c\jcr>"^-mcDr--.oocr)oi-CMco-^-ir) 

Age  (days) 

Figure  10 

Mortality  curve  of  S.  dumerili  based  upon  regression  plot  of  loge 

abundance  on  age  of  individuals  collected  in  2000. 

Acknowledgments 

We  thank  J.  Harper,  M.  Lowe,  B.  Geary,  J.  Turner, 
and  J.  Wells  for  their  assistance  in  the  field.  Fund- 
ing for  this  project  was  provided  by  The  Aquarium  at 
Moody  Gardens  (grant  479005  to  JRR).  Top  Hatt  char- 
ters provided  boat  time  offshore,  and  Kirk  Winemiller 
and  Jaime  Alvarado  offered  constructive  criticism  and 
suggestions. 


Literature  cited 

Bortone,  S.  A..  P.  A.  Hastings,  and  S.  B.  Collard. 

1977.     The  pe\agic-Sargassum  ichthyofauna  of  the  eastern 
Gulf  of  Mexico.     Northeast  Gulf  Sci.  1:60-67. 
Bricelj,  V.  M. 

1993.  Aspects  of  the  biology  of  the  northern  quahog, 
Mercenaria  mercenaries,  with  emphasis  on  growth  and 
survival  during  early  life  history.  In  Proceedings  of 
the  2nd  Rhode  Island  shellfish  industry  conference, 
p.  29-48.  Sea  Grant  Progress  report,  Narragansett. 
RI. 
Chen.  C,  R.  Ji,  J.  Huang,  H.  He,  and  Z.  Liao. 

1997.     The    relationship    between   the    salinity    and 
the  embryonic,  early  larval  development  in  Seriola 
dumerili.     J.  Shanghai  Fish.  Univ.  6:5-10. 
Cho,  S-H..  J-G.  Myoung,  and  J-M.  Kim. 

2001.     Fish  fauna  associated  with  drifting  seaweed  in  the 
coastal  area  of  Tongyeong,  Korea.     Trans.  Am.  Fish. 
Soc.  130:1190-1202. 
Cole,  J. 

1999.     Environmental  conditions,  satellite  imagery,  and 
clupeoid  recruitment  in  the  northern  Benguela  upwell- 
ing  system.     Fish.  Oceanogr.  8:25-38. 
Deegan,  L.  A. 

1990.     Effects  of  estuarine  environmental  conditions 
on   population   dynamics  of  young-of-the-year  gulf 
menhaden.     Mar.  Ecol.  Prog.  Ser.  68:195-205. 
Dooley,  J.  K. 

1972.     Fishes  associated  with  the  pelagic  Sargassum 


complex,  with  a  discussion  of  the  Sargassum   com- 
munity.    Contrib.  Mar.  Sci.  16:1-32. 
Fahay,  M.  P. 

1975.  An  annotated  list  of  larval  and  juvenile  fishes 
captured  with  surface-towed  meter  net  in  the  South 
Atlantic  Bight  during  four  RV  Dolphin  cruises  between 
May  1967  and  February  1968.  NOAA  Tech.  Rep.  NMFS. 
SSRF-685:l-39. 
Hildebrand,  S.  F.,  and  L.  E.  Cable. 

1930.     Development  and  life  history  of  fourteen  teleostean 
fishes  at  Beaufort,  N.C.     Fish.  Bull.  46:383-488. 
Houde,  E.  D. 

1987.  Fish  early  life  dynamics  and  recruitment  vari- 
ability. Am.  Fish.  Soc.  Symp.  2:17-29. 
1996.  Evaluating  stage-specific  survival  during  the  early 
life  offish.  In  Survival  strategies  in  early  life  stages 
of  marine  resources  (Y.  Watanabe,  Y.  Yamashita,  and  Y. 
Oozeki.  eds.l.  p.  51-66.  AA  Balkema.  Rotterdam. 
Leak,  J.  C. 

1981.     Distribution  and  abundance  of  carangid  fish  larvae 
in  the  eastern  Gulf  of  Mexico,  1971-1974.     Biol.  Ocean- 
ogr. 1(11:1-28. 
Manooch,  C.  S.,  HI.  and  J.  C.  Potts. 

1997a.     Age,  growth  and  mortality  of  greater  amber- 
jack  from  the  southeastern  United  States.     Fish.  Res. 
30:229-240. 
1997b.     Age,  growth,  and  mortality  of  greater  amberjack, 
Seriola  dumerili,  from  the  U.S.  Gulf  of  Mexico  headboat 
fishery.     Bull.  Mar.  Sci.  61:671-683. 
Marino,  G.,  A.  Mandich,  A.  Massari,  F.  Andaloro,  S.  Porrello, 
M.  G.  Finoia,  and  F.  Cevasco. 

1995.  Aspects  of  reproductive  biology  of  the  Mediter- 
ranean amberjack  (Seriola  dumerili  Risso)  during  the 
spawning  period.     J.  Appl.  Ichthyol.  11:9-24. 

Meekan,  M.  G.,  and  L.  Fortier. 

1996.  Selection  for  fast  growth  during  the  larval  life  of 
Atlantic  cod  Gadus  morhua  on  the  Scotian  Shelf.  Mar. 
Ecol.  Prog.  Ser.  137:25-37. 

Nelson,  G.  A. 

1998.  Abundance,  growth,  and  mortality  of  young-of- 
the-year  pinfish,  Lagodon  rhomboides,  in  three  estu- 
aries along  the  gulf  coast  of  Florida.  Fish.  Bull.  96: 
315-328. 


554 


Fishery  Bulletin  102(3) 


NOAA  (National  Oceanic  and  Atmospheric  Administration). 

1996.  Magnuson-Stevens  Fishery  Conservation  and  Man- 
agement Act.  as  amended  through  Oct.  11,  1996.  NOAA 
Tech.  Mem.  NMFS-F/SPO-23.  121  p. 

2000.     Stock  assessments  of  Gulf  of  Mexico  greater  amber- 
jack  using  data  through  1998.     Sustainable  Fisheries 
Division,  Southeast  Fisheries  Science  Center,  NMFS 
report  99/00-100,  27  p. 
Ott,  R.  L. 

1993.     An  introduction  to  statistical  methods  and  data 
analysis,  4th  ed.,  1051  p.     Duxbury  Press,  Belmont, 
CA. 
Paperno,  R..  T.  E.  Targett,  and  P.  A.  Grecay. 

2000.     Spatial  and  temporal  variation  in  recent  growth, 
overall  growth,  and  mortality  of  juvenile  weakfish 
(Cynoscion  regalis)  in  Delaware  Bay.     Estuaries  23:10-20. 
Pihl,  L.,  J.  Modin,  and  H.  Wennhage. 

2000.     Spatial  distribution  patterns  of  newly  settled  plaice 
iPleuronectes  platessa  L.)  along  the  Swedish  Skagerrak 
archipelago.     J.  Sea.  Res.  44:65-80. 
Pipitone,  C,  and  F.  Andaloro. 

1995.     Food  and  feeding  habits  of  juvenile  greater  amber- 
jack,  Seriola  dumerili  (Osteichthyes,  Carangidae)  in 
inshore  waters  of  the  central  Mediterranean  Sea.     Cy- 
bium  19:305-310. 
Plaganyi,  E.  E.,  L.  Hutchings,  and  J.  G.  Field. 

2000.     Anchovy  foraging:  simulating  spatial  and  temporal 
match/mismatches  with  zooplankton.     Can.  J.  Fish. 
Aquat.  Sci.  57:2044-2053. 
Planes,  S.,  E.  Macpherson,  F  Biagi,  A.  Garcia-Rubies, 
J.  Harmelin,  M.  Harmelin-Vivien,  J.  Y.  Jouvenel,  L.  Tunesi, 
L.  Vigliola,  and  R.  Galzin. 

1999.     Spatio-temporal  variability  in  growth  of  juvenile 
sparid  fishes  from  the  Mediterranean  littoral  zone.     J. 
Mar.  Biol.  Assoc.  U.K.  79:137-143. 
Rilling,  G.  C,  and  E.  D.  Houde. 

1999.     Regional  and  temporal  variability  in  growth  and 
mortality  of  bay  anchovy,  Anchoa  mitchilli,  larvae  in 
Chesapeake  Bay.     Fish.  Bull.  97:555-569. 
Rooker.  J.  R.,  and  S.  A.  Holt. 

1997.  Utilization  of  subtropical  seagrass  meadows  by 
newly  settled  red  drum  Sciaenops  ocellatus:  patterns 
of  distribution  and  growth.  Mar.  Ecol.  Prog.  Ser. 
158:139-149. 

Rooker,  J.  R„  S.  A.  Holt,  G.  J.  Holt,  and  L.  A.  Fuiman. 

1999.  Spatial  and  temporal  variability  in  growth,  mortal- 
ity, and  recruitment  potential  of  postsettlement  red  drum. 


Sciaenops  ocellatus,  in  a  subtropical  estuary.     Fish. 
Bull.  97:581-590. 

Sahl,  L.  E.,  W.  J.  Merrell,  and  D.  C.  Biggs. 

1993.  The  influence  of  advection  on  the  spatial  variabil- 
ity of  nutrient  concentrations  on  the  Texas-Louisiana 
continental  shelf.     Cont.  Shelf.  Res.  13:233-251. 

Sakakura.  Y..  and  K.  Tsukamoto. 

1997.  Age  composition  in  the  schools  of  juvenile  yellowtail 
Seriola  quinqueradmta  associated  with  drifting  seaweeds 
in  the  East  China  Sea.     Fish.  Sci.  63:37-41. 

Sano,  M. 

1997.  Temporal  variation  in  density  dependence:  Recruit- 
ment and  postrecruitment  demography  of  a  temperate 
zone  sand  goby.     J.  Exp.  Mar.  Biol.  Ecol.  214:67-84. 

Schnack,  D.,  St.  John,  R.  Schneider,  B.  Klenz,  A.  Nissling,  and 
E.  Aro. 

1998.  3rd  European  Mar.  Sci.  and  Tech.  Conf.  (MAST). 
Fisheries  and  Aquaculture  5:181-185. 

Settle,  L.  R. 

1993.     Spatial  and  temporal  variability  in  the  distribution 
and  abundance  of  larval  and  juvenile  fishes  associated 
with  pelagic  Sargassum.     M.S.  thesis,  64  p.     Univ.  North 
Carolina-Wilmington.  Wilmington.  NC. 
Smith,  N.  P. 

1980.     On  the  hydrography  of  shelf  waters  off  the  central 
Texas  coast.     J.  Phys.  Oceanogr.  10:806-813. 
Spurr,  A.  R. 

1969.     A  low-viscosity  epoxy  resin  embedding  medium  for 
electron  microscopy.     J.  Ultrastruct.  Res.  26:31-43. 
Sullivan,  M.  C,  R.  K.  Cowen,  K.  W.  Able,  and  M.  P.  Fahay. 

2000.     Spatial  scaling  of  recruitment  in  four  continental 
shelf  fishes.     Mar.  Ecol.  Prog.  Ser.  207:141-154. 
Thomas,  L.  M.,  S.  A.  Holt,  and  C.  R.  Arnold. 

1995.  Chemical  marking  techniques  for  larval  and  juve- 
nile red  drum  [Sciaenops  ocellatus)  otoliths  using  dif- 
ferent fluorescent  markers.  In  Recent  developments  in 
fish  otolith  research,  p.  703-717.  The  Belle  W.  Baruch 
Library  in  Marine  Science,  number  19.  LIniv.  South 
Carolina  Press,  Columbia,  SC. 
Thompson.  B.  A.,  M.  Beasley,  and  C.  A.  Wilson. 

1999.  Age  distribution  and  growth  of  greater  amber- 
jack,  Seriola  dumerili,  from  the  north-central  Gulf  of 
Mexico.     Fish.  Bull.  97:362-371. 

Tzeng,  W.  N„  Y  T.  Wang,  and  Y  T.  Chern. 

1997.  Species  composition  and  distribution  offish  larvae 
in  Yenliao  Bay.  northeastern  Taiwan.  Zool.  Stud. 
36:146-158. 


555 


Identification  of  formalin-preserved  eggs  of 
red  sea  bream  iPagrus  major)  (Pisces:  Sparidae) 
using  monoclonal  antibodies 


Shingo  Hiroishi 

Yasutaka  Yuki 

Eriko  Yuruzume 

Faculty  of  Biotechnology 

Fukui  Prefectural  University 

1-1  Gakuen-cho 

Obama  City,  917-0003  Fukui,  Japan 

E-mail  address  (for  S  Hiroishi),  hiroishi@fpu.ac.|p 

Yosuke  Onishi 

Tomoji  Ikeda 

Hironobu  Komaki 

Kansai  Environmental  Engineer  Center 
1-3-5  Azuchi-cho,  Chuo-ku 
Osaka  City,  541-0052  Osaka,  Japan 

Muneo  Okiyama 

Ocean  Research  Institute 
University  of  Tokyo 
1-15-1  Minamidai,  Nakano-ku, 
Tokyo,  Japan 


Catches  of  important  commercial  fish 
such  as  red  sea  bream,  fiat  fish,  and 
yellowtail  are  decreasing  in  Japan. 
In  order  to  sustain  these  species  it  is 
especially  important  that  their  distri- 
bution and  biomass  at  all  life  stages 
are  known.  However,  information  on 
the  early  life  stages  of  these  species  is 
limited  because  identifying  the  eggs 
and  larvae  of  such  fish  is  sometimes 
extremely  difficult. 

Mito  (1960,  1979)  and  Ikeda  and 
Mito  (1988)  developed  methods  for 
identifying  pelagic  fish  eggs  based  on 
morphological  features.  However,  their 
methods  have  limitations  because 
many  unidentified  eggs  have  similar 
features.  In  addition,  eggs  are  usu- 
ally fixed  in  formaldehyde  solution 
just  after  collection  in  the  field.  This 
procedure  may  alter  several  egg  char- 
acteristics and  therefore  prevent  iden- 
tification (Ikeda  and  Mito,  1988),  or 
make  identification  difficult  when  the 
egg  diameter  measures  0.8-1.0  mm 
because  so  many  kinds  of  eggs  fall  in 
that  range.  Thus,  an  alternative  iden- 
tification method  would  be  useful. 


Effective  genetic  analyses  for  iden- 
tifying fish  eggs  or  larvae  (or  both) 
have  been  developed  by  Graves  et  al. 
(1989),  Daniel  and  Graves  (1994), 
and  Shao  et  al.  (2002).  However, 
their  methods  may  have  limitations 
if  samples  are  preserved  in  formal- 
dehyde for  several  years  or  if  DNA 
must  be  extracted  from  numerous 
samples.  In  addition,  we  are  lacking 
the  DNA  sequences  for  many  species 
sequences  that  are  necessary  for  iden- 
tifying eggs  in  the  field. 

We  have  successfully  produced 
monoclonal  antibodies  to  differenti- 
ate harmful  marine  phytoplankton 
species  from  morphologically  simi- 
lar harmless  species  (Hiroishi  et 
al.,  1988;  Nagasaki  et  al.,  1991b; 
Sako  et  al.,  1993;  Vrieling  et  al., 
1993;  Hiroishi  et  al.,  2002)  as  well 
as  Microcystis,  a  toxic  fresh  water 
bloom-forming  cyanobacteria  (Kondo 
et  al.,  1998).  These  antibodies  were 
obtained  from  a  culture  supernatant 
solution  of  hybridoma  cells  that  was 
produced  by  a  cell  fusion  procedure 
between  myeloma  cells  and  antibody- 


producing  spleen  cells.  The  specific 
antibodies  described  above  could  be 
used  to  detect  and  quantify  harmful 
bloom-forming  microorganisms  that 
react  with  the  monoclonal  antibod- 
ies and  that  secondarily  react  with 
fluorescein  isothiocyanate  conjugated 
goat  anti-mouse  Ig(G+M)  antibody. 
With  fluorescence  microscopy  with 
B-exciting  light,  yellowish  fluorescein 
coronas  around  the  cells  of  the  toxic 
species  were  observed,  confirming  a 
positive  reaction.  These  antibodies 
can  recognize  different  molecules 
distributed  on  the  cell  surface,  even 
when  the  organisms  have  similar 
morphological  features.  One  of  the 
molecules  distributed  on  Chattonella 
was  determined  to  be  glycoprotein 
(Nagasaki  et  al.,  1991a).  This  method 
would  help  us  to  differentiate  small 
marine  organisms  like  fish  eggs. 

Red  sea  bream  (Pagrus  major) 
(Table  1)  eggs  can  easily  be  distin- 
guished from  those  of  other  sparids 
also  found  in  Japan,  such  as  Acan- 
thopagrus  latus,  by  differences  in  egg 
size  and  spawning  seasons,  and  from 
those  of  Evynnis  japonica  by  differ- 
ences in  spawning  seasons  (Ikeda 
and  Mito,  1988;  Kinoshita,  1988; 
Hayashi.  2000).  However,  eggs  of 
some  sparids,  such  as  Aeanthopagrus 
schlegeli,  Sparus  sarba,  and  Dentex 
tumifrons  are  extremely  difficult  to 
distinguish  from  eggs  of  P.  major. 
Therefore,  we  developed  monoclonal 
antibodies  that  allow  P.  major  eggs 
to  be  clearly  identified  by  immuno- 
staining,  thus  differentiating  them 
from  other  similar  sparids. 

This  technique  may  be  a  useful 
new  tool  for  identifying  fish  eggs. 
Here,  we  report  a  method  for  identi- 
fying P.  major  eggs  using  monoclonal 
antibodies  developed  to  react  specifi- 
cally with  the  eggs. 


Materials  and  methods 

Eggs  of  P.  major  were  obtained  from 
adult  female  fish  that  had  spawned  in 


Manuscript  submitted  4  April  2003 
to  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
2  March  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:555-560  (20041. 


556 


Fishery  Bulletin  102(3) 


Table  1 

Characteristics  of  Sparidae  distributed  in  Japan. 

Egg  oil  globule 

Suborder 

Species 

Distribution 

Spawning  season 

Egg  size  (mmi 

size  (mm) 

Pagrinae 

Pagrus  major 

South  of  Hokkaido  I  Coastal) 

Mar-May 

0.90-1.03 

0.19-0.25 

Evynnis  jappon  ica 

South  of  Hokkaido  (Coastal) 

Oct-Dec 

0.89-0.98 

0.19-0.21 

Sparinae 

Acanthopagrus  schlegi 

li         South  of  Hokkaido  (Coastal) 

Mar-Jun 

0.83-0.91 

0.20-0.22 

Acanthopagrus  latus 

South  Japan  (Coastal) 

Oct-Nov 

0.76-0.81 

0.2 

Sparus sarba 

South  Japan  (Coastal) 

Apr-Jun 

0.88-0.92 

0.19-0.22 

Denticinae 

Dentex  tumifrons 

South  Japan  (Oceanic) 

May-Jun 

0.90-0.93 

0.19 

isolation  tanks  at  several  sea  farming  centers  described 
in  Table  2.  Immediately  after  collection,  fish  eggs  were 
fixed  in  a  solution  of  57c  formaldehyde  to  sea  water  solu- 
tion and  stored.  Before  use,  the  eggs  were  thoroughly 
washed  with  distilled  water  and  suspended  in  phosphate 
buffered  saline  (PBS)  solution. 

Monoclonal  antibodies  were  developed  according  to 
the  methods  of  Kdhler  and  Milstein  (1975),  Garfre  and 
Milstein  (1981),  and  Hiroishi  et  al.  (1984,  1988):  0.5  mL 
of  egg  suspension  (200  eggs/PBS  solution  from  Fukui 
Prefectural  Sea  Farming  Center,  Obama  City,  Fukui 
Prefecture)  was  mixed  with  0.5  mL  Freund's  complete 
adjuvant  (Nacalai  Tesque,  Inc.,  Kyoto,  Japan).  The  mix- 
ture were  then  injected  subcutaneously  into  BALB/c 
female  mice  (4  weeks  of  age).  The  female  mice  received 
second  and  third  injections  at  2-week  intervals.  For  the 
final  immunization,  P.  major  eggs  collected  in  the  sea 
farming  center  of  Kansai  Environmental  Engineering 
Center  Co.  (Miyazu  City,  Kyoto,  Japan)  were  injected 
into  the  mouse  after  being  emulsified  with  Freund's 
incomplete  adjuvant  (Nacalai  Tesque,  Inc.).  Three  days 
after  the  final  immunization,  the  spleens  of  the  mice 
were  removed  and  passed  through  a  mesh  (mesh  size: 
100  urn).  The  spleen  cells  obtained  by  this  procedure 
were  fused  with  the  myeloma  cell  line  X63-AG8.653 
at  a  ratio  of  10:1  with  50%  polyethylene  glycol.  After 
cell  fusion,  hybrid  cells  were  incubated  in  a  selective 
hypoxanthine-aminopterin-thymidine  medium  (Kohler, 
1979;  Garfre  and  Milstein,  1981). 

The  reactivity  of  the  antibodies  produced  by  the  hy- 
bridomas  was  then  determined.  Eggs  fixed  with  57c 
formaldehyde  in  seawater  were  washed  with  PBS  solu- 
tion in  a  96-well  plate.  Throughout  the  experiments, 
the  principal  eggs  used  were  from  the  Fukui  Prefec- 
tural Sea  Farming  Center.  Normal  horse  serum  solution 
(200  jUL),  diluted  100-fold  with  PBS,  was  added  to  the 
wells  to  prevent  any  nonspecific  reactions.  After  incuba- 
tion at  room  temperature  for  20  minutes,  the  eggs  were 
washed  with  200  pL  of  PBS.  After  removing  the  PBS, 
200  fih  of  the  hybridoma  culture  supernatant  solution 
was  added  to  the  wells  and  incubated  at  room  tempera- 
ture for  30  minutes.  After  washing  with  PBS  (100  /jL), 
biotinylated  horse  anti-mouse  IgG  (100  pL)  was  added 


to  the  wells  and  incubated  at  room  temperature  for  20 
minutes.  After  the  incubation,  VECTASTAINR  ABC  re- 
agent (avidin  DH  +  biotinylated  horseradish  peroxidase/ 
PBS,  100  jjL)  was  added  according  to  the  direction  of 
VECTASTAINR  Elite  ABC  kit  (ABC  Mouse  IgG  Kit,  Fu- 
nakoshi  Co.,  Tokyo,  Japan).  After  immunostaining  the 
eggs  were  observed  by  stereoscopic  microscopy  ( SMZ-2T, 
Nikon  Co.,  Tokyo,  Japan).  In  a  positive  reaction,  the 
surface  of  the  fish  egg  was  stained  brown  as  a  result 
of  the  oxidation  of  3,3'-diaminobenzidine  (substrate) 
by  horseradish  peroxidase  bound  to  the  egg  surface  by 
the  antibody. 

Unidentified  pelagic  fish  eggs  from  open  water  were 
collected  by  using  a  plankton  net  (MTD  net,  NGG54 
with  mesh  size  of  0.344  mm,  Rigo  Co.,  Tokyo,  Japan) 
from  Wakasa  Bay  (Fukui  Prefecture,  Japan)  in  May 
1997.  They  were  fixed  with  59c  formaldehyde  in  sea 
water,  either  immediately  or  after  incubation  in  seawa- 
ter in  finger  bowls  at  20°C  for  24  hours,  and  identified 
by  careful  observation  as  described  by  Ikeda  and  Mito 
(1988)  and  Ikeda  et  al.  (1991).  The  fixed  eggs  were 
transferred  to  net  wells  (mesh  size  200  ^m,  diameter  24 
mm,  Corning  Incorporated,  Corning,  NY)  and  washed 
with  10  mL  of  distilled  water  three  times.  Then  the 
eggs  in  the  netwells  were  immersed  in  100  mL  of  PBS 
in  a  polystyrene  tray  (Corning  Incorporated,  Corning, 
NY)  for  5  minutes.  The  egg  suspension  was  placed  into 
the  wells  of  a  six-well  plate  and  incubated  with  10  mL 
of  normal  horse  serum  solution  for  20  minutes.  After 
incubation,  the  eggs  were  incubated  with  10  mL  of  MT-1 
antibody  solution  (hybridoma  culture  supernatant)  and 
then  incubated  with  10  mL  of  biotinylated  horse  anti- 
mouse  IgG.  The  subsequent  procedure  was  performed 
as  described  above. 

The  immunoglobulin  subclass  of  monoclonal  antibod- 
ies was  determined  according  to  the  directions  of  the 
mouse  monoclonal  antibody  isotyping  kit  (Amersham 
Pharmacia  Biotech  Co.,  Uppsala,  Sweden)  as  follows: 
3  mL  of  monoclonal  antibodies  solution  (hybridoma 
supernatant  solution)  obtained  in  this  study  was  added 
to  0.3  mL  of  horseradish  peroxidase-conjugated  anti- 
mouse  IgG  in  the  kit.  An  isotyping  stick  in  the  kit 
was  incubated  with  the  above  solution  at  room  tern- 


NOTE     Hiroishi  et  al.:  Identification  of  Pagus  ma/or  eggs  using  monoclonal  antibodies 


557 


perature  for  15  minutes.  Then 
the  stick  was  washed  with  0.1% 
Tween  20/PBS,  and  incubated 
with  4-chloro-l-naphthol  solution 
(substrate  of  horseradish  peroxi- 
dase in  the  kit)  containing  0.1% 
H202  at  room  temperature  for 
15  minutes.  The  immunoglobu- 
lin subclass  of  the  monoclonal 
antibodies  was  determined  by 
observing  the  positions  of  bands 
that  appeared  on  the  stick. 


Results  and  discussion 

After  cell  fusion,  hybridomas 
were  grown  in  42  wells  of  96- 
well  plates.  Supernatant  solu- 
tions of  the  cultures  were  used 
for  the  immunostaining  assay 
to  select  hybridomas  producing 
antibodies  reactive  to  P.  major 
eggs.  After  the  assay,  positive 
reactions  were  observed  in  six 
wells.  These  hybridomas  were 
cloned  by  the  limiting  dilution 
method,  and  finally  three  clones 
producing  monoclonal  antibod- 
ies reactive  with  P.  major  were 
obtained.  Those  antibodies  were 
named  MT-1,  MT-2,  and  MT-3. 
The  subclass  of  all  antibodies 
was  IgGj.  Specificity  of  the  anti- 
bodies was  examined  by  using 
the  eggs  shown  in  Table  2.  As  a 
result,  the  antibodies  were  reac- 
tive with  all  the  P.  major  eggs 
in  both  the  early  and  late  stages 
(before  or  after  tail-bud  stage), 
but  not  with  eggs  of  other  species 
(Table  3,  Fig.  1).  Thus,  it  becomes 
possible  to  identify  P.  major  eggs. 
The  immunostaining  assay  took 
2.5  hours. 

The  oldest  eggs  of  P.  major  (20 
April,  1995)  could  react  with  the 
antibodies  obtained  as  clearly 
as  the  recently  collected  eggs 
of  P.  major,  indicating  that  egg 
samples  preserved  for  up  to  7 
years  could  be  analyzed  by  this 
method. 

The  method  was  also  success- 
ful with  102  eggs  collected  from 
Wakasa  Bay  (Table  4),  which 
had  been  immediately  fixed  with 
5%  formaldehyde  in  seawater. 
Among  them,  only  11  eggs  were 
identified  as  Callionymoidei  spp. 


P1 

CO 

m 

co 

OS 

CO 
C35 

CO 

1a 

en 

CD 

05 

CO 

en 

CD 

cn 

cn 

cn 

jn 

cn 

bo 

c 

k 

O) 

P5 

>. 

S-. 

-. 

>i 

>. 

>> 

05 

>* 

>. 

> 

3 
3 

CO 

34 

>! 

3 
CO 
>-3 

k. 

en 

>, 

CO 

s 

"o. 

a. 
< 

< 

c 

3 

c 

3 

CO 

CO 

-. 

c 

3 

CO 

-. 

CO 

co 

C 
3 

cO 

CO 

0 
2 

■-3 

jn 

4^3 

a, 
< 

CO 

0. 
< 

e 

o 

■-3 

>-5 

,_, 

^_( 

■-3 

^_, 

,_( 

cn 

!-J 

0) 

-. 

,_( 

1 

u 

0 

~ 

i-H 

CO 

CO 

CO 

cm 

<N 

CN 

CM 

CO 

CO 

CO 

CN 

^H 

co 

CM 

CO 

CN 

CN 

CM 

CO 

CO 

00 

.«-> 

4J 

bo 
bo 

01 

_ffl 

£ 

TJ 

c 

C 

— 

c 

<4_ 

o 

B 

CO 

0 

S 

; 

B 
CO 

0 

Co 

_>, 

>, 

_>. 

_>> 

_>> 

_>. 

^ 

_>> 

_>> 

B 

B 

_>> 

>. 

c 

bo 

CO 

"k 

CD 

-4-i 

"C 

CO 

4J 

c 

■~ 

■~ 

- 

^ 

CO 

4-> 

- 

T^ 

CO 
4-3 

^ 

CO 
4J 

-i! 

'- 

CO 
4-3 

01 
4-3 

k 

S£ 

CO 

crj 

CO 

CO 

CO 

CO 

CO 

co 

CO 

CO 

CO 

CO 

CO 

c 

a 

c 

CO 

CO 

cO 

CO 

C 

co 

fa 

J 

fa 

►J 

H 

fa 

fa 

fa 

W 

CO 

►J 

CO 

- 

CO 

fa 

CiJ 

fa 

C4-: 

CO 

fa 

00 

►J 

c*_: 

CO 

u 

- 

CO 

ID 

ID 

fa 

J 

<4^ 

CO 

(4 

fa 

fa 

Id 

CO 

CO 

CO 

0 

CO 

CO 

CO 

t- 

CO 

CO 

u> 

CO 

CO 

t- 

CO 

o 

CM 

O, 

0 

0 

fa 

0 

O 

fa 

a 

CO 

GO 

<d 

CO 

CO 

CO 

£ 

£ 

<4-i 

C4-. 

— 

^ 

0 

U-4 

u 

CO 

i- 

01 

O 

t- 

00 

01 

: 

5? 

o 

C 

o 

■4-3 
O 

co 

k 

PL, 

'3 

0) 

kl 

fa 
'3 

CO 

k 

- 

o 

CO 

u 

- 

0 

3 
j: 

3 
fa 

CO 

}-• 

fa 

5 

4J 

c 

CO 

0 

bo 

a 

CO 

0 

bo 

3 
3 

fa 

c 

CO 

0 

bo 

4-3 

c 

CO 

0 

bo 

3 

3 
fa 

1 

o 

3 

N 

CO 

>> 

3" 

N 
CO 
>> 

-* 

M 

- 

- 

CO 

_<: 

C 

C 

co" 

B 

c 

CO 

c*J 

<*-: 

0 

C(-4 

B 
u 

CD 

3 
[i, 

co" 

3 

fa 

co" 

0 
3 

3 
fa 

co" 

a 

CO 

a 
k> 

CO 

O 

3 
.a 

£ 

CO 

a 

CO 

CO 

0 

3 
M 

CO 

u 

fa 

0 

00 
kl 

— 

0 

0 

a 

_E 

CO 
kl 

fa 

0 

X 

g 

g 

£ 

£ 

3 

g 

fa 

fa 

3 

fa 

fa 

— 

3 

4-> 

0 

- 

CO 

3 

- 

CD 

CO 

CO 

cO 

a 

fa 

CO 

CO 

CO 

fa 

CO 

CO 

fa 

>> 

>> 

>> 

5     c 

.a    = 

"i    -a 

t—       01 

CO 

3 

k 

k 

j= 

^ 

>» 

>i 

' — ' 

fi 

CO 

CO 

1 — ' 

CO 

CO 

0 

* — ■ 

w 

w 

w 

c 
o 

01 
4-3 
C 

CO 

S3 

o 

O 

g 

g 

>> 

4J 

Z 

CO 
T3 

CO 

T3 

4-3 

CO 
13 

CO 
T3 

4J 

0 

£ 

CO 

£ 
3 

4J 

3" 

3" 

>> 

4-3 

'm 

00 
> 

'2 

3" 

4-3 
SI 

co 

_o 

bo 

a 

"3. 
6 

CO 

co 

CO 

O 

bo 
C 
k 

CO 

o 

bo 

6 

u 

0) 

c 

(V 

CJ 

u 

0) 
fc3 

c 
CO 
O 

k 

CO 

4J 

B 

CO 

0 

co 

4J 

c 

CO 

0 

CO 

CO 

_> 

E 

u 

CO 

C 

CO 

0 

a 

CO 

c 
_o 

B 

CO 

c 
_o 
4J 

CO 

u 

CO 

> 
C 
ID 

B 

CO 

s 
0 

"+3 

c 

CO 

c 
0 

4J 

00 

> 

'c 

CO 

cO 

g 

N 
CO 

m 

bo 

bo 

CO 

01 
0) 

c 
'bo 

co 

01 

c 
'5b 

bo 
C 

B 

k 

CO 

fa 

CO 
CO 
CO 

"3 

k 

3 
4J 

bo 

b 
1 

k 

CO 

fa 

CO 
0) 

en 

~. 
u 
3 
+^ 

bo 
C 

B 

bo 

c 

B 

3 

X. 

11 

3 

bo 
C 

E 

CO 

fa 

CO 
CO 
CO 

73 

kl 

3 

CO 

4-> 
CO 

CO 
CO 

CO 

CO 

CO 
CO 

3 

CO 

3 

cO 

4-3 
CO 

m 

.2 

CO 

4-3 

CO 
OQ 
CO 

0 

4-3 
O 

£ 

3 
J3 

CO 

3 

c 

4J 

C 
O 

4-3 

93 

Id 

B 

c 
0 

40 
CO 

03 

co 
fa 

c 
fa 

"3 
c 

01 

a 

c 
o 

k 

c 
w 

■3 

4J 

c 
co 

E 

c 
o 

- 

k 

CO 

fa 

CO 
CO 
VI 

- 
u 
3 

4-3 

CO 

fa 

CO 
CO 
CO 

~z 
u 
3 

4J 

of 
u 

3 

0 
'C 

bo 

- 
CO 

'/. 

fa 

kl 
3 
^-^ 
0 

CO 

X 

CO 

fa 
"3 

t- 

3 
*j 

u 

CO 

3 

0 

bO 

- 
CO 

X. 
m 

fa 

73 

3 
4-3 
u 

- 
CO 
X, 

CO 

fa 

u 

3 

4-3 

CO 

CO 

£ 
3 
US 

co" 

GO 

3 

CO 

3 

— 3 

3 

u 

bo 

us 

CO 

34 
fa 

CO 

— 
0 

cO 

*G 
0 

CO 
CO 

< 

bo 

B 

a 

34 

CO 

c^ 
O 

CO 
CO 

< 

bo 

C 

a 

cO 

2 

C*4 
O 

>3 

ki 
O 

— 3 

cO 
ki 
0 

CO 

0 

CO 

to 

< 

bo 

B 

a 

kl 
cO 

CO 
k 

fa 

'3 

3 
fa 

> 

'> 

J 

J 

O 

«2 

u 

< 

eg 

.2 

=2 

< 

=2 

c2 

£ 

< 

X 

fa 

fa 

CO 

►J 

fa 

>> 

CO 

CO 

a 

c 

0) 

01 

CO 

k 
fa 

CO 

fa 

Cm 

CO 

<D 

(D 

**-" 

CO 

CO 

< 

t*4 

-E 

03 

CO 

cb 

a 

fa 

k 
fa 

- 

O 
>> 

- 

fa 

fa 

fa 

O 

fa 

fa 

0 

O 

O 

CO 

CO 

CO 

CO 

CO 

CO 

CO 

CO 

'3 

GQ 

C 

'3 

CO 

c 

5 

5 

O 
O 

O 
O 

— 

5 

CO 
M 

CO 

CO 
CO 

— 
P 

u 

CO 
CO 

CO 
CO 

CO 

a 

— 

3 

CO 

co" 
CO 

c 

CO 

3 

CO 

k 
CO 

J3 

CO 

3 
CO 

a 

CO 

CO 

M 

CO 

CO 

CO 

3 

3 

>? 

>> 

CS 

3 

CO 

CO 

CO 

CO 

CO 

CO 

— 

CO 

•-3 

cO 

>"3 

CO 

>"3 

£ 

« 

fa 

■— 

a 

» 

ta 

fa 

O 

O 

fe 

O 

0 

S 

fa 

Z 

fa 

03 

CO 

3 

a 

CO 

5 

CO 

a 

03 

.= 

CO 

C 

e 

_o 

CD 
CO 
CO 
01 

o 
3> 

to 

k 
bo 
0 

3 

a 
-0 

!3 

CO 

"0 

CO 

1 

0 

_co 

"a 

C3 

a. 

CO 

s 

§■ 

? 

3 

CO 

-g 

^ 

CO 

X 

"~ 

3 

3 

CO 

k 

c 

c 

k. 

C3 

k. 

bo 

cs 

cs 

CS 

cr 

k 

bo 

53 

0 

co 

D. 

CO 

e 

a, 

< 

1: 

ay 

Q 

a, 

K3 

6 
c 
bo 

bo 

T—t 

CM 

CO 

■* 

in 

CD 

i> 

CO 

05 

S 

1— 1 

CN 

CO 

-0* 

m 

■r 

t- 

X 

cn 

r 

i—l 

fa 

CM 

CS 

558 


Fishery  Bulletin  102(3) 


(type  II).  The  remaining  91  unidentified  eggs  were  di- 
vided into  three  groups  (types  I,  III,  and  IV)  based  on 
diameters.  Of  these  91,  51  type-II  eggs  reacted  with 
MT-1.  This  finding  is  compatible  with  the  possibility 
that  the  eggs  were  P.  major,  because  the  size  was  simi- 
lar to  that  of  P.  major  and  each  contained  a  single  oil 
globule  of  a  similar  size  (Tables  1  and  4).  Another  43 
eggs  were  collected  from  another  area  of  Wakasa  Bay 
(Table  5).  None  of  the  eggs  fixed  just  after  collection 


Table  3 

Reactivity  of  monoclonal  antibodies  to  fish  eggs.  +  repre- 
sents positive  reaction;  -  represents  negative  reaction. 


Reactivity 


Egg  no. 


Species 


MT-1     MT-2     MT-3 


1 

Pagrus  major 

+ 

2 

+ 

3 

+ 

4 

+ 

5 

+ 

6 

+ 

7 

+ 

8 

+ 

9 

Acanthopagrus  schlegeli 

- 

10 

- 

11 

- 

12 

- 

13 

- 

14 

Acanthopagrus  latus 

- 

15 

Sparus sarba 

- 

16 

Dentex  tumifrons 

- 

17 

Paralichthys  olivaceus 

- 

18 

- 

19 

- 

20 

- 

21 

Engraulis  japonica 

- 

were  morphologically  identifiable.  But,  after  incuba- 
tion at  20°C  for  24  hours  until  the  late  stage,  all  six 
eggs  identified  as  P.  major  were  reactive  with  the  an- 
tibody MT-1,  whereas  the  others  were  not.  These  find- 
ings strongly  suggest  that  the  method  developed  in  this 
study  is  useful  for  identifying  P.  major  eggs  in  seawater. 
Although  only  late  stage  eggs  were  used  in  this  experi- 
ment, early  stage  eggs  are  also  detectable  because  the 
antibody  recognized  both  stages  of  P.  major  eggs  from 
several  sea  farming  centers  (Table  2). 

Compared  to  genetic  analysis  of  fish  eggs,  this  method 
has  the  advantage  of  being  able  to  assay  many  eggs 
simultaneously  without  the  need  to  separate  individual 
eggs  in  tubes  and  without  extracting  DNA  from  the  in- 
dividual egg  in  each  tube.  Further,  this  method  works 
with  formalin-fixed  eggs,  whereas  extraction  of  DNA 
from  formalin-fixed  material  is  problematic.  Plankton 
samples  from  field  studies  are  typically  fixed  in  forma- 
lin-seawater  solution. 

There  was  no  problem  obtaining  a  large  amount  of 
the  monoclonal  antibody  required  when  identifying  P. 
major  eggs.  The  antibody  can  be  easily  obtained  by 
large-scale  cultures  of  hybridoma  cells.  About  50  mL 
of  antibody  solution  was  obtained  after  two  weeks  of 
cultivation.  There  was  no  technical  problem  assaying  43 
or  102  eggs  from  natural  waters.  However,  one  assay  of 
a  field  sample  cost  about  20  U.S.  dollars.  To  keep  costs 
down  an  assay  kit  cheaper  than  the  VECTASTAINR 
Elite  ABC  kit  is  needed  when  a  large  number  of  field 
samples  are  analyzed. 


Acknowledgments 

We  would  like  to  thank  the  following  sea  farming  centers 
and  universities  for  providing  the  fish  eggs  used  in  this 
study:  Fukui  Prefectural  Sea  Farming  Center;  Kyoto 
Prefectural  Sea  Farming  Center;  Faculty  of  Agriculture, 
Kyushu  University;  Osaka  Prefectural  Fisheries  Station; 
Sea  Farming  Center  of  the  Japan  Sea-Farming  Associa- 
tion; Fisheries  Laboratory  of  Kinki  University.  We  also 
thank  Jeffrey  M.  Leis,  Australian  Museum,  Sydney, 
Australia,  for  his  kind  advice  during  the  writing  of  this 
manuscript. 


Table  4 

Reactivity  of  monoclonal  antibody  MT-1  to  the  pelagic  eggs  fixed  with  formaldehyde  just  after  collection  from  Wakasa  Bay. 
O.G.  diameter  =  oil  globule  diameter 


Fish  egg  type 


Egg  diameter  imm) 


O.G.  diameter  imm) 


Reactivity  (%) 
( positive  egg  no./  total  egg  no.) 


I 
II 

III 
IV 


0.72-0.79 
0.75-0.82 
0.81-1.02 
1.07 


0.16-0.19 
no  oil  globule 
0.19-0.28 
0.21 


0(0/2) 

0(0/11) 

58(51/88) 

0(0/1) 


NOTE     Hiroishi  et  al.:  Identification  of  Pagus  ma/or  eggs  using  monoclonal  antibodies 


559 


Figure  1 

Reactivity  of  monoclonal  antibody  MT-1  to  fish  eggs  detected  by  immunos- 
taining.  (A)  Pagrus  major  (positive  reaction);  (B)  Pagrus  major  (negative 
control  reaction  without  primary  antibody);  (C)  Acanthopagrus  shlegeli; 
(D)  Sparus  sarba;  (E)  ParaUchthys  olivaceus;  (F)  Engraulis  japonica. 
Bar  represents  1  mm.  Only  the  P.  major  eggs  stained  brown  and  showed 
a  positive  reaction. 


Table  5 

Reactivity  of  monoclonal  antibody  MT-1  to  the  pel 

agic  eggs  reared  for  24  hours 

after  collection  from  Wakasa  Bay. 

Egg 

Species 

Reactivity  (%)  (positive  egg  no. /total  egg  no.) 

Fish 

Pagrus  major 
Acanthopagrus  shlegeli 
ParaUchthys  olivaceus 
Triglidae  sp. 
Konosirus  punctutus 
Soleoidei  sp. 
Englauris  japonicus 

100(6/6) 
0(0/8) 
0(0/1) 
0(0/1) 
0(0/2) 
0(0/7) 
0(0/13) 

Decapod 

Enploteuthidae  sp. 

0(0/5) 

560 


Fishery  Bulletin  102(3) 


Literature  cited 

Daniel  III,  L.  B„  and  J.  E.  Graves. 

1994.     Morphometric  and  genetic  identification  of  eggs 
of  spring-spawning  sciaenids  in  lower  Chesapeake 
Bay.     Fish.  Bull.  92:254-261. 
Garfre,  G.,  and  C.  Milstein. 

1981.     Preparation  of  monoclonal  antibodies:  strategies 
and  procedures.     In  Methods  in  enzymology  (J.  J.  Lan- 
gone  and  H.  V.  Vunakis,  eds.),  vol.  73,  p. 3-46.     Academic 
press,  New  York,  NY. 
Graves,  J.  E.  M.  J.  Curtis,  P.  S.  Oeth,  and  R.  S.  Waples. 

1989.     Biochemical   genetics   of  southern   California 
basses  of  the  genus  Paralabrax:  specific  identification 
of  fresh  and  ethanol-preserved  individual  eggs  and  early 
larvae.     Fish.  Bull.  88:59-66. 
Hayashi,  K. 

2000.     Sparidae.     In    Fishes   of  Japan   with   pictorial 
keys  to  the  species  (second  ed.),  T.  Nakabo  (ed.),  p. 
857-858.     Tokai  University  Press,  Tokyo.     [In  Japanese.] 
Hiroishi,  S.,  S.  Matsuyama,  T.  Kaneko,  Y  Nishimura, 
and  J.  Arita. 

1984.     Inhibition  of  cytotoxicity  for  screening  a  monoclo- 
nal antibody  to  HLA  antigen.  Preparation  of  a  highly 
specific  monoclonal  antibody  to  HLA  antigen.     Tissue 
Antigens  24:307-312. 
Hiroishi,  S.,  R.  Nakai,  H.  Seto,,  T  Yoshida,  and  I.  Imai. 

2002.     Identification  of  Heterocapsa  circularisquama  by 
means  of  antibody.     Fisheries  Sci.  68:627-628. 
Hiroishi,  S.,  A.  Uchida,  K.  Nagasaki,  and  Y.  Ishida. 

1988.     A  new  method  for  identification  of  inter-  and 
intra-species  of  the  red  tide  algae  Chattonella  antiqua 
and  Chattonella  marina  (Raphidophyceae)  by  means  of 
monoclonal  antibodies.     J.  Phycol.  24:442-443. 
Ikeda,  T,  S.  Chuma,  and  M.  Okiyama. 

1991.     Identification  of  pelagic  eggs  of  marine  fishes  by 
rearing  method.     Jap.  J.  Ichthyol.  38:199-206.     [In 
Japanese.] 
Ikeda,  T,  and  S.  Mito. 

1988.     Identification  of  eggs  and  hatched  larvae.     In  An 
atlas  of  the  early  stage  fishes  in  Japan  (M.  Okiyama, 
ed.),   p.  999-1119.     Tokai  Univ.   Press,  Tokyo.     [In 
Japanese.] 
Kinoshita,  I. 

1988.     Sparidae.     In  An  atlas  of  the  early  stage  fishes  in 


Japan  (M.  Okiyama,  ed.),  p.  527-536.     Tokai  University 
Press,  Tokyo.     ]In  Japanese.] 
Kohler,  G 

1979.     Fusion  of  lymphocytes.     In  Immunological  methods 
(I.  Lefkovits  and  B.  Pernis,  eds.),  p.  391-395.     Academic 
press,  New  York. 
Kohler,  G.,  and  C.  Milstein. 

1975.     Continuous  cultures  of  fused  cells  secreting  anti- 
body of  predefined  specificity.     Nature  (Lond.)  256: 
495-497. 
Kondo,  R.,  G.  Kagiya,  Y  Yuki,  S.  Hiroishi,  sand  M.  Watanabe. 
1998.     Taxonomy  of  a  bloom-forming  cyanobacterial  genus 
Microcystis.     Nippon  Suisan  Gakkaishi  64:291-292.     [In 
Japanese.] 
Mito,  S. 

1960.     Keys  to  the  pelagic  fish  eggs  and  hatched  larvae 

found  in  the  adjacent  waters  of  Japan.     The  Science 

Bulletin  of  the  Faculty  of  Kyushu  University  18:71-94, 

pis.  2-17.     [In  Japanese.] 

1979.     Fish  egg.     Kaiyokagaku  11:126-130.     [In  Japanese.] 

Nagasaki,  K,  A.  Uchida,  S.  Hiroishi,  and  Y.  Ishida. 

1991a.     An  epitope  recognized  by  the  monoclonal  antibody 
MR-21  which  is  reactive  with  the  cell  surface  of  Chat- 
tonella marina  type  II.     Fish.  Sci.  57:885-890. 
Nagasaki,  K,  A.  Uchida,  and  Y.  Ishida. 

1991b.     A  monoclonal  antibody  which  recognizes  the  cell  sur- 
face of  red  tide  alga  Gymnodinium  nagasakiense.     Fish. 
Sci.  57:1211-1214. 
Sako,  Y,  M.  Adachi,  Y  Ishida,  C.  Scholin,  and  D.  M.  Anderson. 
1993.     Preparation  and  characterization  of  monoclonal 
antibodies  to  Alexandrium  species.     In  Toxic  phytoplank- 
ton  blooms  in  the  sea  (T.  J.  Smayda  and  Y.  Shimizu, 
eds.)  p.  87-93.     Elsevier,  New  York,  NY. 
Shao,  K.-T,  K.-C.  Chen,  and  J.-H.  Wu. 

2002.     Identification  of  marine  fish  eggs  in  Taiwan  using 
light  microscopy,  scanning  electric  microscopy  and  mt 
DNA  sequencing.     Mar.  Freshw.  Res.  53:355-365. 
Vrieling,  E.,  A.  Draaijer,  L.  Van  Zeiljl,  W.  Gieskes,  and 
M.  Veenhuis. 

1993.  The  effect  of  labeling  intensity,  estimated  by 
realtime  confocal  laser  scanning  microscopy,  on  flow 
cytometric  appearance  and  identification  of  immuno- 
chemical labeled  marine  dinoflagellates.  J.  Phycol. 
29:180-188. 


Fishery  Bulletin  102(3) 


561 


Superintendent  of  Documents  Publications  Order  Form 
*5178 


YES,  please  send  me  the  following  publications: 


Subscriptions  to  Fishery  Bulletin 
for  $55.00  per  year  ($68.75  foreign) 


The  total  cost  of  my  order  is  $  . 


.  Prices  include  regular  domestic 


postage  and  handling  and  are  subject  to  change. 


(Company  or  Personal  Name) 


( Please  type  or  print ) 


(Additional  address/attention  line i 


( Street  address ) 


(City,  State,  ZIP  Code) 


(Daytime  phone  including  area  code) 


( Purchase  Order  No. 


Charge 
your 
order. 


■MasterCard 


Please  Choose  Method  of  Payment: 

]  Check  Payable  to  the  Superintendent  of  Documents 


]  GPO  Deposit  Account 

h 

]  VISA  or  MasterCard  Account 

I    (Credit  card  expiration  date) 

To  fax 

your  orders 

(202)  512-2250 


(Authorizing  Signature) 

Mail  To:    Superintendent  of  Documents 

P.O.  Box  371954,  Pittsburgh,  PA  15250-7954 


Thank  you  for 
your  order! 


WH    i 


Fishery  Bulletin 

Guidelines  for  contributors 


Content  of  papers 

Articles 

Articles  arc  reports  of  10  to  30  pages  (double 
spaced)  that  describe  original  research  in  one  or 
a  combination  of  the  following  fields  of  marine 
science:  taxonomy,  biology,  genetics,  mathematics 
(including  modeling),  statistics,  engineering,  eco- 
nomics, and  ecology. 

Notes 

Notes  are  reports  of  5  to  10  pages  without  an 
abstract  that  describe  methods  and  results  not 
supported  by  a  large  body  of  data.  Although  all 
contributions  are  subject  to  peer  review,  responsi- 
bility for  the  contents  of  articles  and  note 
upon  the  authors  and  not  upon  the  editor  or  the 
publisher.  It  is  therefore  important  that  authors 
consider  the  contents  of  their  manuscripts  care- 
fully Submission  of  an  article  is  un-derstood  to 
imply  that  the  article  is  original  and  is  not  being 
considered  for  publication  elsewhere.  Manuscripts 
must  be  written  in  English.  Authors  whose  native 
language  is  not  English  are  strongly  advised  to 
have  their  manuscripts  checked  for  fluency  by 
English-speaking  colleagues  prior  to  submission. 

Preparation  of  papers 

Text 

Title  page  should  include  authors'  full  nami 
mailing  addresses  (street   address  required)  and 

telephone,  fax  riumbei 
address  alistofkeywi  ribethe 

nuscript.     Abstract  must 
thanom 

if   emphasize  the 

culated  by  abstract! 
tanl  that  tin 

!        General  text  double- 

iefintroductii 

I  methods   R 

within  be  short,  refl 

sequence,  and  follow  the  rules  of  mu 

in  be  no  subdi  hout  at 

!  he  entire  text  should  be 

d  all  lesser-known  technic  ould  be 

une  they  are   mentioned.  The 
scientific  species  must  he  written 

first  time  tin  . 

of  scientific  names  may  be  abbn  ow  Sci- 

ld  format:  CBE  manual  foi  ..' 
editors,  and  publishers  (6th  ed.)  for  editorial  style 
and  the  most  current  issue  of  the  American  Fish- 
common    and 

■In  for 
fish  nomenclature.  Dates  should  In-  written  as  fol- 
lows: 11  November  1991.  Measurements  should  be 
expressed  in  metric  units,  e.g.  metric  tons  it).  The 
numeral  one  1 1 1  should  be  typed  as  a  one,  not  as  a 
lower-case  el  'I  I. 

Footnotes 

Use  footnotes  to  add  editorial  comments  regarding 
claims  made  in  the  text  and  to  document  unpub- 


lished works  or  works  with  local  circulation.  Foot- 
notes should  be  numbered  with  Arabic  numerals 
and  inserted  in  10-point  font  at  the  bottom 
first  page  on  which  they  are  cited.  Footnotes  should 
be  formatted  in  the  same  manner  as  citations 
If  a  manuscript  is  unpublished,  in  the  \ 
of  review,  or  if  the  information  provided  in  the 
footnote  has  been  conveyed  verbally,  please  state 
this  information  as  "unpuhl.  data."  "manuscript 
in  review,"  and  "personal  commun.,"  respt 
Authors  are  adi  ised  wherever  possible  to  avoid  ref- 
erences to  nonstandard  lib  n  published  lit- 
erature that  is  difficult  to  obtain,  such  as  internal 
reports,  processed  reports,  administrative  reports, 
ICES  council  minutes.  IWC  minutes  or  working 
papers,  any  "research"  or  "working"  documents. 
laboratory  reports,  contract  reports,  and  manu- 
scripts in  review).  If  these  references  are  used, 
please  indicate  whether  they  are  available  from 
NTIS  (National  Technical  Information  See 
from  some  other  public  depi  isitor]  Fi  mtnote  format: 
author  (last  name,  followed  bj  first-name  initials': 
year:  title  of  report  or  manuscript;  type  of  report 
and  its  administrai  i  number;  name  and 
address  i  n  institution  where  the  n 
filed. 

Literature  cited 

The  literature  cited  sect  ii  ks  that 

or  pub- 
lication   '  '.Mu  | 

iks.  Follow  tl 

th  and 
Jones  (1977  nation  takes 

the  form  of  p  Smith 

and  J.i  Mure  cited  si 

list  citations  alpha  I 

For  example,  Alston,  1! 
Smith     I 

lould  conform 

for  the  BIOSl 

all  citai  mi  format: 

title  of  repi  t  of  the 

journal  in  which  lit  published,  volume 

provide 

publish' 

Tables 

Tables  in  size  and  m 

.    numerical  order  in  the  text.  Headings  in 
ll  ample  enough  to 

All  unusual 
symbols  must  he  explained  m  the  table  legend. 
Other  incidental  com]  ntnoted  (use 

iks  onlj  in  in"  lability  in  stal 

data.  Place  table!  as  the 

table  data.  We  accept  tabli  spread- 

sheet software  programs  (e.g.   Microsoft  Excel' 
Please  note  the  following: 

•  tTse  a  comma  in  numbers  of  five  digits  or  more 
(e.g.  13,000  but  3000). 

•  Use  zeros  before  all  decimal  points  for  values 
less  than  one  (e.g.  0.3] 

Figures 

Figures  include  line  illustrations,  computer-gener- 
ated line  graphs,  and  photographs  ( or  slides  I.  They 


'1  in  numerical  order  in  the  text    Line 
tire  best  submitted  as  original  draw- 
iler-generated  line  graphs  should  be 
printed  on  laser-quality  paper.  Photographs  should 
be  submitted  on  glossy  paper  with  gi 
All  figures  are  to  be  labeled  with  sei 
name  and  the  number  of  the  figure  (e.g    Smith, 
Fig.  -4 1    Use  Helvetica  or  Arial  font  to  label  ana- 
tomical parts  iline  draw  in  graphs) 
within  figures;  use  Times  Roman  bold  fonl  in  label 
the  dil                     ions  of  a  figure  (e.g.  A,  B,  Cl. 
Figure  legends  should  explain  all 
abbreviations  seen  within  the  figure  and  should  be 
typed  m  double-spaced  format  on  a 
at  the  end  of  the  manuscript,  We  advise  authors  to 
peruse  a  recent  issue  of  /                              for  stan- 
dard formats.  Please  note  the 

•  Capitalize  the  first  letter  off  ord  of 

abels. 

•  Do  not  use  overly  large  font  sizes  to  label  axes 
or  parts  within  figures. 

•  Do  not  use  boldface  fon 
•Do 

•Do  I.  sontal  lines 

•  Do  mu  use  large 
longitude  and  latitude  on  mi 

•Indii  urn  of  degree  and 

latitu  ■ 

to  mark 
of  bar  graphs 

Copyright  law 

reprod 
work,  i 

Submission  of  papers 

iscript. 
hen  the 
lication. 

Dr.  Norman  Bartoo 

Marine  Fisherie- 

'037 

Once  the  manuscript  has  been  publi- 

sked  to  subi 
manuscript.  The  mid  be 

submiii.  irdPerfect   or   Word   Ion; 

Word,  save  as  Rich  Text  Format  I,  Plei 
we  do  not  accept  ASCII 
I YK  files. 

Reprints 

Copies  of  published  articles  and  notes  are 
able  free  of  charge  to  the  senior  author  (50  copies) 
and  to  his  or  her  laboratory  (50  copies).  Addiliona! 
copies  may  be  purchased  in  lots  of  100  when  tin 
author  receives  page  proofs. 


tftf^oi 


U.S.  Department 
of  Commerce 

Volume  102 
Number  4 
October  2004 


— 

c 

o 

■~-  3 

S  «> 

«• 

&  c 

«■ 

2o 
°  S. 

■«=»■ 
C3 
CD 

-> 

CSJ 

O* 

M 

H- 

CJ 

■ 

o 

—■» 

Fishery 
Bulletin 


U.S.  Department 
of  Commerce 

Donald  L.  Evans 
Secretary 


National  Oceanic 
and  Atmospheric 
Administration 

Vice  Admiral 

Conrad  C.  Lautenbacher  Jr., 

USN  (ret.) 

Under  Secretary  for 
Oceans  and  Atmosphere 


National  Marine 
Fisheries  Service 

William  T.  Hogarth 
Assistant  Administrator 
for  Fisheries 


Sr4T£S  0* 


The  Fishery  Bulletin  I  ISSN  0090-0656) 
is  published  quarterly  by  the  Scientific 
Publications  Office,  National  Marine  Fish- 
eries Service,  NOAA,  7600  Sand  Point  Way 
NE,  BIN  C15700,  Seattle.  WA98115-0070. 
Periodicals  postage  is  paid  at  Seattle.  WA, 
and  at  additional  mailil  POST- 

MASTER: Send  address  changes  tor  sub- 
<ns  to  Fishery  Bulletin.  Superin- 
tendent of  Documents.  Aim.:  Chief,  Mail 
List  Branch.  Mail  Stop  SSOM.  Washing- 
ton. DC  20402-9373. 

Although  the  contents  of  this  publica- 
tion have  not  been  copyrighted  and  may 
be  reprinted  entiri  Ij  reference  ti>  source 
is  appreciated. 

The  Secretary  ol  e  has  deter- 

mined  that    the  publication  of  this    pi  ri 
odical  is  necessary  according  to  law  for 
the  tr:i:  of  this 

f  funds  lor  printing  of 

this  periodical  has  been  approved  h\  the 
i  ol  the  Office  of  Management  and 
Budget. 

For  sale  by  the  Superintendent  of 
Documi  ment    Printing 

Office,  Washington,  DC  20402.  Sub 
lion  price  per  year:  $55.00  domes!  i 

foreign.     I  ingle    issue: 

$28.00  domestic  and  $35.00  foreign.  See 
back  for  order  form. 


Scientific  Editor 

Norman  Bartoo,  PhD 

Associate  Editor 

Sarah  Shoffler 

National  Marine  Fisheries  Service,  NOAA 
8604  La  Jolla  Shores  Drive 
La  Jolla,  California  92037 

Managing  Editor 

Sharyn  Matriotti 

National  Marine  Fisheries  Service 
Scientific  Publications  Office 
7600  Sand  Point  Way  NE,  BIN  C15700 
Seattle,  Washington   981 15-0070 


Editorial  Committee 

Harlyn  O.  Halvorson,  PhD 
Ronald  W.  Hardy,  PhD 
Richard  D.  Methot,  PhD 
Theodore  W.  Pietsch,  PhD 
Joseph  E.  Powers,  PhD 
Harald  Rosenthal,  PhD 
Fredric  M.  Serchuk,  PhD 
George  Waiters,  PhD 


University  of  Massachusetts,  Boston 
University  of  Idaho,  Hagerman 
National  Marine  Fisheries  Service 
University  of  Washington,  Seattle 
National  Marine  Fisheries  Service 
Universitat  Kiel,  Germany 
National  Marine  Fisheries  Service 
National  Marine  Fisheries  Service 


Fishery  Bulletin  web  site:  www.fishbull.noaa.gov 


The  Fishery  Bulletin  carries  original  research  reports  and  technical  notes  on  investigations  in 
fishery  science,  engineering,  and  economics.  It  began  as  the  Bulletin  of  the  United  States  Fish 
Commission  in  1881;  it  became  the  Bulletin  of  the  Bureau  of  Fisheries  in  1904  and  the  Fishery 
Bulletin  of  the  Fish  and  Wildlife  Service  in  194  1  Separates  were  issued  as  documents  tl 
volume  46;  the  last  document  was  No.  1103.  Beginning  with  volume  -17  m  1931  and  continuing 
through  volume  62  in  1963,  each  separate  appeared  as  a  numbered  bulletin  A  new  system 
began  in  1963  with  volume  63  in  which  papers  are  bound  together  in  a  single  issue  of  the 
bulletin.  Beginning  with  volume  70,  number  1.  January  1972,  the  Fishery  Bulletin  became  a 
periodical,  issued  quarterly  In  this  form,  it  is  available  by  subscription  from  the  Superintendent 
of  Documents,  U.S.  Government  Printing  Office,  Washington,  DC  20402.  It  is  also  available  free  in 
limited  numbers  to  libraries,  research  institutions,  State  and  Federal  agencies,  and  m  exchange 
for  other  scientific  publications. 


U.S.  Department 
of  Commerce 

Seattle,  Washington 

Volume  102 
Number  4 
October  2004 


Fishery 
Bulletin 


Contents 


Articles 


563-580  Calambokidis,  John,  Gretchen  H.  Steiger,  David  K.  Ellifrit, 

Barry  L.  Troutman,  and  C.  Edward  Bowlby 
Distribution  and  abundance  of  humpback  whales 
(Megaptera  novaeangliae)  and  other  marine  mammals 
off  the  northern  Washington  coast 


581-592  Danilewicz,  Daniel,  Juan  A.  Claver,  Alejo  L.  Perez  Carrera, 

Eduardo  R.  Secchi,  and  Nelson  F.  Fontoura 
Reproductive  biology  of  male  franciscanas 
(Pontoporia  bloinvillei)  (Mammalia:  Cetacea)  from 
Rio  Grande  do  Sul,  southern  Brazil 


593-603  Fischer,  Andrew  J.,  M.  Scott  Baker  Jr.,  and  Charles  A.  Wilson 

Red  snapper  (Lut/anus  campechanus)  demographic  structure  in 
the  northern  Gulf  of  Mexico  based  on  spatial  patterns 
in  growth  rates  and  morphometries 


The  conclusions  and  opinions  expressed 
in  Fishery  Bulletin  are  solely  those  of  the 
authors  and  do  not  represent  the  official 
position  of  the  National  Marine  Fisher- 
ies Service  'NOAAi  or  any  other  agency 
or  institution. 

The  National  Marine  Fisheries  Service 
NMFS  i  does  not  approve,  recommend,  or 
endorse  any  proprietary  product  or  pro- 
prietary material  mentioned  in  this  pub- 
lication- No  reference  shall  be  made  to 
NMFS,  or  to  this  publication  furnished  by 
NMFS.  in  any  advertising  or  sales  pro- 
motion which  would  indicate  or  imply 
that  NMFS  approves,  recommends,  or 
endorses  any  proprietary  product  or  pro- 
prietary material  mentioned  herein,  or 
which  has  as  its  purpose  an  intent  to 
cause  directly  or  indirectly  the  advertised 
product  to  be  used  or  purchased  because 
of  this  NMFS  publication. 


604-616  FitzGerald,  Jennifer  L,  Simon  R.  Thorrold,  Kevin  M.  Bailey, 

Annette  L.  Brown,  and  Kenneth  P.  Severin 
Elemental  signatures  in  otoliths  of  larval  walleye  pollock 
(Theragra  chalcogramma)  from  the  northeast  Pacific  Ocean 


617-633  Gaughan,  Daniel  J.,  Timothy  I.  Leary,  Ronald  W.  Mitchel 

and  Ian  W.  Wright 

A  sudden  collapse  in  distribution  of  Pacific  sardine 
(Sardinops  sagax)  off  southwestern  Australia  enables 
an  obiective  re-assessment  of  biomass  estimates 


634-647  Griffiths,  Shane  P.,  Ron  J.  West,  Andy  R.  Davis, 

and  Ken  G.  Russell 

Fish  recolonization  in  temperate  Australian  rockpools: 
a  quantitative  experimental  approach 


Fishery  Bulletin  102(4) 


648-660  Hesp,  S.  Alexander,  Ian  C.  Potter,  and  Sonja  R.  M.  Schubert 

Factors  influencing  the  timing  and  frequency  of  spawning  and  fecundity  of  the  goldlined  seabream 
(Rhabdosargus  sarba)  (Spandae)  in  the  lower  reaches  of  an  estuary 

661-670  Maxwell,  Michael  R.,  Annette  Henry,  Christopher  D.  Elvidge,  Jeffrey  Safran,  Vinita  R.  Hobson, 

Ingrid  Nelson,  Benjamin  T.  Tuttle,  John  B.  Dietz,  and  John  R.  Hunter 

Fishery  dynamics  of  the  California  market  squid  (Loligo  opalescens),  as  measured  by  satellite  remote  sensing 

671-681  Murray,  Kimberly  T. 

Magnitude  and  distribution  of  sea  turtle  bycatch  in  the  sea  scallop  (Placopecten  magellamcus)  dredge  fishery 
in  two  areas  of  the  northwestern  Atlantic  Ocean,  2001-2002 

682-692  Snover,  Melissa  L,  and  Aleta  A.  Hohn 

Validation  and  interpretation  of  annual  skeletal  marks  in  loggerhead  (Caretta  caretta)  and 
Kemp's  ridley  (Lepidochelys  kempii)  sea  turtles 

693-710  Stehlik,  Linda  L,  Robert  A.  Pikanowski,  and  Donald  G.  McMillan 

The  Hudson-Raritan  Estuary  as  a  crossroads  for  distribution  of  blue  (Callmectes  sapidus), 
lady  (Ovalipes  ocellatus),  and  Atlantic  rock  (Cancer  irroratus)  crabs 

711-722  Stevens,  Melissa  M.,  Allen  H.  Andrews,  Gregor  M.  Cailliet,  Kenneth  H.  Coale,  and  Craig  C.  Lundstrom 

Radiometric  validation  of  age,  growth,  and  longevity  for  the  blackgill  rockfish  (Sebastes  melanostomus) 

723-732  Tolan,  James  M.,  and  David  A.  Newstead 

Descriptions  of  larval,  preiuvenile,  and  |uvenile  finescale  menhaden  {Brevoortia  gunteri)  (family  Clupeidae), 
and  comparisons  to  gulf  menhaden  (6.  patronus) 

733-739  Uchikawa,  Kazuhisa,  John  R.  Bower,  Yasuko  Sato,  and  Yasunori  Sakurai 

Diet  of  the  minimal  armhook  squid  (Berryteuthis  anonychus)  (Cephalopoda:  Gonatidae)  in  the  northeast 
Pacific  during  spring 

740-749  Weinberg,  Kenneth  L,  Robert  S.  Otto,  and  David  A.  Somerton 

Capture  probability  of  a  survey  trawl  for  red  king  crab  (Paralithodes  camtschaticus) 

Notes 

750-756  Kerstetter,  David  W.,  Jeffery  J.  Polovina,  and  John  E.  Graves 

Evidence  of  shark  predation  and  scavenging  on  fishes  equipped  with  pop-up  satellite  archival  tags 

757-759  Vladimir  V.  Laptikhovsky 

Survival  rates  of  rays  discarded  by  the  bottom  trawl  squid  fishery  off  the  Falkland  Islands 


760  Acknowledgment  of  2004  reviewers 

761  2004  indexes 
770  Subscription  form 


563 


Abstract— We  examined  the  summer 
distribution  of  marine  mammals 
off  the  northern  Washington  coast 
based  on  six  ship  transect  surveys 
conducted  between  1995  and  2002, 
primarily  from  the  NOAA  ship 
McArthur.  Additionally,  small  boat 
surveys  were  conducted  in  the  same 
region  between  1989  and  2002  to 
gather  photographic  identification 
data  on  humpback  whales  iMegap- 
tera  novaeangliae)  and  killer  whales 
(Orcinus  orca)  to  examine  movements 
and  population  structure.  In  the  six 
years  of  ship  survey  effort.  706  sight- 
ings of  15  marine  mammal  species 
were  made.  Humpback  whales  were 
the  most  common  large  cetacean  spe- 
cies and  were  seen  every  year  and  a 
total  of  232  sightings  of  402  animals 
were  recorded  during  ship  surveys. 
Highest  numbers  were  observed  in 
2002,  when  there  were  79  sightings  of 
139  whales.  Line-transect  estimates 
for  humpback  whales  indicated  that 
about  100  humpback  whales  inhab- 
ited these  waters  each  year  between 
1995  and  2000;  in  2002,  however,  the 
estimate  was  562  (CV=0.21)  whales. 
A  total  of  191  unique  individuals  were 
identified  photographically  and  mark- 
recapture  estimates  also  indicated 
that  the  number  of  animals  increased 
from  under  100  to  over  200  from  1995 
to  2002.  There  was  only  limited  inter- 
change of  humpback  whales  between 
this  area  and  feeding  areas  off  Oregon 
and  California.  Killer  whales  were 
also  seen  on  every  ship  survey  and 
represented  all  known  ecotypes  of  the 
Pacific  Northwest,  including  southern 
and  northern  residents,  transients, 
and  offshore-type  killer  whales.  Dall's 
porpoise  iPhocoenoides  dalli)  were  the 
most  frequently  sighted  small  ceta- 
cean; abundance  was  estimated  at 
181-291  individuals,  except  for  2002 
when  we  observed  dramatically  higher 
numbers  (876,  CV=0.30i.  Northern 
fur  seals  (Callorhinus  ursinus)  and 
elephant  seals  iMirounga  angustiros- 
tris)  were  the  most  common  pinnipeds 
observed.  There  were  clear  habitat 
differences  related  to  distance  off- 
shore and  water  depth  for  different 
species. 


Manuscript  submitted  25  September  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
4  June  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:563-580  i2004). 


Distribution  and  abundance  of  humpback  whales 

(Megoptera  novaeangliae) 

and  other  marine  mammals 

off  the  northern  Washington  coast 

John  Calambokidis 

Gretchen  H.  Steiger 

David  K.  Ellifrit 

Cascadia  Research  Collective 

Waterstreet  Building 

218V2  West  Fourth  Ave. 

Olympia,  Washington  98501 

E-mail  address  (for  J  Calambokidis)  calambokidis@cascadiaresearch.org 

Barry  L.  Troutman 

Washington  Dept  of  Fish  and  Wildlife 

600  Capitol  Way 

Olympia,  Washington  98501 

C.  Edward  Bowlby 

Olympic  Coast  National  Marine  Sanctuary,  NOAA 
115  Railroad  Ave  E,  Suite  301 
Port  Angeles,  Washington  98362 


Marine  mammals  have  had  an  impor- 
tant role  in  the  history  of  the  Olympic 
Peninsula  for  centuries.  Many  species, 
including  sea  otters  ( En  hydra  lutris). 
harbor  seals  iPhoca  vitulina),  hump- 
back whales  (Megaptera  novaean- 
gliae), and  gray  whales  iEschrichtius 
robustus)  were  hunted  by  the  Makah 
tribe  (Swan,  1868;  Huelsbeck,  1988). 
Much  later,  modern  whalers  targeted 
humpback  whales  in  this  region  from 
stations  at  Bay  City,  Washington 
(1911-25,  Scheffer  and  Slipp,  1948). 
and  southern  Vancouver  Island,  Brit- 
ish Columbia  (1905-43,  Gregr  et  al., 
2000).  A  small  aboriginal  hunt  for 
gray  whales  resumed  in  these  waters 
in  1998,  and  the  Makah  killed  one 
gray  whale  in  May  1999.  Since  the 
end  of  commercial  whaling,  marine 
mammals  have  been  afforded  protec- 
tion under  the  Marine  Mammal  Pro- 
tection Act  of  1972.  In  addition,  the 
waters  off  the  northern  Washington 
coast  were  designated  as  the  Olympic 
Coast  National  Marine  Sanctuary  in 
1994. 


A  number  of  studies  have  docu- 
mented marine  mammals  in  this  re- 
gion. Some  surveys  of  broader  areas 
have  included  the  waters  off  north- 
ern Washington  (Von  Saunder  and 
Barlow,  1999;  Brueggeman1;  Green 
et  al.2).  Species-specific  studies  also 


1  Brueggeman,  J.  J.  1992.  Oregon  and 
Washington  marine  mammal  and  sea- 
bird  surveys.  Final  report  of  OCS 
Study  MMS"  91-0093  by  Ebasco  Envi- 
ronmental, Bellevue,  Washington,  and 
Ecological  Consulting,  Inc.,  Portland. 
Oregon,  for  the  Minerals  Management 
Service  (MMS),  445  p.  MMS,  Pacific 
OCS  Region,  U.S.  Dept.  of  Interior. 
770  Paseo  Camarillo,  Camarillo.  CA 
93010. 

2  Green,  G.  A.,  M.  A.  Smultea,  C.  E.  Bowlby. 
and  R.  A.  Rowlett.  1993.  Delphinid 
aerial  surveys  in  Oregon  and  Washing- 
ton offshore  waters.  Final  report  for 
contract  50ABNF200058  to  the  National 
Marine  Mammal  Laboratory,  National 
Marine  Fisheries  Service,  100  p.  Nat. 
Mar.  Mamm.  Lab.,  NMFS,  7600  Sand 
Point  Way  NE  F/AKC3,  Seattle,  WA 
98115.1 


564 


Fishery  Bulletin  102(4) 


have  been  conducted  on  harbor  porpoise  iPhocoena  pho- 
coena;  Barlow  et  al.,  1988;  Osmek  et  al.,  1996;  Calam- 
bokidis  et  al.3)  and,  to  a  limited  degree,  on  humpback 
whales  ( Calambokidis  et  al.,  1996,  2000)  and  gray 
whales  (Darling,  1984;  Green  et  al.,  1995;  Shelden  et 
al.,  2000;  Calambokidis  et  al.,  2002).  Studies  on  pin- 
nipeds and  sea  otters  have  also  been  conducted  in  this 
region  (Jeffries  et  al.,  2003;  Jameson  et  al.,  1982,  1986; 
Kvitek  et  al.  1992,  1998;  Bowlby  et  al.4). 

Information  on  humpback  whales  is  of  particular 
interest  because  they  were  the  primary  species  hunted 
by  whalers  off  Washington  in  the  early  1900s.  Since 
then,  little  has  been  known  about  their  movements  and 
distribution  in  this  region.  Photo-identification  research 
has  helped  define  the  movements  and  stock  structure  of 
the  humpback  whales  feeding  off  California  (Calamboki- 
dis et  al.,  1990.  1996,  2000).  Calambokidis  et  al.  (1996) 
suggested  that  a  demographic  boundary  exists  between 
humpback  whales  that  feed  off  the  coasts  of  California, 
Oregon,  and  Washington  and  humpback  whales  feeding 
farther  north  off  British  Columbia  and  Alaska.  The 
identity  and  degree  of  interchange  of  the  whales  that 
feed  in  this  boundary  area  have  been  unclear. 

Similarly  for  killer  whales,  photo-identification  stud- 
ies have  revealed  much  about  whale  groups  that  fre- 
quent the  inland  waters  of  Washington  and  British 
Columbia  (Bigg  et  al.,  1990;  Ford  et  al.,  1994).  Very 
little  is  known  about  their  occurrence  off  the  coast,  in 
particular,  about  the  "offshore"  groups  that  are  believed 
to  be  a  distinct  race  (Ford  et  al.,  1994)  that  are  seen 
primarily  offshore  but  occasionally  also  enter  inland 
waterways. 

We  report  here  on  the  summer  distribution  of  marine 
mammals  off  the  northern  Washington  coast  based  on 
six  ship  line-transect  surveys  conducted  between  1995 
and  2002.  These  surveys  were  initiated  to  understand 
marine  mammal  distribution  and  abundance  in  the 
newly  designated  Olympic  Coast  National  Marine  Sanc- 
tuary, as  well  as  to  collect  information  on  seabirds, 
oceanographic  conditions,  and  juvenile  fish.  Each  ship 
survey  was  conducted  between  mid-June  and  late  July. 
Density  estimates  were  made  for  the  two  most  common 
species:  humpback  whales  and  Dall's  porpoise.  In  ad- 
dition, photo-identification  data  gathered  during  these 
ship  surveys  and  from  supplemental  small  boat  surveys 


:i  Calambokidis,  J.,  J.  C.  Cubbage,  J.  R.  Evenson,  S.  D.  Osmek, 
J.  L.  Laake,  P.  J.  Gearin,  B.  J.  Turnock,  S.  J.  Jeffries,  and  R. 
F.  Brown.  1993.  Abundance  estimates  of  harbor  porpoise 
in  Washington  and  Oregon  waters.  Report  to  the  National 
Marine  Mammal  Laboratory.  National  Marine  Fisheries  Ser- 
vice, 55  p.  Nat.  Mar.  Mamm.  Lab.,  NMFS,  7600  Sand  Point 
Way  NE  F/AKC3,  Seattle,  WA  98115. 

4  Bowlby,  C.  E.,  B.  L.  Troutman,  and  S.  J.  Jeffries.  1988.  Sea 
otters  in  Washington:  distribution,  abundance,  and  activ- 
ity patterns.  Final  report  to  National  Coastal  Resources 
Research  and  Development  Institute,  Hatfield  Marine  Sci- 
ence Center,  2030  S.  Marine  Dr.,  Newport,  Oregon  97365, 
131  p.  Cascadia  Research  Collective,  Wash.  State  Dept.  of 
Wildlife,  Olympia,  WA. 


within  the  same  area  between  1989  and  2002  provided 
information  on  humpback  and  killer  whale  movements 
and  stock  structure. 


Materials  and  methods 

Ship  surveys 

Generally,  ship  surveys  covered  the  area  between  the  20-m 
isobath  and  the  landward  margin  of  the  continental  shelf 
i200-m  isobath)  from  the  entrance  to  Strait  of  Juan  de 
Fuca  to  the  mouth  of  the  Copalis  River  to  include  the 
boundaries  of  the  Olympic  Coast  National  Marine  Sanc- 
tuary (Fig.  1).  Although  the  northern  extent  of  these 
waters  is  off  southern  British  Columbia  (Vancouver 
Island),  the  entire  overlapping  region  will  be  referred 
to  as  northern  Washington. 

Fourteen  east-west  tracklines  were  selected,  follow- 
ing permanent  tracklines  established  by  the  NOAA 
ship  Miller  Freeman  in  1989.  Tracklines  were  spaced 
at  5-nmi  intervals  and  were  surveyed  each  year  ex- 
cept in  2002,  when  only  ten  lines  were  surveyed  (four 
southernmost  lines  were  not  included).  Extra  ship  time 
allowed  for  replicate  surveys  of  the  northern  survey 
legs  in  1995,  a  short  offshore  extension  of  two  lines 
in  1996  and  2000  (up  to  17  nmi  in  1986),  the  addition 
of  three  short  east-west  lines  off  southern  Vancouver 
Island  around  La  Perouse  Bank  in  1997,  and  one  ad- 
ditional line  that  was  surveyed  south  of  the  study  area 
in  2000  (Fig.  1). 

Ship  surveys  were  conducted  over  a  two-week  period 
in  late-June  and  July  1995,  1996,  1997.  1998,  and  2000 
(Table  1).  In  2002,  a  shorter,  one-week  survey  was  done 
in  mid-June.  The  marine  mammal  ship  surveys  were 
conducted  by  a  single  primary  observer  from  the  vessel's 
flying  bridge  (the  sighting  platform)  with  a  viewing 
height  of  10  m  above  the  water  level.  All  surveys  were 
conducted  from  the  NOAA  ship  Mc Arthur  (55  m)  except 
during  2000,  when  the  naval  ship  Agate  Passage  (33  m) 
was  used.  From  these  platforms,  the  primary  observer 
scanned  a  180-degree  arc  encompassing  the  area  ahead 
of  the  ship  and  abeam  to  either  side.  Observers  used 
reticle  binoculars  when  possible  and  obtained  measure- 
ments of  distance  to  a  sighting  derived  from  the  angle 
below  the  horizon  (measured  with  graded  reticles  in  the 
binoculars)  and  the  known  platform  height.  For  sight- 
ings where  the  species  could  not  be  determined  by  the 
observer,  animals  were  identified  to  a  general  taxonomic 
level  (e.g.,  unidentified  pinniped). 

Photo-identification  surveys 

In  addition,  photo-identification  data  were  examined  that 
had  been  gathered  within  the  survey  area.  Research- 
ers took  photographs  directly  from  the  survey  ship,  or 
from  a  Zodiac  rigid-hulled  inflatable  that  was  launched 
when  animals  were  sighted.  In  1996,  the  last  two  days 
of  vessel  time  on  the  McArthur  were  used  to  photograph 
whales  for  identification. 


Calambokidis  et  al.:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


565 


1 26°0'0"W 
1 


125:'0'CrW 

I 


Vancouver 
Island 


Barkley 
Sound 


N  3  lines  only  in  1997 

La  Perouse  

Bank 


t 


Swiftsure 
Bank      . 


Strait  of 
Juan  de  Fuca 


Cape 
Flattery 


Olympic 
Peninsula 


Sanctuary  Boundary 


Figure  1 

On-effort  ship  survey  tracklines  (horizontal  lines)  off  the  northern  coast  of  Wash- 
ington between  1995  and  2002.  The  Olympic  Coast  National  Marine  Sanctuary 
boundary  is  delineated  and  labeled.  Dashed  and  dotted  lines  show  three  northern 
lines  surveyed  only  in  1997,  the  western  extension  of  two  lines  surveyed  only  in 
1996,  and  the  southern  four  lines  missed  in  2002. 


In  addition,  dedicated  photo-identification  surveys 
were  conducted  by  Cascadia  Research  scientists  us- 
ing a  5.3-m  Novurania  rigid-hulled  inflatable  that  was 
launched  from  nearby  ports  and  operated  in  areas 
where  whales  were  concentrated.  Photo-identification 
data  in  the  present  study  includes  data  collected  off 
the  northern  Washington  coast  between  1989  and  2002 
(Table  2).  It  also  includes  photographs  contributed  by 


other  researchers  and  boat  operators  taken  in  the  area 
during  this  time  (Table  2). 

Generally,  photographs  were  taken  with  Nikon  8008 
35-mm  cameras  equipped  with  300-mm  Nikkor  telepho- 
to  lenses.  High-speed  black-and-white  film  (Ilford  HP 
5+)  was  pushed  IV2  stops  so  that  exposure  times  were 
generally  1/1000  or  1/2000  of  a  second.  Identification 
photographs  were  taken  with  standard  procedures  used 


566 


Fishery  Bulletin  102(4) 


Table  1 

Summa 

ry 

of  ship  survey 

effort  off 

northern  Washin 

gton (does 

not  include  small  boat 

surveys). 

Dates  of  effort 

No.  of 

nmi  on 

Year 

Start 

End 

legs 

Effort  (h) 

effort 

Ship 

Observers 

1995 

21  Jul 

27  Jul 

10 

46 

546 

Mc  Arthur 

Troutman.  Ellifrit 

1996 

28  Jun 

5  Jul 

14 

46 

540 

Mc  Arthur 

Troutman,  Ellifrit 

1997 

9  Jul 

18  Jul 

17 

52 

513 

Mc  Arthur 

Troutman,  Ellifrit 

1998 

25  Jun 

4  Jul 

14 

55 

572 

McArthur 

Troutman,  Quan 

2000 

16  Jun 

24  Jun 

14 

60 

589 

Agate  Passage 

Rowlett,  Nelson 

2002 

12  Jun 

18  Jun 

10 

32 

315 

McArthur 

Troutman,  Douglas 

All  years 

291 

3075 

Table  2 

Photo-identification  effort  off  the 

coast  of 

northern 

Washington 

bet 

ween 

1989  and  2002. 

These 

data 

include  whales  identified 

from  the  ship  or 

small  boats  launched  from  the  sh 

ip,  dedicated  small  boat  surveys, 

and 

opportunistic  photographs  taken  by 

others.  Unique  = 

number  of  different  animals. 

Days  IDs  obtained 

Humpback  whales  identified 

Other  sources  of  photographs 

Year                No 

First 

Last 

No. 

Unique 

No. 

of  mothers       No 

of  calves 

1989                 1 

lOct 

1  Oct 

1 

1 

0 

0 

1990                 3 

25  Aug 

6  Sep 

10 

10 

1 

1 

Balcomb/Bloedel' 

1991                 4 

23  Aug 

4  Sep 

14 

13 

0 

0 

Balcomb/Bloedel' 

1993                 1 

15  Jul 

15  Jul 

3 

3 

0 

0 

1994                 3 

25  Jun 

15  Jul 

20 

16 

0 

0 

G.  Ellis.-  R.  Baird 

1995                 7 

14  Jul 

25  Jul 

50 

35 

4 

2 

S.  Mizroch3 

1996                 9 

29  Jun 

6  Oct 

55 

34 

1 

0 

1997                 9 

13  Jul 

18  Oct 

25 

23 

2 

0 

1998               19 

28Mav 

16  Oct 

71 

48 

1 

1 

V.  Deeke,  B.  Gisborne 

1999               28 

20  May 

20  Oct 

103 

60 

2 

0 

B.  Gisborne 

2000               12 

2  Jun 

4  Oct 

56 

40 

2 

1 

B.  Gisborne 

2001               15 

8  Jun 

5  Oct 

59 

41 

2 

1 

SWFSC.J  B.  Gisborne 

2002                 9 

13  Jun 

5  Sep 

41 

32 

0 

0 

Total             120 

508 

356 

15 

6 

Unique 

191 

;  Center  for  Whale 

Research,  P.O.  Bo> 

1577,  Fn 

day  Harboi 

,  WA  98250. 

2  Dept.  of  Fisheries  and  Oceans,  Pacific  Biological  Station 

Nanaimo,  BC 

,  V9T  6N7, 

Canada. 

3  National  Marine 

Mammal  Laboratory,  NMFS 

7600  Sand  Point  Way  NE,  Seattle, 

WA  98115. 

4  Southwest  Fisheries  Science  Center 

8604  La 

lolla  Shores  Dr.,  La  Jolla 

CA  92037. 

in  past  research  (Calambokidis  et  al.,  1990).  For  hump- 
back whales,  photographs  were  taken  of  the  ventral 
side  of  the  tail  flukes.  For  killer  whales,  the  dorsal  fin 
and  surrounding  saddle-patch  area  were  photographed 
from  both  sides. 

Photographs  of  individuals  were  first  compared  to 
those  identified  in  the  same  region.  To  analyze  inter- 
change with  other  regions,  we  compared  these  individu- 
als with  existing  catalogs  to  obtain  sighting  histories. 
For  humpback  whales,  a  catalog  was  used  of  over  1000 
humpback  whales  identified  since  1986  along  the  West 


Coast.  The  regions  used  for  comparison  were  Oregon, 
northern  California  (Oregon-California  border  to  Pt. 
Arena i,  northern  central  California  (Pt.  Arena  to  north 
of  Monterey  Bay),  southern  central  California  (north 
of  Monterey  Bay  to  Pt.  Conception!  and  southern  Cali- 
fornia (southern  California  Bight).  For  killer  whales, 
whales  were  matched  to  existing  catalogs  (Bigg  et  al., 
1987;  Ford  et  al.,  1994;  Black  et  al.,  1997).  All  iden- 
tifications and  group  determinations  were  confirmed 
by  one  of  the  authors  tDKEi  or  Graeme  Ellis  (Dept.  of 
Fisheries  and  Oceans.  Nanaimo.  British  Columbia). 


Calambokidis  et  al.:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


567 


Data  analysis 

For  ship  surveys  between  1995  and  2000,  position  and 
oceanographic  data  (including  depth,  sea  surface  tem- 
perature) logged  by  the  ship's  computer  were  later  rec- 
onciled with  the  sighting  and  effort  data  recorded  by 
the  observers.  Sighting  positions  were  analyzed  for  each 
species  for  water  depth,  distance  from  shore,  distance 
from  shelf  edge  (200-m  depth  contour)  and  sea  surface 
temperature.  Data  analysis  and  mapping  were  conducted 
by  using  a  geographic  information  system  (GIS)  with 
Arclnfo  software  (ESRI,  Redlands,  CA).  Data  from  the 
shorter  2002  ship  survey  were  included  in  the  summary 
of  sightings  but  were  not  available  for  the  analyses  of 
sightings  related  to  oceanographic  features. 

Line-transect  analysis  to  determine  density  and  abun- 
dance was  conducted  for  the  two  species  with  more  than 
30  sightings  (humpback  whales  and  Dall's  porpoise). 
We  used  the  program  (Distance,  version  3.5,  Research 
Unit  for  Wildlife  Population  Assessment,  University  of 
St.  Andrews,  St.  Andrews.  UK)  to  conduct  analyses. 
For  these  analyses,  we  used  only  effort  and  sightings 
from  the  regular  east-west  transect  lines  and  did  not 
include  on-effort  data  from  opportunistic  lines  or  cross- 
tracks.  We  included  sightings  made  by  secondary  as 
well  as  the  primary  observer.  Although  whales  were 
reportedly  seen  out  to  6  nmi,  we  truncated  the  sight- 
ings at  3  nmi  for  humpback  whales  and  2.5  nmi  for 
Dall's  porpoise.  For  humpback  whales  we  included  16 
sightings  of  unidentified  whales  (unidentified  mainly 
because  of  distance).  These  were  probably  humpback 
whales  because  the  only  other  large  whales  that  were 
seen  in  the  surveys  were  a  few  gray  whales  seen  close 
to  shore.  Distance  position  data  were  incomplete  for  13 
of  the  188  whale  sightings  and  14  of  82  Dall's  porpoise 
sightings;  for  these  the  missing  value  was  randomly 
selected  from  the  observed  measurements. 

The  Distance  program  was  used  to  select  the  best 
model  for  sighting  probability  in  relation  to  distance  off 
the  transect.  We  allowed  the  program  to  select  among 
models  (half-normal,  uniform,  hazard-rate,  and  nega- 
tive exponential)  and  varying  numbers  of  adjustment 
terms  (cosine  and  simple  polynomials)  based  on  lowest 
Akaike's  information  criterion  (AIC)  score.  All  years 
were  pooled  for  the  model  of  sighting  probability,  but 
encounter  rate  and  group  size  were  calculated  by  year. 
An  adjustment  to  group  size  was  calculated  if  there 
was  a  significant  group  size  bias  with  distance  from 
the  track  line,  which  was  not  the  case  for  humpback 
whales  but  was  present  in  some  years  (1996  and  1997) 
for  Dall's  porpoise. 

Area  was  calculated  for  abundance  estimation  based 
on  the  zone  covered  by  the  regularly  scheduled  transect 
lines  covered  in  most  years  (study  area  was  considered 
to  encompass  waters  2.5  nmi  north  of  the  northernmost 
line  and  2.5  nmi  south  of  the  southernmost  line).  The 
only  annual  adjustment  for  area  was  for  humpback 
whales  in  2002.  Surveys  in  that  year  did  not  cover 
the  southern  end  of  the  study  area  (because  of  limited 
ship  time),  an  area  with  a  typically  lower  abundance 


of  whales.  To  avoid  extrapolating  the  higher  density 
of  whales  from  the  northern  portion  of  the  study  area 
to  this  region,  we  excluded  this  missed  area  from  the 
abundance  estimates. 

Estimates  of  abundance  for  humpback  whales  were 
also  calculated  by  using  capture-recapture  models  (Se- 
ber,  1982;  Hammond,  1986).  We  used  identifications 
obtained  in  pairs  of  adjacent  years  taken  from  1994  to 
2002  to  generate  Petersen  capture-recapture  estimates. 
The  Chapman  modification  of  the  Petersen  estimate 
(Seber,  1982)  was  used  because  it  was  appropriate  for 
sampling  without  replacement  (Hammond,  1986). 


Results 

In  total,  there  were  706  sightings  of  2467  animals  over 
the  six  ship  surveys  combined  (Table  3).  Fifteen  differ- 
ent marine  mammal  species  were  seen:  nine  cetacean 
species,  five  pinniped  species,  and  the  sea  otter  were 
identified.  Each  year,  9  to  12  different  species  were  seen, 
except  in  2002  when  only  six  species  were  observed. 
This  2002  survey,  although  shorter  than  those  of  the 
other  years,  showed  a  dramatic  change  in  the  species 
diversity  and  numbers  of  animals.  We  saw  many  more 
humpback  and  Dall's  porpoise  than  in  previous  years. 
We  also  noted  the  absence  of  six  regularly  observed  spe- 
cies: harbor  porpoise,  gray  whales,  Pacific  white-sided 
dolphins  (Lagenorhynchus  obliquidens),  Risso's  dolphin 
(Grampus griseus),  harbor  seals,  and  California  sea  lions 
iZalophus  californianus). 

Humpback  whales 

Of  the  large  cetaceans,  humpback  whales  were  the  most 
common  species  seen;  there  were  232  sightings  of  402 
animals  during  ship  surveys  (Table  3).  Largest  numbers 
of  humpback  whales  were  seen  in  2002,  when  there 
were  79  sightings  of  139  individuals  during  the  one- 
week  survey.  Group  sizes  ranged  from  1  to  8  animals 
(mean=1.7,  SD=1.1).  Only  six  calves  were  recorded  from 
the  ship  surveys — probably  because  it  was  difficult  to 
identify  calves  at  the  distance  at  which  most  sightings 
were  made.  Of  these  six  sightings  of  mothers  with  calves, 
four  sightings  were  outside  the  primary  areas  where 
other  humpback  whale  groups  were  seen. 

Sightings  were  concentrated  in  the  northern  part  of 
the  study  area  between  Juan  de  Fuca  Canyon  and  the 
outer  edge  of  the  continental  shelf,  an  area  known  as 
"the  Prairie"  (Fig.  2).  A  small  area  east  of  the  mouth  of 
Barkley  Canyon  and  north  of  the  Nitnat  Canyon  where 
the  water  depth  was  125-145  m  had  a  high  density  of 
sightings  in  all  years.  A  smaller  number  of  humpback 
whales  were  also  seen  on  Swiftsure  Bank.  Sightings  in 
2002  were  not  only  more  numerous  but  more  broadly 
distributed;  sightings  were  recorded  in  the  areas  de- 
scribed above  and  also  farther  south  and  closer  to  shore 
than  those  seen  in  previous  years. 

Line-transect  estimates  for  humpback  whales  were 
very  consistent  in  the  first  five  surveys  (1995  to 


568 


Fishery  Bulletin  102(4) 


CO 

Cu 

O 

TO 

a 

CN 

CO 

CO 

CO                              ^f             CN 

co                   as 

— 

t> 

d 

o 

CO 

o 

co                    cn         as 

—i                               CN 

CD 

CO 

Z 

ct3 

Tf 

T 

rH 

CN 

CO 

CD 

>> 

to 

-3 

_ 

ti_ 

bjD 

CD 

< 

O 

C 

CN 

CD 

LO 

05 

00                           -*            CO 

in                     x 

CO 

CO 

-a 

*p 

co 

CN 

l — l 

00 

CN                              -^             CN 

— i                        o^ 

1 — 1 

c 

_3 

c 

z 

J= 

CN 

1 — i 

l> 

u 

.SP 

a 

'to 

CO 

*~ 

Cm 

CO 

o 

TO 

OS    t> 

CD 

CO 

00 

as                    oo         oo 

O                    CD 

o 

co 

CO 

CO 

CN 

CO 

+a 

d 

T— 1 

CO 

2: 

c 

co 

CN 

C 

o 

IS 

o 

to 

c 
o 

e 

CN 

Cm 

o 

as   m 

■* 

m 

-d-              as   cn         ^t- 

O                    CD 

in 

d 

CO 

t> 

CN 

t> 

CN 

CN 

to 

bo 

a 

CO 

C|_) 

bp 

CO 

O 

TO 

» 

CO 

00 

1 — 1 

as              as   co 

i — i                               i — i 

i-H 

m 

d 

2; 

6 
'3 

lO 

Tf 

co                      co 

CO 

m 

_CJ 

o 
o 

CO 

+3 

to 

o 

CN 

<*- 

CO 

be 

3 

s- 

o 

o 

C 

•* 

CO 

t 

r- 

tJ-             CO             CO 

^H                                                   t— < 

^H 

CO 

d 

be 

CN 

CD 

a 

CO 

ft 

o 

t_- 

J2 

c 

O 

TO 

CO 

^H 

CO 

CO 

h   a   w         m 

t-~   cn   ^t   co   as 

■* 

as 

o 

d 

z 

£ 

CO 

t 

■tf 

i> 

bo 

'S 

^H 

S 

00 

TO 

IS 

CO 

OS 
OS 

CO 

CO 

*~* 

c*— 

bo 

5: 

c 

O 

c 

c- 

t— 1 

CO 

■* 

r+     CN             CO 

c^   o-i   tt   ^   as 

■^ 

rH 

d 

.-C 
.5P 

<N 

^H 

1-1 

as 

aj 

"co 

C>    J3 

+-> 

01     - 

—      o 

■' 

.o    e 
£  is 

o 

TO 

a 

'5 

ti-   co 

■*     CN 

CO 

to 

CD     >-l     —<             Tf 

CO     CN                      CN 

CN 

CO 

d 

;z; 

Tf 

1-1 

-* 

>-"     CN 

t 

CO 

t> 

TO 

CD 

OS 

> 

OS 

CO 

[~ 

1 — ' 

«-_, 

be 

3 

CO 

o 

_c 

CO     CM 

■>*     iH 

!D 

CN 

rf     CN     CN             CM 

CN     CN                      iH 

^H 

in 

g. 

d 

•5P 

CN 

•"^ 

^^ 

CD 

2 

CO 

CO 

bo 

C 

CO 

t-4 

3 

-a 

o 

TO 

s 

'S 

CD     CO 

rH     CO 

"* 

O 

as              m   x         cd 

co   o   co   --^   m   co 

CO 

m 

d 

2: 

00 

CO 

CN 

^f                    co         in 

^^ 

as 

CO 

so 

TO 

CD 

OS 

>. 

O) 

CO 

>> 

.— 1 

i;,   , 

bo 

-o 

o 

C 

■*    CN 

^h     CO 

o 

^H 

CO                     ^H     CN             00 

co   o   co   ^   in   co 

CO 

„ 

CO 

bo 

c 

d 

be 

CO 

CN 

"" 

^H 

" 

■>* 

"53 

■U 

X 

bo 

CO 

CO 

o 

TO 

a 

o    — • 

CO    CO 

CN 

o 

co             t--  co        as 

CM     CO     i-t     iO     CD     0? 

tC 

CN 

CO 

d 

■* 

C~ 

as              in   <-i 

t^ 

£ 

2 

s 

in 

t> 

£ 

OS 

CO 

CO 

OS 

£ 

1-1 

CO 

bo 

0) 

o 

C 

3 

m   .-i 

CO     T 

c~ 

■* 

■^                   IC    CO           CD 

O)     CO     ^h     CO     CO     CO 

■■# 

T* 

c 

CO 

d 

2: 

be 

CN 

CM 

as 

s 

'co 

T3 

c*h 

CD 

O 
>> 

CO 

— 

CD 

CO 

_ 

-a 

CD 

3        ■    c 

to                    - 

(-                5         — 

ft 

B 
- 

s 

S 

CD 
CO 

3 
ft 

u 

o 
a 

?       x.  ■£  Z 

°      -s  n  §  2 

3. 

£ 

3 
CO 

CO        S 

_os  _^.  ,£ 
*ra     O     CO 

ii  -a 

CO      CD 

*  '5 

'8 

«  £■ 

2    o 

as          »  ax    o,  tj 

--             'C    "5      D-  "^      CD    "3 

■go    c  ■§  £  *  S   c 

^    a  j£    co  -g    u    .cd    a 

_               CO      CD     .-      1- 
^    —      CD      CO    —      3 

«    S    "    S    g    c    ^ 

to      W      C      £      «      £      o; 

■a 

CD 

cc 

— 

to 
bo 

3 
-t-> 
-C 

bo 
to 

"m 

+3 
O 

CO 

CD 

'Zi 

CD 

*  -a  s 

S   3  2 

s   § 

IS     c 

CD      D. 

1     ■- 

C      TO 

u 
: 
_= 
u 

CO 

T3      fe      CO      t.      i.      CD     *j 

a-g  -R  ^3^5    = 

'3     £     CD    "C3     CD     U     w 
3      "    ~      CO     ^S      3      CD 

c 

CD 

-a 

3 

■a  k  o 

S  D 

•§  a 

X 

cC      z      cs  fcc;  d 

3   ac   a   O   0Q   Z   00 

D 

cn 

pq 

o 

£ 

H 

Calambokidis  et  al.:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


569 


Figure  2 

Locations  (by  year)  for  humpback  whales  (Megaptera  novaeangliae)  seen  during  ship 
surveys  off  the  northern  Washington  coast  between  1995  and  2002. 


2000,  Table  4,  Fig.  3).  The  encounter  rate  of  groups 
(0.046-0.053  sightings  per  nmi),  density  (0.034-0.050 
whales  per  nmi2),  and  abundance  (85-125  individuals) 
were  similar  among  these  years.  These  data  indicate 
that  about  100  humpback  whales  used  the  study  area 
during  this  period. 

The  sighting  rate  of  humpback  whales  was  dramati- 
cally higher  in  2002  than  in  all  previous  years  and 
was  reflected  in  the  line-transect  estimates  (Fig.  3). 
Estimated  density  (0.23  whales  per  nmi2)  was  more 
than  four  times  higher  than  any  previous  year.  Apply- 


ing this  density  to  only  the  reduced  area  surveyed  in 
2002  (1953  instead  of  2505  nmi2)  still  yielded  much 
higher  estimates  of  abundance  (562,  CV=0.21)  than  in 
any  previous  year.  These  higher  abundance  estimates 
could  not  have  been  an  artifact  of  random  variation;  the 
lower  bound  of  the  95%  confidence  interval  for  the  2002 
estimates  was  well  above  the  upper  confidence  interval 
of  any  of  the  previous  years  (Table  4). 

Of  the  humpback  whales  photographed  during  small 
boat  surveys  off  the  northern  Washington-BC  border 
between  1989  and  2002,  508  individuals  were  success- 


570 


Fishery  Bulletin  102(4) 


Table  4 

Results 

of  line-transeet  analysis  for  h 

umpback  whales  off  northern  Wash 

ngton.  C 

n-effort  sight 

ngs  of  b 

umpback 

and  un- 

identified  large  whales 

made  during  regular  transects  ( not  including  deadheads  [areas 

between  transect  lines 

and  opportunistic 

sightings)  within  3 

nm 

of  ship 

were  used.  Best  detection  model  fit  (AIC  scores)  was  a  negative  exponential  w 

th  1  cosine  adjust- 

ment  yielding  /'l  0)  = 

1.0E 

.  Effect 

ve  strip 

width  was  0.95 

nmi  w 

ith  CV=0.09. 

Survey  effort 

953  Conf.  int. 

Sightings 

Encounter 

Group 

Density 

Area 

Estimated 

Year 

/! 

lines 

nmi 

rate 

size 

(per  nmi2) 

(nmi2) 

abundance 

CV 

lower 

upper 

1995 

23 

58 

438 

0.053 

1.48 

0.041 

2505 

102 

0.33 

54 

193 

1996 

24 

59 

474 

0.051 

1.54 

0.041 

2505 

103 

0.33 

55 

193 

1997 

26 

92 

493 

0.053 

1.62 

0.045 

2505 

112 

0.3 

63 

199 

1998 

20 

62 

432 

0.046 

1.40 

0.034 

2505 

85 

0.31 

47 

155 

2000 

23 

70 

504 

0.046 

2.09 

0.050 

2505 

125 

0.32 

67 

234 

2002 

72 

43 

305 

0.236 

1.81 

0.224 

1953 

562 

0.21 

375 

841 

Total 

188 

384 

2646 

fully  identified  of  which  191  were  unique  individuals 
i  Table  2).  Of  these  191,  83  (44%)  had  been  seen  in  this 
area  in  more  than  one  year  within  this  time  period.  The 
proportion  of  animals  seen  more  than  one  year  changed 
over  the  course  of  the  study  ( Fig.  4).  The  proportion  of 
whales  identified  each  year  that  had  been  seen  in  others 
years  decreased  annually  (Fig.  4,  regression  r-  =  0.63, 
P=0.002);  the  most  dramatic  drop  occurred  between 
1998  and  1999. 

Photographs  of  humpback  whales  documented  animal 
movements  within  the  study  area  and  provided  some 
insight  into  possible  reasons  for  the  high  sighting  rates 
during  the  2002  ship  surveys.  On  two  occasions,  the 
same  humpback  whale  was  identified  on  different  days 
in  a  slightly  different  area  and  represented  a  duplicate 


600  -1 

•  •♦-  ■  Line-transect 

♦ 

— ■ —  Capture-recapture 

500  ■ 

,' 

400  - 

.' 

O 

§    300- 

c 

200  - 

100  - 

■    m  '-•♦---' 

■              ' 

l<)')4       IW5       1996       1997       1998        1999       2000       2001 

.' 2       2003 

Year 

Figure  3 

Line-transect  Idashed  line i  and  capture-recapture 

(solid  line) 

estimates  for  humpback  whale  iM.   noracannliae) 

abundance 

between  1995  and  2002. 

sighting  of  this  animal  from  the  ship  survey.  It  is  pos- 
sible that  shifting  humpback  whale  distribution  during 
the  course  of  the  2002  survey  could  have  occurred  in 
a  manner  that  resulted  in  the  same  animals  being 
encountered  multiple  times  and  that  elevated  the  sight- 
ing rate  and  line-transect  abundance  estimate  (Fig.  3). 
We  cannot  test  this  hypothesis  because  other  animals 
may  have  shifted  in  a  manner  that  they  avoided  being 
detected  at  all. 

Abundance  of  humpback  whales  from  capture-re- 
capture models  yielded  estimates  of  89  to  343  whales 
(Table  5,  Fig.  3).  These  estimates  tended  to  increase 
over  the  course  of  the  study  from  a  low  of  89  whales 
for  1994-95  to  a  high  of  343  for  2000-2001  and  230 
for  2001-2002  (regression  r2  =  0.60,  P=0.02).  There  was 
fairly  good  agreement  between  the  capture-re- 
capture and  line-transect  estimates  until  2002 
(Fig.  3). 

A  total  of  17  of  the  191  (9%)  whales  that  we 
identified  off  northern  Washington  had  also 
been  photographed  off  California  and  Oregon 
(Table  6).  Interchange  of  whales  seen  off  north- 
ern Washington  and  other  feeding  areas  to  the 
south  decreased  as  distance  among  feeding  ar- 
eas increased.  About  10%  (10  of  105)  of  the 
whales  that  were  identified  off  Oregon  were 
also  photographed  off  northern  Washington. 
This  rate  of  matching  dropped  below  3%  (8  of 
313)  off  northern  California  and  continued  to 
decrease  to  no  interchange  seen  for  whales  pho- 
tographed off  southern  California. 

The  proportion  of  whales  that  were  seen  in 
areas  to  the  south  appeared  to  change  over 
the  course  of  the  study.  From  1989  to  1998, 
when  resighting  rates  between  years  within 
our  study  area  were  highest,  we  also  had  a 
higher  proportion  of  interchange  with  feeding 
areas  to  the  south  (13  of  109  whales  or  12%). 
From  1999  to  2002,  after  resightings  within 
our  region  decreased,  there  was  also  a  decrease 


Calambokidis  et  al.:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


571 


Table  5 

Estimates  of  humpback 
Each  estimate  was  base 

whale  abundance 
d  on  the  identificat 

Est 
ions 

)  off  northern  Washington  obtained 
obtained  (n)  in  each  of  two  adjacent 

with  the  Petersen 
years. 

capt 

are-recapture 

model. 

Period 

Sample  1 

Sc 

mple  2 

Match 

Est. 

CV 

Year 

n 

Year 

n 

1994- 

-95 

1994 

14 

1995 

35 

5 

89 

0.27 

1995- 

-96 

1995 

35 

1996 

34 

11 

104 

0.19 

1996- 

-97 

1996 

34 

1997 

21 

7 

95 

0.24 

1997- 

-98 

1997 

23 

1998 

48 

6 

167 

0.28 

1998- 

-99 

1998 

48 

1999 

60 

13 

213 

0.19 

1999- 

-2000 

1999 

60 

2000 

31 

14 

129 

0.16 

2000- 

-01 

2000 

40 

2001 

41 

4 

343 

0.36 

2001- 

-02 

2001 

41 

2002 

32 

5 

230 

0.32 

in  the  proportion  of  these  whales  that  had  also 
been  seen  off  California  and  Oregon  (7  of  136 
whales  or  5%).  This  difference  falls  just  short 
of  statistical  significance  (j2  =  3.71,  P<0.10)  but 
is  in  the  reverse  direction  from  what  would  be 
expected  if  immigration  from  the  south  were  to 
increase  over  time. 

Between  1989  and  2002,  15  different  mothers 
were  seen  with  16  calves  (one  mother  seen  with  a 
calf  in  two  different  years).  Mothers  with  calves 
represented  4.2%  of  the  individual  whales  iden- 
tified each  year  ( 15  of  356  unique  annual  iden- 
tifications. Table  2).  For  each  year  only  a  small 
proportion  of  the  calves  were  identified  because 
calves  raise  their  flukes  less  often. 


Killer  whales 

One  other  large  cetacean  species  (killer  whales) 
was  also  seen  every  year;  there  were  a  total  of 
14  sightings  of  124  animals  from  ship  surveys 
(Table  3).  Three  of  these  sightings  were  of  large 
groups  between  20  and  35  animals,  and  the 
rest  were  in  groups  fewer  than  ten  (14  sight- 
ings, mean  =  8.9,  SD  =  11.2).  Killer  whales  were 
widely  distributed  across  different  habitats;  there 
were  sightings  of  animals  both  close  to  and  far 
from  shore  and  in  fairly  shallow  and  deep  water 
(Fig.  5). 

All  three  ecotypes  of  killer  whales  (namely, 
1)  southern  and  northern  residents,  2)  transients,  and 
3)  offshore  residents)  were  observed  off  the  northern 
Washington  coast.  Of  the  15  groups  identified  pho- 
tographically between  1989  and  2002,  there  were 
sightings  of  animals  from  the  southern  resident  (2 
groups),  northern  resident  (3),  transient  (5)  and  off- 
shore (3)  groupings  (Table  7).  Other  sightings  appeared 
to  be  northern  residents  (1)  and  offshore  (1)  animals 
but  the  quality  of  the  photographs  were  too  poor  for 


■B       80<r 


i1 

5 
■D 


2091 


X  %y  %  \  x  %  x  \  \  \  \  % 

Year 

Figure  4 

The  proportion  of  humpback  whales  (M.  novaeangliae)  seen 
in  more  than  one  year  during  annual  surveys  off  northern 
Washington  from  1989  to  2002. 


us  to  be  certain.  Large  groups  of  killer  whales  (20-40 
animals)  were  seen  on  five  occasions  during  small  boat 
surveys. 

Dall's  porpoises 

Dall's  porpoises  were  the  most  frequently  sighted  small 
cetacean;  there  were  115  sightings  of  406  animals  and 
Dall's  porpoises  were  observed  every  year  (Table  3).  No 


572 


Fishery  Bulletin  102(4) 


Table  6 

Number  of  humpback 

whales 

identified  in 

different  regions  along  the  U.S.  west  coast  and  the  number  and  percentage  of  these 

that  matched  with  northern  Washington 

For  northern  Washington,  we  report  the  number  of 

whales  that  were  seen  in  that 

region  in  more  than  one  year. 

No.  of 

No.  of  matches 

%  of  whales  that 

Region 

individuals 

with  N.  Wash. 

match  with  those  in  N.  Wash. 

Northern  Washington 

191 

83 

43.5% 

Oregon 

105 

10 

9.5% 

N.  California 

313 

8 

2.6% 

N.  Centra]  California 

921 

13 

1.4% 

S.  Central  California 

666 

3 

0.5% 

S.  California 

303 

0 

0.0% 

Table  7 

Summary  of  killer  wh 

lie  sightir 

gs  off  north 

ern  Washington  between  1989 

and  2002  where  identifiable  photographs  were 

taken. 

No.  of 

animals 

Lat. 

Long. 

Date 

estimated 

°N 

°W 

Community 

Pod  or  ID                                   Comments 

13  Sep 

89 

3 

48  23.0 

124  48.5 

Resident — southern 

L10,  L28.L41 

15  Jul 

94 

4 

48  20.9 

125  20.0 

Transient 

CA195 

25  Jul 

95 

7 

47  49.8 

124  59.5 

Transient 

CA195 

26  Jul 

95 

8 

47  53.7 

125  03.3 

Transient 

CA20.CA27 

17  Mar 

96 

6 

46  58.2 

124  15.7 

Resident — southern 

L26,  L83                            outside  Grays  Harbor 

31  Mar 

96 

7 

46  55.0 

124  09.7 

Transient — probably 

T50?                                     Grays  Harbor  entrance 

5  Jul 

96 

3 

48  13.1 

124  55.0 

Transient 

T185 

6  Jul 

96 

40 

48  26.7 

125  43.2 

Resident — northern 

C,  D.  Gls.  G12s 

15  Jul 

97 

30 

48  19.4 

125  09.5 

Offshore 

18  Jul 

97 

10 

48  18.3 

125  23.6 

Offshore 

CA105 

10  Aug 

97 

8 

48  21.0 

125  34.6 

Transient 

T36,  T99,  T36A?,  T137? 

27  Aug 

98 

40 

48  28.0 

125  17.0 

Offshore 

044.030,  031,0172, 
014,  0158,  0218 

10  Oct 

99 

30 

48  22.0 

125  38.1 

Resident — northern 

111 

18  Jun 

00 

20 

48  03.8 

125  04.3 

Probably  resident — northern 

not  southern  residents 

6  Sep 

01 

12 

47  01.8 

124  46.6 

Resident — northern 

G12s,  G17s,  G29s 

Table  8 

Results  of  line-transect  analysis  for  Dall's  porpoise 

off  northern 

Washingt 

Dn.  All  on 

-effort  sightings 

during 

regular 

transects 

(not  incluc 

ing  deadheads  [areas  between  transect  li 

nes]  and  opportunistic 

sight  ingsi 

within  2.5  nmi 

of  the  s 

lip  were 

included. 

The  best  detection  model  fit  lAIC  scores) 

was  the  hazard  rate  with 

no  cosine  adjustment,  yielding  /"(0)  = 

2.60.  Effective  strip  width 

of  0.38  nmi  with  CV= 

0.12.  The 

group  size  for  96-97 

was  adjusted  to  account  for  a  significant  group  size  bias  with  dist 

a  nee  from 

the  trackline. 

Survey 

effort 

95r;  Conf.  int. 

Sightings 

Encounter 

Group 

Density 

Area 

Estimated 

Year 

n 

lines 

nmi 

rate 

size 

per  nmi'-) 

(nmi-) 

abundance 

CV 

lower 

upper 

1995 

16 

58 

438 

0.037 

2.25 

0.100 

2505 

268 

0.32 

143 

501 

1996 

14 

59 

474 

0.030 

2.65 

0.102 

2505 

255 

0.32 

138 

472 

1997 

13 

92 

493 

0.026 

2.28 

0.078 

2505 

197 

0.38 

95 

405 

1998 

9 

62 

432 

0.021 

2.67 

0.072 

2505 

181 

0.49 

72 

453 

2000 

13 

70 

504 

0.026 

3.46 

0.116 

2505 

291 

0.42 

132 

644 

2002 

17 

43 

305 

0.056 

4.82 

0.350 

2505 

876 

0.3 

487 

1576 

Total 

82 

384 

2646 

Calambokidis  et  al.:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


573 


126°0'0"V\ 

i 

125'0'0"W 
1 

Other  Large  Cetaceans 

• 

Minke  Whale 

Vancouver 

■ 

Gray  Whale 

490'0"N- 

Island 

A 

Killer  Whale 

-49"0'0"N 

Barkley 

Sound 

La 

Perouse 
Bank 

A 

Swifl 
Bg 

sure 
Dk 

▲ 

Strait  of 
^           1     Juan  de  Fuca 

Cape 

■  lattery 

^^-—-^^                            A 

^--^Prairie 

AR'  n'fl'M  — 

c/ 

*             A                                 • 

V 

\              6® 

A 

Olympic 
.              Peninsula 

A                         • 

-48°0,0,,N 

ho  U  U  IN 

^V/ 

A 

i  ^IM 

A 

■ 
■ 

i 

i. 

Sanctuary  Boundary 

N 

47"0'0"N- 

' 1 1 

-47°0'0"N 

Figure  5 

Locations  of  other  large  cetaceans  seen  during  ship  surveys  off  the  northern  Wash- 
ington coast  between  1995  and  2002. 


calves  were  recorded  during  the  surveys.  Dall's  porpoises 
were  widely  distributed  in  the  study  area  but  were  not 
as  commonly  seen  in  more  shallow  coastal  waters  or  in 
the  southern  portion  of  the  study  area  (Fig.  6).  Group 
size  ranged  between  1  and  12  individuals  (mean=3.5, 
SD  =  2.2).  Harbor  porpoises  were  observed  each  year 
(except  2002)  and  there  were  a  total  of  38  sightings  for 
the  entire  study  period.  Group  size  ranged  between  1 
and  6  individuals  except  for  one  sighting  of  a  group  of  20 
animals  (mean=2.3,  SD  =  3.1).  The  distribution  range  for 


harbor  porpoises  was  more  restricted  to  coastal  waters 
and  showed  only  a  small  overlap  with  the  distribution 
range  for  Dall's  porpoises  (Fig.  6). 

Line-transect  analysis  allowed  estimation  of  Dall's 
porpoise  density  and  abundance  (Table  8).  Similar  to 
those  for  humpback  whales,  results  for  Dall's  porpoises 
were  fairly  consistent  for  the  first  five  surveys  (1995 
to  2000):  annual  abundances  were  estimated  between 
181  and  291.  For  2002,  the  encounter  rate  and  corre- 
sponding density  and  abundances  increased  dramati- 


574 


Fishery  Bulletin  102(4) 


126°0'0"W 

125'0,0"W 
i 

Small  Cetaceans 

• 

Risso's  Dolphin 

Vancouver 

■ 

Northern  Right  Whale  Dolphin 

■KV'A'fV'M  — 

Island 

A 

Pacific  White-sided  Dolphin 

4y  U  U  N~ 

* 

Dall's  Porpoise 

Barkley 
Sound 

+ 

Harbor  Porpoise 

La  Perouse 
Bank 

Swiftsure 
Bap1' 

*     *  * 
I 

• 

* 
* 

*+*   +                 *»* 

*  * 

Strait  of 

\*        ***** 

*A* 

*  *  *        >* 

V*^— -<A      *+ 

-^^F'rairie           * 

*  *  a*+*  **  i 

'S^-f  *+     * 

* 

*  +* 

* 

*  *        * 

*Ak          ,             j     Juan  de  Fuca 

A        .      Cape 
"                 Flattery 

+ 

>5 

>* 

**  A*     ** 

** 

<** 

*                      Olympic 

48'0'0"N- 

+ 

\* 

4 

"* 

* 

•  * 

Peninsula 

*         AA± 

A 

• 

** 

A 

A 

• 

A* 

* 

** 

"1 

»          + 

*  £* 

T    * 

**        + 

* 

*  \ 

• 

+++ 

t 

A   1 

*            *         -H- 
■H- 

+ 

N 

Sanctuary  Boundary 

47  O'ff'N- 

1 1 

r- 

Figure  6 

Locations  of  small  cetaceans  seen  during  ship  surveys  off  the  northern  Washington 
coast  between  1995  and  2002. 


cally  yielding  an  estimated  abundance  of  876  porpoises 
(CV=0.30,  Table  8).  Confidence  intervals  for  some  of  the 
annual  estimates  overlapped  among  years. 


of  28  occasions.  All  but  one  of  these  sightings  were  of 
a  single  animal.  Elephant  seals  were  seen  in  all  years 
except  1998  and  2002. 


Pinnipeds 

Pinnipeds  were  not  as  frequently  observed  as  cetaceans 
•  Table  3,  Fig.  7).  The  two  most  pelagic  species  observed 
in  this  region,  northern  fur  seals  and  elephant  seals, 
were  the  most  commonly  seen  pinnipeds.  Northern  fur 
seals  were  observed  every  year  except  2002  on  a  total 


Habitat  differences 

A  number  of  broad  habitat  patterns  emerged  for  differ- 
ent groups  of  species  based  on  their  association  with 
water  depth  and  distance  from  shore  during  the  ship 
surveys  from  1995  to  2000  (Table  9,  data  were  not  avail- 
able for  2002).  Five  species  were  seen  in  shallow  waters 


Calambokidis  et  at:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


575 


125=0'0"W 
I 


Vancouver 
Island 


Barkley 
Sound 


Pinnipeds 

•  Northern  Fur  Seal 
■  Northern  Sea  Lion 

*  Sea  Otter 

a  Elephant  Seal 

+  Harbor  Seal 

♦  California  Sea  Lion 


La  Perouse 
Bank 


Strait  of 
I    Juan  de  Fuca 

Cape 
Flattery 


Olympic 
Peninsula 


1 


Sanctuary  Boundary 


Figure  7 

Locations  of  pinnipeds  and  sea  otters  {Enhydra  lutris)  seen  during  ship  surveys  off 
the  northern  Washington  coast  between  1995  and  2002. 


(<100  m).  Gray  whales  and  sea  otters  were  seen  in  the 
shallowest  water  of  all  species  with  average  water  depths 
of  just  20  and  22  m.  respectively;  they  also  were  the  only 
two  species  for  which  sightings  averaged  less  than  10  km 
from  shore.  The  three  other  species — harbor  porpoise, 
California  sea  lions,  and  northern  sea  lions  (JEumetopias 
jubatus) — were  seen  in  slightly  deeper  waters  (averag- 
ing 34  to  91  m)  and  farther  from  shore  (averaging  11  to 
23  km).  The  five  species  that  were  predominantly  found 
at  mid-shelf  depths  (mean  depths  at  100-200  m)  were 
humpback  whales,  killer  whales,  Dall's  porpoises,  harbor 


seals,  and  minke  whales  (Balaenoptera  acutorostrata). 
Species  seen  far  from  shore  (>40  km)  and  also  in  deepest 
waters  (>200  m)  included  Pacific  white-sided  dolphins, 
Risso's  dolphins,  elephant  seals,  and  northern  fur  seals. 
All  of  these  species  are  known  to  feed  along  the  conti- 
nental slope  or  off  the  shelf. 

Distances  from  the  shelf  break  for  different  species 
did  not  fall  into  as  clear  a  pattern  as  water  depth  and 
distance  from  shore  (Table  9).  This  disparity  may  be 
the  result  of  the  varied  habitat  (with  canyons  cutting 
through  the  study  area)  and  the  lack  of  much  effort  off 


576 


Fishery  Bulletin  102(4) 


Table  9 

Summary  of  habitat  and  oceanograph 

ic  parameters 

for  si 

jhtings 

of  different  species  during  ship 

surveys 

from  1995  to  2000. 

Distance  fr 

om 

Distance  from 

Sea  surface 

Species 

Water  depth 

iml 

shore  (km I 

shelf  l  km) 

temp.  (°C) 

n 

Mean 

SD 

n 

Mean 

SD 

7f 

Mean 

SD 

n 

Mean 

SD 

Baleen  whales 

Humpback  whale 

153 

144 

87 

153 

43.8 

14.9 

153 

8.4 

6.7 

101 

13.9 

1.6 

Gray  whale 

5 

20 

8 

5 

5.0 

2.0 

5 

26.1 

8.1 

5 

14.4 

1.9 

Minke  whale 

3 

106 

67 

3 

41.2 

27.7 

3 

8.0 

6.5 

3 

16.1 

0.9 

Unidentified  large  whale 

21 

189 

280 

21 

40.5 

18.4 

21 

8.0 

7.3 

18 

15.4 

1.3 

Unidentified  whale 

1 

197 

— 

1 

36.3 

— 

1 

0.1 

— 

1 

13.0 

— 

Odontocetes 

Dall's  porpoise 

90 

167 

118 

90 

40.1 

14.9 

90 

5.6 

5.5 

72 

14.3 

1.7 

Harbor  porpoise 

38 

58 

70 

38 

16.3 

15.6 

38 

17.2 

11.6 

29 

13.9 

1.7 

Pacific  white-sided  dolphin 

24 

689 

505 

24 

65.6 

25.7 

24 

8.3 

8.7 

20 

15.0 

0.8 

Northern  right-whale  dolphin 

1 

259 

— 

1 

16.2 

— 

1 

0.7 

— 

Risso's  dolphin 

9 

552 

310 

9 

55.4 

21.4 

9 

4.9 

5.2 

8 

14.4 

1.3 

Killer  whale 

12 

148 

58 

12 

28.8 

15.0 

12 

5.9 

4.7 

7 

14.1 

1.1 

Unidentified  delphinid 

19 

219 

253 

19 

37.4 

17.4 

19 

5.7 

6.7 

19 

14.5 

1.5 

Pinnipeds  and  otters 

Harbor  seal 

15 

102 

154 

15 

17.3 

11.0 

15 

15.5 

12.0 

14 

14.2 

1.4 

Elephant  seal 

20 

466 

370 

20 

46.2 

18.5 

20 

3.8 

5.0 

16 

14.7 

1.8 

California  sea  lion 

4 

91 

74 

4 

22.8 

15.2 

4 

9.3 

14.2 

1 

13.9 

— 

Steller  sea  lion 

4 

34 

18 

4 

11.3 

5.4 

4 

18.5 

6.6 

3 

13.6 

0.4 

Northern  fur  seal 

22 

382 

349 

22 

47.1 

17.1 

22 

3.1 

3.7 

21 

14.3 

1.4 

Sea  otter 

3 

22 

1 

3 

8.9 

0.5 

3 

25.5 

18.1 

3 

12.6 

0.4 

Unidentified  pinniped 

13 

170 

144 

13 

30.5 

18.4 

13 

8.0 

8.1 

11 

14.5 

1.9 

All  sightings 

457 

205 

251 

457 

39.1 

20.1 

457 

8.4 

8.4 

352 

14.3 

1.6 

the  continental  shelf.  Despite  most  of  our  effort  being  on 
the  continental  shelf,  the  presence  of  several  deep  can- 
yons in  addition  to  the  shelf  edge,  resulted  in  all  species 
being  an  average  of  less  than  11  km  from  the  200  m 
depth  contour.  The  average  surface  water  temperature 
for  species  that  were  seen  also  varied  and  was  likely 
both  a  function  of  distance  from  shore  and  association 
with  upwelling  areas  (Table  9).  Sea  otters  were  seen  in 
the  coldest  waters  (12.6°C)  where  they  are  predominant- 
ly found.  Among  the  more  offshore  species,  humpback 
whales,  tended  to  be  seen  in  colder  waters  (13.9°C)  than 
most  other  offshore  species,  probably  because  of  their 
association  with  offshore  upwelling  areas. 


Discussion 

Although  humpback  whales  were  the  most  abundant 
large  cetacean  seen  in  our  study,  their  numbers  of  a 
few  hundred  still  appear  to  be  substantially  lower  than 
numbers  found  prior  to  whaling.  Commercial  hunting 
of  humpback  whales  occurred  in  the  1900s  from  coastal 
whaling  stations  in  northern  California,  Washington, 


and  British  Columbia.  In  these  areas,  thousands  of 
humpback  whales  were  killed  over  a  relatively  short 
time  period  (less  than  10  years)  before  catches  dropped 
precipitously  with  the  depletion  of  the  population.  At 
the  south  end  of  our  study  area,  1933  humpback  whales 
were  taken  from  a  station  at  Bay  City  (in  Grays  Harbor), 
Washington,  from  1911  to  1925  (Scheffer  and  Slipp, 
1948).  To  the  north,  5638  humpback  whales  were  taken 
from  British  Columbia  stations  from  1908  to  1967,  of 
which  60f'f  (3393)  were  taken  from  1908  to  1917  from  the 
two  southernmost  whaling  stations  on  Vancouver  Island 
closest  to  our  study  area  (Gregr  et  al.,  2000;  Nichol  et 
al.,  2002).  Additionally,  1871  humpback  whales  were 
taken  from  two  stations  in  northern  California  from 
1919  to  1926  (Clapham  et  al.,  1997).  Although  these 
hunts  encompassed  areas  larger  than  our  study  area, 
the  number  killed  in  short  periods  dwarfs  even  the  sum 
of  our  abundance  estimates  for  Washington  and  British 
Columbia  and  the  estimate  of  under  1000  whales  esti- 
mated in  the  1990s  for  California.  Oregon,  and  Wash- 
ington (Calambokidis  and  Barlow,  2004).  Moreover, 
humpback  whales  have  not  returned  to  some  of  the 
areas  where  they  were  once  found  prior  to  commercial 


Calambokidis  et  al.:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


577 


whaling;  humpback  whales  were  commonly  observed  in 
the  inside  waters  of  Washington  and  British  Columbia 
(Scheffer  and  Slipp,  1948;  Webb,  1988)  and  have  not 
returned  to  these  areas  in  any  numbers  (Calambokidis 
and  Steiger,  1990). 

The  distribution  of  humpback  whales  within  our  study 
area  was  not  uniform  and  indicated  that  some  specific 
areas  were  important  feeding  habitat  for  this  recovering 
species.  The  region  between  the  Juan  de  Fuca  Canyon 
and  the  shelf  edge  (the  Prairie) — the  mouth  of  Bark- 
ley  Canyon  and  Swiftsure  Bank — was  the  area  where 
humpback  whales  were  concentrated.  In  monthly  aerial 
surveys  in  1989-90  by  Green  et  al.,5  there  were  only  a 
total  of  13  sightings  of  25  humpback  whales  along  the 
entire  Washington  coast  between  July  and  September. 
Over  half  of  those  sightings  were  in  the  Prairie  area. 

Our  line-transect  estimates  revealed  that  about  100 
humpback  whales  inhabit  the  northern  Washington 
coast  waters  each  summer;  substantially  more  (over 
500),  however,  were  present  in  2002.  Although  this  is  a 
small  number  compared  to  estimates  of  just  under  1000 
humpback  whales  for  California,  Oregon,  and  Wash- 
ington (Calambokidis  and  Barlow,  2004),  our  study 
area  encompasses  a  relatively  small  area  and  reflects  a 
high  density  of  animals.  Additionally  our  line-transect 
estimates  were  not  corrected  for  any  missed  animals; 
therefore  they  are  probably  biased  slightly  downward. 

Despite  the  relatively  high  density  of  humpback 
whales  in  this  region,  the  photographic  identification 
data  indicated  that  a  relatively  small  number  of  indi- 
viduals use  the  area  consistently.  Both  the  line-transect 
and  the  photographic  identification  data  (increasing 
capture-recapture  estimates,  as  well  as  decreased  pro- 
portions of  animals  sighted  multiple  years)  showed  that 
the  number  of  whales  using  this  region  has  increased  in 
recent  years.  The  growing  number  of  whales  in  this  re- 
gion could  be  either  the  result  of  births  or  immigration 
into  this  area.  Births  alone  could  not  account  for  this  in- 
crease, especially  because  the  proportion  of  whales  that 
were  mothers  with  calves  seen  in  this  region  was  not 
high.  There  did  not  appear  to  be  a  shift  in  distribution 
of  animals  from  areas  to  the  south  because  interchange 
with  those  areas  dropped  from  1999  to  2002.  The  most 
likely  explanation  for  these  changes  is  that  there  was  a 
shift  of  animals  from  feeding  areas  from  the  north  into 
this  region  beginning  in  the  late  1990s. 

This  interchange  of  humpback  whales  with  feeding  ar- 
eas to  the  south  provides  new  insight  into  the  structure 
of  humpback  whale  feeding  aggregations.  In  a  study 
that  examined  interchange  rates  of  humpback  whales 


6  Green,  G.  A.,  J.  J.  Brueggeman,  R.  A.  Grotefendt,  and  C. 
E.  Bowlby.  1992.  Cetacean  distribution  and  abundance 
off  Oregon  and  Washington,  1989-1990.  In  Oregon  and 
Washington  marine  mammal  and  seabird  surveys  (J.  J. 
Brueggeman,  ed.  I,  100  p.  Final  report  of  OCS  Study  MMS 
91-0093  by  Ebasco  Environmental,  Bellevue,  Washington, 
and  Ecological  Consulting,  Inc.,  Portland,  Oregon,  for  the 
Minerals  Management  Service,  Pacific  OCS  Region,  U.S. 
Dept.  of  Interior,  770  Paseo  Camarillo,  Camarillo,  CA 
93010. 


along  the  west  coast,  Calambokidis  et  al.  (1996)  iden- 
tified northern  Washington  as  a  demographic  bound- 
ary between  the  whales  feeding  area  along  California, 
Oregon,  and  Washington  and  those  to  the  north.  The 
larger  sample  reported  here  shows  the  same  general 
pattern  of  decreasing  interchange  with  distance  from 
a  feeding  area  as  that  reported  previously  for  whales 
off  California  (Calambokidis  et  al.,  1996).  The  decreas- 
ing rate  of  interchange  with  distance  among  feeding 
areas  does  not  allow  for  a  clear  demarcation  between 
feeding  areas,  however,  as  suggested  by  Calambokidis 
et  al.  (1996).  Although  humpback  whales  demonstrate 
site  fidelity  to  specific  feeding  locations,  their  feeding 
aggregations  may  not  have  clear  boundaries  and  may 
occupy  overlapping  ranges. 

The  commercial  whaling  data  also  tended  to  support 
the  existence  of  somewhat  discrete  feeding  areas  off  the 
west  coast  of  the  United  States  and  British  Columbia. 
Commercial  whaling  resulted  in  the  depletion  of  hump- 
back whales  off  British  Columbia  by  1917,  whereas  the 
numbers  taken  off  Washington  and  California  did  not 
decline  until  the  mid-1920s  (Scheffer  and  Slipp,  1948; 
Clapham  et  al.,  1997;  Gregr  et  al,  2000). 

The  relatively  small  proportion  of  mothers  with  calves 
identified  in  our  study  is  consistent  with  findings  off 
California  and  Oregon  (Steiger  and  Calambokidis, 
2000).  Steiger  and  Calambokidis  reported  reproductive 
rates  along  the  California,  Oregon,  and  Washington 
coasts  that  are  lower  than  those  reported  for  other  re- 
gions in  southeastern  Alaska  and  the  North  Atlantic 
(Clapham  and  Mayo,  1987,  1990;  Baker  et  al.,  1992; 
von  Ziegesar  et  al.,  1994).  In  aerial  transect  surveys, 
no  humpback  whale  calves  were  seen  among  the  68 
humpbacks  observed  off  the  Oregon  and  Washington 
coasts  in  1989-90  (Green  et  al.5).  If  geographic  segre- 
gation is  occurring  by  humpback  mothers  and  calves, 
as  was  suggested  by  Steiger  and  Calambokidis  (2000), 
this  northern  region  is  not  the  area  where  mothers  and 
calves  are  congregated.  It  is  interesting  to  note,  how- 
ever, that  mothers  and  calves  were  distributed  around 
the  periphery  of  the  main  feeding  region — a  finding  that 
suggests  that  a  more  local  segregation  may  be  occurring. 
A  bias  in  sampling  would  occur  if  large  concentrations  of 
whales  are  targeted  and  mother  with  calves  feeding  on 
the  perimeter  of  these  groups  were  underrepresented. 

In  contrast  to  humpback  whales,  no  other  large  ror- 
quals (blue,  fin,  or  sei  whales)  were  observed  during 
any  of  our  ship  or  small  boat  surveys.  Likewise,  these 
species  were  absent  in  other  recent  surveys  of  Wash- 
ington waters  (Wahl,  1977;  Von  Saunder  and  Barlow. 
1999;  Shelden  et  al.,  2000;  Green  et  al.5),  although 
they  were  seen  in  surveys  farther  offshore  in  surveys  in 
July  1994  (Thomason  et  al.6).  Fin  whales  were  common 


,;  Thomason,  J.,  M.  Dahlheim,  S.  E.  Moore,  J.  Braham,  K. 
Stafford,  and  C.  Fox.  1997.  Acoustic  investigations  of 
large  cetaceans  off  Oregon  and  Washington:  NOAA  ship 
Surveyor  (21  July-1  August  1994),  27  p.  Final  report  by 
the  National  Marine  Mammal  Laboratory,  7600  Sand  Point 
Way  NE  F/AKC3,  Seattle.  WA  98115. 


578 


Fishery  Bulletin  102(4) 


in  Washington  waters  in  the  early  1900s  when  they 
were  the  second  most  commonly  killed  species  by  Bay 
City  whalers  (Scheffer  and  Slipp,  1948).  Blue  and  sei 
whales  were  less  common,  although  they  were  present 
historically  (Scheffer  and  Slipp.  1948).  Although  Bay 
City  whaling  stations  (in  Grays  Harbor,  Washington) 
were  closed  after  humpback  whales  were  depleted,  se- 
rial depletion  of  whale  populations  continued  off  British 
Columbia  waters,  beginning  with  humpback  and  blue 
whales,  then  with  fin  and  sperm  whales,  and  finally 
with  sei  whales  (Gregr  et  al.,  2000). 

No  sperm  whales  or  beaked  whales  were  seen  during 
our  surveys,  although  our  study  area  did  not  include  the 
deeper  waters  where  we  would  expect  to  find  these  spe- 
cies. Most  of  the  sperm  whales  (90%)  seen  by  Green  et 
al.5  off  Washington  and  Oregon  were  present  in  deeper 
offshore  waters  outside  of  our  study  area. 

The  other  cetacean  species  not  seen  in  our  surveys 
that  have  been  reported  to  occur  off  Washington  his- 
torically included  northern  right  whale  (Eubalaena 
japonica),  pygmy  sperm  whale  iKogia  breviceps),  false 
killer  whale  iPseudorea  er-assidens),  short-finned  pilot 
whale  (Globicephala  macrorhynchus),  and  striped  dol- 
phin (Stenella  coeruleoalba)  (Scheffer  and  Slipp,  1948). 
Sightings  of  northern  right  whales  throughout  the  east- 
ern North  Pacific  are  scarce;  there  have  been  only  a 
small  number  of  sightings  since  the  1960s  (Brownell 
et  al.,  2001).  Several  of  these  sightings,  however,  have 
been  off  the  northern  Washington  coast  (Fiscus  and 
Niggol,  1965;  Osborne  et  al.,  1988;  Rowlett  et  al.,  1994). 
The  primary  reason  for  the  paucity  of  sightings  in  the 
eastern  North  Pacific  in  recent  decades  is  due  to  the  il- 
legal take  of  372  right  whales  in  the  early  to  mid-1960s 
by  the  USSR  (Brownell  et  al.,  2001;  Doroshenko7). 

Although  some  small  cetacean  species  such  as  Pacific 
white-sided  dolphins  and  Risso's  dolphins  were  sighted 
frequently  on  our  surveys,  they  were  not  as  common 
as  in  some  previous  surveys  (Green  et  al.5),  probably 
because  our  coverage  was  concentrated  in  shallower 
waters  inside  the  shelf  break.  In  contrast  to  our  find- 
ings of  a  number  of  species  seen  near  the  shelf  edge. 
Wahl  (1977)  reported  that  most  marine  mammal  species 
off  central  Washington  tended  to  be  in  either  inshore 
or  in  deeper  offshore  waters  and  only  killer  whales  and 
Dall's  porpoises  regularly  used  the  slope  waters  (13- 
45  km  offshore). 

It  is  difficult  to  make  abundance  estimates  of  Dall's 
porpoise  because  of  their  proclivity  to  approach  ships 
(Buckland  and  Turnock,  1992).  If  they  begin  to  ap- 
proach the  ship  before  the  observer  sights  them,  the  es- 
timate is  biased  upwards,  which  would  be  the  case  with 
our  estimate.  Our  estimate  would  also  have  a  downward 
bias  because  we  did  not  attempt  to  adjust  for  animals 
missed  even  if  they  were  on  the  track  line. 


Doroshenko,  N.  V.  2000.  Soviet  whaling  for  blue,  gray, 
bowhead  and  right  whales  in  the  North  Pacific  Ocean. 
1961-1979.  In  Soviet  whaling  sata  (1949-1979),  p. 
96-103.  Center  for  Russian  Environmental  Policv.  Vavilov 
St.  26,  Moscow  117071,  Russia. 


All  three  types  of  killer  whales  (residents  [both 
northern  and  southern],  transients,  and  offshore  type) 
were  identified  in  the  waters  off  northern  Washington. 
These  sightings  are  interesting  because  of  concerns 
about  killer  whale  populations,  especially  the  southern 
resident  community  that  has  declined  in  recent  years. 
Although  killer  whales  have  been  intensely  studied  in 
inside  waters  of  the  Pacific  Northwest,  little  has  been 
known  about  their  use  of  outside  waters,  where  they 
may  spend  large  portions  of  their  lives.  Little  is  known 
about  the  offshore  type  of  killer  whales,  which  is  be- 
lieved to  be  a  distinct  race  of  killer  whale  that  has  only 
recently  been  described.  These  whales  are  believed  to 
be  found  usually  in  large  groups  along  the  continental 
shelf  but  also  have  been  seen  in  inland  waters  (Ford  et 
al.,  1994;  Dahlheim  et  al.,  1997).  All  three  sightings  of 
the  offshore  form  were  just  west  of  the  Juan  de  Fuca 
canyon  on  the  Prairie;  the  closest  sighting  to  shore  was 
37  km  (30  animals  on  15  July  1997). 


Acknowledgments 

We  are  grateful  to  those  who  assisted  with  this  study. 
This  work  was  supported  by  the  Olympic  Coast  National 
Marine  Sanctuary  and  Southwest  Fisheries  Science 
Center  (Jay  Barlow,  COTR).  Many  people  contributed 
to  this  study.  Jennifer  Quan,  Richard  Rowlett,  Anne 
Nelson,  and  Annie  Douglas  worked  on  the  ship  surveys. 
We  thank  the  ship  personnel  on  board  the  McArthur  and 
Agate  Passage.  Researchers  who  helped  with  small  boat 
work  included  Joe  Evenson  and  Todd  Chandler.  Photo- 
graphs of  whales  from  this  area  were  also  contributed  by 
L.  Baraff,  R.  Baird,  P.  Bloedel,  V.  Deeke.  P.  Ellifrit,  G. 
Ellis,  J.  Evenson,  B.  Gisborne.  B.  Halliday,  H.  Hunt,  S. 
Mizroch,  K.  Rasmussen,  J.  Wilson  and  SWFSC  research- 
ers. Permission  to  survey  in  Canadian  waters  was  given 
by  the  Dept.  of  Fisheries  and  Oceans.  Lisa  Schlender, 
Kristin  Rasmussen,  and  Annie  Douglas  organized  and 
conducted  the  photographic  matching  with  the  help  of 
many  interns  at  Cascadia  Research.  DKE  and  Graeme 
Ellis  identified  the  killer  whales;  Oscar  Torres  assisted 
with  the  photographic  matching.  Data  analyses  and 
mapping  were  conducted  with  the  help  of  Scot  McQueen 
at  ESRI  and  Tom  Williams. 


Literature  cited 

Baker,  C.  S.,  A.  Perry,  and  L.  M.  Herman. 

1992.     Population  characteristics  of  individually  identi- 
fied humpback  whales  in  southeastern  Alaska:  summer 
and  fall  1986.     Fish.  Bull.  90:429-437. 
Barlow,  J.,  C.  W.  Oliver,  T.  1).  Jackson,  and  B.  L.  Taylor. 

1988.     Harbor  porpoise.  Phocoena  phocoena,  abundance 
estimate  for  California,  Oregon  and  Washington:  II. 
Aerial  surveys.      Fish.  Bull.  86:433-444. 
Bigg,  M.  A..  G.  M.  Ellis,  J.  K.  B.  Ford,  and  K.  C.  Balcomh. 

1987.     Killer  whales.   A  study  of  their  identification, 
genealogy  and  natural  history  in  British  Columbia  and 


Calambokidis  et  al.:  Distribution  and  abundance  of  marine  mammals  off  the  northern  Washington  coast 


579 


Washington  State,  79  p.     Phantom  Press  and  Publish- 
ers, Nanaimo,  British  Columbia,  Canada. 
Bigg,  M.  A.,  P.  F.  Olesiuk,  G.  M.  Ellis,  J.  K.  B.  Ford,  and 
K.  C.  Balcomb  III. 

1990.     Social  organization  and  genealogy  of  resident  killer 
whales  (Orcinus  orca)  in  the  coastal  waters  of  British 
Columbia  and  Washington  State.     Rep.  Int.  Whal.  Coram. 
(spec,  issue  121:383-405. 
Black,  N.  A.,  A.  Schulman-Janiger,  R.  L.  Ternullo,  and 
M.  Guerrero-Ruiz. 

1997.     Killer  whales  of  California  and  western  Mexico: 
a  catalog  of  photo-identified  individuals.     NOAA  Tech. 
Memo.  NOAA-TM-NMFS-SWFSC-247,  174  p. 
Brownell,  R.  L.,  Jr.,  P.  J.  Clapham,  T.  Miyashita,  and  T.  Kasuya. 

2001.  Conservation  status  of  North  Pacific  right  whales.  J. 
Cet.  Res.  Manag.  (spec,  issue  21:269-286. 

Buckland.  S.  T„  and  B.  J.  Turnock. 

1992.     A  robust  line  transect  method.     Biometrics  48:901- 
909. 
Calambokidis,  J.,  and  J.  Barlow. 

2004.     Abundance  of  blue  and  humpback  whales  in  the 
eastern  North  Pacific  estimated  by  capture-recapture  and 
line-transect  methods.     Mar.  Mamm.  Sci.  20:63-85. 
Calambokidis,  J.,  J.  C.  Cubbage,  G.  H.  Steiger,  K.  C.  Balcomb, 
and  P.  Bloedel. 

1990.     Population  estimates  of  humpback  whales  in  the 
Gulf  of  the  Farallones,  California.     Rep.  Int.  Whal. 
Commn  (spec,  issue  12):325-333. 
Calambokidis,  J.,  J.  D.  Darling,  V.  Deeke,  P.  Gearin,  M.  Gosho, 
W.  Megill,  C.  M.  Tombach,  D.  Goley,  C.  Toropova,  and 

B.  Gisborne. 

2002.  Abundance,  range  and  movements  of  a  feeding 
aggregation  of  gray  whales  (Eschrichtius  robustus)  from 
California  to  southeastern  Alaska  in  1998.  J.  Cet.  Res. 
Manag.  4:267-276. 

Calambokidis.  J.,  and  G.  H.  Steiger. 

1990.     Sightings  and  movements  of  humpback  whales  in 
Puget  Sound.  Washington.     Northwestern  Nat.  71:45- 
49. 
Calambokidis,  J.,  G.  H.  Steiger,  J.  R.  Evenson,  K.  R.  Flynn, 
K.  C.  Balcomb,  D.  E.  Claridge,  P.  Bloedel.  J.  M.  Straley, 

C.  S.  Baker,  O.  von  Ziegesar,  M.  E.  Dahlheim.  J.  M.  Waite, 
J.  D.  Darling,  G.  Ellis,  and  G.  A.  Green. 

1996.     Interchange  and  isolation  of  humpback  whales  off 
California  and  other  North  Pacific  feeding  grounds.     Mar. 
Mamm.  Sci.  12:215-226. 
Calambokidis,  J.,  G.  H.  Steiger,  K.  Rasmussen,  J.  Urban  R., 
K.  C.  Balcomb,  P.  Ladron  de  Guevara  P..  M.  Salinas  Z., 
J.  K.  Jacobsen,  C.  S.  Baker,  L.  M.  Herman,  S.  Cerchio,  and 
J.  D.  Darling. 

2000.     Migratory  destinations  of  humpback  whales  that 
feed  off  California,  Oregon  and  Washington.     Mar.  Ecol. 
Prog.  Ser.  192:295-304. 
Clapham,  P.  J.,  and  C.  A.  Mayo. 

1987.     Reproduction  and  recruitment  of  individually 
identified  humpback  whales,  Megaptera  novaeangliae, 
observed  in  Massachusetts  Bay,  1979-85.     Can.  J.  Zool. 
65:2852-2863. 
1990       Reproduction  of  humpback  whales  [Megaptera 
novaeangliae)  observed  in  the  Gulf  of  Maine.     Rep. 
Int.  Whal.  Comm.  (spec,  issue  12C171-175. 
Clapham,  P.  J..  S.  Leatherwood,  I.  Szczepaniak,  and 
R.  L.  Brownell,  Jr. 

1996.  Catches  of  humpback  and  other  whales  from  shore 
stations  at  Moss  Landing  and  Trinidad,  California, 
1919-1926.     Mar.  Mamm.  Sci.  13:368-394. 


Dahlheim,  M.  E„  D.  K.  Ellifrit,  J.  D.  Swenson. 

1997.  Killer  whales  of  southeast  Alaska:  a  catalogue  of 
photo-identified  individuals,  79  p.  Day  Moon  Press, 
Seattle,  WA. 

Darling,  J.  D. 

1984.  Gray  whales  (Eschrichtius  robustus)  off  Vancou- 
ver Island,  British  Columbia.  In  The  gray  whale, 
Eschrichtius  robustus  (M.  L.  Jones,  S.  Swartz,  and 
S.  Leatherwood,  eds.l,  p.  267-287.  Academic  Press, 
Orlando,  FL. 

Fiscus,  C.  H.,  and  K.  Niggol. 

1965.  Observations  of  cetaceans  off  California,  Oregon, 
and  Washington.  U.S.  Fish  and  Wildl.  Serv.,  Special 
Scientific  Report-Fisheries  498,  27  p.  U.S.  Fish  and 
Wildl.  Serv.,  Washington,  D.C. 

Ford,  J.  K.  B.,  G.  M.  Ellis,  and  K.  C.  Balcomb. 

1994.  Killer  whales,  102  p.  UBC  Press,  Vancouver, 
British  Columbia. 

Green,  G.  A.,  J.  J.  Brueggeman,  R.  A.  Grotefendt,  and 
C.  E.  Bowlby. 

1995.  Offshore  distances  of  gray  whales  migrating  along 
the  Oregon  and  Washington  coasts,  1990.  Northwest 
Sci.  69:223-227. 

Gregr,  E.  J.,  L.  Nichol,  J.  K.  B.  Ford,  G.  Ellis,  and  A.  W.  Trites. 
2000.     Migration  and  populations  structure  of  northeastern 
Pacific  whales  off  coastal  British  Columbia:  an  analysis 
of  commercial  whaling  records  from  1908-1967.     Mar. 
Mamm.  Sci.  16:699-727. 
Hammond,  P.  S. 

1986.     Estimating  the  size  of  naturally  marked  whale 
populations  using  capture-recapture  techniques.     Rep. 
Int.  Whal.  Comm.  (spec,  issue  81:253-282. 
Huelsbeck,  D.  R. 

1988.     Whaling  in  the  precontact  economy  of  the  central 

northwest  coast.     Arctic  Anthropol.  25:1-15. 
Jameson,  R.  J.,  K.  W.  Kenyon,  S.  Jeffries,  and  G.  R.  Van- 

Blaricom. 
1986.     Status  of  a  translocated  sea  otter  population  and 
its  habitat  in  Washington.     Murrelet  67:84-87. 
Jameson,  R.  J.,  K.  W.  Kenyon,  A.  M.  Johnson,  and 
H.  M.  Wright. 

1982.     History  and  status  of  translocated  sea  otter  pop- 
ulations in  North  America.     Wildl.  Soc.  Bull.  10:100- 
107. 
Jeffries,  S.,  H.  Huber,  J.  Calambokidis  and  J.  Laake. 

2003.     Trends  and  status  of  harbor  seals  in  Washington 
State:  1978-1999.     J.  Wildl.  Manag.  67:201-219. 
Kvitek,  R.  G.,  P.  J.  Iampietro,  and  C.  E.  Bowlby. 

1998.  Sea  otters  and  benthic  prey  communities:  a  direct 
test  of  the  sea  otter  as  keystone  predator  in  Washington 
state.     Mar.  Mamm.  Sci.  14:895-902. 

Kvitek,  R.  G.,  J.  S.  Oliver,  A.  R.  DeGange,  and  B.  S.  Anderson. 
1992.     Sea  otters  and  benthic  prey  communities  in  Wash- 
ington State.     Mar.  Mamm.  Sci.  5:266-280. 
Nichol,  L.  M..  E.  J.  Gregr,  R.  Flinn,  J.  K.  B.  Ford,  R.  Gurney, 
L.  Michaluk,  and  A.  Peacock. 

2002.     British  Columbia  commercial  whaling  catch  data 
1908  to  1967:  a  detailed  description  of  the  B.C.  histori- 
cal whaling  database.     Can.  Tech.  Rep.  Fish.  Aquat. 
Sci.  2396,  vii  +76  p. 
Osborne,  R.,  J.  Calambokidis,  and  E.  M.  Dorsey. 

1988.     A  guide  to  marine  mammals  of  greater  Puget 
Sound,  191  p.     Island  Publishers,  Anacortes,  WA. 
Osmek,  S.,  J.  Calambokidis,  J.  Laake,  P.  Gearin,  R.  DeLong, 
S.  Jeffries,  and  R.  Brown. 

1996.  Assessment  of  the  status  of  harbor  porpoises  (Pho- 


580 


Fishery  Bulletin  102(4) 


coena  phocoena)  in  Oregon  and  Washington  waters. 
U.S.  Dep.  of  Commerce,  NOAA  Tech.  Memo.,  NMFS- 
AFSC-76,  46  p. 
Rowlett,  R.  A.,  G.  A.  Green,  C.  E.  Bowlby,  and  M.  A.  Smultea. 
1994.     The  first  photographic  documentation  of  a  north- 
ern right  whale  off  Washington  State.     Northwestern 
Nat.  75:102-104. 
Scheffer,  V.  B„  and  J.  W.  Slipp. 

1948.     The  whales  and  dolphins  of  Washington  State 
with  a  key  to  the  cetaceans  of  the  west  coast  of  North 
America.     Am.  Midi.  Nat.  39:257-337. 
Seber,  G.  A.  F. 

1982.     The  estimation  of  animal  abundance  and  related 
parameters,  2nd  ed.,  654  p.     Griffin,  London. 
Shelden,  K.  E.  W.,  D.  J.  Rugh,  J.  L.  Laake,  J.  M.  Waite, 
P.  J.  Gearin,  and  T.  R.  Wahl. 

2000.     Winter  observations  of  cetaceans  off  the  northern 
Washington  coast.     Northwestern  Nat.  81:54-59. 
Steiger,  G.  H.,  and  J.  Calambokidis. 

2000.     Reproductive   rates   of  humpback   whales   off 
California.     Mar.  Mamm.  Sci.  16:220-239. 


Swan,  J.  G. 

1868.     The  Indians  of  Cape  Flattery,  109  p.     Shorey  Pub- 
lications, Seattle,  WA. 
Von  Saunder,  A.,  and  J.  Barlow. 

1999.     A  report  of  the  Oregon,  California  and  Washington 
line-transect  experiment  lORCAWALE)  conducted  in 
West  Coast  waters  during  summer/fall  1996.     NOAA 
Tech.  Memo.  NOAA-TM-NMFS-SWFSC-264,  49  p. 
von  Ziegesar,  O..  E.  Miller,  and  M.  E.  Dahlheim. 

1994.     Impacts  of  humpback  whales  in  Prince  William 
Sound.     In  Marine  mammals  and  the  Exxon  Valdez  IT. 
R.  Loughlin.  ed.),  p.   173-191.     Academic  Press,  San 
Diego,  CA. 
Wahl,  T.  R. 

1977.     Sight  records  of  some  marine  mammals  offshore 
from  Westport,  Washington.     Murrelet  58:21-23. 
Webb,  L.  W. 

1988.  On  the  Northwest.  Commercial  whaling  in  the 
Pacific  Northwest  1790-1967,  425  p.  Univ.  British 
Columbia  Press.  Vancouver,  Canada. 


581 


Abstract — The  reproductive  biology 
of  male  franciseanas  iPontoporia 
blainvillei),  based  on  121  individu- 
als collected  in  Rio  Grande  do  Sul 
State,  southern  Brazil,  was  studied. 
Estimates  on  age,  length,  and  weight 
at  attainment  of  sexual  maturity  are 
presented.  Data  on  the  reproductive 
seasonality  and  on  the  relationship 
between  some  testicular  characteris- 
tics and  age.  size,  and  maturity  status 
are  provided.  Sexual  maturity  was 
assessed  by  histological  examina- 
tion of  the  testes.  Seasonality  was 
determined  by  changes  in  relative  and 
total  testis  weight,  and  in  seminifer- 
ous tubule  diameters.  Testis  weight, 
testicular  index  of  maturity,  and 
seminiferous  tubule  diameters  were 
reliable  indicators  of  sexual  maturity, 
whereas  testis  length,  age,  length, 
and  weight  of  the  dolphin  were  not. 
Sexual  maturity  was  estimated  to  be 
attained  at  3.6  years  (CI  957c  =  2.7- 
4.5)  with  the  DeMaster  method  and 
3.0  years  with  the  logistic  equation. 
Length  and  weight  at  attainment  of 
sexual  maturity  were  128.2  cm  (CI 
95<£  =  125.3-131.1  cm)  and  26.4  kg  (CI 
95^=24.7-28.1  kg),  respectively.  It 
could  not  be  verified  that  there  was 
any  seasonal  change  in  the  testis 
weight  and  in  the  seminiferous  tubule 
diameters  in  mature  males.  It  is  sug- 
gested that  at  least  some  mature  males 
may  remain  reproductively  active 
throughout  the  year.  The  extremely 
low  relative  testis  weight  indicates 
that  sperm  competition  does  not  occur 
in  the  species.  On  the  other  hand, 
the  absence  of  secondary  sexual  char- 
acteristics, the  reversed  sexual  size 
dimorphism,  and  the  small  number 
of  scars  from  intrassexual  combats  in 
males  reinforce  the  hypothesis  that 
male  combats  for  female  reproductive 
access  may  be  rare  for  franciscana. 
It  is  hypothesized  that  P.  blainvillei 
form  temporary  pairs  (one  male  copu- 
lating with  only  one  female)  during 
the  reproductive  period. 


Reproductive  biology  of  male  franciseanas 
(Pontoporio  blainvillei)  (Mammalia:  Cetacea) 
from  Rio  Grande  do  Sul,  southern  Brazil* 


Daniel  Daniiewicz 

Grupo  de  Estudos  de  Mamiferos  Aquaticos  do  Rio  Grande  do  Sul  (GEMARS) 

Rua  Felipe  Nen,  382/203 

Porto  Alegre  90440-150,  Brazil 

Present  address:    Laboratono  de  Dinamica  Populacional-Pontilicia  Universidade  Catolica  do 

Rio  Grande  do  Sul  (PUCRS) 

Av.  Ipiranga,  6681 

Porto  Alegre  90619-900,  Brazil 
Email  address  (for  D  Daniiewicz):  Daniel. Danilewicza1  terra  com  br 


Juan  A.  Claver 

Alejo  L.  Perez  Camera 

Area  Histologia  y  Embnologia  Facultad  de  Ciencias  Vetermarias 
Universidad  de  Buenos  Aires 
Ar  Chorroarin  280  C1427CWO 
Buenos  Aires,  Argentina 

Eduardo  R.  Secchi 

Laboratono  de  Mamiferos  Mannhos,  Museu  Oceanografico  "Prof.  Eliezer  C.  Rios" 
Fundacao  Universidade  Federal  do  Rio  Grande,  Cx  P.  379 
Rio  Grande  96200,  Brazil 

Nelson  F.  Fontoura 

Laboratono  de  Dinamica  Populacional-Pontificia  Universidade  Catolica  do 

Rio  Grande  do  Sul  (PUCRS) 

Av.  Ipiranga,  6681 

Porto  Alegre  90619-900,  Brazil 


Manuscript  submitted  4  October  2002 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
18  May  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:581-592  (2004). 


The  franciscana  (Pontoporia  blainvil- 
lei) is  a  small  dolphin  endemic  to  the 
coastal  waters  of  the  southwestern 
Atlantic  Ocean.  The  distribution  of 
this  species  ranges  from  Golfo  Nuevo 
(42°35'S;  64°48'W),  Chubut  Province, 
Argentina  (Crespo  et  al.,  1998)  to 
Itaunas  (18°25'S;  30°42'W),  Espirito 
Santo,  southeastern  Brazil  (Moreira 
and  Siciliano,  1991)  (Fig.  1). 

The  franciseanas  coastal  habitat 
makes  it  vulnerable  to  being  caught 
as  incidental  catch  in  gill  nets  and 
trammel  nets  throughout  most  of  the 
species  range  (e.g.,  Praderi  et  al., 
1989;  Corcuera  et  al.,  1994;  Secchi  et 
al.,  2003).  Because  of  its  vulnerability 
as  bycatch,  the  franciscana  has  been 
considered  the  most  impacted  small 


cetacean  in  the  southwestern  Atlantic 
Ocean  (Secchi  et  al.,  2002).  In  the 
Rio  Grande  do  Sul  coast,  southern 
Brazil,  this  species  has  been  subject 
to  an  intense  bycatch  in  gill  nets  for 
at  least  three  decades  (Moreno  et  al., 
1997;  Secchi  et  al.,  1997;  Ott,  1998; 
Ott  et  al.,  2002).  The  annual  mor- 
tality of  franciseanas  in  this  region 
was  estimated  to  range  from  several 
hundred  up  to  about  a  thousand  indi- 
viduals (Ott  et  al.,  2002).  Simulations 


^Contribution  012  from  the  Grupo  de 
Estudos  de  Mamiferos  Aquaticos  do  Rio 
Grande  do  Sul  (GEMARS),  Rua  Felipe 
Neri,  382/203,  Porto  Alegre  90440-150, 
Brazil. 


582 


Fishery  Bulletin  102(4) 


Figure  1 

Map  of  the  study  area  showing  the  locations  along  the  southern  coast  of  Brazil 
where  franciscanas  were  caught  as  bycatch  between  1992  and  1998. 


studies  on  the  effects  of  incidental  captures  on  francis- 
canas in  Rio  Grande  do  Sul  were  carried  out  by  using 
available  data  on  vital  rates,  stock  size,  and  bycatch 
estimates  (e.g.,  Secchi,  1999;  Kinas,  2002).  All  these 
studies  showed  that  there  is  a  decline  in  franciscana 
abundance  in  this  region. 

Although  the  reproductive  biology  of  the  female  fran- 
ciscanas have  been  studied  in  detail  in  Uruguay  (Ka- 
suya  and  Brownell,  1979;  Harrison  et  al.,  1981),  Rio 
Grande  do  Sul  (Danilewicz  et  al.,  2000;  Danilewicz, 
2003),  and  Rio  de  Janeiro  (Ramos,  1997),  there  are  few 
data  about  male  reproduction.  Kasuya  and  Brownell 
(1979)  presented  information  on  male  reproduction  for 
Uruguay,  although  their  small  sample  size  precluded 
them  from  estimating  age  and  size  at  attainment  of 
sexual  maturity. 

In  the  Rio  Grande  do  Sul  coast,  franciscanas  are 
known  to  reproduce  seasonally;  births  occur  from  Oc- 
tober to  early  February  (about  75"7r  from  October  to  De- 
cember). Because  the  gestation  period  was  estimated  to 
last  about  11.2  months,  mating  and  conception  may  take 
place  between  November  and  early  March  I  Danilewicz, 
2003).  Seasonal  changes  in  testicular  size  and  activity 
have  been  used  to  infer  or  corroborate  mating  seasons 
in  some  cetacean  species  (e.g.,  Neimanis  et  al.,  2000  I. 
Nevertheless,  it  is  not  known  if  male  franciscanas  also 
undergo  seasonal  changes  in  the  testicular  activity. 

In  this  study,  we  describe  the  reproductive  biology  of 
male  franciscanas  from  Rio  Grande  do  Sul  and  present 
evidences  for  the  species'  mating  system. 


Materials  and  methods 

Sampling  procedures 

Data  and  samples  collected  from  121  specimens  inci- 
dentally caught  (889c)  or  beached  (12%)  along  the  Rio 
Grande  do  Sul  coast  between  1992  and  1998  were  used 
for  the  analysis  on  reproduction  of  male  franciscanas. 
The  sampling  of  the  incidentally  caught  animals  was 
carried  out  through  the  monitoring  of  the  commercial 
fishery  fleet  from  Rio  Grande  (32°08'S;  52°05'W)  and 
Tramandai/Imbe  (29°58'S,  50°07'W).  Stranded  dolphins 
were  sampled  from  systematic  beach  surveys  conducted 
in  an  area  with  an  extension  of  270  km  of  sandy  beaches, 
between  Torres  (29°19'S,  49°43'W)  and  Lagoa  do  Peixe 
(31°15'S,  50°54'W). 

Not  all  information  could  be  collected  from  each  car- 
cass; therefore  sample  sizes  varied  among  parameters. 
Standard  length  (SL,  n  =  118)  was  measured  by  following 
the  guidelines  established  by  the  American  Society  of 
Mammalogy  (1961).  The  animals  were  weighed  (rc  =  97) 
and  teeth  were  extracted  and  preserved  dried  or  in  a 
1:1  mix  of  glycerin  and  alcohol  (70%).  Testes  and  epi- 
didymis were  removed  and  fixed  in  10'>  formalin. 

Age  determination 

Age  was  estimated  by  counting  the  growth  layer  groups 
(GLGs)  in  thin,  longitudinal  sections  of  teeth  (ra=47). 
The  teeth  were  decalcified  in  nitric  acid  or  in  RDO 


Danilewicz  et  al.:  Reproductive  biology  of  male  Pontopona  blainvillei  from  Rio  Grande  do  Sul,  southern  Brazi 


583 


(a  commercial  mixture  of  acids)  and  sectioned  on  a  freez- 
ing microtome.  The  15-20  fim  sections  were  stained  with 
Mayer's  hematoxylin  and  mounted  on  microscope  slides 
with  Canadian  balsam  or  in  glycerin.  Poor  and  off-center 
sections  were  discarded  in  favor  of  new  preparations. 
Three  readers  counted  independently  the  number  of 
growth  layer  groups  in  both  the  dentine  and  cementum. 
When  reader  estimates  differed,  the  sections  were  reex- 
amined together  and  a  best  estimate  was  agreed  upon. 
In  this  study,  we  considered  one  GLG  to  represent  one 
year  of  age,  which  is  the  accepted  model  for  the  francis- 
cana  (Kasuya  and  Brownell,  1979;  Pinedo,  1991;  Pinedo 
and  Hohn;  2000). 

Reproduction 

In  the  laboratory,  the  testes  were  separated  from  the 
epididymis,  weighed  to  the  nearest  0.01  g  (n  =  107),  and 
measured  in  three  dimensions  (length  and  two  diam- 
eters perpendicular  to  each  other  in  the  middle  of  the 
testis)  to  the  nearest  0.1  mm  (;?  =  104).  The  mean  of 
these  two  diameters  was  called  mean  testis  diameter. 
The  weight  of  one  of  the  gonads  could  not  be  recorded  on 
some  occasions  («  =  8)  and  we  assumed  that  both  testes 
had  the  same  weight.  Then,  relative  testis  weight  was 
determined  as  the  ratio  of  the  combined  testis  weight  to 
the  animal  weight. 

A  1-cm3  subsample  of  each  testis  from  the  central 
portion  of  the  organ  was  removed  and  examined  by 
using  standard  histological  preparations.  The  tissue 
was  embedded  in  paraffin,  sectioned  in  4-10  fim  thick 
slides  through  a  manual  microtome,  and  stained  with 
hematoxylin  and  eosin  (H&E).  Male  sexual  maturity 
status  was  determined  by  examining  the  testicular 
sections  at  a  magnification  of  lOOx.  In  this  study,  we 
followed  the  classification  criteria  suggested  by  Hohn 
et  al.  (1985): 

1  Immature — seminiferous  tubules  containing  main- 
ly spermatogonias.  Abundant  interstitial  tissue 
present  between  the  seminiferous  tubules  and  lu- 
men totally  closed. 

2  Pubertal — seminiferous  tubules  containing  sper- 
matogonias and  spermatocytes.  Less  interstitial 
tissue  present  between  the  seminiferous  tubules 
than  in  immature  animals.  The  lumen  is  partially 
opened. 

3  Mature — seminiferous  tubules  containing  sper- 
matogonias, spermatocytes,  spermatids  and,  in 
many  cases,  spermatozoa.  Interstitial  tissue  al- 
most nonexistent  between  the  seminiferous  tu- 
bules. The  lumen  is  totally  opened. 

The  diameters  of  ten  random  circular  seminiferous 
tubules  were  measured  for  each  specimen  («  =  93)  with 
a  scale  present  in  the  lens  of  the  microscope  in  order 
to  calculate  the  seminiferous  tubule  mean  diameter.  A 
maturity  index  (MI)  was  calculated  as  the  ratio  of  the 
combined  testes  weight  by  the  combined  testes  length 

aw/iL). 


An  analysis  of  the  variation  along  the  year  of  the  val- 
ues of  relative  and  combined  testes  weight,  and  seminif- 
erous tubule  mean  diameter,  was  employed  to  assess  re- 
productive seasonality.  Values  of  these  parameters  were 
compared  between  months  when  mating  and  conception 
occur  ("reproductive  months":  November-March)  and 
months  when  they  not  occur  ( "nonreproductive  months": 
April-October).  In  order  to  increase  the  sample  size  of 
mature  animals  collected  in  reproductive  months,  data 
on  testes  weight  from  mature  male  franciscanas  from 
Uruguay  were  included  in  the  analysis  (data  supplied 
by  Kasuya1). 

The  mean  age  at  attainment  of  sexual  maturity  (ASM) 
was  estimated  through  the  DeMaster  (1978)  method  and 
the  logistic  regression. 

The  DeMaster  (1978)  equation  computes  the  mean 
as 

ASM  =  £a(/Q  -f^), 


where  fa   =    the  fraction  of  sexually  mature  animals  in 
the  sample  with  age  a; 
j     =    the  age  of  the  youngest  sexually  mature 

animal  in  the  sample;  and 
k    =    the  age  of  the  oldest  sexually  immature 
animal  in  the  sample. 

The  variance  of  the  DeMaster  method  estimate  is  cal- 
culated as 

k 
variASM)=^[ifaa-fa)/Na-D], 
«=j 

where  N   =  the  total  number  of  animals  aged  a. 

The  logistic  regression  approach  fits  a  sigmoid  curve 
representing  the  probability  that  a  franciscana  of  age 
a  is  sexually  mature  to  the  distribution  of  sexually 
mature  and  immature  animals  by  age  as 

Y  =  l/(l+e°+/,v)    or   In  (1/1-1)  =  a  +  bx, 

where   x  =   the  age  of  the  dolphin; 

b  =  the  slope  of  the  regression;  and 
a  =  the  intercept. 

To  obtain  the  age  when  50%  of  the  animals  are  sexu- 
ally mature  (Y=0.5),  the  last  equation  is  simplified  as 
ASM  =  -alb. 

Mean  length  and  weight  at  sexual  maturity  was  also 
estimated  by  the  DeMaster  (1978)  method,  by  substitut- 
ing age  for  length  and  weight,  respectively.  The  meth- 
od was  slightly  modified,  as  suggested  by  Ferrero  and 
Walker  (1993),  and  was  calculated  as 


1  Kasuya,  T.  1970-73.  Unpubl.  data.  Teikyo  University  of 
Science  and  Technology.  Uenohara,  Yamanashi  Prefecture, 
409-0193,  Japan. 


584 


Fishery  Bulletin  102(4) 


Cmax 

LSM=  X  Uft-ft-i), 

Cmin 

where  Cmax   =  the  length  or  weight  class  of  the  largest  or 
heaviest  sexually  immature  animal; 
Cmin    =  the  length  or  weight  class  of  the  smallest 
or  lightest  sexually  mature  animal; 
L   =  the  lower  value  of  the  length  or  weight 

class  t;  and 
ft    =  fraction  of  mature  animals  in  the  length 
or  weight  class  t. 

The  specimens  were  pooled  into  length  and  weight  inter- 
vals of  4  cm  and  4  kg,  respectively. 

The  estimated  variance  of  this  method  is  also  modi- 
fied and  is  calculated  as 


var(MS)  =  M>2  £  [(/; (1 -/",)/ A^,  -l]. 


(x=33.6  mm),  respectively.  The  weight  and  length  of  the 
right  testes  ranged  from  0.17  to  9.98  g  (.v=2.62  g)  and 
from  17.9  to  60.0  mm  (.v=34.5  mm),  respectively.  The 
relationship  of  testes  weight  and  testes  length  resulted 
in  significant  regression  (P<0.0001)  and  correlation 
<r2=0.91;  F=823.9;  P<0.0001;  y=0.000012x333).  The  male 
with  the  heaviest  relative  testes  weight  was  141.6  cm  in 
length  and  31.2  kg  in  weight,  and  its  combined-testes 
weight  was  20.1  g,  which  is  0.064%  of  its  total  weight. 
The  mean  of  the  relative  testes  weight  from  23  mature 
males  was  0.036%  of  their  total  weight. 

The  testes  of  the  franciscana  are  characterized  by  a 
high  lateral  symmetry.  There  was  no  statistically  dif- 
ference in  weight  (£=-0.09;  P=0.93;  n=ll)  and  length 
(*=-0.4;  P=0.69;  «  =  100)  between  testes  of  the  same 
animal.  A  strong  correlation  was  found  between  left 
and  right  testes  length  (6  =  0.95;  F=1073.0;  r2  =  0.92; 
P<0.0001;  ra=100;  y  =  1.232  x095)  and  between  left 
and  right  testes  weight  (6  =  0.99;  F=7262.8;  r2  =  0.99; 
P<0.0001;  ra=71;  y  =  1.02.v1  °i,  where  x  and  y  represent 
values  of  the  left  and  right  testis,  respectively. 


where  N:   =   the  number  of  specimens  in  the  length  or 
weight  class  f;  and 
w    =  the  interval  width,  a  constant  equal  to  4  in 
these  cases. 

For  estimating  age,  length,  and  weight  at  sexual  matu- 
rity, pubertal  animals  were  grouped  together  with  imma- 
ture animals. 


Results 

The  weight  and  length  of  the  left  testes  ranged  from 
0.23  to  10.42  g  (.r=2.60  g)  and  from  15.7  to  59.7  mm 


§       200  -i 

yi 

A                                                _^_ '        A 

<»       175- 

A                                        &- — A~^      A 

3 
J3 

Aaa              a, ■& 

3        150  ■ 

-^'"~&A 

tfl 

^** — '             A 

Zi 

2        125- 
0) 

_D/-D 

•p       100  - 

□    y^ 

E 

/S      m 

(D 

ofs 

en 

D^Jn          D 

S         50- 

/£]      □ 

a> 

E 

ra        25  - 

73 

c 

0) 

5               0             2             4             6             8            10           12           14           16           18           20           22 

Combined  testes  weight  (g) 

Figure  2 

Relationship  between  combined-testes  weight  and  mean  seminiferous 

tubule  diameter  in  immature  (open  boxesi,  pubertal  (filled  boxes),  and 

mature  (triangles)  male  franciscanas  (Pontoporia  blainvillei)  from  Rio 

Grande  do  Sul  (re=59). 

Seminiferous  tubule  diameter 

A  nonlinear  regression  demonstrated  positive  allometry 
(6>0.333)  of  the  seminiferous  tubule  diameter  to  the 
combined  testicular  weight  (6  =  0.39;  95%  CI=0.35-0.44) 
(Fig.  2),  and  a  strong  correlation  between  these  two  vari- 
ables (F=343.6;  r2  =  0.86;  P<0.0001;  y=59.4.x°  39). 

The  relationship  between  the  seminiferous  tubule  di- 
ameter and  testes  length  is  shown  in  Figure  3  and  the 
relationship  between  the  seminiferous  tubule  diameter 
and  standard  length  is  shown  in  Figure  4.  In  immature 
males,  there  was  almost  no  increase  in  the  seminif- 
erous tubule  diameter  with  the  increase  of  standard 
length  (0.26  jim/cm)  and  total  weight  (0.5  ^m/kg).  In 
mature  males,  however,  seminiferous  tu- 
bule diameter  was  significantly  correlat- 
ed with  standard  length  (6  =  1.06;  F=4.4; 
r2  =  0.18;  P=0.048;  y  =  1.4775.v-43.572) 
and  there  was  no  correlation  with  total 
weight  (6  =  0.23;  P=1.28;  r2=0.07;  P=0.27; 
y  =  1.6132.\-+108.54). 

The  differences  of  the  seminiferous  tu- 
bule mean  diameters  were  statistically 
significant  between  immature,  pubertal, 
and  mature  male  franciscanas  (ANOVA, 
Fs=255.4;  df=87;  P<0.001). 

Combined-testes  weight  and  length, 
and  sexual  maturity 


There  was  almost  no  increment  in  mass 
of  the  combined-testes  weight  in  imma- 
ture dolphins.  An  increment  of  only  about 
2.0  g  in  the  combined-testes  mass  was 
observed  in  animals  of  70.0  to  125.0  cm  in 
length.  For  dolphins  about  120.0-130.0  cm 
in  length,  the  combined-testes  mass  sud- 
denly increased  (Fig.  5),  indicating  the 


Danilewicz  et  al.:  Reproductive  biology  of  male  Pontopona  blainvillei  from  Rio  Grande  do  Sul,  southern  Brazil 


585 


attainment  of  sexual  maturity.  The  rate  of 
testes-mass  gain  was  0.05  g/cm  for  imma- 
ture and  0.28  g/cm  for  mature  dolphins. 
All  animals  with  combined-testes  weight 
higher  than  5.0  g  were  sexually  mature, 
and  this  finding  may  indicate  that  this 
parameter  can  be  used  as  a  reliable  indi- 
cator of  sexual  maturity  in  male  fran- 
ciscanas  from  Rio  Grande  do  Sul  (Table 
1).  However,  the  large  variation  in  testes 
weight  after  the  attainment  of  sexual 
maturity  precludes  a  correlation  between 
testes  mass  and  standard  length. 

Although  testes  length  increases  pro- 
gressively as  standard  length  increases 
(Fig.  6),  there  is  no  abrupt  increase  in 
testes  length  at  the  moment  of  attain- 
ment of  sexual  maturity,  as  observed  in 
the  testes  mass.  A  nonlinear  regression 
(exponential)  best  fits  this  relationship 
(y =4. 5444  e0.0163.vi.  As  opposed  to  testes 
mass,  there  is  a  considerable  overlap  in 
the  values  of  testes  length  of  immature, 
pubertal,  and  mature  franciscanas  (Table 
1),  which  makes  testes  length  a  less  reli- 
able predictor  of  sexual  maturity  than 
testes  mass. 

Age,  length,  and  weight  at  sexual  maturity 

Forty-seven  specimens  in  the  sample  pro- 
vided information  on  age  and  reproduc- 
tive status  (35  immature  or  pubertal,  and 
12  mature).  The  oldest  immature  animal 
was  5  years  old  and  the  youngest  mature 
was  2  years  old  (Table  1).  Average  age  at 
attainment  of  sexual  maturity  was  esti- 
mated to  be  3.6  years  by  the  DeMaster 
method  (SD  =  0.47;  95%  CI  =2.7-4.5)  and 
3.0  years  by  the  logistic  equation  Y  =  1/(1+ 
eo.74-2  23i)  The  age  structure  of  the  sample 
studied  is  presented  in  Figure  7. 

Sexual  maturity  in  relation  to  standard 
length  was  estimated  for  110  males.  The 
smallest  mature  and  the  largest  immature 
males  were  120.5  and  137.5  cm  long,  respectively.  The  av- 
erage length  at  sexual  maturity  was  128.2  cm  (SD=1.49; 
959r  CI=125. 3-131.1  cm).  Sexual  maturity  in  relation  to 
total  weight  was  estimated  for  90  males.  The  lightest 
mature  and  the  heaviest  immature  males  were  20.3  and 
29.7  kg,  respectively.  The  average  weight  at  sexual  matu- 
rity was  26.4  kg  (SD  =  0.88;  95%  01=24.7-28.1  kg). 

Index  of  testicular  maturity 

The  differences  of  the  mean  index  of  testicular  maturity 
between  immature  (0.03),  pubertal  (0.04),  and  mature 
(0.11)  dolphins  were  statistically  significant  (ANOVA, 
Fs=210.0  df=101,  P<0.001).  There  was  almost  no  overlap 
in  the  values  of  this  index  between  mature  specimens 


225  -I 

~      200 

of     17S 

E  c 

fS        150' 

E  c/> 

i     O         125' 

f  E     too. 

c 

°          50  - 

A 

A                          A^&— - & — 

^       A       A  A 
A 

■            A 

16    18    20    22    24    26    28    30    32    34    36    38    40    42    44    46    48    50    52    54    56    58    60    62 

hnsrns  hnlers  (ww) 

Figure  3 

Relationship  between  testes  length  and  mean  diameter  of  seminifer- 

ous tubules  in  immature  (open  boxes),  pubertal  (filled  boxes),  and 

mature  (trianglesi  male  franciscanas  (Pontoporia  blainvillei  i  in  Rio 

Grande  do  Sul  (ra=54).  Data  from  pubertal  animals  are  not  included 

in  the  curves. 

5  g 


205  ■ 

A 

185  - 
165  - 
145  - 

A                  3 

A^      AAA         A^ 

A1       A           A  A 
A                  A         A 

A 

A 

125  ■ 

*f» 

105  - 

85  ■ 
65  ■ 

45  ■ 

D 

D 

rjPrP 

D 

1 

is? 

crP^1 

25  - 

70   75   80   85   90   95  100  105  110  115  120  125  130  135  140  145  150  155  160 

itanoaro  lentth  (nm) 

Figure  4 

Relationship  between  standard  length  and  mean  diameter  of  seminif- 
erous tubules  in  immature  (open  boxes),  pubertal  (filled  boxesl.  and 
mature  (trianglesi  male  franciscanas  (Pontoporia  blainvillei)  in  Rio 
Grande  do  Sul  l;i=91>. 


and  immature  and  pubertal  specimens  (Table  1).  All 
males  with  an  index  value  higher  than  0.07  were  sexu- 
ally mature.  These  results  indicate  that  the  index  of 
testicular  maturity  is  a  very  good  indicator  of  sexual 
maturity  for  franciscanas. 

Reproductive  seasonality 

The  null  hypothesis  that  the  combined  and  relative  testis 
weight  would  be  higher  in  the  months  when  the  females 
are  reproductively  active  is  rejected.  No  increase  in  the 
testes  weight  was  observed  during  the  months  when 
most  mating  occurs  (Fig.  8).  There  were  no  statistically 
significant  differences  in  the  combined-testes  weight 
(ANOVA,  Fs=2.28;  df=34;  P=0.48;  n=35)  and  relative 


586 


Fishery  Bulletin  102(4) 


£ 
o 
O 


V     a 


nnj 


60      70      80      90     100     110     120     130     140     150     160 

Standard  length  V;ma 

Figure  5 

Relationship  between  standard  length  and  combined-testes  weight  in 
immature  (open  boxes),  pubertal  (filled  boxes),  and  mature  (triangles) 
male  franciscanas  ^Pontoporia  blainvillei)  in  Rio  Grande  do  Sul  (tt=79). 
Data  from  pubertal  animals  are  not  included  in  the  curves. 


testes  weight  (ANOVA,  Fs=2.42;  df=29; 
P=0.76;  /?  =  30)  between  reproductive  and 
nonreproductive  months. 

The  analyses  of  the  variation  in  the 
diameter  of  the  seminiferous  tubules 
throughout  the  year  also  did  not  indicate 
that  the  testes  undergo  seasonal  changes. 
However,  it  is  important  to  view  this  re- 
sult with  caution  because  the  sample  size 
of  mature  males  (and  therefore  the  infor- 
mation on  tubules  diameter  collected  in 
reproductive  months)  was  small  (n=3). 
However,  the  presence  of  spermatids  or 
spermatozoa  (or  both)  in  the  seminiferous 
tubules  may  be  also  regarded  as  a  direct 
evidence  of  testicular  activity.  Three  ma- 
ture males  (119<-)  in  the  sample  presented 
seminiferous  tubules  with  spermatids  or 
spermatozoa  (or  both)  and  were  collected 
in  nonreproductive  months  (May,  June, 
and  August).  Although  the  epididymis  of 


Table  1 

Summarized  information  on  age 

length 

mass,  and  testicul 

ar  characteristics  for  ma 

e  franciscanas  iPontoporia 

blainvillei)  in  the 

Rio  Grande  do  Sul  at  different  sexual  m 

aturity  stages. 

Characteristics  and  maturity  state 

;? 

Mean 

Standard  deviation 

Range 

Age  (years) 

Immature 

31 

1.29 

1.01 

0-5 

Pubertal 

4 

2.0 

0.82 

1-3 

Mature 

12 

3.8 

1.14 

2-6 

Standard  length  (cm) 

Immature 

62 

111.2 

13.62 

70.0-137.5 

Pubertal 

7 

118.5 

9.75 

107.8-132.5 

Mature 

37 

133.7 

7.71 

120.5-155.0 

Total  mass  (kg) 

Immature 

53 

19.0 

5.6 

4.95-29.7 

Pubertal 

6 

21.4 

4.62 

17.1-28.0 

Mature 

30 

29.9 

5.22 

20.25-41.5 

Mean  diameter  of  seminiferous  t 

ubules 

(/./m) 

Immature 

54 

69.6 

12.2 

50.0-105.0 

Pubertal 

6 

95.0 

19.2 

74.5-121.2 

Mature 

33 

154.1 

21.7 

113.0-197.0 

Combined  testes  mass  (g) 

Immature 

63 

1.59 

0.84 

0.33-4.78 

Pubertal 

7 

2.73 

1.28 

1.30-4.8 

Mature 

37 

10.24 

3.94 

4.27-20.08 

Testes  length  (mm) 

Immature 

62 

27.2 

4.9 

15.7-35.5 

Pubertal 

7 

32.6 

6.2 

25.0-41.0 

Mature 

35 

45.4 

5.6 

31.6-59.7 

Index  of  testicular  maturity 

Immature 

61 

0.03 

0.01 

0.01-0.06 

Pubertal 

7 

0.04 

0.01 

0.02-0.06 

Mature 

36 

0.11 

0.03 

0.05-0.18 

Danilewicz  et  al.:  Reproductive  biology  of  male  Pontoporia  blainvillei  from  Rio  Grande  do  Sul,  southern  Brazil 


587 


a  subsample  of  10  mature  males  were 
examined  histologically,  we  did  not  find 
any  sign  of  spermatozoa. 


Discussion 

The  high  bilateral  uniformity  in  tes- 
ticular weight  and  length  presented 
by  the  franciscana  is  a  characteristic 
shared  with  many  other  cetacean  spe- 
cies. Studies  on  the  striped  dolphin, 
Stenella  coeruleoalba  (Miyazaki,  1977), 
the  common  dolphin,  Delphinus  delphis 
(Collet  and  Saint  Girons,  1984),  the 
sperm  whale,  Physeter  macrocephalus 
(Mitchell  and  Kozicki,  1984),  and  the 
dusky  dolphin,  Lagenorhynchus  obscu- 
rus  (van  Waerebeek  and  Read,  1994), 
among  others,  demonstrate  the  same 
pattern  of  testis  symmetry.  Given  the 
similar  dimensions  of  both  testes  in 
franciscanas,  it  is  possible  to  extrapo- 
late the  combined-testes  weight  by 
weighing  only  one  testis  without  intro- 
ducing bias  in  the  analysis.  It  is  recom- 
mended, however,  that  the  weight  of 
the  testes  should  be  presented  without 
the  epididymis  weight,  as  it  was  pre- 
sented in  the  most  extensive  compara- 
tive study  on  the  subject  (Kenagy  and 
Trombulak,  1986). 

There  is  a  negative  allometry  of  the 
seminiferous  tubule  diameter  in  rela- 
tion to  testis  length,  standard  length, 
and  total  weight.  This  pattern  is  ac- 
centuated in  immature  males,  in  which 
the  tubule  diameters  remain  almost 
unchanged  with  the  increase  of  the 
other  variables.  The  lack  of  values  for 
tubule  diameters  in  the  testes  weight 
interval  (2.5-6.0  g)  and  testes  length 
interval  (34-42  mm)  just  before  the 
attainment  of  sexual  maturity  (Figs. 
2  and  3)  indicates  that  the  increase  in 
tubule  size  in  relation  to  sexual  matu- 
rity must  occur  very  quickly,  probably 
when  the  tubules  are  between  85  and 
125  urn  in  diameter. 

Attainment  of  sexual  maturity 

Length  and  weight  at  attainment  of 
sexual  maturity  of  male  franciscanas 
in  Rio  Grande  do  Sul  are  very  similar 
to  those  values  estimated  in  previous 
estimates  for  Uruguay  (Table  2).  In  con- 
trast to  the  present  study,  Kasuya  and 
Brownell  (1979)  calculated  mean  length 
at  sexual  maturity  for  Uruguay  as  the 


65  - 

60  ■ 
CO 
E        55. 

E 

<        50- 
'2       « 

*  c  a  <,,  -0,0163* 

y=  4,5444e 

r!  =  0.72                             A                   . 

A             ^    4  A    &/*     A 

< 

O)       35  ■ 
c/1 

8)       30- 

n 

25 
20 

°    DEh  3- — °          □ 

65       70       75       80       85       90       95      100     105     110     115     120     125     130     135     140     145     150     155     160 

atcnecre  Ignhth  Ydma 

Figure  6 

Relationship  between  standard  length  and  testes  length  in  immature  (open 

boxes),  pubertal  (filled  boxes),  and  mature  (triangles)  male  franciscanas 

(Pontoporia  blainvillei)  in  Rio  Grande  do  Sul  i;;  =  99l. 

60 -I 

50- 

o 

c 

CD 

g  "»■ 

CD 

< 

CD       30- 

O 

CD 
EC 

1 

10- 

.      1 

1    I    ■    .    _ 

0  year                  1  year 

2  years                3  years                4  years               5  years               6  years 

Figure  7 

Age  structure  of  male  franciscanas  (Pontoporia  blainvillei )  collected  in 

Rio  Grande  do  Sul  (n=48). 

Figure  8 

Relationship  between  month  and  relative  testes  weight  in  mature  male 
franciscanas  (Pontoporia  blainvillei)  (7j=31)  (l  =  January,  12  =  December; 
filled  boxes=the  nonreproductive  months,  open  boxes=reproductive  months). 
Bars  indicate  25^  and  75%  percentiles. 


588 


Fishery  Bulletin  102(4) 


Table  2 

Comparison  between  average  age,  weight,  and  length  at 
sexual  maturity  between  male  and  female  franciscanas 
from  Rio  Grande  do  Sul  and  Uruguay.  The  means  of  the 
animals  from  Rio  Grande  do  Sul  were  estimated  by  using 
the  DeMaster  method  (modified!  and  those  from  Uruguay- 
were  estimated  by  using  a  linear  regression  to  determine 
the  moment  when  50(7(  of  the  animals  are  mature. 


Rio  Grande  do  Sul 

Uruguay' 

Males     Females2 

Males          Females 

Age 

Weight  (kg) 
Length  (cm) 

3.6              3.7 

26.6           32.6 

127.4         138.9 

2-4                  2.7 

25.0-29.0      33.0-34.0 

131.4              140.3 

'  Data  from  Uruguay  were  compiled  from  Kasuya  and  Brownell 

11979' 
2  Data  from  Damlewicz  (2003). 


length  where  50r*  of  the  dolphins  were  mature  through 
a  linear  regression.  Applying  this  same  approach,  a  LSM 
of  125.4  cm  was  estimated  for  Rio  Grande  do  Sul — a 
value  still  very  similar  to  the  Uruguay  estimate.  Male 
franciscanas  attain  sexual  maturity  at  less  length  and 
weight  than  do  females  in  Rio  Grande  do  Sul  (Danile- 
wicz,  2003),  as  observed  previously  in  Uruguay  (Kasuya 
and  Brownell,  1979)  and  Rio  de  Janeiro  (Ramos  et  al., 
2000). 

This  is  the  first  estimate  of  mean  age  at  sexual  ma- 
turity presented  for  male  franciscanas.  Kasuya  and 
Brownell  (1979)  could  not  calculate  ASM  for  Uruguay 
because  of  their  small  sample  size  («=25).  Nevertheless, 
Kasuya  and  Brownell  suggested  that  sexual  maturity 
is  attained  when  males  are  between  2  and  4  years  of 
age.  Franciscanas  from  Rio  de  Janeiro  were  considered 
mature  when  they  were  older  than  2  years  of  age  and 
larger  than  115.0  cm  in  length  (Ramos  et  al.,  2000). 
However,  histological  analysis  of  the  testes  was  not 
performed  and  Ramos  et  al.  employed  indirect  methods 
to  determine  sexual  maturity.  Nevertheless,  despite  the 
uncertainties  produced  by  the  use  of  different  criteria 
to  determine  sexual  maturity,  it  was  evident  that  there 
was  substantial  difference  in  the  size  at  maturity  be- 
tween males  from  Rio  de  Janeiro  and  those  from  Rio 
Grande  do  Sul  and  Uruguay.  This  difference  is  probably 
the  result  of  the  well-known  distinct  growth  patterns  of 
the  franciscanas  from  these  two  regions  (Pinedo,  1991) 
and  does  not  necessarily  reflect  an  early  attainment 
of  sexual  maturity  in  males  from  the  Rio  de  Janeiro 
population. 

The  trade-off  between  growth  and  reproduction  is 
the  best-documented  phenotypic  trade-off  in  nature 
(Stearns,  1992)  and  has  been  studied  in  a  wide  range 
of  taxa.  Because  animals  from  Rio  de  Janeiro  invest 
less  in  growth  than  do  animals  from  Rio  Grande  do 
Sul,  it  is  still  an  open  question  whether  the  francisca- 
nas from  the  Rio  de  Janeiro  have  higher  reproductive 


rates  or  start  reproducing  earlier  than  those  from  Rio 
Grande  do  Sul. 

The  oldest  male  and  female  franciscana  ever  aged 
were  16  and  21  years-old,  respectively  (Kasuya  and 
Brownell,  1979;  Pinedo,  1994).  These  ages  contrasts 
with  the  age  distribution  found  in  the  present  study, 
where  the  oldest  specimen  analyzed  was  6  years  old. 
Similar  to  what  is  observed  in  catches  for  several  other 
small  cetacean  species  (e.g..  Hector's  dolphins — Slooten 
and  Lad,  1991;  harbor  porpoise — Read  and  Hohn.  1995), 
a  general  feature  of  incidental  catches  for  these  spe- 
cies is  the  high  entanglement  rate  of  immature  ani- 
mals. In  all  fishing  communities  studied  in  Argentina, 
Uruguay,  and  Brazil,  a  large  proportion  (>50%)  of  the 
specimens  caught  were  less  than  three  years  old  (e.g., 
Kasuya  and  Brownell,  1979;  Corcuera  et  al..  1994;  Ott, 
1998;  Di  Beneditto  and  Ramos,  2001).  Although  the 
precise  reason  for  biased  catch  rates  towards  imma- 
ture individuals  is  not  well  understood,  it  could  be  a 
combination  of  factors,  including  the  imbalanced  age- 
structure  of  local  populations  (where  there  are  fewer 
older  individuals  because  of  an  extensive  history  as 
bycatch)  and  a  behavior-related  higher  vulnerability  to 
bycatch  for  immature  individuals  (i.e.,  juveniles  can  be 
more  inquisitive  and  have  less  ocean  experience  so  that 
they  rove  into  the  area  increasing  the  chances  of  being 
entangled).  The  typically  low  proportion  of  old  animals 
in  bycatches  may  explain  the  characteristics  of  the  data 
used  in  this  study. 

Index  of  testicular  maturity 

An  index  of  testicular  maturity  may  be  very  useful 
in  studies  where  it  is  necessary  to  know  the  sexual 
maturity  of  a  large  sample  of  animals  without  the  need 
of  histological  analysis,  which  is  time  consuming  and 
requires  expertise.  Although  Hohn  et  al.  (1985)  recom- 
mended the  investigation  of  the  applicability  of  this 
indirect  index  of  sexual  maturity  for  male  cetaceans, 
the  research  on  this  subject  has  shown  no  progress.  To 
date,  the  index  of  sexual  maturity  has  been  calculated 
only  for  the  common  dolphin,  Delphinus  delphis  (Collet 
and  Saint  Girons,  1984),  and  from  the  pantropical  spot- 
ted dolphin,  Stenella  attenuate!  (Hohn  et  al.,  1985).  For 
both  species,  this  index  distinguished  satisfactorily  the 
mature  from  the  immature  and  pubertal  dolphins.  Given 
the  results  presented,  we  also  recommend  the  use  of  the 
index  of  testicular  maturity  as  an  alternative,  nonhisto- 
logical  method,  to  determine  the  sexual  maturity  of  male 
franciscanas.  Males  with  index  values  lower  than  0.05 
can  be  safely  classified  as  immature,  and  males  with 
index  values  above  0.08  can  be  classified  as  mature.  It  is 
recommended  that  for  animals  with  intermediate  values 
their  testes  be  analyzed  histologically  so  that  their 
reproductive  status  may  be  determined  definitively. 

Besides  making  intra-  and  inter-populational  com- 
parisons possible,  the  index  of  testicular  maturity  also 
permits  interspecific  comparisons  because  size  differ- 
ences between  species  are  eliminated.  The  mean  index 
of  testicular  maturity  of  mature  franciscanas  (0.12)  is 


Danilewicz  et  al.:  Reproductive  biology  of  male  Pontoporia  blamvillei  from  Rio  Grande  do  Sul,  southern  Brazil 


589 


considerably  lower  than  mature  pantropical  spotted 
dolphins  (1.9)  (Hohn  et  al.,  1985).  This  difference  is 
a  consequence  of  the  relatively  small  increase  of  the 
testes  weight  of  male  franciscanas  when  sexual  matu- 
rity is  attained.  Although  male  spotted  dolphin  show 
a  marked  increase  of  about  25-fold  in  testes  weight  at 
this  moment,  franciscanas  show  an  increment  in  testes 
weight  of  about  ninefold  only. 

Reproductive  seasonality 

The  reproductive  activities  in  male  mammals  are  usually 
restricted  to  the  periods  when  the  females  are  in  estrus 
(Lincoln,  1992).  Reproductive  seasonality  in  males  has 
been  reported  for  several  cetacean  species  and  popula- 
tions through  the  identification  of  temporal  variations 
in  the  testes  weight  and  histological  characteristics.  In 
species  where  the  reproductive  period  is  restricted  for 
a  few  months,  as  with  the  dusky  dolphin  (Lagenorhyn- 
chus  obscurus)  and  the  harbor  porpoise  (Phocoena  pho- 
coena),  the  testes  weight  presents  marked  fluctuations 
accompanying  the  reproductive  period  (Read,  1990;  van 
Waerebeek  and  Read,  1994;  Neimanis  et  al.,  2000).  Even 
in  species  with  a  diffuse  reproductive  period  (i.e.,  with 
more  than  one  peak  for  births  per  year)  as  in  the  case 
of  dolphins  of  the  genus  Stenella  in  the  tropical  Pacific, 
it  was  possible  to  detect  seasonal  variation  in  the  male 
reproductive  rhythm  (Perrin  et  al.,  1976,  Hohn  et  al., 
1985). 

Because  of  the  known  seasonality  for  births  for  fran- 
ciscana  (Kasuya  and  Brownell,  1979,  Harrison  et  al., 
1981,  Danilewicz,  2003),  it  would  be  expected  that  the 
males  would  accompany  the  female  rhythm,  decreas- 
ing or  even  ceasing  testicular  activity  in  autumn  and 
winter  months.  Kasuya  and  Brownell  (1979)  examined 
the  seasonal  change  in  testes  weight  in  the  months  of 
January,  June,  and  December.  From  our  knowledge  of 
the  species'  reproduction  period,  testes  weight  would 
be  expected  to  be  higher  in  December  and  January. 
However,  the  authors  could  not  confirm  this  predic- 
tion and  attributed  the  lack  of  seasonality  to  the  small 
sample  size  of  mature  animals.  Nevertheless,  the  lack 
of  seasonality,  even  when  the  testes  weight  of  the  ma- 
ture males  from  Rio  Grande  do  Sul  are  included,  may 
indicate  what  is  occurring  in  the  population,  and  not  be 
a  bias  introduced  by  a  small  sample  size. 

In  species  that  possess  small  testes,  as  in  the  case  of 
the  franciscana,  the  variation  in  the  testicular  activity 
may  be  better  reflected  by  changes  in  the  diameter  of 
the  seminiferous  tubules  and  the  rate  of  spermatogen- 
esis rather  than  by  changes  in  the  testes  weight.  Nev- 
ertheless, the  preliminary  results  about  these  charac- 
teristics (mature  males  with  spermatids  or  spermatozoa 
[or  both]  in  the  seminiferous  tubules  in  nonreproductive 
months  and  little  monthly  variation  in  the  diameter 
of  the  seminiferous  tubules)  also  do  not  support  the 
hypothesis  of  a  male  reproductive  seasonality.  The  com- 
bination of  results  presented  here  indicates  that  testicu- 
lar activity  is  not  completely  interrupted  in  all  males 
within  the  population,  and  that  at  least  some  of  them 


may  remain  capable  of  fertilizing  females  during  the 
year.  This  conclusion  is  supported  by  the  observation 
of  pregnancies  outside  the  normal  gestation  season  and 
that  the  births  resulting  from  these  pregnancies  were 
estimated  to  take  place  in  September  and  in  late  March 
(Danilewicz,  2003). 

The  hormone  and  sperm  production  by  the  testes  dur- 
ing periods  when  the  females  are  not  able  to  reproduce 
may  represent  an  unnecessary  energetic  expense  by 
the  male  (Dewsbury,  1982)  and  may  be  an  explanation 
for  the  period  of  reproductive  inactivity  for  males  of 
several  mammal  species.  In  species  with  large  relative 
testes  weight,  the  maintenance  of  high  levels  of  sperm 
production  in  the  testes  is  a  considerable  energetic  cost 
for  the  individual.  However,  as  discussed  earlier,  this 
is  definitely  not  the  case  for  the  franciscana.  For  this 
reason,  we  suggest  that  the  small  energy  investment 
in  producing  sperm  all  over  the  year,  due  to  the  small 
testicular  mass,  may  be  an  evolutionary  advantage  for 
male  franciscanas  in  case  of  the  appearance  of  off-sea- 
son reproductive  females. 

Franciscana  reproductive  strategy 

Although  important  advances  in  the  knowledge  of  fran- 
ciscana behavior  in  the  wild  have  been  made  (e.g.,  Bor- 
dino  et  al.,  1999;  Bordino,  2002),  there  is  no  information 
on  the  species'  reproductive  behavior  and  its  mating 
strategy  remains  unknown.  Relative  testis  weight, 
sexual  size  dimorphism,  and  secondary  sexual  charac- 
teristics may  provide  indirect  clues  regarding  mating 
strategy  in  franciscana  and  are  discussed  below. 

Relative  testis  weight  In  mammals,  there  is  a  func- 
tional relationship  between  relative  testis  weight  and 
the  species'  mating  system  (Kenagy  and  Trombulak, 
1986).  Testes  are  relatively  small  in  species  presenting 
monogamy  or  extreme  poliginy  (several  females  +  few 
males),  i.e.,  where  a  male  copulates  with  all  females  of  a 
group  or  harem.  Comparative  studies  have  demonstrated 
that  males  tend  to  be  larger  than  females  and  show 
secondary  sexual  characteristics  in  species  present- 
ing extreme  poliginy.  On  the  other  hand,  the  relative 
testis  weight  is  high  and  the  sexual  size  dimorphism  is 
reduced  or  nonexistent  in  species  where  several  males 
copulate  with  only  one  estrus  female  (polyandry).  In 
this  case,  the  evolution  for  a  large  testis  is  attributed 
to  the  sperm  competition  in  a  system  where  different 
males  attempt  to  fertilize  the  same  female  and  where  a 
higher  copulatory  frequency  and  higher  levels  of  sperm 
production  are  required  (Harcourt  et  al.,  1981;  Kenagy 
and  Trombulak,  1986). 

Using  the  data  on  133  mammal  species,  Kenagy  and 
Trombulak  (1986)  presented  a  function  describing  the 
relationship  between  body  weight  and  combined-testes 
weight  without  epididymis.  Applying  their  equation 
for  the  adult  male  franciscanas,  we  discovered  that 
mature  franciscanas  have  testes  3  to  12  times  lighter 
than  expected  (mean  =  6  times)  for  a  mammal  of  its 
body  weight.  Indeed,  among  the  133  species  analyzed, 


590 


Fishery  Bulletin  102(4) 


the  relative  testes  weight  of  the  franciscana  is  heavier 
than  that  of  gorilla  {Gorilla  gorilla),  humpback  whales 
(Megaptera  novaeangliae),  and  fin  whales  iBalaenoptera 
physalus),  indicating  that  sperm  competition  does  not 
occur  in  franciscanas. 

Sexual  size  dimorphism  Males  are  larger  than  females 
in  most  mammal  species.  Nevertheless,  the  reversed 
sexual  size  dimorphism  (RSSD)  (i.e.,  females  are  larger 
than  males)  is  more  common  than  previously  thought 
and  has  been  documented  for  12  out  of  the  20  orders  of 
mammals  (Ralls,  1976,  1977).  Among  the  odontocetes, 
four  (Ziphidae,  Pontoporiidae,  Phocoenidae,  and  Delph- 
inidae)  out  of  the  eight  families  present  RSSD. 

Although  sexual  selection  may  be  the  main  reason 
why  males  are  the  larger  sex  in  most  mammal  species, 
it  has  been  systematically  refused  as  an  explanation  in 
the  cases  where  females  are  the  larger  sex  (Ralls,  1976, 
1977;  Andersson,  1994).  In  species  with  RSSD,  females 
do  not  mate  with  many  males,  they  are  not  dominant, 
and  are  not  more  aggressive  than  males  of  the  same 
species.  Moreover,  they  do  not  show  secondary  sexual 
characteristics  associated  with  intrasexual  selection 
(e.g.,  horns  in  Artiodactyla  and  large  canine  teeth  in 
Primates).  Therefore,  the  occurrence  of  RSSD  in  mam- 
mals may  be  explained  more  satisfactorily  by  natural 
selection  (Andersson,  1994). 

Slooten  (1991)  proposed  an  interesting  hypothesis 
for  the  occurrence  of  RSSD  in  cetaceans,  suggesting 
that  a  minimum  size  may  be  necessary  for  a  newborn 
cetacean  to  survive.  In  odontocetes,  the  smallest  mean 
sizes  at  birth  are  about  70-80  cm.  Because  the  size  of 
the  newborn  is  directly  related  to  the  size  of  the  moth- 
er, in  species  of  small  dimensions  the  females  would 
suffer  a  selective  pressure  to  be  a  larger  size,  so  that 
they  could  produce  offspring  with  the  minimum  viable 
size.  This  hypothesis  is  reinforced  by  the  fact  that  most 
of  the  odontocete  species  with  RSSD  (e.g.,  Pontoporia 
blainvillei,  Cephalorhynchus  hectori,  Cephalorhynchus 
commerssoni,  Phocoena  phocoena,  Phocoena  sinus)  are 
the  smallest  species  within  the  group.  Moreover,  spe- 
cies presenting  RSSD  also  have  larger  relative  size  at 
birth  than  the  other  species  within  the  taxonomic  group 
(Ralls,  1976). 

The  degree  and  direction  of  SSD  (sexual  size  dimor- 
phism) in  mammals  is  the  result  of  the  difference  of 
the  sum  of  all  selective  pressures  affecting  the  female's 
size  and  the  sum  of  all  selective  pressures  affecting  the 
male's  size  (Ralls,  1976).  Thus,  it  is  very  probable  that 
more  than  one  factor  may  act  selectively  on  animals 
of  both  sexes  in  Pontoporia,  molding  the  degree  and 
direction  of  SSD.  We  propose  that  the  requirement  of 
a  neonate  minimum  viable  size  (70-80  cm  in  length) 
is  one  of  the  main  selective  pressures  acting  on  female 
franciscanas.  It  is  important  to  emphasize  that  other 
factors  may  also  be  influencing  SSD  in  franciscana, 
and  in  some  species  it  was  evident  that  different  selec- 
tive pressures  could  affect  body  size  in  opposing  direc- 
tions in  males  and  females  and  in  different  age  classes 
(Grant,  1986;  Andersson,  1994).  Among  the  factors  that 


may  be  simultaneously  acting  on  franciscana  body  size 
are  intrinsic  genetic  and  physiological  limitations,  and 
the  requirement  of  maintaining  an  optimum  size  for  the 
species'  ecological  niche. 

Secondary  sexual  characteristics  The  presence  and 
intensity  of  secondary  sexual  characteristics  in  males 
is  a  more  precise  indication  of  the  degree  of  intrasexual 
selection  than  is  body  size  (Andersson,  1994).  In  odonto- 
cete males,  these  characteristics  are  present  in  the  form 
of  "weapons,"  such  as  the  tusk  of  the  narwhal  (Monodon 
monoceros)  and  the  teeth  in  species  of  the  genus  Gram- 
pus, Physeter,  Berardius,  Hyperoodon,  and  Mesoplodon 
used  in  male-male  combats  (MacLeod.  1998).  In  spe- 
cies of  these  genus,  the  teeth  were  reduced  in  number, 
enlarged  in  size,  and  their  form  was  modified  (specially 
in  males  of  Ziphiidae).  The  teeth  of  these  species  also 
lost  their  function  in  feeding  because  of  a  diet  comprising 
almost  exclusively  cephalopods  and  were  used  uniquely 
in  intrasexual  combats.  There  is  no  evidence  that  the 
same  evolutionary  process  occurred  in  male  francisca- 
nas because  their  teeth  are  very  small  and  numerous 
(around  200),  their  diet  is  primarily  fish,  and  the  number 
of  combat  scars  is  apparently  low.  These  characteristics 
support  the  hypothesis  that  male-male  combat  must  be 
very  rare  or  even  nonexistent  in  franciscanas. 

The  sexual  features  presented  in  this  study  (extremely 
low  testis  weight,  reversed  sexual  size  dimorphism,  ab- 
sence of  secondary  sexual  characteristics  in  males,  and 
a  low  number  of  scars  in  males)  indicate  the  absence  of 
sperm  competition  in  the  franciscana,  and  these  features 
differ  drastically  from  those  characteristics  of  odonto- 
cete species  where  males  combat  each  other  for  copula- 
tion. This  finding  may  indicate  that  franciscanas  form 
temporary  reproductive  pairs  during  the  reproductive 
period,  where  a  male  pairs  and  copulate  with  only  one 
female.  Recently,  Valsecchi  and  Zanelatto  (2003)  pro- 
vided molecular  evidence  suggesting  that  franciscanas 
may  travel  in  kin  groups  that  include  mothers  with  their 
calves  and  the  father  of  the  youngest  offspring.  The  au- 
thors also  suggested  that  male  franciscanas  may  prolong 
their  bond  with  their  reproductive  partner,  providing 
some  form  of  paternal  care.  For  a  better  understanding 
of  franciscana  social  structure  and  mating  system,  the 
following  suggestions  are  proposed:  1)  an  increase  in  the 
efforts  of  behavioral  studies  of  free-ranging  francisca- 
nas; 2)  quantification  of  the  intraspecific  teeth  scars  in 
franciscanas  of  different  sexes  and  reproductive  status 
in  order  to  confirm  the  absence  of  intrassexual  aggres- 
sions among  males;  3)  investigation  of  the  relationship 
of  relative  testis  weight,  SSD.  and  reproductive  strate- 
gies in  cetaceans,  by  phylogenetic  methods  (see  Harvey 
and  Pagel.  1991)  to  understand  the  evolution  of  these 
characters  in  this  group. 


Acknowledgments 

This  study  could  not  be  made  without  the  cooperation 
and  friendship  of  the  fishermen  from  Tramandai/Imbe 


Danilewicz  et  al.:  Reproductive  biology  of  male  Pontopona  blainvillei  from  Rio  Grande  do  Sul,  southern  Brazil 


591 


and  Rio  Grande.  Many  people  collaborated  in  the  collec- 
tion and  necropsy  of  the  dolphins,  and  the  authors  wish 
to  thank  Paulo  Ott,  Ignacio  Moreno,  Marcio  Martins, 
Glauco  Caon,  Larissa  Oliveira,  Manuela  Bassoi,  Alexan- 
dre Zerbini,  Luciana  Moller,  Luciano  Dalla  Rosa,  Lilia 
Fidelix,  and  numerous  volunteers  for  helping  in  this 
task.  Gonad  histology  was  partially  done  in  Laboratorio 
de  Histologia  e  Embriologia  Comparada  of  Universi- 
dade  Federal  do  Rio  Grande  do  Sul  and  we  thank  Sonia 
Garcia  and  Nivea  Lothhammer  for  their  instructions 
and  encouragement  on  this  subject.  Paulo  Ott,  Enrique 
Crespo  and  Silvana  Dans  participated  in  the  age  deter- 
mination procedures.  Part  of  the  age  determination  was 
also  done  in  the  Rio  Grande  and  the  first  author  thanks 
Cristina  Pinedo  and  Fernando  Rosas  for  their  instruc- 
tion. The  authors  also  thanks  Norma  Luiza  Wiirdig, 
Iraja  Damiani  Pinto  (CECLIMAR-UFRGS),  and  Lauro 
Barcellos  (Director  of  the  Museu  Oceanografico)  for  their 
constant  logistical  support  and  for  encouraging  marine 
mammal  studies  in  southern  Brazil.  Marcio  Martins, 
Ignacio  Moreno,  Luiz  Malabarba,  and  Monica  Muelbert 
reviewed  an  early  draft  of  this  paper.  We  wish  to  thank 
Renata  Ramos  and  three  anonymous  reviewers  for  their 
suggestions  regarding  the  manuscript.  Financial  support 
was  given  by  Cetacean  Society  International,  Fundacao 
O  Boticario  de  Protecao  a  Natureza,  The  MacCarthur 
Foundation,  World  Wildlife  Fund,  CNPq,  UNEP,  Yaqu 
Pacha  Organization,  and  Whale  and  Dolphin  Conser- 
vation Society.  This  paper  is  part  of  the  first  author's 
M.S.  thesis,  and  The  Coordenagao  de  Aperfeicoamento 
de  Pessoal  de  Nivel  Superior  (CAPES)  has  granted 
him  a  graduate  fellowship.  The  Conselho  Nacional  de 
Desenvolvimento  Cientifico  e  Tecnologico  of  the  Brazilian 
Government  (CNPq)  has  granted  graduate  fellowships 
to  E.R.  Secchi  (Grant  no.  200889/98-2). 


Literature  cited 

American  Society  of  Mammalogy. 

1961.     Standardized  methods  for  measuring  and  record- 
ing data  on  the  smaller  cetaceans.     J.  Mamm.  42: 
471-476. 
Andersson,  M. 

1994.     Sexual  selection.  Monographs  in  behavior  and  ecol- 
ogy, 599  p.     Princeton  Univ.  Press.  Princeton,  NJ. 
Bordino,  P. 

2002.     Movement  patterns  of  franciscana  dolphin  {Pon- 
toporia  blainvillei)  in  Bahia  Anegada,  Buenos  Aires, 
Argentina.     LAJAM  (special  issue)  1:71-76. 
Bordino,  P.,  G.  Thompson,  and  M.  Iniguez. 

1999.     Ecology  and  behaviour  of  the  franciscana  {Pon- 
toporia  blainvillei)  in  Bahia  Anegada,  Argentina.     J. 
Cetacean  Res.  Manag.  1  (2):213-222. 
Collet,  A.,  and  H.  Saint  Girons. 

1984.     Preliminary  study  of  the  male  reproductive  cycle 
in  common  dolphins,  Delphinus  delphis,  in  the  eastern 
North  Atlantic.     Rep.  Int.  Whal.  Comm.  Ispecial  issue) 
6:355-360. 
Corcuera,  J..  F.  Monzon,  E.  A.  Crespo,  A.  Aguilar.  and 
J.  A.  Raga. 

1994.     Interactions  between  marine  mammals  and  the 


coastal  fisheries  of  Necochea  and  Claromeco  (Buenos 
Aires  Province,  Argentina).  Rep.  Int.  Whal.  Comm. 
(special  issue)  15:283-290. 

Crespo,  E.  A.,  G.  Harris,  and  R.  Gonzalez. 

1998.  Group  size  and  distribution  range  of  the  fran- 
ciscana, Pontoporia  blainvillei.  Mar.  Mamm.  Sci.  14 
i  4  i:845-849. 

Danilewicz,  D. 

2003.  Reproduction  of  female  franciscana  {Pontoporia 
blainvillei )  in  Rio  Grande  do  Sul,  southern  Brazil. 
LAJAM  2  (2):67-78. 

Danilewicz,  D..  E.  R.  Secchi.,  P.  H.  Ott.  and  I.  B.  Moreno. 

2000.  Analyses  of  the  age  at  sexual  maturity  and  repro- 
ductive rates  of  franciscanas  {Pontoporia  blainvillei) 
from  Rio  Grande  do  Sul,  Southern  Brazil.  Comum. 
Mus.  Cienc.  Tecnol.  PUCRS  13(11:89-98. 

DeMaster,  D.  P. 

1978.  Calculation  of  the  average  age  of  sexual  matu- 
rity in  marine  mammals.  J.  Fish.  Res.  Board  Can. 
35  (6):912-15. 

Dewsbury,  D.  A. 

1982.     Ejaculate  cost  and  male  choice.     Am.  Nat.  119: 
601-610. 
Di  Beneditto,  A.  P.  M„  and  R.  M.  A.  Ramos. 

2001.  Biology  and  conservation  of  the  franciscana  {Pon- 
toporia blainvillei)  in  the  north  of  Rio  de  Janeiro  State, 
Brazil.     J.  Cetacean  Res.  Manag.  3:185-192. 

Ferrero  R.  C,  and  W.  A.  Walker. 

1993.     Growth  and  reproduction  of  the  northern  right 
whale  dolphin.  Lissodelphis  borealis,  in  the  offshore 
waters  of  the  North  Pacific  Ocean.     Can.  J.  Zool.  71 
(121:2335-2344. 
Grant,  P.  R. 

1986.     Ecology  and  evolution  of  Darwin's  finches.     Princ- 
eton Univ.  Press,  Princeton.  NJ. 
Harcourt,  A.  H.,  P.  H.  Harvey.  S.  G.  Larson,  and  R.  V.  Short. 
1981.     Testis  weight,  body  weight  and  breeding  system 
in  primates.     Nature  293:55-57. 
Harrison.  R.  J.,  M.  M.  Bryden,  D.  A.  McBrearty,  and 
R.  L.  Brownell  Jr. 

1981.     The  ovaries  and  reproduction  in  Pontoporia  blain- 
villei (Cetacea:  Platanistidae).     J.  Zool.  193:563-580. 
Harvey,  P.,  and  M.  Pagel. 

1991.  The  comparative  method  in  evolutionary  biology, 
239  p.  Oxford  series  in  ecology  and  evolution.  Oxford 
Univ.  Press,  Oxford,  England. 

Hohn,  A.,  S.  J.  Chivers.  and  J.  Barlow. 

1985.  Reproductive  maturity  and  seasonality  of  male 
spotted  dolphins,  Stenella  attenuata,  in  the  eastern 
tropical  Pacific.     Mar.  Mamm.  Sci.,  1(41:273-293. 

Kasuya,  T.,  and  R.  L.  Brownell  Jr. 

1979.  Age  determination,  reproduction,  and  growth  of 
the  franciscana  dolphin,  Pontoporia  blainvillei.  Sci. 
Rep.  Whales  Res.  Inst.  31:45-67. 

Kinas,  P.  G. 

2002.  The  impact  of  the  incidental  kills  by  gillnets  on  the 
franciscana  dolphin  {Pontoporia  blainvillei)  in  southern 
Brazil.  Bull.     Mar.  Sci.  70:409-421. 

Kenagy  ,  G.  K.,  and  S.  C.  Trombulak. 

1986.  Size  and  function  of  mammalian  testes  in  relation 
to  body  size.     J.  Mamm.  67(l):l-22. 

Lincoln,  G.  A. 

1992.  Biology  of  seasonal  breeding  in  deer.  In  The  biol- 
ogy of  deer  (R.  D.  Brown,  ed.),  p.  565-574.  Springer- 
Verlag,  New  York,  NY. 


592 


Fishery  Bulletin  102(4) 


MacLeod.  CD. 

1998.     Intraspecific  scarring  in  odontocete  cetaceans:  an 
indicator  of  male  'quality'  in  agressive  interactions?     J. 
Zool.  244:71-77. 
Mitchell,  E.  D.,  and  V.  M.  Kozicki. 

1984.     Reproductive  condition  of  male  sperm  whales, 
Physeter  macrocephalus,  taken  off  Nova  Scotia.     Rep. 
Int.  Whal.  Coram,  (special  issue)  6:243-48. 
Miyazaki,  N. 

1977.     Growth  and  reproduction  of  Stenella  coeruleoalba 
off  the  Pacific  coast  of  Japan.     Sci.  Rep.  Whales  Res. 
Inst.  29:21-48. 
Moreira,  L.  M.,  and  S.  Siciliano. 

1991.     Northward  extension  range  for  Pontoporia  blainvil- 
lei, p.  8.     Abs.  Ninth  Bienn.  Conf.  Biol.  Mar.  Mamm. 
Chicago,  IL. 
Moreno,  I.  B.,  P.  H.  Ott,  and  D.  Danilewicz. 

1997.  Analise  preliminar  do  impacto  da  pesca  artesanal 
costeira  sobre  Pontoporia  blainvillei  no  litoral  norte  do 
Rio  Grande  do  Sul,  sul  do  Brasil.  In  Proceedings  of 
the  second  workshop  for  the  research  coordination  and 
conservation  of  the  franciscana  (Pontoporia  blainvillei) 
in  the  Southwestern  Atlantic  (M.  C.  Pinedo  A.  and  Bar- 
reto,  eds.),  p.  31-41.  Editora  da  FURG,  Rio  Grande, 
Rio  Grande  do  Sul,  Brazil. 

Neimanis,  A.  S.,  A.  J.  Read,  R.  A.  Foster,  and  D.  E.  Gaskin. 

2000.  Seasonal  regression  in  testicular  size  and  his- 
tology in  harbour  porpoises  (Phocoena  phocoena,  L.) 
from  the  Bay  of  Fundy  and  Gulf  of  Maine.  J.  Zool. 
250:221-229. 

Ott,  P.  H. 

1998.  Analise  das  capturas  acidentais  da  toninha,  Ponto- 
poria blaivillei,  no  litoral  norte  do  Rio  Grande  do  Sul,  sul 
do  Brasil.  MSc.  Thesis,  120  p.  Pontificia  Universidade 
Catolica  do  Rio  Grande  do  Sul,  Porto  Alegre,  Brazil. 

Ott,  P.  H.,  E.  R.  Secchi.  I.  B.  Moreno,  I.B.,  D.  Danilewicz, 
E.  A.  Crespo,  P.  Bordino,  R.  Ramos,  A,  P.  Di  Beneditto, 
C.  Bertozzi,  R.  Bastida,  R.  Zanelatto,  J.  Perez,  and  P.  G.  Kinas. 
2002.     Report  of  the  working  group  of  fishery  inter- 
actions.    LAJAM  (special  issue)  1:55-64. 
Perrin,  W.  F.,  J.  M.  Coe,  and  J.  R.  Zweifel. 

1976.     Growth  and  reproduction  of  the  spotted  porpoise, 
Stenella  attenuata,  in  the  offshore  eastern  tropical 
Pacific.     Fish.  Bull.  74:229-269. 
Pinedo,  M.  C. 

1991.  Development  and  variation  of  the  franciscana, 
Pontoporia  blainvillei.  Ph.D.  diss.,  406  p.  Univ.  Cali- 
fornia, Santa  Cruz,  CA. 
1994.  Impact  of  incidental  fishery  mortality  on  the  age 
structure  of  Pontoporia  blainvillei  in  southern  Brazil 
and  Uruguay.  Rep.  Int.  Whal.  Comm.  (spec,  issue 
151:261-264. 
Pinedo,  M.  C,  and  A.  Hohn. 

2000.     Growth  layer  patterns  in  teeth  from  the  fran- 
ciscana, Pontoporia  blainvillei:  developing  a  model 
for  precision  in  age  estimation.     Mar.  Mamm.  Sci.  16 
(l):l-27. 
Praderi,  R.,  M.  C.  Pinedo,  and  E.  A.  Crespo. 

1989.  Conservation  and  management  of  Pontoporia  bla- 
invillei in  Uruguay,  Brazil  and  Argentina.  In  Biol- 
ogy and  conservation  of  river  dolphins  (W.  F.  Perrin, 
R.  L.  Brownell,  Z.  Kaya,  and  L.  Jiankang  (eds.),  p. 
52-56.     Allen  Press,  Lawrence,  KS. 


Ralls.  K. 

1976.  Mammals  in  which  females  are  larger  than  males. 
Q.  Rev.  Biol.  51:245-276. 

1977.  Sexual  dimorphism  in  mammals:  avian  models  and 
unanswered  questions.     Am.  Nat.  111:917-937. 

Ramos,  R. 

1997.  Determinacao  de  idade  e  biologia  reprodutiva  de 
Pontoporia  blainvillei  e  da  forma  marinha  de  Sotalia 
fluviatilis  no  litoral  norte  do  Rio  de  Janeiro.  M.S.  thesis, 
95  p.  Universidade  Estadual  do  Norte  Fluminense,  Rio 
de  Janeiro,  Brazil. 

Ramos,  R.,  A.  P.  Di  Beneditto,  and  N.  R.  W.  Lima. 

2000.  Growth  parameters  of  Pontoporia  blainvillei  and 
Sotalia  fluviatilis  (Cetacea)  in  northern  Rio  de  Janeiro, 
Brazil.     Aquat.  Mamm.  26  (11:65-75. 

Read,  A. 

1990.  Reproductive  seasonality  in  harbour  porpoises, 
Phocoena  phocoena,  from  the  Bay  of  Fundy.  Can.  J. 
Zool.  68:284-288. 

Read,  A.,  and  A.  Hohn. 

1995.     Life  in  the  fast  lane:  the  life  history  of  the  har- 
bour porpoises  from  the  Gulf  of  Maine.     Mar.  Mamm. 
Sci.  11:423-440. 
Secchi,  E.  R. 

1999.  Taxa  de  crescimento  potencial  intrinseco  de  um 
estoque  de  franciscanas,  Pontoporia  blainvillei  (Ger- 
vais  &  D'Orbigny,  1846)  (Cetacea,  Pontoporiidae)  sob  o 
impacto  da  pesca  costeira  de  emalhe.  M.S.  thesis,  152 
p.  Fundacao  Universidade  Federal  do  Rio  Grande,  Rio 
Grande,  do  Sul,  Brazil.  [In  Portuguese.] 
Secchi,  E.  R.,  A.  N.  Zerbini,  M.  Bassoi,  L.  Dalla  Rosa, 
L.  M.  Moller,  and  C.  C.  Rocha-Campos. 

1997.     Mortality  of  franciscanas,  Pontoporia  blainvillei, 
in  coastal  gillnetting  in  southern  Brazil.     Rep.  Int. 
Whal.  Comm.  47:653-658. 
Secchi,  E.  R.,  Ott,  P.  H.,  and  Danilewicz,  D. 

2002.  Report  of  the  fourth  workshop  for  the  coordi- 
nated research  and  conservation  of  the  franciscana 
dolphin  [Pontoporia  blainvillei)  in  the  western  South 
Atlantic.     LAJAM  (special  issue)  1:11-20. 

2003.  Effects  of  fishing  bycatch  and  the  conservation 
status  of  the  franciscana  dolphin,  Pontoporia  blainvillei. 
In  Marine  mammals:  fisheries,  tourism  and  manage- 
ment issues  (N.  Gales,  M.  Hindell,  and  R.  Kirkwood, 
eds.),  p.  174-191.  CSIRO  Publishing,  Collingwood, 
Victoria,  Australia. 

Slooten. E. 

1991.  Age,  growth,  and  reproduction  in  Hector's  dolphins. 
Can.  J.  Zool.  69:1689-1700. 

Slooten,  E„  and  F.  Lad. 

1991.  Population  biology  and  conservation  of  Hector's 
dolphin.     Can.  J.  Zool.  69:1701-1707. 

Stearns,  S.  C. 

1992.  The  evolution  of  life  histories,  249  p.  Oxford. 
Univ.  Press,  New  York,  NY. 

Valsecchi,  E..  and  R.  C.  Zanelatto. 

2003.     Molecular  analysis  of  the  social  and  population 
structure  of  the  franciscana  (Pontoporia  blainvillei): 
conservation  implications.     J.  Cetacean  Res.  Manag. 
5il):69-75. 
Van  Waerebeek,  K..  and  A.  Read. 

1994.  Reproduction  of  dusky  dolphins,  Lagenorhyn- 
chus  obscurus,  from  coastal  Peru.  J.  Mamm.  75 
i  11:1054-1062. 


593 


Abstract— Red  snapper  (Lutjanus 
campechanus)  in  the  United  States 
waters  of  the  Gulf  of  Mexico  ( GOM ) 
has  been  considered  a  single  unit 
stock  since  management  of  the  spe- 
cies began  in  1991.  The  validity  of  this 
assumption  is  essential  to  manage- 
ment decisions  because  measures  of 
growth  can  differ  for  nonmixing  popu- 
lations. We  examined  growth  rates, 
size-at-age,  and  length  and  weight 
information  of  red  snapper  collected 
from  the  recreational  harvests  of  Ala- 
bama (n=2010),  Louisiana  (re=1905), 
and  Texas  (re=1277)  from  1999  to 
2001.  Ages  were  obtained  from  5035 
otolith  sections  and  ranged  from  one 
to  45  years.  Fork  length,  total  weight, 
and  age-frequency  distributions  dif- 
fered significantly  among  all  states; 
Texas,  however,  had  a  much  higher 
proportion  of  smaller,  younger  fish. 
All  red  snapper  showed  rapid  growth 
until  about  age  10  years,  after  which 
growth  slowed  considerably.  Von  Ber- 
talanffy  growth  models  of  both  mean 
fork  length  and  mean  total  weight- 
at-age  predicted  significantly  smaller 
fish  at  age  from  Texas,  whereas  no 
differences  were  found  between  Ala- 
bama and  Louisiana  models.  Texas 
red  snapper  were  also  shown  to  differ 
significantly  from  both  Alabama  and 
Louisiana  red  snapper  in  regressions 
of  mean  weight  at  age.  Demographic 
variation  in  growth  rates  may  indicate 
the  existence  of  separate  management 
units  of  red  snapper  in  the  GOM.  Our 
data  indicate  that  the  red  snapper 
inhabiting  the  waters  off  Texas  are 
reaching  smaller  maximum  sizes  at 
a  faster  rate  and  have  a  consistently 
smaller  total  weight  at  age  than  those 
collected  from  Louisiana  and  Alabama 
waters.  Whether  these  differences  are 
environmentally  induced  or  are  the 
result  of  genetic  divergence  remains 
to  be  determined,  but  they  should  be 
considered  for  future  management 
regulations. 


Red  snapper  (Lutjanus  campechanus) 
demographic  structure  in  the  northern 
Gulf  of  Mexico  based  on  spatial  patterns 
in  growth  rates  and  morphometries 


Andrew  J.  Fischer 

Coastal  Fisheries  Institute 

School  of  the  Coast  and  Environment 

Louisiana  State  University 

Baton  Rouge,  Louisiana  70803-7503 

E-mail  address:  afische  g  Isu  edu 


M.  Scott  Baker  Jr. 

Coastal  Fisheries  Institute 

School  of  the  Coast  and  Environment 

Louisiana  State  University 

Baton  Rouge,  Louisiana  70803-7503 


Charles  A.  Wilson 

Coastal  Fisheries  Institute  and 

Department  of  Oceanography  and  Coastal  Sciences 

School  of  the  Coast  and  Environment 

Louisiana  State  University 

Baton  Rouge,  Louisiana  70803-7503 


Manuscript  submitted  6  May  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
19  April  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:593-603  12004). 


Red  snapper  {Lutjanus  campechanus) 
in  the  United  States  waters  of  the  Gulf 
of  Mexico  (GOM)  are  heavily  exploited 
by  both  recreational  and  commercial 
fishermen  (Wilson  and  Nieland,  2001; 
Shirripa  and  Legault1).  Harvest,  how- 
ever, has  not  proceeded  without  det- 
rimental affects  on  the  population. 
Commercial  landings  have  declined 
substantially  from  6048  metric  tons  (t) 
in  1964  to  1207  t  in  1990;  recreational 
landings  exhibited  similar  declines 
from  1937  t  in  1981  to  481  t  in  1990 
(NMFS'2).  In  1991,  harvest  restrictions 
including  reef  fish  permits,  seasonal 
fishing,  fish  quotas,  creel  limits,  and 
minimum  size  limits  were  placed  upon 
the  red  snapper  fishermen  by  the  Gulf 
of  Mexico  Fishery  Management  Council 
(GMFMC3)  to  increase  the  spawning 
potential  ratio  to  20%,  which  is  indic- 
ative of  recovery.  These  regulations 
have  also  been  adopted  for  state  waters 
in  Alabama,  Louisiana,  and  Texas. 
Despite  the  management  actions,  GOM 
red  snapper  remain  overfished  (Good- 
year4; Schirripa  and  Legault1). 


1  Shirripa,  M.  J.,  and  C.  M.  Legault. 
1999.  Status  of  the  red  snapper  in  the  U. 
S.  waters  of  the  Gulf  of  Mexico;  updated 
through  1998,  44  p.  +  appendices.  Con- 
tribution rep.  SFD-99/00-75  from  Sus- 
tainable Fisheries  Division.  Miami 
Laboratory,  Southeast  Fisheries  Science 
Center,  National  Marine  Fishery  Ser- 
vice, 75  Virginia  Beach  Drive,  Miami, 
FL  33149-1099.  [Not  available  from 
NTIS], 

2  NMFS  (National  Marine  Fisheries 
Service).  2003.  Fisheries  Statistics 
and  Economics  Division.  Website:  www. 
nmfs.noaa.gov. 

3  GMFMC  (Gulf  of  Mexico  Fishery  Man- 
agement Council).  1991.  Amendment 
3  to  the  reef  fishery  management  plan 
for  the  reef  fish  resources  of  the  Gulf 
of  Mexico,  38  p.  Gulf  of  Mexico  Fish- 
ery Management  Council,  3018  N.  U.S. 
Hwy  301  Suite  1000,  Tampa,  FL.  33619- 
2272.     [Not  available  from  NTIS], 

4  Goodyear,  C.  P.  1995.  Red  snapper  in 
U.S.  waters  of  the  Gulf  of  Mexico.  Stock 
assesment  report  MIA-95/96-05,  171  p. 
Miami  Laboratory,  Southeast  Fisheries 
Science  Center,  National  Marine  Fish- 
eries Service,  75  Virginia  Beach  Dr., 
Miami,  FL,  33149-1099.  [Not  available 
from  NTIS]. 


594 


Fishery  Bulletin  102(4) 


96c 

94c 

92 ; 

90 

88 :               86 :               84 

82 ; 

80 

32° 

? 

/>  ^M,  — , 

30° 

Fourchon, 

Daulphin  Island,  AL      , 

28° 

7* 

LA 

26° 

S  Port  Aransas 

TX 

-**-^ 

24° 

N 

400 

0 

400 

800  Kilometers 

A 

22° 

32 


30 


28 


26 


24 


22 


96 


94° 


92° 


90 


88°  86° 


84 


82  = 


80 


Figure  1 

Map  of  the  northern  Gulf  of  Mexico  showing  the  three  red  snapper  iLutjanus  campecha- 
nus)  sampling  locations. 


An  underlying  assumption  crucial  to  a  fishery  man- 
agement plan  is  that  the  fish  species  being  managed 
is  a  unit  stock  (Gulland,  1965).  A  stock  is  defined  as 
the  part  of  a  fish  population  that  is  under  consideration 
as  an  actual  or  potential  resource  (Ricker,  1975).  Since 
management  began  in  1991,  red  snapper  in  the  north- 
ern GOM  have  been  considered  a  unit  stock.  Genetic 
studies  to  date  have  shown  that  there  is  little  evidence 
to  dispute  this  assumption  (Camper  et  al.,  1993;  Gold 
et  al.,  1997,  Heist  and  Gold,  2000).  On  the  other  hand, 
tag-recapture  studies  indicate  that  red  snapper  have 
the  capacity  to  move  great  distances,  making  it  pos- 
sible for  separate  stocks  to  develop  (Patterson  et  al., 
2001). 

The  validity  of  an  assumption  of  a  single  stock  of  red 
snapper  is  essential  to  management  decisions  because 
measures  of  growth,  natural  mortality,  reproductive 
capacity,  and  recruitment  can  differ  among  nonmix- 
ing  populations.  Should  separate  red  snapper  stocks 
exist,  management  plans  would  have  to  be  enacted 
for  each  defined  stock  in  order  to  follow  federal  guide- 
lines. Even  if  a  single  large  red  snapper  stock  exists, 
management  should  be  sensitive  to  both  the  diversity 
of  habitats  and  user  groups  within  the  species  area  of 
occurrence.  Because  red  snapper  are  arguably  the  most 
important  recreational  and  commercial  offshore  fishery 
from  Florida  to  southern  Texas,  every  effort  should  be 
undertaken  to  develop  the  most  effective  and  productive 
management  plan. 

The  objective  of  this  study  was  to  evaluate  the  stock 
structure  of  GOM  red  snapper  based  on  growth  rates 
and  size-at-age  information.  We  hypothesized  that  red 
snapper  sampled  from  across  the  northern  GOM  would 
be  indistinguishable  in  their  growth  rates  and  size  at 
age — a  uniformity  indicative  of  a  single  unit  stock. 


Methods  and  materials 

Red  snapper  were  collected  from  the  recreational  har- 
vests of  1999,  2000,  and  2001  from  the  northern  GOM 
at  Dauphin  Island,  Alabama,  at  Port  Fourchon,  Louisi- 
ana, and  at  Port  Aransas,  Texas  (Fig.  1).  A  maximum 
of  75  fish  were  randomly  selected  and  sampled  from 
the  daily  catch  of  each  charter  boat  or  head  boat  while 
the  captains  and  deck  hands  cleaned  fish.  These  fish 
were  not  selected  by  size.  Larger  individuals  (>6.8  kg) 
were  opportunistically  sampled  from  spear  fishing  and 
hook-and-line  fishing  tournaments  in  Alabama  and 
Louisiana.  In  addition,  a  number  of  smaller  fish  (<406 
mm.  <457  mm  during  summer  1999)  were  randomly 
sampled  during  red  snapper  tagging  cruises  in  Alabama. 
Morphometric  measurements  were  recorded  I  fork  length 
[FL]  in  mm,  total  weight  [TW]  in  kg,  and  eviscerated 
body  weight  |BW|  in  kg),  sex  was  determined  by  macro- 
scopic examination  of  gonads,  and  both  sagittal  otoliths 
were  removed,  rinsed,  and  stored  in  coin  envelopes  until 
processed.  Fish  weights  were  not  recorded  for  1999  Texas 
samples. 

A  transverse  thin  section  (containing  the  core)  was 
taken  from  the  left  sagittal  otolith  of  each  individual. 
Sections  were  made  with  the  Hillquist  model  800  thin- 
sectioning  machine  equipped  with  a  diamond  embedded 
wafering  blade  and  precision  grinder  (Cowan  et  al., 
1995).  When  the  left  otolith  was  unavailable,  the  right 
otolith  was  sectioned.  Examinations  of  otolith  sections 
were  made  with  a  dissecting  microscope  with  transmit- 
ted light  and  polarized  light  filter  at  20x  to  64x  mag- 
nification Opaque  annulus  counts  were  made  along  the 
ventral  side  of  the  sulcus  acousticus  from  the  core  to 
the  proximal  edge  (Wilson  and  Nieland.  2001).  Annulus 
counts  were  performed  by  two  independent  readers  (AJF 


Fischer  et  al.:  Demographic  structure  of  Lutjanus  campechanus  in  the  northern  Gulf  of  Mexico 


595 


and  MSB)  without  knowledge  of  either  date  of  capture 
or  morphometric  data.  The  appearance  of  the  otolith 
section  edge  condition  was  coded  as  opaque  or  translu- 
cent after  Beckman  et  al.  (1989).  Annuli  were  counted  a 
second  time  when  initial  counts  disagreed.  In  instances 
where  a  consensus  between  the  two  readers  could  not  be 
reached,  annulus  counts  of  the  more  experienced  reader 
(AJF)  were  used.  Between-reader  differences  in  annulus 
counts  were  evaluated  with  the  coefficient  of  variation 
(CV),  index  of  precision  (D)  (Chang,  1982),  and  average 
percent  error  (APE)  (Beamish  and  Fournier,  1981).  The 
periodicity  of  opaque  zone  formation  was  verified  for 
each  sampling  location  with  edge  analysis  after  Wilson 
and  Nieland  (2001).  Ages  of  red  snapper  were  estimated 
from  opaque  annulus  counts  and  capture  date  with  the 
equation  described  by  Wilson  and  Nieland  (2001): 


Day  age=  -182  +  (opaque  increment  count 
((»;-!)  x  30)  +  d, 


365)  + 


where  m  =  the  ordinal  number  (1-12)  of  month  of  cap- 
ture; and 
d    =  the  ordinal  number  (1-31)  of  the  day  of  the 
month  of  capture. 

The  182  days  subtracted  from  each  age  estimate  are  to 
account  for  the  uniform  hatching  date  assigned  for  all 
specimens  (Render,  1995;  Wilson  and  Nieland,  2001). 
Age  in  years  was  assigned  by  dividing  day  age  by  365. 

Fork  length-TW  relationships  were  fitted  with  lin- 
ear regression  to  the  model  FL  =  a  TWb  from  log1(l- 
transformed  data  for  Alabama,  Louisiana,  and  Texas 
specimens.  Analysis  of  covariance  (ANCOVA)  was  used 
to  compare  slopes  and  intercepts  among  sampling  lo- 
cations (SAS,  1985).  Variability  in  age.  FL,  and  TW 
frequency  distributions  of  red  snapper  were  compared 
among  states  with  the  Komolgorov-Smirnov  two-sample 
test  (Tate  and  Clelland,  1957). 

Growth  of  red  snapper  was  modeled  for  FL  and  TW 
with  the  von  Bertalanffy  growth  equations.  Because 
of  differences  in  sample  population  size  among  states, 
weighted  mean  FL  and  mean  TW  at  age  were  fitted  for 
each  state  with  nonlinear  regression  in  the  forms: 

FL,  =  LJ1  -e'-*"11) 
TW,  =  Wjl-el-'''"'!)''. 

where  FL,    =  FL  at  age  t\ 
TW,  =  TW  at  age  /; 
L„     =  the  FL  asymptote; 
W„    =  the  TW  asymptote; 
k        =  the  growth  coefficient; 
t        =  age  in  years;  and 
b        =  exponent  derived  from  our  length-weight 

regressions  (SAS,  version  5,  1985,  SAS 

Inst,  Cary,  NO. 

Because  of  a  lack  of  smaller  individuals  in  all  sample 
populations,  no  y-intercepts  for  t0  were  specified  and 
models  were  forced  through  0.  Larger  individuals  and 


Table  1 

Numbers  of  red  snapper  (Lutjanus  camped 

lanus) 

sampled 

from  recreational  sources  by  stai 

e  and  year. 

State 

Males 

Females 

Jnknown 

sex 

Total 

Alabama 

1999 

434 

396 

5 

835 

2000 

355 

415 

7 

111 

2001 

189 

209 

0 

Total 

398 
2010 

Louisiana 

1999 

367 

339 

31 

737 

2000 

399 

397 

8 

804 

2001 

160 

179 

25 

Total 

364 
1905 

Texas 

1999 

268 

293 

14 

575 

2000 

278 

284 

22 

584 

2001 

52 

56 

10 

Total 

118 
1277 

juveniles  selectively  sampled  by  size  were  excluded  from 
the  models  to  more  accurately  reflect  a  random  sample. 
Likelihood  ratio  tests  (Cerrato,  1990)  were  used  to  test 
for  differences  among  states  in  models  and  in  growth 
parameter  estimates.  Differential  growth  was  evalu- 
ated for  red  snapper  in  the  first  10  years  of  life  when 
somatic  growth  is  most  rapid  (Szedlmayer  and  Shipp, 
1994;  Patterson  et  al.,  2001;  Wilson  and  Nieland,  2001). 
Linear  regressions  of  mean  FL  and  mean  TW  at  age  for 
fishes  aged  1  to  10  years  were  compared  among  states 
with  analysis  of  covariance  (ANCOVA)  and  tested  for 
homogeneity  of  slopes. 


Results 

During  the  three-year  study  period,  5192  red  snapper 
were  sampled  from  the  recreational  harvest  of  the  north- 
ern GOM  (Table  1):  642  individuals  from  fishing  tourna- 
ments, 71  undersize  fish  from  tagging  cruises,  and  4479 
random  samples  from  recreational  catches.  The  samples 
included  2502  males,  2568  females,  and  122  individu- 
als of  undetermined  sex.  The  resultant  male-to-female 
ratios  were  0.96:1  for  Alabama,  1:0.99  for  Louisiana, 
0.94:1  for  Texas,  and  0.97:1  for  all  states  combined.  A 
chi-square  test  indicated  no  significant  difference  in 
the  number  of  males  to  females  (j2=0.78,  P=0.38).  Fork 
lengths  ranged  from  237  to  916  mm  (Fig.  2A).  Speci- 
mens from  Alabama  ranged  from  237  to  916  mm  FL, 
Louisiana  specimens  ranged  from  282  to  913  mm  FL, 
and  Texas  specimens  ranged  from  266  to  846  mm  FL. 
The  FL  frequency  distributions  of  the  random  samples 
were  different  among  all  states  (AL  and  LA,  maximum 
difference  (MD)=5.26;  AL  and  TX,  MD  =  51.86;  LA  and 
TX,  MD  =  51.77)(Fig.  2A). 


596 


Fishery  Bulletin  102(4) 


30  -> 


25 


20 


15 


10 


n  ,nnM 


DAL 
■  LA 
DTX 


tlk^^idirftr^.,. 


125  250  300  350  400  450  500  550  600  650  700  750  800  850  900 
Fork  length  (m) 


50 
45 
40 
35 
30 
25 
20 
15 
10 
5 
0 


B 


:; 


IV ryfk  .ryryry  r*  n  .rim  m  ri  r.  .  .     .fl 


5  6  7 

Total  weight  (kg) 


10 


>10 


Figure  2 

Distributions  of  (A)  fork  length  in  mm  (/!  =  5177)  and  (B)  total  weight  in 
kg  (n  =  4531l  for  red  snapper  (Lutjanus  campechanus)  sampled  from  the 
1999-2001  recreational  harvests  of  Alabama.  Louisiana,  and  Texas. 


Total  weights  of  all  fish  sampled  ranged  from  0.11 
to  17.35  kg  (Fig.  2B).  Specimens  from  Alabama  ranged 
from  0.22  to  15.42  kg  TW,  Louisiana  specimens  were 
0.42  to  17.35  kg  TW,  and  Texas  specimens  ranged  from 
0.33  to  9.42  kg  TW.  Total  weight-frequency  distributions 
(in  0.5  kg  increments)  differed  significantly  between  all 
states  (AL  and  LA,  MD  =  5.37;  AL  and  TX,  MD  =  53.68; 
and  LA  TX,  MD  =  52.28)(Fig.  2B).  Significant  differ- 
ences in  red  snapper  FL-TW  regression  models  were 
detected  among  states  (ANCOVA  test  of  homogeneity  of 
slopes,  F5  4522=23.36;  P<0.001;  r2=0.98;  ANCOVA  test 
for  equal  intercepts,  F5  4522=22.77,  P<0.001,  r2=0.98); 
therefore,  separate  models  were  fitted  for  each  state. 
The  resultant  equations  were 

AL  TW=  1.51  x  10-"'  iFL:,,i:i) 

(F1;1965=102740;  P<0.0001;  r2=0.98); 


TX  TW  =  2.88  x  10-5  (PL2-92) 


LA  TW  =  1.02  x  10-r'  iFL 


') 


^1; 


=  13345;  P<0.0001;  r2=0.95). 


(P 


1;1856 


=77981;  P<0.0001;  r2=0.98); 


Ages  were  obtained  from  5035  transverse  otolith  sec- 
tions. Thirty  fish  had  otolith  sections  deemed  unread- 
able by  both  readers.  The  age  estimates  determined  by 
the  two  readers  were  evaluated  for  reader  agreement, 
precision,  and  average  percent  error  for  first  and  sec- 
ond readings  of  otolith  sections  by  sample  year.  Table  2 
gives  APE,  CV,  D.  percentage  agreement  (O),  and  per- 
centages of  differences  in  age  estimates  (±1,  2,  and  3 
years).  The  readers  agreed  on  age  estimates  for  4053 
otoliths  (80.5%)  after  the  initial  reading.  Re-examina- 
tion of  the  982  otolith  sections  for  which  annulus  counts 
differed  produced  agreement  for  5007  individuals. 

We  compared  the  timing  of  opaque  annulus  formation 
among  red  snapper  sample  sites  by  plotting  the  monthly 
occurrence  of  maximum  and  minimum  proportions  of 
opaque  otolith  edges.  Sample  limitations  of  red  snapper 
in  Texas,  however,  prevented  meaningful  comparisons  of 


Fischer  et  al.:  Demographic  structure  of  Lut/anus  campechanus  in  the  northern  Gulf  of  Mexico 


597 


opaque  annulus  formation  for  this  state.  However,  mini- 
mum proportions  of  opaque  edges  during  the  months  of 
April  through  October  may  indicate  that  red  snapper 
from  Texas  form  an  opaque  annulus  during  the  winter 
months.  Proportions  of  opaque  edges  for  Alabama  and 
Louisiana  were  essentially  the  same:  maximum  propor- 
tions of  opaque  edges  during  the  months  of  February 
and  March  followed  by  a  decrease  to  minimum  propor- 
tions during  the  months  of  May  through  November 
(Fig.  3).  These  findings  are  consistent  with  previous 
age  and  growth  studies  on  red  snapper  in  the  northern 
COM  (Patterson  et  al,  2001;  Wilson  and  Nieland,  2001), 
indicating  that  the  formation  of  one  opaque  annulus  in 
the  winter  months  is  followed  by  the  formation  of  one 
translucent  annulus  in  summer.  Annulus-based  age 
estimates  of  red  snapper  from  the  northern  GOM  have 
also  been  validated  to  55  years  with  otolith  radiocarbon 
chronologies  based  on  accelerator  mass  spectrometry 
14C  measurements  (Baker  and  Wilson,  2001). 

Red  snapper  ages  ranged  from  1  to  45  years  and  the 
majority  (90%)  of  individuals  were  between  2  and  6 
years  (Fig.  4).  Alabama  fish  ranged  from  1  to  35  years 
(re=1985),  Louisiana  fish  ranged  from  2  to  37  years 
(n=1864).  and  Texas  fish  ranged  from  1  to  45  years 
(rc=1186).  Modal  ages  were  4  years  for  Alabama  and  3 
years  for  Louisiana  and  Texas  red  snapper.  We  found 
significant  differences  among  age-frequency  distribu- 
tions from  all  states  (AL  and  LA,  MD  =  9;  AL  and  TX, 
MD  =  33.84;  and  LA  and  TX,  MD=24.84).  Texas  had  a 
much  higher  proportion  of  younger  individuals;  63%  of 
sampled  fish  were  aged  at  3  years  or  less  compared  to 
only  30%  of  Alabama  and  39%  of  Louisiana  fish  aged 
at  3  years  or  less. 

Red  snapper  growth  was  modeled  from  weighted  mean 
FL  at  age  and  mean  TW  at  age  by  using  the  von  Berta- 
lanffy  growth  equation  (Fig.  5,  A  and  B).  Resultant  von 
Bertalanffy  growth  equations  were 


Table  2 

Differences  between  two  readers  in  average  percent  error 

(APE). 

?oefficient  of  variation  (CV),  index  of 

precision  (D). 

and  in  percentages  of  agreement  ( O )  for  counts  of  opaque 

annuli 

in  red  snapper 

(Lutjanus  campechanus)  otoliths 

after  fi 

rst  and  second 

readings  for  each 

sample  year. 

«=number  of  otoliths  sa 

mpled. 

Year 

1st  reading 

2nd  reading 

1999  in 

=2100) 

APE 

0.483 

0.499 

CV 

0.014 

0.0008 

D 

0.010 

0.0006 

O 

89.48% 

99.43% 

±1 

8.62% 

0.48% 

±2 

1.19% 

0.095% 

±3 

0.71'. 

2000  (n 

=2069) 

APE 

0.487 

0.499 

CV 

0.034 

0.0006 

D 

0.024 

0.0004 

O 

73.79f; 

99.47% 

±1 

22.49', 

0.53% 

±2 

1.78% 

±3 

1.93% 

2001  in 

=  866) 

APE 

0.459 

0.498 

CV 

0.032 

0.0005 

D 

0.023 

0.0003 

O 

74.73* 

99.42^ 

±1 

22.06% 

0.58% 

±2 

2.27% 

±3 

0.94% 

ALFL.    =  839(1 -e1-0381"1) 

(F,  15=  2824.9;  P<0.0001;  r2=0.95); 

LAFL„  =  847.8(1  -  e'-° •25"") 

(F1;  13=5024.4;  P<0.0001;  r2=0.76); 

TXFL„  =  778.2(1  -  e<-o.49tt)>) 

(F1;19=1452.1;  P<0.001;  r2=0.85); 


AL  TW 


17.05(1  -e 


(-0.15inii3.03 


(F1;15=457.9;  P<0.0001;  r2=0.89); 

LA  TW_._  =  12.61(1  -  ec-o.32(»))3.03 

(F114=122.02;  P<0.0001;  ;-2  =  0.18); 

TX  TWrr,  =  8.89(1  -  e'-0-21"")2  84 

(F1;12= 613.01;  P<0.0001;  r2=0.96). 

Models  of  mean  red  snapper  FL  at  age  for  Alabama  and 
Louisiana  were  markedly  similar  with  likelihood  ratio 
tests  indicating  no  significant  differences  between  red 
snapper  from  the  two  states  (Table  3).  However,  the 


Texas  model  differed  from  both  Alabama  and  Louisiana 
models.  The  Texas  model  displayed  significant  differ- 
ences from  the  other  models  in  both  Lm  and  in  k.  A 
comparison  of  the  models  of  mean  TW  at  age  indicated 
no  significant  differences  between  Alabama  and  Loui- 
siana red  snapper  (Table  3).  Differential  growth  in  TW 
was  found  when  comparing  Alabama  and  Louisiana  with 
the  Texas  model;  significant  differences  were  manifested 
in  both  WM  and  in  k.  The  model  failed  to  converge  for 
estimating  a  common  value  of  k  for  both  Louisiana  and 
Texas. 

We  recognized  that  the  larger  red  snappers  from 
Louisiana  might  bias  the  data;  therefore  we  compared 
growth  for  fish  from  2  to  10  years  of  age — a  time  pe- 
riod when  red  snapper  have  demonstrated  rapid  linear 
growth  (Szedlmayer  and  Shipp,  1994;  Patterson  et  al., 
2001;  Wilson  and  Nieland,  2001).  Linear  regressions 
of  mean  FL  at  age  for  all  individuals  2  to  10  years 
(Fig.  6A)  were  compared  among  states.  We  found  no 
significant  differences  among  states  (ANCOVA  test  of 
homogeneity  of  slopes,  F2;28=2.7;  P=0.08;  ANCOVA  test 
for  equal  intercepts,  F2.28=0.52;  P=0.6). 


598 


Fishery  Bulletin  102(4) 


Mean  TW  at  age  was  also  examined  among  states  for 
red  snapper  2  to  10  years  in  age  as  above  (Fig.  6B).  No 
significant  differences  were  found  between  Alabama  and 
Louisiana  (ANCOVA  test  of  homogeneity  of  slopes,  F117= 
0.1;  P=0.75;  ANCOVA  test  for  equal  intercepts,  F1;'17= 
0.26;  P=0.66  for  intercepts).  However,  a  significant 
difference  between  slopes  was  detected  when  compar- 
ing Alabama  and  Texas  red  snapper  (ANCOVA  test  of 
homogeneity  of  slopes,  F1;16=19.68;  P<0.0007;  ANCOVA 
test  for  equal  intercepts,  F1;16=2.74;  P<0.12).  The  same 
was  found  when  comparing  slopes  for  Louisiana  and 
Texas  red  snapper  I  ANCOVA  test  of  homogeneity  of 


Figure  3 

Marginal  increment  analysis  of  red  snapper  {Lutjanus  campecha- 
nus)  otoliths  for  specimens  from  Alabama  (n  =  1985l,  Louisiana 
(n  =  1864),  and  Texas  <n  =  1186>. 


45  -, 

40  ■ 

J 

DAL 

35  - 

■  LA 

£.    30  - 

DTX 

Frequency 

o        en 

| 

- 

15  - 

n 

10  - 

1 

5  - 

JJ 

1 

flflr^m-rm.^         .     _       r». 

o  - 

1 

2       3 

4       5      6       7       8      9      10     11     12     13     14    15    >15 

Age  (yr) 

Figure  4 

Age  distri 

butions  for  red  snapper  (Lutjanus  campechanus)  sam- 

pled  from 

the  1999-2001  recreational  harvests  from  Alabama, 

Louisiana 

,  and  Texas. 

slopes,  Fl   16=9.62;  P<0.008)  but  not  when  comparing 
intercepts'  (Fh  16  =  0.64;  P<0.44). 


Discussion 

Demographic  variations  in  growth  rates  and  in  size- 
frequency  distributions  may  indicate  the  existence  of 
isolated  management  units  of  red  snapper  in  the  north- 
ern GOM.  The  recreational  harvests  of  Alabama  and 
Louisiana  red  snapper  were  dominated  by  individuals 
ranging  from  375  to  425  mm  FL,  whereas  the  majority 
of  Texas  fish  (69%)  were  375  mm  FL  or  less.  It 
was  within  this  size  range  (375-400  mm  FL) 
that  the  significant  differences  in  red  snapper 
among  states  were  detected.  The  FL  distribu- 
tion of  red  snapper  sampled  in  Texas  also  dif- 
fered from  those  for  Alabama  and  Louisiana; 
there  were  very  few  large  fish  represented  in 
the  Texas  sample  population,  partly  because 
fishing  tournaments  (where  larger  individuals 
are  targeted)  were  not  sampled  in  Texas.  Signifi- 
cant differences  in  TW  frequencies  among  states 
were  also  detected  at  approximately  1  kg  (the 
approximate  weight  of  a  red  snapper  375-400 
mm  FL);  86%  of  Texas  fish  weighed  1  kg  or  less, 
compared  to  only  27%  of  Alabama  fish  and  28% 
of  Louisiana  fish  in  this  size  range. 

One  factor  possibly  contributing  to  the  modal 
size  class  difference  was  the  type  of  fishing 
vessel  used  to  catch  the  fish.  The  majority  of 
Texas  specimens  I~95fr<)  were  sampled  from 
headboats;  whereas  Louisiana  and  Alabama 
fish  were  obtained  almost  exclusively  from  char- 
terboats.  This  is  not  to  say  that  charterboats 
were  purposely  excluded  from  the  Texas  sur- 
vey. On  the  contrary,  red  snapper  were  sampled 
from  any  and  all  available  recreational  fishing 
parties  at  the  three  individual  sampling  loca- 
tions. Differences  in  modal  size  and  number  of 
red  snapper  caught  per  person  onboard  charter- 
boats  versus  headboats  may  be  inconsequential 
considering  that  both  trip  types  used  similar 
gear  and  targeted  similar  or  the  same  fishing 
locations.  It  should  be  noted  however  that  in 
the  Texas  study  area,  charterboats  routinely 
frequented  a  wider  array  of  fishing  spots  (rigs, 
hardbottom.  wrecks,  etc.)  than  did  headboats, 
which  typically  return  to  the  same  few  rigs  and 
large  structures  over  and  over  again  iTolan 5), 
Our  von  Bertalanffy  growth  models  on 
FL  at  age  showed  that  red  snapper  from  all 
three  states  exhibit  a  pattern  of  rapid,  linear 
growth  to  approximately  10  years,  after  which 
maximum  theoretical  (asymptotic)  FL  is  soon 


Tolan,  J.  2003.  Personal  commun.  Texas  Parks 
and  Wildlife  Department,  Resource  Protection,  6300 
Ocean  Dr.,  Corpus  Christi,  TX  78412. 


Fischer  et  al.:  Demographic  structure  of  Lut/anus  campechanus  in  the  northern  Gulf  of  Mexico 


599 


Table  3 

Chi-square  if),  degrees  of  freedom  (df ),  and  P-values  for  likelihood  ratio  tests  for  comparing  FL  and  TW  von  Bertalanffy  growth 
models  and  parameters  among  sample  locations  (states).  AL=  Alabama;  LA=  Louisiana:  and  TX=Texas.  n/a=not  available. 


r 

df 

P 


r 

df 

p 


AL-LA 


FL  model 


2.54 
1,28 
0.11 


TW  model 


2.15 
1,29 
0.14 


FL  model 


5.14 
1,34 
0.023 


TW  model 


38.8 

1,27 

4.7x10" 


AL-TX 


LA-TX 


k 


FL  model 


13.67 

1,34 

0.0002 


21.53 

1,34 

3.48xl0"6 


5.8 
1,32 
0.015 


k 


TW  model 


21.3 

1,27 
3.9x10" 


37.8 

1,27 

7.97xl0-10 


16.77 

1.26 

4.2xl0"5 


10.16 
1,32 
0.001 


15.1 
1,26 
0.001 


k 


9.8 
1,32 
0.002 


n/a 
n/a 
n/a 


reached  and  growth  in  length  becomes  negligi- 
ble. This  pattern  of  rapid  growth  was  similar  to 
that  reported  in  previous  studies  (Szedlmayer 
and  Shipp,  1994;  Manooch  and  Potts,  1997; 
Patterson,  1999;  Wilson  and  Nieland,  2001). 
However,  our  models  predicted  smaller  L  r  and 
higher  values  of  It.  Because  of  the  minimum 
size  limits  on  the  recreational  fishery,  very  few 
fish  under  age  2  years  (>300  mm  FL)  were  in- 
cluded in  our  sample  populations.  We  forced  our 
models  through  t0  =  0  to  more  accurately  pre- 
dict juvenile  growth,  which  in  turn  increased 
our  estimates  of  k.  In  addition,  we  had  a  much 
larger  sample  population  that  included  more 
older,  larger  fish  than  most  of  the  previously 
cited  studies.  These  larger  fish  pulled  the  curve 
down,  driving  the  lesser  estimations  of  LM.  The 
lack  of  significant  differences  in  growth  param- 
eters between  the  Alabama  and  Louisiana  mod- 
els supports  the  findings  of  previous  research, 
which  indicates  that  Alabama  and  Louisiana 
red  snapper  grow  at  similar  rates  and  reach 
comparable  sizes  (Patterson  et  al.,  2001).  How- 
ever, values  of  LM  for  Texas  red  snapper  were 
significantly  smaller  than  parameters  predicted 
for  Alabama  and  Louisiana  red  snapper.  In- 
terestingly, Texas  had  a  value  of  k  that  was 
significantly  larger  then  that  for  Alabama  and 
Louisiana  and  this  would  indicate  that  Texas 
fish  obtain  a  smaller  maximum  theoretical  FL 
but  reach  it  at  a  faster  rate  then  fish  from  Ala- 
bama and  Louisiana. 

Von  Bertalanffy  growth  models  of  mean 
weight  at  age  produced  similar  results,  in- 
dicating that  Texas  red  snapper  obtain  sig- 
nificantly smaller  maximum  theoretical  TW 
than  fish  from  Alabama  and  Louisiana.  Fish 
sampled  from  tournaments  were  excluded  from 
all  growth  models  to  more  accurately  reflect 


30    35    40    45 


B 


15  ■ 


12  - 


0     5     10    15    20    25    30    35    40    45 
Age  (yr) 

Figure  5 

Observed  (A)  mean  fork  length  (mml  at  age  and  (B)  mean  total 
weight  (kg)  at  age  for  red  snapper  iLutjanus  campechanus)  from 
Alabama,  Louisiana,  and  Texas.  Plotted  lines  are  weighted  von 
Bertalanffy  growth  functions  fitted  to  the  data. 


600 


Fishery  Bulletin  102(4) 


growth  of  a  random  population.  Tournament  anglers 
target  large  fish,  possibly  the  fastest  growing  individu- 
als at  a  given  age,  and  their  catches  may  bias  growth 
estimates  (Ottera,  1992;  Vaughan  and  Burton,  1993; 
Goodyear,  1995).  Without  these  tournament  fish,  how- 
ever, the  Alabama  red  snapper  TW  model  did  not  reach 
an  asymptote.  Therefore  the  growth  parameters  for 
that  model  were  poorly  estimated.  Notwithstanding, 
Alabama  and  Louisiana  models  did  not  differ  signifi- 
cantly. Estimates  of  Wm  and  k  predicted  for  Louisiana 
red  snapper  were  slightly  larger  than  previously  re- 
ported for  fish  from  the  Louisiana  commercial  and 
recreational  catches  (Render,  1995).  Although  the  Texas 
model  predicted  a  value  of  Wc.  that  was  significantly 
less  than  those  for  both  Alabama  and  Louisiana  red 
snapper,  Texas  had  a  growth  coefficient  (k)  that  was 
larger  then  that  for  Alabama.  It  appears  that,  as  in  the 
length  models,  Texas  fish  reach  a  smaller  theoretical 
maximum  weight  but  at  a  faster  rate  than  Alabama 
fish.  Louisiana  fish  attained  maximum  weight  at  a 
faster  rate  than  Alabama  or  Texas  red  snapper.  Our 
growth  models  indicate  that  although  Texas  red  snap- 


900 

800 

E 
E 

700 
600 

*— 
c 

CD 

500 
400 

O 

u. 

300 

200 

100 

0 

12 

10 

A 

j 

A 

X 

'.     ^T^' 

r                 \_ 

•  AL 

1 

LA 
XTX 

— i 

10 


B 


01  23456789         10 

Age  (yr) 

Figure  6 

Scattergram  with  linear  regression  lines  for  relationships 
(A)  between  age  (yr)  and  mean  fork  length  (mm)  and  (B)  age 
(yr)  and  mean  total  weight  (kg)  for  red  snapper  (Lutjanus 
campechanus)  aged  1  to  10  years  from  the  1999-2001  rec- 
reational harvests  of  Alabama.  Louisiana,  and  Texas.  Krror 
bars  represent  standard  deviations  from  the  mean. 


per  grow  in  mass  at  a  faster  rate  than  Alabama  fish. 
Texas  red  snapper  are  consistently  smaller  at  age  and 
reach  smaller  maximum  sizes  than  those  from  Alabama 
and  Louisiana  and  that  there  is  a  veritable  difference 
in  size  at  age  and  growth  rates  among  regions.  Similar 
demographic  variations  in  growth  rates  among  popula- 
tions have  been  previously  noted  for  other  marine  fish 
species  of  the  South  Atlantic  and  GOM,  such  as  gray 
snapper  (Johnson  et  al.,  1994;  Burton  2001),  and  king 
mackerel  (DeVries  et  al,  1990;  DeVries  and  Grimes, 
1997). 

Linear  regressions  of  mean  FL  and  mean  TW  at  age 
for  red  snapper  aged  one  to  10  years  indicated  that  only 
TW  was  significantly  different  among  sample  regions. 
Texas  red  snapper  were  shown  to  differ  significantly 
from  both  Alabama  and  Louisiana  red  snapper  in  re- 
gressions of  mean  weight  at  age.  Although  comparisons 
of  FL  at  age  for  all  regions  were  not  significantly  differ- 
ent, Texas  fish  were  significantly  smaller  in  mass  (TW) 
at  age  than  fish  from  Alabama  and  Louisiana.  This 
difference  was  observed  in  all  age  classes. 

Our  research  efforts  indicate  that  there  is  mounting 
evidence  for  discrete  differences  in  size  at  age 
and  in  overall  growth  rates  between  red  snapper 
sampled  from  the  north  central  GOM  (Louisiana 
and  Alabama)  and  the  southwest  GOM  (Texas). 
Texas  red  snapper  are  clearly  reaching  smaller 
maximum  sizes  and  are  consistently  smaller  (TWi 
at  age  than  those  collected  from  Louisiana  and 
Alabama  waters.  Although  the  reasons  behind 
these  differences  remain  uncertain,  logic  indicates 
that  factors  such  food  availability,  habitat  prefer- 
ence, and  actual  population  size  may  cause  these 
differences  between  regions. 

The  more  productive,  nutrient-rich  waters  of 
the  Mississippi  River  and  north-central  GOM  off 
Louisiana  and  Alabama  may  be  more  conducive 
to  faster  growth  than  the  less  fertile  waters  off 
Texas.  Approximately  70-80%  of  GOM  fishery 
landings  come  from  the  waters  surrounding  the 
Mississippi  River  delta  (Grimes,  2001).  The  west- 
ern GOM  (including  the  sampling  area  of  Port 
Aransas,  TX)  is  devoid  of  a  contributing  river  sys- 
tem anything  remotely  similar  to  the  Mississippi 
River.  Draining  43%  of  the  continental  United 
States,  the  Mississippi  River  is  the  largest  river 
system  in  North  America  and  provides  an  enor- 
mous amount  of  nutrient-laden  fresh  water  to  the 
shallow  continental  shelf  of  the  northern  GOM. 
Although  the  mechanics  by  which  the  Mississippi 
River  enhances  fishery  production  remain  uncer- 
tain. Grimes  (2001)  postulated  that  the  discharge 
from  the  Mississippi  primarily  influences  recruit- 
ment m  the  plume  field.  Increased  growth  rates 
associated  with  the  Mississippi  River  plume  com- 
pared with  other  regions  of  the  GOM  have  been 
noted  for  a  number  of  species,  such  as  gulf  menha- 
den (Warlen,  1988),  king  mackerel  (DeVries  et  al., 
1990),  striped  anchovy  (Day,  1993),  and  yellowfin 
tuna  (Lang  et  al.,  1994). 


Fischer  et  al.:  Demographic  structure  of  Lut/anus  campechanus  in  the  northern  Gulf  of  Mexico 


601 


In  addition  to  increased  food  availability  off  of  the 
north-central  GOM,  the  amount  and  condition  of  pre- 
ferred habitat  may  have  some  effect  on  the  observed 
differences  in  growth  rates  for  Texas  and  those  for 
Louisiana  and  Alabama.  Approximately  95"%  of  all 
Louisiana  fishes  sampled  in  this  study  were  harvested 
from  waters  surrounding  nearshore  (<50  km)  oil  and 
gas  platforms.  Similarly,  about  95%  of  all  Alabama 
fishes  sampled  were  caught  over  artificial  reef  sites. 
The  fact  that  there  was  no  detectable  difference  in  size 
at  age  and  overall  growth  rates  between  Louisiana  and 
Alabama  red  snapper  therefore  is  not  surprising,  given 
the  similarity  in  the  habitats  sampled  and  the  proxim- 
ity of  both  locations  to  the  Mississippi  River  discharge 
plume.  Texas  was  the  only  area  in  which  samples  were 
routinely  obtained  from  natural  hard  bottoms  (40%), 
as  well  as  from  oil  and  gas  platforms  and  artificial 
reefs  (60%).  Given  that  more  than  half  of  the  Texas 
specimens  were  captured  in  the  waters  immediately 
surrounding  artificial  structures  (i.e.,  oil  and  gas  plat- 
forms), we  can  assume  that  habitat  type  is  not  be  the 
sole  source  for  the  observed  differences  in  growth  rates 
among  regions. 

Despite  the  current  acceptance  of  a  unit  stock  hypoth- 
esis for  GOM  red  snapper,  the  species  is  not,  and  to  our 
knowledge  never  has  been,  uniformly  distributed  across 
the  northern  GOM.  The  fishery  for  red  snapper  began 
in  northwest  Florida  approximately  20  years  before  the 
Civil  War  (Collins,  1887)  and  during  that  time  period 
was  centered  between  Mobile,  AL,  and  Fort  Walton,  FL 
(Camber,  1955).  One  hundred  years  of  landings  data 
indicate  that  the  fishery,  and  possibly  the  population, 
has  undergone  a  major  shift  from  the  natural  outcrop- 
pings  of  the  West  Florida  Shelf  to  oil  and  gas  platforms 
of  the  north-central  portion  of  the  GOM  (Shirripa  and 
Legault1).  Fishery-dependent  data  indicate  that  cur- 
rently there  is  a  center  of  abundance  of  red  snapper  off 
southwest  Louisiana  and  a  second,  smaller  center  off 
Alabama  (Patterson  et  al.,  2001;  Goodyear4;  Shirripa 
and  Legault1).  Patterson  et  al.  (2001)  stated  that  Loui- 
siana and  Alabama  accounted  for  32.6%  and  11.4%, 
respectively,  of  the  combined  recreational  and  commer- 
cial GOM  landings  from  1981  to  1998.  This  is  especially 
surprising  for  Alabama,  considering  that  its  coastline 
accounts  for  only  3%  of  the  GOM  coastline  from  the 
Texas-Mexico  border  to  the  southern  tip  if  Florida  (Pat- 
terson et  al.,  2001). 

Red  snapper  have  never  been  reported  to  be  plentiful 
in  Texas  waters,  despite  the  availability  of  suitable  hab- 
itat in  the  form  of  natural  hard  bottom  and  the  cur- 
rent high  concentration  of  oil  and  gas  platforms.  In  a 
historical  report  on  red  snapper  fishing  in  the  GOM, 
Camber  ( 1995 )  reported  that  although  a  few  red  snap- 
per were  taken  from  the  "Galveston  Lumps"  or  the 
"Western"  fishing  grounds  off  Texas,  the  fishery  never 
fully  developed  in  this  region  during  the  latter  part  of 
the  nineteenth  century.  Commercial  landings  for  red 
snapper  from  the  GOM  indicated  that  Texas  accounted 
for  approximately  only  18%  of  the  total  catch  during 
the  time  period  1981-95  (Goodyear6).  In  a  recent  fish- 


ery-dependent survey  of  recreational  headboat  discards 
and  landings  in  Texas  coastal  waters,  red  snapper  less 
than  the  minimum  legal  size  (15  inches)  made  up  64% 
of  the  catch  (Dorf.  2000).  In  the  latter  study,  Galveston, 
Port  Aransas,  and  Port  Isabel  were  surveyed  to  canvas 
a  large  portion  of  the  Texas  coast.  Discard-to-landing 
ratios  were  as  high  as  211:1  in  the  waters  off  Galveston 
and  were  possibly  indicative  of  the  paucity  of  legal-size 
red  snapper  in  Texas  waters.  Of  the  three  sampling 
locations.  Port  Aransas  had  the  lowest  discard-to-land- 
ing ratio  (5.2:1)  and  the  largest  mean  fish  length  and 
weight  (387  mm,  0.9  kg) — length  and  weight  data  that 
are  consistent  with  a  3-yr-old  fish  from  our  Texas  (Port 
Aransas)  specimens.  The  majority  of  Texas  fish  (63%) 
were  aged  at  3  years  or  less.  Age  distribution,  along 
with  FL  and  TW  distributions,  may  indicate  that  red 
snapper  are  being  harvested  from  Texas  waters  just 
as  they  reach  legal  size.  Given  the  vast  differences 
in  historical  landings  data  between  the  northern  and 
southwest  GOM,  the  highly  disproportionate  discard- 
to-landing  ratio  reported  for  headboats  in  Texas  wa- 
ters (Dorf,  2000),  and  the  large  number  of  young  fish 
sampled  in  Texas,  it  is  not  inconceivable  to  speculate 
that  there  are  fewer  red  snapper  available  for  harvest 
in  Texas  waters. 

Demographic  variation  in  growth  rates  may  indicate 
the  existence  of  separate  management  units  of  red  snap- 
per in  the  GOM.  Our  data  indicate  that  the  red  snapper 
inhabiting  the  waters  off  Texas  are  reaching  smaller 
maximum  sizes  at  a  faster  rate,  but  are  consistently 
smaller  (TW)  at  age  than  those  collected  from  Louisi- 
ana and  Alabama  waters.  Whether  these  differences  are 
environmentally  induced  or  result  from  genetic  diver- 
gence remains  to  be  determined.  The  more  productive, 
nutrient-rich  waters  of  the  Mississippi  River  and  north- 
central  GOM  off  Louisiana  and  Alabama  may  be  more 
conducive  to  faster  growth  than  the  less  fertile  waters 
off  Texas.  Fishing  pressure  and  its  effects  on  population 
size  may  also  be  leading  to  the  observed  differences  in 
growth  rates.  Fishery-dependent  landing  data  and  dis- 
proportionate discard-to-landing  ratios  in  Texas  waters 
loosely  support  the  concept  that  fewer  red  snapper  are 
available  for  harvest  in  the  southwest  GOM.  Regardless 
of  the  cause,  the  existence  of  demonstrable  demographic 
differences  argues  for  the  delineation  of  multiple  red 
snapper  management  units  in  the  GOM. 


Acknowledgments 

Funding  for  this  research  was  provided  by  the  U.S. 
Department  of  Commerce  Marine  Fisheries  Initiative 
(MARFIN)  program  (grant  number  NA87FF0424).  We 


6  Goodyear,  C.  P.  1996.  An  update  of  red  snapper  harvest 
in  U.S.  waters  of  the  Gulf  of  Mexico.  Report  MIA-95/96- 
60,  21  p.  Miami  Laboratory,  Southeast  Fisheries  Center, 
National  Marine  Fisheries  Service.  75  Virginia  Beach  Dr. 
Miami,  FL.,  33149-1099.     [Not  available  from  NTIS]. 


602 


Fishery  Bulletin  102(4) 


thank  Forrest  Davis,  John  Gold,  Jessica  McCawley, 
Linda  Richardson,  Jim  Tolan,  Melissa  Woods,  Candace 
Aiken,  and  many  others  for  help  with  sampling  red  snap- 
per. We  also  thank  Josh  Maier  and  Brett  Blackman  for 
otolith  sectioning.  We  thank  Steve  Tomeny  and  his  boat 
captains  and  crew  (Port  Fourchon,  LA),  as  well  as  all 
boat  captains  and  crews  in  Dauphin  Island,  Alabama, 
and  Port  Aransas,  Texas,  for  graciously  allowing  us  to 
sample  fish  from  their  charter  fishing  vessels.  We  also 
thank  Yvonne  Allen  for  providing  the  map  in  Figure  1. 
Finally  we  would  like  to  thank  Dave  Nieland  for  time 
spent  fielding  questions  concerning  statistical  analysis 
and  for  a  constructive  review  of  this  manuscript. 


Literature  cited 

Baker,  M.  S.  Jr.,  and  C.  A.  Wilson. 

2001.  Use  of  bomb  radiocarbon  to  validate  otolith  sec- 
tion ages  of  red  snapper  Lutjanus  campechanus  from 
the  northern  Gulf  of  Mexico.  Limnol.  Oceanogr. 
46:1819-1824. 

Beamish,  R.  J.,  and  D.  A.  Fournier. 

1981.  A  method  for  comparing  the  precision  of  a  set 
of  age  determinations.  Can.  J.  Fish.  Aquat.  Sci. 
38:982-983. 

Beckman,  D.  W.,  C.  A.  Wilson,  and  A.  L.  Stanley. 

1989.  Age  and  growth  of  red  drum,  Sciaenops  ocel- 
latus,  from  offshore  waters  of  the  northern  Gulf  of 
Mexico.     Fish.  Bull.  87:17-28. 

Burton,  M.  L. 

2001.     Age.  growth,  and  mortality  of  gray  snapper,  Lut- 
janus griseus,  from  the  east  coast  of  Florida.     Fish. 
Bull.  99:254-265. 
Camber,  C.  I. 

1955.     A  survey  of  the  red  snapper  fishery  of  the  Gulf 
of  Mexico,  with  special  reference  to  the  Campeche 
banks.     Fla.  Board  Conserv.  Tech.  Ser.  No.  12:1-64. 
Camper,  J.  D.,  R.  C.  Barber,  L.  R.  Richardson,  and  J.  R.  Gold. 
1993.     Mitochondrial  DNA  variation  among  red  snapper 
(Lutjanus  campechanus)  from  the  Gulf  of  Mexico.     Mol. 
Mar.  Biol.  Biotech.  2:154-161. 
Cerrato,  R.  M. 

1990.  Interpretable  statistical  tests  for  growth  com- 
parisons using  parameters  in  the  von  Bertalanffy 
equation.     Can.  J.  Fish.  Aquat.  Sci.  47:1416-1426. 

Chang,  W.  B. 

1982.  A  statistical  method  for  evaluating  the  repro- 
ducibility of  age  determination.  Can.  J.  Aquat.  Sci. 
39:1208-1210. 

Collins,  J.  W. 

1887.     Report  on  the  discovery  and  investigation  of  fishing 
grounds  made  by  the  Fish  Commision  Steamer  Albatross 
during  a  cruise  along  the  Atlantic  Coast  and  in  the  Gulf 
of  Mexico.     Rep.  U.S.  Comm.  Fish.  13:217-311. 
Cowan,  J.  H.  Jr.,  R.  L.  Shipp,  H.  K.  Bailey  IV,  and 
I).  W.  Hnywick. 

1995.     Procedure  for  rapid  processing  of  large  otoliths. 
Trans.  Am.  Fish.  Soc.  124:280-282. 
Day,  G   K 

1993.  Distribution,  abundance,  growth  and  mortality  of 
striped  anchovy,  Anchoa  hepsetus,  about  the  discharge 
plume  of  the  Mississippi  River.  M.S.  thesis,  45  p.  Univ. 
West  Florida,  Pensacola.  FL. 


DeVries,  D.  A.,  and  C.  B.  Grimes. 

1997.     Spatial  and  temporal  variation  in  age  and  growth  of 
king  mackerel,  Scomeromorus  cavalla.  1977-1992.     Fish. 
Bull.  95:694-708. 
Devries,  D.  A.,  C.  B.  Grimes.  K.  C.  Lang,  and  D.  B.  White. 

1990.     Age  and  growth  of  king  and  Spanish  mackerel 
larvae  and  juveniles  from  the  Gulf  of  Mexico  and  U.S. 
south  Atlantic.     Environ.  Biol.  Fishes  29:135-143. 
Dorf,  B.  A. 

2003.  Red  snapper  discards  in  Texas  coastal  waters:  a 
fishery-dependent  onboard  survey  of  recreational  head- 
boat  discards  and  landings.  In  Fisheries,  reefs,  and 
offshore  development  (D.  R.  Stanley  and  A.  Scarborough- 
Bull,  eds.i,  p.  155-166.  Am.  Fish.  Soc.  Symposium  36. 
Am.  Fish.  Soc,  Bethesda.  MD. 
Gold  J.  R.,  F.  Sun,  and  L.  R.  Richardson. 

1997.     Population  structure  of  red  snapper  from  the  Gulf 
of  Mexico  as  inferred  from  analysis  of  mitochondrial 
DNA.     Trans.  Am.  Fish.  Soc.  126:386-396. 
Goodyear,  C.  P. 

1995.     Mean  size  at  age:  an  evaluation  of  sampling  strat- 
egies using  simulated  red  grouper  data.     Trans.  Am. 
Fish.  Soc.  124:746-755. 
Grimes,  C.  B. 

2001.     Fishery  production  and  the  Mississippi  River 
discharge.     Fisheries  26:7-26. 
Gulland,  J.  A. 

1965.     Fish  stock  assessment:  a  manual  of  basic  methods, 
223  p.     John  Wiley  &  Sons,  New  York,  NY. 
Heist,  E.  J,  and  J.  R.  Gold. 

2000.  DNA  microsatellite  loci,  and  genetic  structure  of 
red  snapper  in  the  Gulf  of  Mexico.  Trans.  Am.  Fish. 
Soc.  129:469-475. 

Johnson,  A.  G,  L.  A.  Collins,  and  C.  P.  Kelm. 

1994.  Age-size  structure  of  gray  snapper  from  the  South- 
eastern United  States:  a  comparison  of  two  methods  of 
back-calculating  size  at  age  from  otolith  data.  Proc. 
Annu.  Conf.  Southeast  Assoc.  Fish  and  Wildl.  Agencies 
48:592-600. 

Lang,  K.  L..  C.  B.  Grimes,  and  R.  F.  Shaw. 

1994.  Variations  in  the  age  and  growth  of  yellowfin  tuna 
larvae.  Thunnus  albacares,  collected  about  the  Mississippi 
River  plume.     Environ.  Biol.  Fishes  39:259-270. 

Manooch,  C.  S.,  Ill,  and  J.  C.  Potts 

1997.     Age  and  growth  of  red  snapper.  Lutjanus  campecha- 
nus, Lutjanidae,  collected  along  the  southeastern  United 
States  from  North  Carolina  through  the  east  coast  of 
Florida.     J.  Elisha  Mitchell  Sci.  Soc.  113:111-122. 
Ottera.  H. 

1992.     Bias  in  calculating  growth  rates  in  cod  (Gadus 
morhua  L.)  due  to  size  selective  growth  and  mortality.     J. 
Fish  Biol.  40:465-467. 
Patterson,  W.  F.  III. 

1999.     Aspects  of  the  population  ecology  of  red  snapper 
Lutjanus  campechanus  in  an  artificial  reef  area  off 
Alabama.     Ph.D.  diss..  164  p.     Univ.  South  Alabama, 
Mobile.  AL. 
Patterson.  W.  F  III.  James  H.  Cowan  Jr..  Charles  A.  Wilson, 
and  Robert  L.  Shipp. 

2001.  Age  and  growth  of  red  snapper,  Lutjanus  campecha- 
nus, from  an  artificial  reef  area  off  Alabama  in  the 
northern  Gulf  of  Mexico.     Fish.  Bull.  99:617-627. 

Render.  .Ill 

1995.  The  life  history  (age,  growth,  and  reproduction) 
of  red  snapper  [Lutjanus  campechanus)  and  its  affinity 


Fischer  et  al.:  Demographic  structure  of  Lut/anus  campechanus  in  the  northern  Gulf  of  Mexico 


603 


for  oil  and  gas  platforms.     Ph.D.  diss.,  76  p.     Louisiana 
State  Univ.,  Baton  Rouge,  LA. 

Ricker,  W.  E. 

1975.  Computation  and  interpretation  of  biological  statis- 
tics offish  populations,  382  p.  Bull.  Fish.  Res.  Board 
Can.,  Ottawa. 

Szedlmayer,  S.  T.,  and  R.  L.  Shipp. 

1994.  Movement  and  growth  of  red  snapper,  Lutja- 
nus  campechanus.  from  an  artificial  reef  area  in  the 
northeastern  Gulf  of  Mexico.  Bull.  Mar.  Sci.  55(2-3): 
887-896. 

Tate.  M.  W..  and  R.  C.  Clelland. 

1957.  Non-parametric  and  shortcut  statistics  in  the  social, 
biological,  and  medical  sciences,  p.  93-94.  Interstate 
Printers  and  Publishers,  Inc.,  Danville,  IL. 


Vaughan,  D.  S.,  and  M.  L.  Burton. 

1993.     Estimation  of  von  Bertalanffy  growth  parameters 
in  the  presence  of  size-selective  mortality:  a  simulated 
example  with  red  grouper.     Trans.  Am.  Fish.  Soc. 
123:1-8. 
Warlen,  S.  M. 

1988.     Age  and  growth  of  larval  gulf  menhaden,  Brevoortia 
patronis,  in  the  northern  Gulf  of  Mexico.     Fish.  Bull. 
86:77-90. 
Wilson,  C.  A.,  and  D.  L.  Nieland. 

2001.  Age  and  Growth  of  red  snapper,  Lutjanus 
campechanus,  from  the  northern  Gulf  of  Mexico  off 
Louisiana.     Fish.  Bull.  99:653-664. 


604 


Abstract— The  population  struc- 
ture of  walleye  pollock  \Theragra 
chalcogramma)  in  the  northeastern 
Pacific  Ocean  remains  unknown. 
We  examined  elemental  signatures 
in  the  otoliths  of  larval  and  juvenile 
pollock  from  locations  in  the  Bering 
Sea  and  Gulf  of  Alaska  to  determine 
if  there  were  significant  geographic 
variations  in  otolith  composition 
that  may  be  used  as  natural  tags  of 
population  affinities.  Otoliths  were 
assayed  by  using  both  electron  probe 
microanalysis  (EPMA)  and  laser 
ablation  inductively  coupled  plasma 
mass  spectrometry  iICP-MS).  Ele- 
ments measured  at  the  nucleus  of 
otoliths  by  EPMA  and  laser  abla- 
tion ICP-MS  differed  significantly 
among  locations.  However,  geographic 
groupings  identified  by  a  multivariate 
statistical  approach  from  EPMA  and 
ICP-MS  were  dissimilar,  indicating 
that  the  elements  assayed  by  each 
technique  were  controlled  by  sepa- 
rate depositional  processes  within  the 
endolymph.  Elemental  profiles  across 
the  pollock  otoliths  were  generally- 
consistent  at  distances  up  to  100  fim 
from  the  nucleus.  At  distances  beyond 
100  /im,  profiles  varied  significantly 
but  were  remarkably  consistent 
among  individuals  collected  at  each 
location.  These  data  may  indicate 
that  larvae  from  various  spawning 
locations  are  encountering  water 
masses  with  differing  physicochemical 
properties  through  their  larval  lives, 
and  at  approximately  the  same  time. 
Although  our  results  are  promising, 
we  require  a  better  understanding  of 
the  mechanisms  controlling  otolith 
chemistry  before  it  will  be  possible 
to  reconstruct  dispersal  pathways  of 
larval  pollock  based  on  probe-based 
analyses  of  otolith  geochemistry. 
Elemental  signatures  in  otoliths  of 
pollock  may  allow  for  the  delineation 
of  fine-scale  population  structure  in 
pollock  that  has  yet  to  be  consistently- 
revealed  by  using  population  genetic 
approaches. 


Elemental  signatures  in  otoliths  of 

larval  walleye  pollock  (Theragra  chalcogramma) 

from  the  northeast  Pacific  Ocean* 


Jennifer  L.  FitzGerald 

Simon  R.  Thorrold 

Biology  Department,  MS  35 

Woods  Hole  Oceanographic  Institution 

Woods  Hole,  Massachusetts  02543 

E-mail  address  (for  J  L.  FitzGerald):  ifitzgerald  awhoi  edu 

Kevin  M.  Bailey 
Annette  L.  Brown 

NOAA  Alaska  Fisheries  Science  Center 
7600  Sand  Point  Way  NE 
Seattle,  Washington  91185 

Kenneth  P.  Severin 

Department  of  Geology  and  Geophysics 

University  of  Alaska  Fairbanks 

P.O.  Box  755780 

Fairbanks,  Alaska  99775-5780 


Manuscript  submitted  4  August  2003 
to  the  Sceintific  Editor's  <  X'fice. 

Manuscript  approved  for  publication 
28  May  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:604-616(20nl 


The  "stock"  concept  is  a  central  tenet 
of  modern  fisheries  science  because  it 
represents  the  fundamental  manage- 
ment unit  of  marine  fisheries  (Begg 
and  Waldman,  1999).  This  emphasis, 
in  turn,  places  a  premium  on  accu- 
rate identification  of  groups  of  fish 
whose  population  statistics  are  largely 
independent  of  other  groups.  However, 
stock  identification  has  often  proved 
problematic  in  marine  fishes.  For 
instance,  the  stock  structure  of  wall- 
eye pollock  {Theragra  chalcogramma* 
across  the  North  Pacific  Ocean  has 
been  a  topic  of  investigation  for  many 
years.  Early  studies  were  based  on 
phenotypic  characteristics  of  pol- 
lock, such  as  meristics  and  morpho- 
metries (Serobaba.  1977;  Hinckley, 
1987;  Temnykh,  1994).  Other  studies 
have  focused  on  genotypic  markers, 
such  as  DNA  and  allozyme  analyses 
(Grant  and  Utter,  1980;  Mulligan  et 
al.,  1992;  Shields  and  Gust,  1995). 
These  approaches  resulted  in  only  the 
broadest  characterization  of  pollock 
stock  structure  but  have  been  able 
to  distinguish  populations  from  the 
eastern  and  western  Pacific  (Bailey  et 
al.,  1999).  Quasi-isolated  subpopula- 


tions  may  be  at  least  demographicallv 
isolated  on  smaller  spatial  scales.  For 
instance,  within  the  Gulf  of  Alaska, 
spawning  pollock  aggregate  at  specific 
locations  in  Shelikof  Strait,  Prince 
William  Sound,  and  in  the  Shumagin 
Islands  region  (Bailey  et  al.,  1999). 
However,  the  extent  of  larval  dis- 
persal from  the  spawning  sites  and 
the  degree  of  spawning  site  fidelity 
of  adult  pollock  to  these  locations 
remains  unknown. 

The  difficulties  associated  with  de- 
termining stock  structure  in  fishes 
are  essentially  the  same  ones  that 
currently  limit  our  ability  to  deter- 
mine population  connectivity  in  ma- 
rine systems  (Thorrold  et  al..  2002). 
Tag-recapture  studies  using  tags  have 
limited  applicability  in  the  case  of 
pollock.  Adults  are  located  deep  in 
the  water  column  and  are  sensitive 
to  barotrauma  during  the  process  of 
being  caught,  brought  to  the  surface, 
and  tagged.  Traditional  population 
genetic  approaches  may  be  similarly 


Contribution  11219  from  the  Woods  Hole 
Oceanographic  Institution.  Woods  Hole, 
MA  02543. 


Fitzgerald  et  al.:  Elemental  signatures  in  otoliths  of  larval  Theragra  chakogramma 


605 


ineffective  because  of  the  low  level  of  exchange  required 
to  maintain  genetic  homogeneity,  at  least  over  ecological 
time  scales,  and  the  low  level  of  genetic  drift  associated 
with  large  populations  (Waples,  1998;  Hellberg  et  al., 
2002).  However,  preliminary  studies  have  indicated  that 
otolith  geochemistry  may  prove  to  be  a  useful  natural 
tag  of  population  structure  in  walleye  pollock  (Severin 
et  al.,  1995).  Otoliths  are  accretionary  crystalline  struc- 
tures located  within  the  inner  ear  of  teleost  fish.  They 
are  formed  through  concentric  additions  of  alternating 
protein  and  aragonite  layers  around  a  central  nucleus. 
The  use  of  otoliths  as  natural  geochemical  tags  is  con- 
tingent on  the  metabolically  inert  nature  of  the  otolith 
and  the  fact  that  once  deposited,  otolith  material  is 
neither  resorbed  nor  metabolically  reworked  (Campa- 
na  and  Neilson,  1985;  Campana,  1999).  The  chemical 
composition  of  otoliths  also  reflects  to  some  degree  the 
physicochemical  characteristics  of  the  ambient  water 
(Bath  et  al.,  2000).  If  the  water  where  pollock  reside 
has  distinct  oceanographic  characteristics,  then  many  of 
the  elements  incorporated  into  the  otoliths  should  differ 
among  locations.  Migrations  between  water  masses  at 
some  age  will,  therefore,  be  recorded  in  the  chemical 
composition  of  the  otolith  at  the  appropriate  daily  incre- 
ment. Natural  geochemical  signatures  in  otoliths  may 
therefore  be  useful  markers  of  environmental  history 
throughout  the  life  of  the  individual  and  in  turn,  fish 
stock  composition  (e.g.,  Campana  et  al.,  1995). 

The  use  of  geochemical  signatures  in  otoliths  as  natu- 
ral tags  requires  accurate  and  precise  assays  of  otolith 
composition.  Electron  probe  micro-analysis  (EPMA)  has 
been  commonly  used  for  probe-based  analyses  of  otolith 
chemistry  (Gunn  et  al.,  1992).  However,  detection  lim- 
its of  approximately  100  ,ug/g  limit  the  technique  to  a 
relatively  small  number  of  minor  elements  in  otoliths, 
including  Na,  CI,  K,  and  Sr  (Campana  et  al.,  1997). 
Most  of  the  elements  measured  by  EPMA  are  probably 
controlled  by  physiological  rather  than  environmen- 
tal factors,  which  may  limit  their  usefulness  in  stock 
identification  studies  (Campana,  1999).  Nonetheless,  a 
number  of  researchers  using  EPMA  have  reported  geo- 
graphic differences  in  otolith  chemistry  (e.g.,  Thresher 
et  al.,  1994).  More  recently,  attention  has  focused  on 
inductively  coupled  plasma  mass  spectrometry  (ICP-MS) 
to  assay  elements  that  are  typically  below  the  detection 
limits  of  EPMA.  Laser  ablation  ICP-MS  uses  focused 
Nd:YAG  or  excimer  lasers  to  ablate  specific  locations  on 
the  otolith.  The  vaporized  material  is  then  swept  up  by 
a  carrier  gas  into  a  plasma  torch  and  analyzed  by  mass 
spectrometry.  Limits  of  detection  of  the  technique  are 
typically  on  the  order  of  0.1-l^g/g,  allowing  for  quan- 
tification of  several  elements  that  cannot  be  assayed  by 
using  EPMA  including  Mg,  Mn,  Ba,  and  Pb  (Thorrold 
et  al.,  1997;  Thorrold  and  Shuttleworth,  2000).  These 
observations  led  Campana  et  al.  (1997)  to  conclude  that 
EPMA  and  laser  ablation  ICP-MS  were  complementary 
and  that  there  is  little  overlap  in  the  elements  that  are 
accurately  measured  by  the  two  techniques.  Yet  few 
studies  of  otolith  geochemistry  have  attempted  to  use 
both  approaches  on  the  same  samples. 


The  objectives  of  this  study  are  to  determine  if  larval 
walleye  pollock  from  different  geographic  localities  can 
be  distinguished  based  on  elemental  signatures  in  their 
otoliths.  By  analyzing  sagittal  otoliths  with  both  EPMA 
and  laser  ablation  ICP-MS,  we  hoped  to  identify  greater 
differences  among  locations  than  would  have  been  pos- 
sible by  using  either  technique  in  isolation.  If  success- 
ful, the  study  may  provide  a  powerful  tool  for  determin- 
ing stock  structure  and  tracing  migration  pathways  of 
walleye  pollock  in  the  north  Pacific.  These  data  could 
then  be  used  by  managers  of  one  of  the  world's  largest 
single  species  fisheries  to  direct  the  sustainable  harvest 
of  this  considerable  natural  resource. 


Materials  and  methods 

All  fish  used  in  the  study  were  collected  in  the  spring  and 
summer  of  1999  from  Alaska  Fisheries  Science  Center 
research  cruises  in  the  Bering  Sea  and  Gulf  of  Alaska 
(Fig.  1,  Table  1).  Fish  of  birth  year  1999  were  collected 
within  three  months  of  spawning  time  to  minimize  the 
likelihood  of  larval  transport  from  other  regions.  In  the 
case  of  the  Yakutat  samples,  fresh  juvenile  pollock  were 
removed  from  Pacific  cod  guts.  Samples  were  collected 
only  when  the  pollock  were  readily  identifiable  and 
heads  were  intact.  Otoliths  showed  no  visible  sign  of 
degradation  from  digestive  processes.  Juvenile  pollock 
were  frozen  whole  and  transported  to  the  laboratory  for 
otolith  removal. 

Otoliths  were  removed  from  the  fish  and  mounted  on 
petrographic  slides  in  LR  White  resin  (acrylic,  hard- 
grade).  Larval  otoliths  were  ground  on  one  side  to 
expose  the  nucleus  by  using  500-grit  paper  and  were 
polished  with  0.25-um  grit  diamond  paste.  Juvenile 
otoliths  were  ground  and  polished  in  the  sagittal  plane 
on  both  sides  to  maximize  clarity  of  the  nucleus  during 
microanalysis. 

Electron  probe  microanalysis 

After  having  been  polished,  the  otoliths  were  cleaned 
with  Formula  409®  and  coated  with  a  30-nm  layer 
of  carbon.  They  were  subsequently  analyzed  with  a 
Cameca  SX-50  electron  microprobe  equipped  with  four 
wavelength  dispersive  spectrometers.  A  15keV,  10  nA, 
4-/jm  diameter  beam  was  used  for  all  analyses.  Counting 
times,  standards,  detection  limits,  and  analytical  errors 
are  summarized  in  Table  2.  Although  Mg  was  analyzed 
in  all  otoliths,  in  most  cases  it  was  below  detection 
limits  and  was  therefore  not  used  in  the  statistical 
analysis. 

Laser  ablation  ICP-MS 

After  having  been  ground  and  polished,  otolith  sections 
were  decontaminated  before  elemental  analysis  by  using 
laser  ablation  ICP-MS.  Sections  were  rinsed  in  ultra- 
pure  water,  scrubbed  with  a  nylon  brush  in  a  solution 
of  ultrapure  H,,0,  triple  rinsed  with  ultrapure  1%HN03, 


606 


Fishery  Bulletin  102(4) 


•;"- 


North 

Bering 

Sea 


1$' 


■"\T7 


60N 


Bristol    15V"^C' 

bay  •     j     *?p* 


Prince 
William 
Sound 


-55N 


SE  Bering 
Sea 


*>•.. 


Shelikof 
Strait 


^  A 


Yakutat 


•**>&* 


160W 

I 


Figure  1 

Locations  of  sampling  sites  for  larval  and  juvenile  walleye  pollock  (Theragra  chalcogramma 
the  Gulf  of  Alaska  and  Bering  Sea. 


)  in 


sonified  for  5  minutes  in  ultrapure  H90,  and  finally  triple 
rinsed  again  with  Milli-Q  water.  The  section  was  dried 
under  a  positive  flow  hood  for  24  hours  and  stored  in  a 
polyethylene  bag. 

Elemental  analyses  were  conducted  with  a  Finnigan 
MAT  Element2  magnetic  sector  field  ICP-MS  and  Mer- 
chantek  EO  LUV266X  laser  ablation  system  (Thorrold 
and  Shuttleworth,  2000).  Instrument  set-up  was  simi- 
lar to  that  outlined  by  Giinther  and  Heinrich  (1999). 
An  Ar  gas  stream  was  used  to  carry  ablated  material 
from  the  laser  cell  to  the  ICP-MS.  The  carrier  gas  was 
then  mixed  with  the  Ar  sample  gas  and  a  wet  aerosol 
(1%  HN03)  in  the  concentric  region  of  the  quartz  dual 
inlet  spray  chamber.  The  wet  aerosol  was  supplied  by 
a  self-aspirating  PFA  micro-flow  (20  /./L/min)  nebulizer 
attached  to  a  CETAC  ASX100  autosampler.  Diameter 
of  the  266-nm  laser  beam  was  nominally  5  j.im,  repeti- 
tion rate  was  5  Hz,  and  the  scanning  rate  was  set  at 
5  /im/sec. 

A  typical  run  for  an  individual  otolith  consisted  of  a 
blank  sample  (l%HNO:!  only),  a  standard  sample,  five 
laser  samples,  and  then  another  blank  and  standard. 
The  number  of  laser  samples  in  a  run  ranged  from  5 
to  15,  depending  upon  the  size  of  the  otolith.  All  laser 
runs  began  with  a  70  fim  x  70  /.im  raster,  centered  on 
the  otolith  nucleus.  The  laser  software  was  then  used 
to  trace  out  concentric  lines,  720  /jm  in  length  and  ap- 
proximately 40  /jm  apart,  which  followed  the  contour 
of  individual  growth  increments  from  the  raster  to  the 
otolith  edge.  This  approach  produced  reasonably  high 
spatial  resolution  (30-50  u,m)  for  life  history  scans 
across  otoliths  while  allowing  sufficient  acquisition  time 
to  maintain  measurement  precision. 


We  examined  Mn/Ca,  Sr/Ca,  and  Ba/Ca  ratios  in  the 
pollock  otoliths  by  monitoring  48Ca,  55Mn,  86Sr,  and 
138Ba.  Quantification  followed  the  approach  outlined 
by  Rosenthal  et  al.  (1999)  for  precise  element/Ca  ra- 
tios using  sector  field  ICP-MS  (Thorrold  et  al.,  2001). 
Quality  control  was  maintained  by  assaying  a  dissolved 
aragonite  standard  (Yoshinaga  et  al.,  2000)  every  five 
samples.  The  standard  was  introduced  at  the  appropri- 
ate time  by  moving  the  autosampler  probe  from  the 
solution  containing  the  1%  HNO;j  to  the  standard  solu- 
tion, while  maintaining  the  carrier  gas  flow  through 
the  ablation  cell.  Elemental  mass  bias  was  calculated 
by  reference  to  known  values  of  the  standard,  and  a 
correction  factor  was  then  interpolated  and  applied  to 
the  laser  samples  bracketed  between  adjacent  standard 
measurements.  Average  u;  =  40)  within-run  precisions 
(RSD)  of  the  standard  measurements  were  all  less  than 
1%  (Mn/Ca:  0.16%,  Sr/Ca:  0.16%,  and  Ba/Ca:  0.33%). 
Long-term  (5-month)  estimates  of  the  standard  mea- 
surements (n=40),  again  uncorrected  for  changes  in 
mass  bias  over  time,  were  less  precise  (Mn/Ca:  5.6%, 
Sr/Ca:  3.7%,  and  Ba/Ca:  5.6%).  However,  laser  samples 
were  corrected  for  changes  in  mass  bias  by  using  the 
laboratory  standard.  Precision  of  the  technique  was  ap- 
proximately 1%  for  all  the  ratios  that  we  measured. 

Statistical  analyses 

All  elemental  data  were  initially  examined  for  nor- 
mality and  homogeneity  of  variance  by  using  residual 
analysis  (Winer,  1971)  and  were  found  to  conform  to 
the  assumptions  of  ANOVA  without  the  need  for  data 
transformation.  We  therefore  assumed  that  require- 


Fitzgerald  et  al.:  Elemental  signatures  in  otoliths  of  larval  Theragra  cha/cogramma 


607 


Table  1 

Location,  collection 

date,  standard  length  range 

(mm),  and  sample 

sizes 

(n)  of  larval 

and  juvenile  walleye  pollock  iTheragra 

chalcogramma) 

capt 

ured  from  the  southeast  Bering  Sea  (SE  Bering) 

North  Bering  Sea 

I N  Ber 

ing),  Bristol  Bay.  Shelikof  Strait. 

Prince  William 

Sound  (PWSi,  and  Yakutat,  and  analyzed  by  laser  ablation  ICP-MS  (ICP-MS)  and  electron  probe  microanalysis 

(EPMA). 

Area 

Date 

SL  range  (mm) 

n  (total) 

in  ICP-MS )                   n  (EPMA) 

SE  Bering 

23  May-27  July  1999 

5.3-42.2 

117 

8                                30 

N  Bering 

18-23  July  1999 

15.6-30.7 

45 

9                                25 

Bristol  Bay 

22-24  July  1999 

85.1-135.7 

75 

11                                28 

Shelikof 

27-28  May  1999 

3.6-7.9 

46 

25 

PWS 

7  July-19  August  1999 

35.2-66.0 

11 

4                                  6 

Yakutat 

15  July,1999 

— 

50 

6                                24 

Table  2 

Counting  times  for  each  element  itime 

seconds],  standards,  limits  of  detection  [LOD. 

7rweight,  99<7f 

confidence  limits]  and  ana- 

lytical  errors 

1  Error.  '/c  weight. 

1  stand 

ird  deviation])  for  electron  probe 

microanalysis  (EPMA).  Detection  limits 

and 

analytical 

errors  were  calculated  by  following  the  procedures  of  Scott  et  al.  (19951. 

N/A  = 

not  applicable. 

Element 

Time 

Standard 

LOD 

Error 

Na 

60 

Halite  ( CM  Taylor) 

0.029 

0.023 

Mg 

60 

OsumilitelUSNM  143967) 

0.019 

0.022 

P 

60 

Apatite  (Wilberforce) 

0.036 

0.027 

S 

60 

Gypsum  (CM  Taylor) 

0.023 

0.017 

CI 

46 

Halite  1  CM  Taylor) 

0.027 

0.015 

K 

46 

OsumilitelUSNM  143967) 

0.019 

0.012 

Ca 

20 

CalciteiNMNH  136321) 

N/A 

0.245 

Sr 

120 

Strontianite  (Smithsonian  R-10065) 

0.036 

0.019 

merits  for  the  MANOVA  were  also  met  by  the  data. 
Among-location  differences  in  the  elemental  composition 
of  larval  pollock  in  specific  regions  of  the  otoliths  were 
compared  by  using  one-factor  multivariate  analysis  of 
variance  (MANOVA)  and  one-factor  analysis  of  variance 
(ANOVA).  We  treated  location  as  a  fixed  factor  in  both 
MANOVA  and  ANOVA  tests.  Because  of  difficulties  col- 
lecting pollock  larvae,  we  were  unable  to  achieve  equal 
replication  of  sites  within  locations.  We  therefore  pooled 
samples  from  collections  within  a  location  by  randomly 
selecting  fish  from  each  location  for  subsequent  analysis. 
However,  the  lack  of  replication  at  the  within-location 
level  necessarily  restricted  our  ability  to  draw  general 
conclusions  concerning  spatial  variability  in  otolith 
composition  beyond  the  samples  analyzed  in  the  pres- 
ent study.  All  a  posteriori  comparisons  among  locations 
were  performed  by  using  Tukey's  honestly  significant 
difference  (HSD)  test  (experimentwise  error  rate  =  0.05). 
Multivariate  differences  in  elemental  signatures  from 
the  MANOVA  were  visualized  by  using  canonical  dis- 
criminant analyses  (CDA).  All  analyses  were  conducting 
by  using  the  SAS  statistical  program  (SAS,  version  6, 
1990,  SAS  Inst.  Inc.,  Cary,  NO. 

Comparisons  of  elemental  profiles  across  otoliths  were 
made  with  repeated  measures  ANOVA.  We  tested  the 


following  null  hypotheses:  1)  there  was  no  variation  in 
trace  element  profiles  across  individual  otoliths  (i.e., 
from  the  nucleus  to  the  edge),  2)  there  were  no  differ- 
ences in  mean  element  concentrations  among  locations, 
determined  by  averaging  data  across  individual  otolith 
profiles,  and  3)  there  were  no  differences  in  the  pat- 
tern of  element  profiles  across  otoliths  among  locations. 
Otolith  profiles  with  missing  values  were  removed,  and 
therefore  we  were  able  to  use  MANOVA  for  the  repeated 
measures  analysis.  The  multivariate  approach  to  re- 
peated measures  is  generally  more  conservative  than 
univariate  repeated  measures  analysis.  However,  the 
multivariate  test  does  not  assume  sphericity  of  orthogo- 
nal components,  requiring  only  that  the  data  conform  to 
multivariate  normality  with  a  common  covariance  ma- 
trix for  individual  larvae  at  each  location  (Littell  et  al., 
1991).  The  approach  still  requires  that  adjacent  points 
on  the  trajectories  be  equidistant.  Therefore  samples 
from  EPMA  were  assigned  to  a  distance  category  at 
intervals  of  15  ftm  (0  //m,  15  fim,  30  /.im  ,  45  /jm,  60  f/m, 
75  /im,  and  90  j/m)  across  the  otolith,  to  a  distance  of 
90  ^m  from  the  nucleus.  Samples  were  averaged  when 
more  than  one  measurement  was  available  within  a 
distance  category.  Laser  ablation  ICP-MS  samples  were 
assigned  to  a  distance  category  at  intervals  of  approxi- 


608 


Fishery  Bulletin  102(4) 


Table  3 

EPMA  results  of  one-factoi 

ANOVA  l  degrees 

of  freedom  |df];  sums  of  squares  [SS] 

mean  square 

[MS])  at  two  positions  (0- 

20  ,um  and  20-45  pm  from  the  nuc 

leus)  in  otoliths  of  larval 

walleye  pollock  (Theragra  c 

hal 

■ogramma  i 

collected  from  six  locations: 

three  locations  in  the  Berin 

I  Sea 

southeast  B 

;ring  Sea  [SB];  North  Bering  Sea  |NB] 

Br 

stol  Bay  [BB]  I  and  three  in  the  Gulf  of 

Alaska  (Prince  William  Sound  [PW];  Shelikof  Strait  [SH]- 

and  Yakutat  |YK1 

***=  sign 

ficant  at  a 

=  0.05;  ns  =  nonsignificant. 

A  posteriori 

multiple  comparisons 

among  locat 

ons  were  conducted  by  using  Tukey's  hone 

3tly  significant  difference  ( HSD  I.  Loca- 

tions  are  ordered  I  left  to  right  I  from  lowest  to  highest  concentrations,  and  lines  link  locations  that  are  not  significantly  different 

(experimentwise  error  rate  = 

=  0.05) 

Element 

Source 

df 

SS 

.    MS 

F 

P<F 

Tukey's  HSD 

0-20 nm 

Na 

Location 
Error 

5 
113 

24.7 
34.1 

4.9 
3.0x10-1 

16.37 

PW  BB  YK  NB  SB  SH 

P 

Location 
Error 

5 
113 

18.3 
27.7 

3.7 
2.5xl0-i 

14.93 

BB   SH  NBSB  PWYK 

S 

Location 
Error 

5 
113 

2.0 
8.8 

3.9x10-1 
7.8xl0-2 

5.04 

**v 

PW  BB  NB  YK  SH  SB 

CI 

Location 

5 

10.2 

2.0 

3.93 

*** 

PW  NB  BB  YK  SB  SH 

Error 

113 

58.6 

5.2xl0-i 

K 

Location 
Error 

5 
113 

2.4 
13.51 

4.8x10  i 
1.2x10-1 

4.01 

*=*-:- 

SH  NB  SB  PW  BB  YK 

Sr 

Location 
Error 

5 
113 

7.1 
24.1 

1.4 
2.1x10-1 

6.66 

*** 

YK  NB  BB  SH  PWSB 

20-45  pm 

Na 

Location 
Error 

4 
93 

15.9 
13.9 

4.0 
1.5xl0-i 

26.66 

*## 

BB   PWYK  SB   NB 

P 

Location 
Error 

4 
93 

14.9 
20.3 

3.7 
2.2x10-1 

17.07 

*** 

BB   NB  SB  PW  YK 

S 

Location 
Error 

4 
93 

1.1 
5.7 

2.8xl0-i 
6.1xl0-2 

4.62 

BB   PW  SB  NB  YK 

CI 

Location 
Error 

4 
93 

8.4x10- 
12.1 

1         2.1x10' 
1.3x10-1 

1.61 

ns 

PW  SB   NB  BB  YK 

K 

Location 
Error 

4 
93 

1.9 

7.7 

4.7x10-' 
8.3x10-2 

5.65 

*** 

NB  PW  BB  SB  YK 

Sr 

Location 
Error 

4 
93 

9.0x10- 
5.6 

1         2.2x10-1 
6.0x10-2 

3.75 

*** 

NB  YK  BB  PW  SB 

mately  40  |im  (nucleus,  40-80  Jim,  80-120  ^m,  120-160 
urn,  and  160-200  fim)  across  the  otolith,  to  a  distance 
of  200  ^m  from  the  nucleus. 


Results 

Electron  probe  microanalysis 

A  total  of  six  elements  (Na,  P,  S,  CI,  K,  and  Sr)  were 
quantified  in  the  otoliths  of  pollock  larvae  by  using 
EPMA.  Average  concentrations  of  the  elements  ranged 
from  approximately  4  mg/g  (otolith  weight)  for  Na  to 
less  than  1  mg/g  for  CI,  S,  and  K.  We  found  significant 
differences  in  the  elemental  composition  both  among 
sampling  locations  and  across  positions  on  the  oto- 
liths. Multivariate  analyses  of  elemental  signatures 
revealed  significant  differences  among  locations  from 
samples  0-20  urn  from  the  nucleus  (M ANOVA;  Pillai's 
trace=1.18;  F30 ,560=5.74;  P<0.0001),  and  20-45  fim  from 


the  nucleus  (MANOVA;  Pillai's  trace  =  1.47;  F., 


,  =  S,Sli; 


P<0.0001l. 

Analysis  of  variance  and  Tukey's  HSD  a  posteriori 
multiple  comparison  tests  identified  the  individual  ele- 
ments contributing  to  differences  in  the  multivariate 
signatures  among  locations.  All  six  elements  showed 
significant  differences  among  locations  at  0-20  /tm  from 
the  otolith  nucleus  (Table  3).  Multiple  comparisons  for 
each  of  the  elements  indicated  relatively  subtle  differ- 
ences among  locations.  Phosphorus  concentrations  were, 
however,  significantly  higher  in  samples  from  Yakutat 
than  any  of  the  other  locations.  Results  of  the  ANOVA 
from  samples  at  distances  20-45  j<m  from  the  nucleus 
were  generally  comparable  with  samples  closer  to  the 
nucleus  (Table  3).  Although  only  CI  showed  no  signifi- 
cant variation  among  locations  (Table  3),  multiple  com- 
parisons of  mean  values  for  each  element  revealed  little 
geographic  patterns  among  locations.  Note  that  Shelikof 
Strait  samples  were  removed  from  this  analysis  because 
of  a  small  sample  size  in  this  distance  category. 


Fitzgerald  et  al.:  Elemental  signatures  in  otoliths  of  larval  Theragra  chakogramma 


609 


o 


.A 

ft 

a 
ft 

o 

o 
ft 

ft 

ft         +           o 

o 

ft 

\*t%K* 

o 

4-2024 
D  ft  ft      o 

ft        M° 

ft 
ft 

*  o 

<     *     a 

°  * 

ft  #A*°  ^o 

ft  o° 

>    *>■*      *°     *o      *      * 

_•'  "    O      .  "         A         4     .»-* 

II       '     t  o    °    *4     »        • 


•4-20246 

Canonical  variate  1 

Figure  2 

Plot  of  first  two  canonical  variates  con- 
trasting multivariate  elemental  sig- 
natures in  otoliths  of  walleye  pollock 
(Theragra  chalcogramma)  determined 
by  using  electron  probe  microanaly- 
sis, at  0-20  (im  from  the  nucleus  (A) 
and  20-40  Jim  from  the  nucleus  (B). 
Larvae  were  collected  from  the  North 
Bering  Sea  (▲),  southeast  Bering  Sea 
(O),  Bristol  Bay  IB),  Shelikof  Strait 
!♦),  Prince  Wiiliam  Sound  lO>.  and 
Yakutat  (ft). 


■  »\ 


J8       o    - 

°    .  ° 

0OA 

*   *   i 

*tAcf 


H 


-  ■  A.         ° 

.  ^      A      V 


-3-113 

Canonical  variate  1 


Figure  3 

Plot  of  first  two  canonical  variates 
contrasting  multivariate  elemental 
signatures  in  otoliths  of  walleye  pol- 
lock (Theragra  chalcogramma)  deter- 
mined by  electron  probe  microanalysis 
at  0-20  /im  from  the  nucleus  iAi  and 
20-40  jim  from  the  nucleus  IB).  Larvae 
were  collected  from  the  North  Bering 
Sea  (A),  southeast  Bering  Sea  (O).  and 
Bristol  Bay    ■ 


We  used  CDA  to  visualize  multivariate  differences 
among  locations  in  reduced  dimensional  space.  Three 
groups  were  readily  discernible  in  a  plot  of  the  first  two 
canonical  variates  (Fig.  2).  Samples  from  the  North  Ber- 
ing Sea,  the  southeast  Bering  Sea,  and  Shelikof  Strait 
formed  one  group  separated  from  Yakutat,  Bristol  Bay, 
and  Prince  William  Sound  samples  along  the  first  ca- 
nonical variate.  The  second  canonical  variate  separated 
Yakutat  samples  from  Bristol  Bay  and  Prince  William 
Sound  individuals.  Elemental  signatures  at  20-45  jum 
from  the  otolith  nucleus  were  distributed  similarly  in 
canonical  space  to  samples  from  the  otolith  nucleus 
(Fig.  2).  Three  groupings  were  apparent  in  the  canoni- 
cal plot,  and  Bering  Sea  larvae  were  separated  from 
Bristol  Bay  and  Prince  William  samples  on  canonical 
variate  one,  and  Yakutat  samples  were  separated  from 
all  other  locations  on  canonical  variate  two.  We  then 


conducted  a  similar  analysis  with  only  samples  from 
the  southeast  and  North  Bering  Sea  and  Bristol  Bay. 
Elemental  signatures  of  larvae  from  the  Bering  Sea 
separated  from  Bristol  Bay  on  canonical  variate  one. 
The  southeast  Bering  Sea  samples  separated  from  the 
North  Bering  Sea  along  canonical  variate  two,  although 
not  as  clearly  as  with  the  elemental  signatures  from  the 
Bering  Sea  and  Bristol  Bay  (Fig.  3). 

Elemental  profiles  across  otoliths  varied  significantly, 
as  determined  by  repeated  measures  ANOVA,  among 
the  five  locations  for  Na,  P,  S,  and  Sr  (Table  4).  Both 
S  and  Sr  concentrations  declined  from  high  values  at 
the  nucleus  to  significantly  lower  values  towards  the 
edge  of  the  otolith  (Fig.  4).  Repeated  measures  ANOVA 
also  provided  a  test  of  the  differences  among  locations 
when  data  were  averaged  over  the  otolith  profiles.  Sig- 
nificant differences  among  locations  were  detected  for 


610 


Fishery  Bulletin  102(4) 


a. 


5.5 
5 

45 
4 

3.5 
3 

25 

2 

1.5 

1 

05 

0 


1.5 


I       5       $ 

I       I       J 


E. 

u 


i   i 


llfl 


f   i 


E_ 


to 


2.5 


E      15 


0      15    30    45    60    75    90 


0      15    30    45    60    75    90 


Distance  from  nucleus  (mm) 

Figure  4 

Profiles  of  elemental  concentrations,  determined  by  electron  probe 
microanalysis,  from  the  nucleus  out  to  a  distance  of  approximately 
90  Jim  in  the  otoliths  of  larval  walleye  pollock  iTheragra  chal- 
cogramma)  collected  from  the  North  Bering  Sea  (A),  southeast 
Bering  Sea  (O),  Bristol  Bay  (■),  Prince  William  Sound  (O),  and 
Yakutat  ("I.  Individual  points  are  mean  (  +  SE)  values  grouped 
at  15-f/m  intervals. 


five  elements  (Sr,  K,  S,  P,  and  Na).  Finally,  the  interac- 
tion term  (positionxlocation)  in  the  repeated  measures 
ANOVA  tested  the  hypothesis  that  the  shape  of  the 
elemental  profiles  differed  among  locations.  There  was 
a  significant  interaction  between  profile  and  location 
for  K. 

Laser  ablation  ICP-MS 

We  quantified  Mn/Ca,  Sr/Ca,  and  Ba/Ca  ratios  in  the 
otoliths  of  larval  walleye  pollock  using  laser  ablation 
ICP-MS.  Both  Mn  and  Ba  were  found  at  trace  levels  in 
otoliths,  with  average  values  of  approximately  3  jimo}/ 
mol  and  6  umol/mol,  respectively.  Strontium  was  present 
in  the  otoliths  at  an  average  concentration  of  approxi- 
mately 2.2  mmol/mol.  A  MANOVA  detected  significant 
differences  among  locations  from  a  raster  centered  on 
the  nucleus  (MANOVA;  Pillai's  trace  =  0.85;  Flz99=3.26; 
P<0.0005),  and  from  the  average  values  of  lines  40-80 
pm  from  the  nucleus  (MANOVA;  Pillai's  trace  =  0.99; 
F1299=4.1;P<0.0001). 

Univariate  ANOVA  and  a  posteriori  multiple  compari- 
sons by  using  Tukey's  HSD  revealed  that  Mn/Ca,  Sr/Ca, 


5    - 

B 

5    - 

0 

o 

0    - 

-"*dy  ■ 

■ 
■  ■ 

5    - 

5    - 

■ 

-4-2024 

deamahcei  eenhe(Jg 

Figure  5 

Plot  of  the  first  two  canonical 
variates  contrasting  multivari- 
ate elemental  signatures  in  oto- 
liths of  walleye  pollock  iTheragra 
chalcogramma)  determined  with 
laser  ablation  ICP-MS,  at  0-40  fan 
from  the  nucleus  (A)  and  40-80 
fim  from  the  nucleus  (Bi.  Larvae 
and  juveniles  were  collected  from 
the  North  Bering  Sea  (At,  south- 
east Bering  Sea  (O).  Bristol  Bay 
■  Prince  William  Sound  (♦),  and 
Yakutat  (     I. 


and  Ba/Ca  ratios  varied  significantly  among  locations 
at  the  otolith  nucleus  and  at  positions  40-80  um  from 
the  nucleus  (Table  5).  Samples  from  the  North  Bering 
Sea  had  consistently  lower  Sr/Ca  and  Ba/Ca  ratios  than 
those  from  the  southeast  Bering  Sea  at  both  positions. 
However,  we  noted  only  subtle  differences  among  the 
Gulf  of  Alaska  and  Bristol  Bay  samples. 

We  found  a  total  of  three  groupings  in  canonical  plots 
of  multivariate  elemental  signatures  from  the  otoliths  of 
larval  walleye  pollock  (Fig.  5).  Samples  from  the  North 
Bering  Sea  and  Bristol  Bay  were  separated  along  ca- 
nonical variate  one.  A  third  grouping,  including  larvae 
from  the  southeast  Bering  Sea,  Prince  William  Sound, 
and  Yakutat,  clustered  together  in  the  center  of  the 
canonical  plot.  Samples  from  the  nucleus  and  40-80 
fim  outside  the  nucleus  showed  very  similar  geographic 
patterns. 

Repeated  measures  ANOVA  detected  significant  dif- 
ferences in  both  Mn/Ca  and  Ba/Ca  profiles  from  the  nu- 


Fitzgerald  et  al.:  Elemental  signatures  in  otoliths  of  larval  Theragra  chalcogramma 


611 


Table  4 

EPMA  results  from  repeated-measures  ANOVA  of  elements 

profiles  across  otoliths 

from  walleye  pollock 

Theragrc 

chalcogramma) 

larvae  collected  at  five  locations 

in  the  Ber 

ing  Sea  and  the  Gulf  of  Alaska. 

With 

in-subject  effects 

(pr 

ofile  and 

profilexlocationl 

tested  bv 

using  MANOVA  (Pillai's  trace  I,  and  between  sub 

ects  effect  (location I  te 

sted  by  using  ANOVA  I  degrees  of  freedom  [df]: 

sums  of  s 

quares  [SS];  mean  squares  [MS]) 

**  =  significant  at  a  =  0.05;  ns 

=  nonsignificant. 

Element 

Source 

df 

Pillai's  trace  or  SS 

MS 

F 

P<F 

Na 

Profile 
Profile*  location 

6,41 
24,  176 

4.9x10"' 
5.3X10-1 

6.67 
1.12 

ns 

Location 

4 

43.8 

10.9 

10.35 

*** 

Error 

46 

48.7 

1.1 

P 

Profile 

Profile  <  location 

6,40 
24, 172 

3.0x10-' 
4.6x10-' 

2.91 
0.93 

*** 
ns 

Location 

4 

55.0 

13.8 

8.07 

*** 

Error 

45 

76.8 

1.7 

S 

Profile 

Profile  y  location 

6,41 
24,  176 

3.6x10"' 
5.3x10"' 

3.92 
1.12 

*** 
ns 

Location 

4 

3.8 

9.6x10-1 

4.56 

*** 

Error 

46 

9.7 

2.1x10"' 

CI 

Profile 

Profile  x  location 

6,39 
24,  168 

1.8x10-' 
3.7x10-1 

1.41 
0.72 

ns 
ns 

Location 

4 

8.0 

2.0 

1.6 

ns 

Error 

44 

55.3 

1.3 

K 

Profile 
Profile*  location 

6,41 
24, 176 

8.3xl0-2 
8.5xl0-i 

0.62 
1.98 

ns 

Location 

4 

4.3 

1.1 

2.61 

*#* 

Error 

46 

19.1 

4.1x10-' 

Sr 

Profile 

Profile  x  location 

6,40 

24, 172 

4.5x10-' 
7.7x10-' 

5.51 
1.72 

ns 

Location 

4 

6.4 

1.6 

3.96 

*** 

Error 

45 

18.2 

4.0x10-' 

Table  5 

Laser  ablation  ICP-MS  results  of  one-factor  ANOVA  (degrees  of  freedom  [df];  sums  of  squares  [SS];  mean  square  [MS])  at  two 
positions  (0-40  /im  and  40-80  /im  from  the  nucleus)  in  walleye  pollock  (Theragra  chalcogramma)  otoliths  collected  at  five  loca- 
tions: the  southeast  Bering  Sea  [SB],  North  Bering  Sea  (NB),  Bristol  Bay  (BB),  Prince  William  Sound  (PW),  and  Yakutat  (YK). 
**=  significant  at  a  =  0.05;  ns=nonsignificant.  Multiple  comparisons  among  locations  were  conducted  by  using  Tukey's  honestly 
significant  difference  (HSD).  Locations  were  ordered  (left  to  right  I  from  lowest  to  highest  ratios;  lines  link  locations  that  were 
not  significantly  different  (a=0.05). 


Element 


Source 


df 


SS 


MS 


P<F 


Tukev's  HSD 


0-40  fim  (nucleus) 

Mn/Ca 

Locations 

4 

19.3 

4.82 

3.5 

Error 

33 

45.5 

1.38 

Sr/Ca 

Locations 

4 

1.07 

0.27 

3.43 

Error 

33 

2.58 

0.08 

Ba/Ca 

Locations 

4 

219 

54.8 

3.35 

Error 

33 

540 

16.4 

40-80  fun 

Mn/Ca 

Locations 

4 

30.7 

7.68 

3.24 

Error 

33 

78.3 

2.37 

Sr/Ca 

Locations 

4 

0.99 

0.25 

3.70 

Error 

33 

2.21 

0.07 

Ba/Ca 

Locations 

4 

397 

99.2 

5.50 

Error 

33 

595 

18.0 

PW 

NB 

SB 

YK  BB 

NB 

YK 

BB 

PW  SB 

NB 

BB 

PW 

YK  SB 

NB 

PW 

SB 

YK  BB 

NB 

YK 

PW 

BB   SB 

NB 

BB 

PW 

YK  SB 

612 


Fishery  Bulletin  102(4) 


■Si        3 


0 
3.2 


"5      2.8 

o 

■o 


2.4 


1.6  J 
30 


20 


10 


O 


0  60        100       140       180 

aartdneg  Yrre  nuelgur  Yeea 

Figure  6 

Profiles  of  elemental  ratios,  deter- 
mined with  laser  ablation  ICP-MS, 
from  the  nucleus  out  to  a  distance 
of  approximately  200  /.mi  in  the  oto- 
liths of  larval  and  juvenile  walleye 
pollock  iTheragra  ehaleogramma) 
collected  from  the  North  Bering 
Sea  (A),  southeast  Bering  Sea 
(O),  Bristol  Bay  (■),  Prince  Wil- 
liam Sound  lOl,  and  Yakutat  I  ■■  >. 
Individual  points  are  mean  (±SE) 
values  grouped  at  40-jAm  intervals. 


cleus  out  to  a  distance  of  approximately  200  jjm  in  the 
walleye  pollock  otoliths  (Fig.  6,  Table  6).  The  univariate 
test  of  location,  averaged  over  the  individual  otolith 
profiles,  was  significant  for  both  Sr/Ca  and  Ba/Ca.  We 
also  found  significant  interactions  between  profile  and 
location  for  Mn/Ca,  Sr/Ca,  and  Ba/Ca  ratios  (Table  6). 
Manganese  values  increased  from  the  nucleus  to  the 
otolith  edge  at  all  locations,  indicating  that  the  signifi- 
cant interaction  was  generated  by  the  observation  that 
the  profile  from  the  North  Bering  Sea  was  considerably 
flatter  than  profiles  from  Bristol  Bay  and  southeast 
Bering  Sea.  Strontium  trajectories  were  more  dynamic; 
profiles  from  some  locations  increased  from  the  nucleus 
to  the  edge  (Bristol  Bay  and  Prince  Williams  Sound), 


4.5 


3.5- 


'     I 


li'f'Mi1 


*    ft    ft 


ft    ft    ft    ft    ft    ft    ft 


z 

CO 

O 


CO 

O 
O 


*     ft 


HHf 


. 


.  I 


ft! 


*;**** 


tp 


0  6  10  14  18  22  26  30  34  38  42  46  50  54  58 

Distance  from  nucleus  (mm  x  10'2) 

Figure  7 

Profiles  of  elemental  ratios,  determined  by  using 
laser  ablation  ICP-MS.  from  the  nucleus  out  to  a 
distance  of  approximately  600  fim  in  the  otoliths  of 
juvenile  walleye  pollock  (Thuragra  ehaleogramma) 
collected  from  Bristol  Bay  ■  and  Yakutat  i  I. 
Individual  points  are  mean  (±  SE)  values  grouped 
at  40-;/m  intervals. 


profiles  from  other  locations  decreased  (North  Bering 
Sea  and  Yukatat),  and  a  single  location  (southeast  Ber- 
ing Sea)  showed  no  obvious  trend.  Finally,  profile  varia- 
tions in  Ba/Ca  ratios  among  locations  were  dominated 
by  a  sharp  increase  in  Ba/Ca  ratios  across  the  otoliths 
in  the  southeast  Bering  Sea  samples.  Profiles  were  ef- 
fectively horizontal  for  the  other  four  locations. 

Otoliths  in  walleye  pollock  collected  from  Bristol  Bay 
and  Yakutat  were  significantly  larger  than  those  from 
the  other  four  locations.  We  were,  therefore,  able  to  con- 
duct extended  profiles  in  these  otoliths  out  to  a  distance 
of  approximately  600  fim  (Fig.  7).  After  starting  at 
similar  values  at  the  nucleus,  Mn/Ca  and  Sr/Ca  profiles 
from  the  two  locations  quickly  diverged  and  appeared 
to  vary  largely  independently  over  the  remaining  time 
periods.  The  Ba/Ca  profiles  also  appeared  to  be  vary- 
ing independently  between  the  two  locations,  although 
the  relative  magnitude  of  differences  between  the  two 
locations  was  smaller  than  for  either  Mn/Ca  or  Sr/Ca 
profiles. 


Fitzgerald  et  al.:  Elemental  signatures  in  otoliths  of  larval  Theragra  chalcogramma 


613 


Table  6 

Laser  ablat 

on  ICP-MS  results  from  repeated-measures  ANOVA  of  elementa 

profiles 

across  otoliths 

from  walleye  pollock  l  Ther- 

agra  chalcogramma  l  larvae  collected  at  fi 

ire  locations 

in  the  Bering  Sea  and  the  Gulf  of  Alaska. 

Within-subject  effects  I 

profile  and 

profile x location)  were  tested  by  using  MANOVA  iPillai's  trace),  and  between  subject 

s  effect 

location)  tested  by  using  ANOVA 

(degrees  of  freedom  [df];  sums  of  squares 

[SS];  mean 

squares  [MS]).  ***  =  significant  at  a  =  0.05;  ns 

=  nonsignificant 

Element 

Source 

df 

Pillai's  trace  or 

SS 

MS 

F 

P<F 

Mn/Ca 

Profile 

4,30 

3.0  x  10"1 

17.1 

*** 

Position  x  location 

16,  132 

7.4  x  10"1 

1.9 

*** 

Location 

4 

1.99  x  102 

49.7 

2.49 

ns 

Error 

33 

6.61  x  10"2 

20.1 

Sr/Ca 

Profile 

4.30 

2.18  x  lO"1 

2.09 

ns 

Position  x  location 

16,  132 

9.19  x  10"1 

2.46 

*** 

Location 

4 

1.05  x  101 

2.62 

6.02 

*** 

Error 

33 

1.44  x  101 

0.44 

Ba/Ca 

Profile 

4,30 

6.92  x  10"1 

16.83 

*** 

Position  x  location 

16, 132 

1.04 

2.9 

*** 

Location 

4 

5.83  x  103 

1459 

15.25 

*** 

Error 

33 

3.16  x  103 

95.6 

Discussion 

Quantifying  dispersal  pathways  of  larval  fishes  in 
marine  environments  is  a  difficult  proposition.  Marine 
fishes  typically  produce  on  the  order  of  lCH-lO6  eggs  in 
a  single  spawning  episode.  These  propagules  are  quickly 
dispersed  in  large  volumes  of  seawater,  making  recovery 
of  marked  individuals  difficult  even  if  it  were  possible  to 
introduce  an  artificial  tag  into  the  larvae  at  the  time  of 
spawning  I  Jones  et  al.,  1999).  Natural  geochemical  sig- 
natures in  otoliths  offer  a  useful  alternative  to  artificial 
tagging  approaches  (Thorrold  et  al..  2002).  The  technique 
relies  upon  the  assumption  that  larvae  spawned  at  any 
given  location  retain  a  unique  elemental  or  isotopic  sig- 
nature in  their  otoliths  that  can  be  recovered  some  time 
afterwards,  and  that  variations  in  otolith  geochemistry 
are  sufficient  to  distinguish  among  geographic  locations 
of  interest.  We  found  that  elemental  signatures  in  the 
otoliths  of  larval  walleye  pollock  differed  significantly 
geographically  and  with  ontogeny.  Samples  at  specific 
points  on  the  otoliths,  at  the  nucleus,  and  shortly  after 
hatching,  showed  very  similar  patterns  of  variability, 
suggesting  that  the  technique  will  likely  be  a  robust 
method  for  identifying  natal  origins  of  walleye  pollock 
after  suitable  groundtruthing  of  known  spawning  loca- 
tions (e.g.,  Thorrold  et  al.,  2001;  Gillanders,  2002). 

Elemental  signatures  in  the  otoliths  of  the  larval 
pollock  were  assayed  by  using  EPMA  and  laser  abla- 
tion ICP-MS.  Campana  et  al.  (1997)  noted  that  the  two 
techniques  were  largely  complementary  in  terms  of  ele- 
ments that  could  be  reliably  assayed,  and  indeed  only 
Sr  was  able  to  be  quantified  by  both  instruments  in  our 
study.  Patterns  of  geographic  variability  presumably 
reflected  the  elements  that  were  used  in  generating  the 
multivariate  signatures  produced  by  each  instrument. 
Elemental  signatures  from  the  EPMA  data  clustered 


into  a  Bering  Sea  group  that  included  Shelikof  Strait 
but  excluded  Bristol  Bay,  a  coastal  grouping  that  in- 
cluded Bristol  Bay  and  Prince  William  Sound,  whereas 
samples  from  Yakutat  were  separated  from  all  other 
locations.  Although  sample  sizes  were  smaller  for  the  la- 
ser ablation  ICP-MS  assays,  the  data  identified  a  group- 
ing of  locations  in  multivariate  space  that  included  the 
southeast  Bering  Sea,  Prince  William  Sound,  and  Ya- 
kutat, whereas  samples  from  both  the  north  Bering  Sea 
and  Bristol  Bay  were  separated  from  each  other  and  the 
other  locations.  The  observation  that  locations  did  not 
cluster  into  similar  geographic  groupings  was  probably 
a  function  of  different  mechanisms  of  elemental  incor- 
poration in  otoliths.  Elements  assayed  by  ICP-MS  in  the 
present  study  substitute  for  Ca  in  the  aragonitic  lattice, 
and  are  believed  to  primarily  reflect  ambient  physico- 
chemical  differences  among  natal  locations  (Bath  et  al., 
2000;  Milton  and  Chenery,  2001;  Bath  Martin  et  al, 
2003).  However,  with  the  exception  of  Sr,  the  elements 
assayed  by  EPMA  are  likely  under  physiological  regula- 
tion and  therefore  probably  do  not  directly  reflect  either 
water  chemistry  or  temperature  (Campana,  1999).  In 
either  case,  the  application  of  elemental  signatures  in 
otoliths  as  natural  tags  of  natal  origins  requires  only 
that  the  signatures  allow  accurate  classification  of  the 
natal  origins  of  an  unknown  fish.  A  final  caveat  is  nec- 
essary because  it  remains  possible  that  preservation 
effects,  particularly  for  labile  elements  that  are  not 
incorporated  into  the  aragonite  lattice,  may  also  have 
contributed  to  at  least  some  of  the  differences  among 
locations  (Milton  and  Chenery,  1998;  Proctor  et  al., 
1998).  If  present,  such  effects  would  clearly  confound 
attempts  to  use  elemental  signatures  as  a  natural  tag 
of  natal  origins  (Thresher,  1999). 

It  is  important  to  note  that  although  EPMA  and  laser 
ablation  ICP-MS  provided  complementary  information 


614 


Fishery  Bulletin  102(4) 


on  elemental  composition,  the  spatial  scale  on  which 
the  data  were  gathered  was  different.  Our  laser  ablation 
ICP-MS  method  required  that  we  ablate  a  70  /im  x  70 
/jm  raster,  or  a  720-/im  line,  in  order  to  enable  sufficient 
time  to  generate  precise  estimates  of  otolith  composition. 
The  EPMA  analysis  was  less  destructive  than  laser  ab- 
lation ICP-MS,  and  therefore  it  was  possible  to  sample 
individual  points  at  a  much  finer  spatial  resolution 
(~5  fim),  albeit  with  considerably  less  sensitivity  and  pre- 
cision. For  instance,  using  EPMA  we  were  able  to  sam- 
ple five  points  across  a  transect  ending  approximately 
90  urn  from  the  nucleus.  In  contrast,  only  a  single  ras- 
ter could  be  sampled  along  this  profile  with  laser  abla- 
tion ICP-MS.  Although  the  diameter  of  laser  probes 
is  approaching  that  of  EPMA,  ICP-MS  is  unlikely  to 
match  the  spatial  resolution  of  EPMA  without  further 
development  of  truly  simultaneous  mass  analyzers  such 
as  time-of-flight  ICP  mass  spectrometry  (Mahoney  et 
al.,  1996).  However,  we  were  able  to  program  the  laser 
probe  to  trace  out  growth  increments  once  the  otolith 
radius  had  reached  120  fim  and  we  found  that  the  total 
length  of  a  daily  ring  was  approximately  700  /jm.  This 
finding,  in  turn,  allowed  us  to  construct  elemental  pro- 
files at  reasonable  spatial  resolution  across  the  otoliths 
of  larval  pollock  without  sacrificing  instrument  preci- 
sion by  limiting  acquisition  times.  Although  it  has  not 
been  used  before  with  otoliths,  our  approach  provides 
significant  advantages  over  previous  methods  of  using 
a  raster  to  create  elemental  profiles  (e.g.,  Thorrold  et 
al.,  1997;  Thorrold  and  Shuttleworth  2000). 

Previous  work  on  pollock  otolith  chemistry  was  some- 
what successful  at  distinguishing  fish  from  locations 
in  the  Bering  Sea  and  the  Gulf  of  Alaska.  Severin  et 
al.  (1995)  used  EPMA  to  sample  the  outer  margin  of 
otoliths  from  juvenile  pollock  collected  along  the  Alaska 
Peninsula  in  the  Gulf  of  Alaska  and  in  Bristol  Bay.  We 
generated  elemental  profiles  across  otoliths  from  the 
nucleus  out  to  approximately  90  /im  for  the  EPMA  sam- 
ples, and  up  to  600  /mi  for  the  laser  ablation  ICP-MS  as- 
says. The  profiles  revealed  some  interesting  differences 
between  the  elements  assayed  by  each  instrument.  For 
instance,  only  one  of  the  elements  (K)  from  the  EPMA 
analysis  showed  a  significant  interaction  between  pro- 
file and  location,  yet  significant  profile  x  location  interac- 
tions were  detected  for  Mn/Ca,  Sr/Ca,  and  Ba/Ca  ratios 
with  laser  ablation  ICP-MS.  We  were  also  struck  by 
the  similarity  of  profiles  from  individuals  sampled  at 
the  same  location,  as  evidenced  by  the  size  of  standard 
errors  around  mean  values  at  specific  distances  across 
the  otolith.  For  instance,  the  extended  profiles  from  pol- 
lock collected  in  Bristol  Bay  and  Yakutat  show  indepen- 
dent patterns  of  variation  for  all  three  elements  from 
the  nucleus  out  to  600  /jm.  Taken  together,  these  data 
indicate  that  larvae  from  several  spawning  locations 
are  indeed  encountering  water  masses  with  differing 
physicochemical  properties  through  their  larval  lives, 
and  at  approximately  the  same  time.  We  lack,  however, 
a  sufficient  understanding  of  the  mechanisms  control- 
ling otolith  chemistry  to  be  able  to  relate  the  profiles 
to  specific  properties  of  different  water  masses  in  the 


study  area.  This  knowledge  will  be  necessary  before  it 
is  possible  to  reconstruct  dispersal  pathways  of  larval 
pollock  based  on  probe-based  analyses  of  otolith  geo- 
chemistry. Nonetheless,  the  among-location  variability 
in  elemental  profiles  revealed  by  both  instruments  is 
encouraging  and  justifies  further  investigations  of  oto- 
lith geochemistry  in  larval  pollock. 

Past  attempts  at  identifying  stock  structure  of  wall- 
eye pollock  in  the  North  Pacific  Ocean  based  on  genetic 
techniques  have  been  inconclusive  (Bailey  et  al.,  1999). 
In  the  most  recent  study,  Olsen  et  al.  (2002)  were  un- 
able to  distinguish  between  pollock  from  the  Kamchat- 
ka Peninsula  and  several  locations  within  the  Gulf  of 
Alaska  based  on  three  polymorphic  microsatellite  loci. 
Allozyme  and  MtDNA  markers  showed  significant  differ- 
ences between  North  American  and  Asian  populations, 
and  among  Gulf  of  Alaska  locations.  These  data  were 
difficult  to  reconcile  because  both  markers  showed  tem- 
poral instability  within  locations.  Adult  tagging  studies 
shed  little  light  on  the  population  structure  of  pollock 
because  they  address  questions  of  repeat  spawning, 
whereby  adult  fish  return  to  the  same  area  to  spawn  in 
subsequent  years,  rather  than  homing  to  natal  spawn- 
ing locations  (Tsugi,  1989).  It  has  proved  impossible, 
except  in  rare  circumstances  (Jones  et  al.,  1999),  to 
artificially  mark  larvae  before  they  are  dispersed  from 
spawning  grounds,  and  therefore  natural  geochemical 
tags  remain  the  most  promising  avenue  for  determining 
natal  origins  in  walleye  pollock.  The  ability  to  determine 
natal  origins  of  individual  fish  is  critical  in  the  case  of 
migratory  marine  fishes  because  it  allows  quantification 
of  population  connectivity  through  straying  of  adults  as 
well  as  through  larval  dispersal  (Thorrold  et  al..  2001). 
These  data,  in  turn,  identify  the  spatial  extent  of  fish 
stocks  that  are  demographically  isolated  or  alternatively 
provide  connectivity  rates  that  are  necessary  to  param- 
eterize spatially  explicit  models  if  the  species  is  usefully 
viewed  as  a  metapopulation  (Hanski  and  Gilpin,  1997; 
Smedbol  and  Wroblewski,  2002). 

In  summary,  the  elemental  composition  of  otolith 
material  deposited  during  early  larval  life  in  walleye 
pollock  differed  significantly  among  locations  in  the 
Gulf  of  Alaska  and  Bering  Sea.  These  results  imply  that 
the  larvae  originated  from  different  spawning  locations, 
not  that  they  constitute  separate  stocks.  Nonetheless, 
these  data  represent  the  necessary  first  steps  in  using 
elemental  signatures  in  otoliths  as  natural  tags  of  natal 
origins  in  walleye  pollock.  Elemental  profiles  across 
otoliths  were  also  unique  to  specific  locations,  suggest- 
ing that  individuals  collected  at  a  location  had  expe- 
rienced similar  environmental  conditions  throughout 
their  larval  lives.  This  observation  raised  the  possibility 
of  reconstructing  larval  dispersal  pathways  based  on 
high-resolution  sampling  of  otolith  chemistry.  Although 
further  work  is  needed  to  understand  the  processes 
influencing  elemental  uptake  in  pollock  otoliths,  we  sug- 
gest that  the  potential  information  available  from  such 
studies  would  be  invaluable  for  effective  management 
of  commercial  pollock  fisheries  (Bailey  et  al.,  1999).  The 
approach  appears  to  be  particularly  appropriate  for  in- 


Fitzgerald  et  al.:  Elemental  signatures  in  otoliths  of  larval  Theragra  chalcogramma 


615 


vestigating  the  potential  existence  of  fine-scale  popula- 
tion structure  throughout  the  species  range.  Significant 
fine-scale  population  structure  has  been  linked  to  the 
failure  of  northern  cod  stocks  to  recover  from  exploita- 
tion, even  in  the  face  of  fishing  moratoriums  (Frank 
and  Brickman,  2000;  Hutchings,  2000).  Analogous  de- 
mographic processes  acting  in  northern  cod  populations 
are  clearly  possible  in  walleye  pollock,  given  the  phyolo- 
genetic  and  life  history  similarities  between  the  two 
species.  The  structure  of  pollock  stock  complexes  within 
the  major  basins  of  the  North  Pacific  Ocean  remains, 
therefore,  a  critical  gap  in  the  knowledge  necessary  for 
the  sustainable  management  of  one  of  the  world's  larg- 
est marine  fisheries. 


Acknowledgments 

This  work  was  funded  by  North  Pacific  Marine  Research 
Program  to  KMB,  SRT  and  KPS,  and  was  supported  in 
part  by  NSF  grants  OCE-9871047  and  OCE-0134998  to 
SRT.  We  thank  the  MACE,  FOCI  and  groundfish  task 
scientists  who  collected  samples,  and  C.  Latkoczy  for 
assistance  with  the  laser  ablation  ICP-MS  analyses.  This 
is  Fisheries-Oceanography  Coordinated  Investigations 
collection  number  0471-00A-0. 


Literature  cited 

Bailey,  K.  M.,  T.  J.  Quinn,  II,  P.  Bentzen,  and  W.  S.  Grant. 

1999.  Population  structure  and  dynamics  of  walleye 
pollock,  Theragra  chalcogramma.  Adv.  Mar.  Biol. 
37:179-255. 

Bath,  G.  E.,  S.  R.  Thorrold,  C.  M.  Jones,  S.  E.  Campana,  J.  W. 
McLaren,  and  J.  W.  H.  Lam. 

2000.  Strontium  and  barium  uptake  in  aragonitic  oto- 
liths of  marine  fish.  Geochim.  Cosmochim.  Acta  64: 
1705-1714. 

Bath  Martin,  G.  E.,  S.  R.  Thorrold,  and  C.  M.  Jones. 

2004.     The  influence  of  temperature  and  salinity  on  stron- 
tium uptake  in  the  otoliths  of  juvenile  spot  \Leiostomus 
xanthurus).     Can.  J.  Fish.  Aquat.  Sci.  61:34-42. 
Begg,  G.  A.,  and  J.  R.  Waldman. 

1999.     An  holistic  approach  to  fish  stock  identification. 
Fish.  Res.  43:1-3. 
Campana,  S.  E. 

1999.     Chemistry  and  composition  offish  otoliths:  path- 
ways, mechanisms  and  applications.     Mar.  Ecol.  Prog. 
Ser.  188:263-297. 
Campana,  S.  E.,  J.  A.  Gagne,  and  J.  W.  McLaren. 

1995.     Elemental  fingerprinting  of  fish  otoliths  using 
ID-ICPMS.     Mar.  Ecol.  Prog.  Ser.  122:115-120. 
Campana,  S.  E.,  and  J.  D.  Neilson. 

1985.     Microstructure  of  fish  otoliths.     Can.  J.  Fish. 
Aquat.  Sci.  42:1014-1032. 
Campana,  S.  E.,  S.  R.  Thorrold,  C.  M.  Jones,  D.  Guenther, 
M.  Tubrett,  H.  Longerich,  S.  Jackson,  N.  M.  Halden, 
J.  M.  Kalish,  P.  Piccoli,  H.  De  Pontual,  H.  Troadec, 
J.  Panfili.  D.  H.  Secor. 

1997.  Comparison  of  accuracy,  precision,  and  sensitivity  in 
elemental  assays  offish  otoliths  using  the  electron  micro- 
probe,  proton-induced  X-ray  emission,  and  laser  ablation 


inductively  coupled  plasma  mass  spectrometry.     Can. 
J.  Fish.  Aquat.  Sci.  54:2068-2079. 
Frank,  K.  T,  and  D.  Brinkman. 

2000.     Allee  effects  and  compensatory  population  dynam- 
ics within  a  stock  complex.     Can.  J.  Fish.  Aquat.  Sci. 
57:513-517. 
Gillanders,  B.  M. 

2002.     Connectivity  between  juvenile  and  adult  fish 
populations:  do  adults  remain  near  their  recruitment 
estuaries?     Mar.  Ecol.  Prog.  Ser.  240:215-223. 
Grant,  W.  S.,  and  F.  M.  Utter. 

1980.     Biochemical  genetic  variation  in  walleye  pollock, 
Theragra  chalcogramma  population  structure  in  the 
southeastern  Bering  Sea  and  the  Gulf  of  Alaska.     Can. 
J.  Fish.  Aquat.  Sci.  37:  093-1100. 
Gunn,  J.  S.  H.,  I.  R.  Harrowfield,  C.  H.  Proctor  and 
R.  E. Thresher. 

1992.     Electron  probe  microanalysis  of  fish  otoliths-- 
evaluation  of  techniques  for  studying  age  and  stock 
discrimination.     J.  Exp.  Mar.  Biol.  Ecol.  158:1-36. 
Gunther,  D.,  and  C.  A.  Heinrich. 

1999.  Enhanced  sensitivity  in  laser  ablation-ICP  mass 
spectrometry  using  helium-argon  mixtures  as  aerosol 
carrier.     J.  Anal.  At.  Spectrom  14:1363-1368. 

Hanski,  I.,  and  M.  Gilpin  (eds.). 

1997.  Metapopulation  dynamics.  Academic  Press.  San 
Diego,  CA. 

Hellberg,  M.  E.,  R.  S.  Burton,  J.  E.  Neigel,  and  S.  R.  Palumbi. 
2002.     Genetic  assessment  of  connectivity  among  marine 
populations.     Bull.  Mar.  Sci.  70273-290. 
Hinckley,  S. 

1987.     The  reproductive  biology  of  walleye  pollock,  Ther- 
agra chalcogramma,  in  Bering  Sea,  with  reference  to 
spawning  stock  structure.     Fish.  Bull.  85:481-498. 
Hutchings,  J.  A. 

2000.  Collapse  and  recovery  of  marine  fishes.  Nature 
406:882-885. 

Jones,  G.  P.,  M.  J.  Milicich,  M.  J.  Emslie,  and  C.  Lunow. 

1999.     Self-recruitment  in  a  coral  reef  fish  population.     Na- 
ture 402:802-804. 
Littel,  R.  C,  R.  J.  Freund,  and  P.  C.  Spector. 

1991.  SAS  systems  for  linear  models.  3rd  ed.  SAS  Inst.. 
Inc.,  Cary,  NC. 

Mahoney,  P.  P.,  G.  Li,  and  G  M.  Hieftje. 

1996.     Laser  ablation-inductively  coupled  plasma  mass 

spectrometry  with  a  time-of-flight  mass  analyzer.     J. 

Anal.  At.  Spectrom.  11:401-405. 
Milton,  D.  A.,  and  S.  R.  Chenery. 

1998.  The  effect  of  otolith  storage  methods  on  the 
concentrations  of  elements  detected  by  laser-ablation 
ICPMS.     J.  Fish  Biol.  53:785-794. 

2001.  Sources  and  uptake  of  trace  metals  in  otoliths  of 
juvenile  barramundi  iLates  calcarifer).  J.  Exp.  Mar. 
Biol.  Ecol.  264:47-65 

Mulligan,  T.  J.,  R.  W.  Chapman,  and  B.  L.  Brown. 

1992.  Mitochondrial  DNA  analysis  of  walleye  pollock, 
Theragra  chalcogramma  ,  the  eastern  Bering  Sea  and 
Shelikof  Strait,  Gulf  of  Alaska.  Can.  J.  Fish.  Aquat. 
Sci.  49:319-326. 

Olsen,  J.  B.,  S.  E.  Merkouris,  and  J.  E.  Seeb. 

2002.  An  examination  of  spatial  and  temporal  genetic 
variation  in  walleye  pollock  [Theragra  chalcogramma) 
using  allozyme,  mitochondrial  DNA.  and  microsatellite 
data.     Fish.  Bull.  100:752-764. 

Proctor,  C.  H.  and  R.  E.  Thresher. 

1998.     Effects  of  specimen  handling  and  otolith  prepara- 


616 


Fishery  Bulletin  102(4) 


tion  on  concentration  of  element  in  fish  otoliths.     Mar. 
Biol.  131:  681-694. 
Rosenthal,  Y.,  P.  M.  Field,  and  R.  M.  Sherrell. 

1999.     Precise  determination  of  element/calcium  ratios 
in  calcareous  samples  using  sector  field  inductively 
coupled   plasma   mass   spectrometry.     Anal.    Chem. 
71:3248-3253. 
Scott,  V.  D.,  G.  Love,  and  S.  J.  B.  Reed. 

1995.     Quantitative  electron-probe  analysis.     Ellis  Hor- 
wood  Limited,  Hemel  Hempstead,  Hertfordshire,  England. 
Serobaba,  I.  I. 

1977.     Data  on  the  population  structure  of  the  walleye  pol- 
lock, Theragra  chalcogramma.  from  the  Bering  Sea.     J. 
Ichthyol.  17:219-231. 
Severin,  K.  P.,  J.  Carroll  and  B.  L.  Norcross. 

1995.     Electron  microprobe  analysis  of  juvenile  walleye 
pollock,  Theragra  chalcogramma,  otoliths  from  Alaska: 
a  pilot  stock  separation  study.     Environ.  Biol.  Fishes 
43:269-283. 
Shields,  G.  F„  and  J.  R.  Gust. 

1995.     Lack  of  geographic  structure  in  mitochondrial 
DNA  sequences  of  Bering  walleye  pollock,  Theragra 
chalcogramma.     Molecular   Mar.    Biol.    Biotechnol. 
4:69-82. 
Smedbol,  R.  K.,  and  J.  S.  Wroblewski. 

2002.     Metapopulation  theory  and  northern  cod  popula- 
tion structure:  interdependency  of  subpopulations  in 
recovery  of  a  groundfish  population.     Fish.  Res.  55: 
161-174. 
Temnykh,  O.  S. 

1994.     Morphological  differentiation  of  walleye  pollock 
Theragra  chalcogramma  the  western  Bering  Sea  and 
Pacific  Kamchatka  waters.     Voprosy  ikhtiologii  34:  no.  2. 
Thorrold,  S.  R.,  C.  M.  Jones,  and  S.  E.  Campana. 

1997.  Response  of  otolith  microchemistry  to  environ- 
mental variations  experienced  by  larval  and  juvenile 
Atlantic  croaker  iMicropogomas  undulatus) .  Limnol. 
Oceanogr.  42:102-111. 


Thorrold,  S.  R.,  G.  P.  Jones,  M.  E.  Hellberg.  R.  S.  Burton, 
S.  E.  Swearer,  J.  E.  Neigel.  S.  G.  Morgan,  and  R.  R.  Warner. 
2002.     Quantifying  larval  retention  and  connectivity 
in  marine  populations  with  artificial   and  natural 
markers.     Bull.  Mar.  Sci.  70:291-308. 
Thorrold,  S.  R.,  C.  Latkoczy,  P.  K.  Swart,  and  C.  M.  Jones. 

2001.     Natal  homing  in  a  marine  fish  metapopulation. 
Science  291:297-299. 
Thorrold,  S.  R.,  and  S.  Shuttleworth. 

2000.     In  situ  analysis  of  trace  elements  and  isotope 
ratios  in  fish  otoliths  using  laser  ablation  sector  field 
inductively  coupled  plasma  mass  spectrometry.     Can. 
J.  Fish.  Aquat.  Sci.  57:232-242. 
Thresher,  R.  E. 

1999.  Elemental  composition  of  otoliths  as  a  stock  delin- 
eator in  fishes.     Fish.  Res.  43:165-204. 

Thresher,  R.  E.,  C.  H.  Proctor,  J.  S.  Gunn,  and  I.  R.  Harrowfield. 

1994.     An  evaluation  of  electron-probe  microanalysis  of 

otoliths  for  stock  delineation  and  identification  of  nursery 

areas  in  a  southern  temperature  groundfish.  Nemadac- 

tylus  macropterus.     Fish.  Bull.  92:817-840. 

Tsugi,  S. 

1989.     Alaska   pollock  population,   Theragra   chalco- 
gramma, of  Japan  and  its  adjacent  waters:  Japanese 
fisheries  and  population  studies.     Mar.  Behav.  Physiol. 
15:147-205. 
Waples,  R.  S. 

1998.     Separating  the  wheat  from  the  chaff:  patterns  of 
genetic  differentiation  in  high  gene  flow  species.     J. 
Heredity  89:438-450. 
Winer,  B. 

1971.     Statistical  principles  in  experimental  design,  2nd 
ed.     McGraw-Hill.  New  York.  NY. 
Yoshinaga,  J.,  A.  Nakama.  M.  Morita,  and  J.  S.  Edwards. 

2000.  Fish  otolith  reference  material  for  quality  assur- 
ance of  chemical  analyses.     Mar.  Chem.  69:91-97. 


617 


Abstract — Fishery-independent  esti- 
mates of  spawning  biomass  (Bsp)  of 
the  Pacific  sardine  {Sardinops  sagax) 
on  the  south  and  lower  west  coasts  of 
Western  Australia  (WA)  were  obtained 
periodically  between  1991  and  1999 
by  using  the  daily  egg  production 
method  (DEPMl.  Ichthyoplankton 
data  collected  during  these  surveys, 
specifically  the  presence  or  absence 
of  S.  sagax  eggs,  were  used  to  investi- 
gate trends  in  the  spawning  area  of  S. 
sagax  within  each  of  four  regions.  The 
expectation  was  that  trends  in  Bsp  and 
spawning  area  were  positively  related. 
With  the  DEPM  model,  estimates  of 
Bsp  will  change  proportionally  with 
spawning  area  if  all  other  variables 
remain  constant.  The  proportion  of 
positive  stations  (PPS),  i.e.,  stations 
with  nonzero  egg  counts — an  objec- 
tive estimator  of  spawning  area — was 
high  for  all  south  coast  regions  during 
the  early  1990s  (a  period  when  the 
estimated  BSP  was  also  high)  and 
then  decreased  after  the  mid-1990s. 
There  was  a  decrease  in  PPS  from 
the  mid-1990s  to  1999.  The  particu- 
larly low  estimates  in  1999  followed  a 
severe  epidemic  mass  mortality  of  S. 
sagax  throughout  their  range  across 
southern  Australia.  Deviations  from 
the  expected  relationship  between 
BSP  and  PPS  were  used  to  identify 
uncertainty  around  estimates  of  Bsp. 
Because  estimation  of  spawning  area 
is  subject  to  less  sampling  bias  than 
estimation  ofBsp,  the  deviation  in  the 
relation  between  the  two  provides  an 
objective  basis  for  adjusting  some  esti- 
mates of  the  latter.  Such  an  approach 
is  particularly  useful  for  fisheries 
management  purposes  when  sampling 
problems  are  suspected  to  be  present. 
The  analysis  of  PPS  undertaken  from 
the  same  set  of  samples  from  which 
the  DEPM  estimate  is  derived  will 
help  provide  information  for  stock 
assessments  and  for  the  management 
of  purse-seine  fisheries. 


A  sudden  collapse  in  distribution  of 

Pacific  sardine  (Sardinops  sagax) 

off  southwestern  Australia  enables 

an  objective  re-assessment  of  biomass  estimates 


Daniel  J.  Gaughan 

Timothy  I.  Leary 

Ronald  W.  Mitchell 

Ian  W.  Wright 

Western  Australian  Marine  Research  Laboratories 

Department  of  Fisheries 

West  Coast  Drive 

Waterman,  Western  Australia  6020,  Australia 

E-mail  address  (for  D  J  Gaughan):  dgaughan  Sfish  wa govau 


Manuscript  submitted  5  December  2002 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
27  April  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:617-633  (2004). 


As  a  stock  of  small  pelagic  fish  de- 
creases, biomass  assessment  becomes 
problematic  because  of  such  factors 
as  patchy  distribution  (Fletcher  and 
Sumner,  1999)  and  continuing  high 
catchability  as  a  result  of  the  schooling 
behavior  of  some  fish  (Uphoff,  1993).  In 
these  circumstances,  ichthyoplankton 
surveys  can  provide  a  useful  means  of 
estimating  spawning  biomass,  Bsp. 
for  some  pelagic  fish  species.  Mangel 
and  Smith  (1990)  used  a  technique 
that  assessed  the  presence  or  absence 
of  sardine  (Sardinops  sagax)  eggs  in 
a  known  spawning  area.  They  found 
that  changes  in  adult  biomass  were 
more  accurately  predicted  by  using 
presence-absence  of  eggs  in  sampling 
surveys  than  mean  egg  abundance 
because  of  misleading  results  arising 
from  the  spatial  patchiness  of  eggs.  In 
their  presence-absence  analysis,  the 
spatial  distribution  of  eggs  is  the  key 
determinant  of  Bsp  estimates  and  is 
used  in  a  model  with  a  series  of  other 
parameters  to  provide  an  estimate 
of  Bsp  (Mangel  and  Smith,  1990). 
Although  this  technique  provides 
an  objective  indication  of  stock  size 
that  is  not  subjected  to  the  inherent 
problems  in  estimating  BSP  with  the 
daily  egg  production  method  (DEPM, 
e.g.,  Ward  et  al.,  2001),  the  modeling 
requires  substantial  prior  knowledge 
of  adult  and  egg  production  param- 
eters. More  recently,  Zenitani  and 
Yamada  (2000)  developed  an  optimal 
relationship  between  Bsp  and  spawn- 


ing area  for  the  Japanese  sardine 
(Sardinops  melanostictus)  using  a 
nonlinear  model  that  assumed  patchy 
egg  distribution.  In  their  case,  bio- 
mass was  estimated  by  using  virtual 
population  analysis  with  catch-at-age 
data  from  the  commercial  fishery. 

The  purse  seine  fishery  for  Sar- 
dinops sagax  in  Western  Australia 
(WA)  operates  along  the  south  coast 
around  the  port  regions  of  Esperance, 
Bremer  Bay,  and  Albany;  and  on  the 
lower  West  Coast  in  the  regions  of 
Fremantle  and  Dunsborough  (Fig.  1). 
A  level  of  spatial  distinctness  among 
adult  Sardinops  populations  neces- 
sitates that  three  south  coast  regions 
and  the  west  coast  region  be  managed 
as  separate  fisheries  (Gaughan  et  al., 
2002).  Unlike  the  case  with  Japanese 
sardine  (Zenitani  and  Yamada,  2000), 
it  has  not  been  possible  to  estimate 
the  Bsp  of  Sardinops  in  each  fish- 
ery in  WA  using  only  an  age-based 
approach.  Although  Gaughan  et  al. 
(2002)  considered  the  catch-at-age 
data  for  the  WA  Sa?-dinops  fisheries 
to  be  reasonable,  the  data  span  a  rel- 
atively short  time  series,  commenc- 
ing in  1988  at  Albany  and  Bremer 
Bay  and  later  at  the  other  regions. 
Therefore,  both  age-structure  data 
and  estimates  of  spawning  biomass 
(Bsp)  obtained  with  the  DEPM  have 
provided  the  biological  basis  for 
managing  the  Sardinops  fisheries  in 
WA  for  over  a  decade  (Fletcher, 
1991,   1995;  Fletcher  et  al.,   1996, 


618 


Fishery  Bulletin  102(4) 


114° 


116° 


118° 


J  Fremantle 


•Dunsborough 
If"  Bremer  Ba 

Albany 


Figure  1 

Map  of  southwestern  Australia  showing  Pacific  sardine  fishing  ports.  North  of  Dunsborough  to 
Fremantle  constitutes  the  west  coast  fishery  and  the  other  regions  constitute  the  south  coast 
fishery.  WA  =  Western  Australia,  SA  =  South  Australia. 


19961;  Cochrane2).  A  population  model  that  integrates 
age-structure  information  and  DEPM-derived  estimates 
of  Bsp  (BSp.DEPM)  has  recently  been  developed  by  Hall 
(2000)  for  each  of  the  three  south  coast  regions. 

Although  the  DEPM  is  able  to  provide  relatively  ro- 
bust estimates  of  BSP  for  a  variety  of  species  (Alheit, 
1993;  Hunter  and  Lo,  1997),  it  is  not  without  problems 
(Cochrane,  1999;  Ward  et  al.,  2001).  The  Bsp_DEPM  es- 
timates for  Sardinops  in  WA  are  presented  at  Manage- 
ment Advisory  Committee  (MAC)  meetings  and,  in  turn, 
are  provided  to  the  relevant  government  minister.  The 
Bsp  DEPM  estimates  therefore  undergo  critical  scrutiny 
by  industry  representatives.  The  shortcomings  of  the 
DEPM  (e.g.,  sensitivity  of  precision  for  small  sample 
sizes)  are  well  understood  by  the  members  of  the  man- 
agement committee;  industry  recognizes  that  onshore 
infrastructure  and  fleet  capacity  must  be  matched  to 
long-term  average  Bsp  and  that  industry  should  not 
capitalize  at  levels  that  require  maximal  stock  sizes  to 
meet  financial  expectations.  Inasmuch,  a  level  of  conser- 
vatism has  been  adopted  by  the  Management  Advisory 
Committee  when  setting  quotas.  Nonetheless,  the  accu- 


1  Fletcher,  W.  J.,  K.  V.  White.  D.  J.  Gaughan,  and  N.  R. 
Sumner.  1996.  Analysis  of  the  distribution  of  pilchard 
eggs  off  Western  Australia  to  determine  stock  identity  and 
monitor  stock  size.  Final  Report  to  Fisheries  Research  and 
Development  Corporation.  Project  No.  92/95.  109  p.  De- 
partment of  Fisheries,  Government  of  Australia,  168-170 
St.  Georges  Tee,  Perth,  WA  6000,  Australia. 

2  Cochrane,  K.  L.  1999.  Review  of  Western  Australia  pil- 
chard fishery,  12-16  April  1999.  Fisheries  Management 
Paper  129,  32  p.  Department  of  Fisheries,  Government 
of  Western  Australia,  168-170  St.  Georges  Tee,  Perth,  WA 
6000,  Australia. 


racy  of  estimates  has  been  a  contentious  issue;  industry 
members  typically  believe  that  the  scientific  advice 
presented  often  underestimates  the  BSP.  Likewise,  wide 
confidence  intervals  around  biomass  estimates  introduce 
doubt  in  the  minds  of  industry  members  regarding  the 
reliability  of  scientific  advice,  which  can  therefore  stall 
the  implementation  of  management  measures.  However, 
the  lack  of  a  formal  and  objective  means  of  dealing  with 
suspect  and  imprecise  Bsp _DEPM  estimates  (e.g.,  because 
of  problems  with  sampling  spawning  fish)  has  previ- 
ously not  been  rigorously  addressed. 

Following  the  progression  along  the  southern  WA 
coast  in  early  1999  of  a  mass  mortality  of  Sardinops, 
estimates  of  the  quantity  killed  at  Albany  appeared 
to  be  very  low  (Gaughan  et  al.,  20001.  That  is,  very- 
few  dead  Sardinops  were  found  in  comparison  to  the 
other  regions  where  fisheries  occur.  Mortality  rates 
for  Esperance  and  Bremer  Bay  were  69. 6rA  and  74.59f 
of  the  Bsp,  respectively,  whereas  that  for  Albany  was 
estimated  to  be  only  2.4'7f .  Estimates  of  the  mortality 
rate  of  Sardinops  in  South  Australia  (SA.  Fig.  1)  for 
the  same  epizootic  event  were  independently  found  to 
also  be  around  70*  (Ward  et  al.,  2001).  The  inconsis- 
tency with  Albany  could  not  be  attributed  to  different 
weather  conditions;  the  weather  conditions  at  Albany- 
were  similar  to  those  at  Esperance  and  Bremer  Bay  and 
would  be  expected  to  result  in  equally  visible  evidence 
of  mortality.  Gaughan  et  al.  (2000)  contended  that  the 
true  epizootic  mortality  rate  of  Sardinops  in  Albany  was 
similar  to  that  for  the  other  regions,  but  that  the  very 
low  mortality  estimate  was  likely  seen  as  such  in  view 
of  the  previous  overestimation  ofBsp- 

In  this  study  we  aimed  to  address  the  problem  of  poor 
precision,  while  also  developing  a  technique  to  identify- 
particularly  poor  estimates  of  BSPmDEPM,  i.e.,  those  for 


Gaughan  et  al.:  Distribution  of  Sardinops  sagax  off  southwestern  Australia 


619 


Table  1 

Estimates  of  s 

Dawning  biomass  (metric  tonsl  of  Sardinops  sagax  obtained  by 

usi 

ng  the  daily 

egg  production 

method  (S^p.qjpv1 

for  each  of  four 

regions  in  southwestern  Australia. 

In  those  cases  where  data  were  sufficient  to  estimate  a  coe 

fficient  of  variation 

(CV).  the  range  around  the  Bsp  DEPM  estimate  (Min 

./Max.!  was  caleu 

lated  as  ± 

1  standard  deviation  (SDl;  otherwise,  the  8Sn  Mp u 

range  was  calculated  by  using  assumed  (AS )  values  for  one  or  more 

oftheDEPM 

parameters 

see  text).  The 

lumbers  of  adult  S. 

sagax  and  plan 

kton  samples  used  in  these  calculations  are  shown. 

Year 

Mm.                "sp-depm 

Max. 

CV 

±1SD 

Adult  n 

Plankton  n 

Albany 

1991 

12,088               19,300 

30,700 

— 

AS 

10 

41 

1992 

9006                16,994 

24,981 

0.44 

±1SD 

10 

31 

1993 

16,402               23,432 

30,4620 

0.30 

±1SD 

9 

61 

1994 

15,440               31,330 

55,000 

— 

AS 

10 

107 

1995 

7720                17.544 

27,368 

0.56 

±1SD 

10 

83 

1997 

13,018               18,597 

24,176 

0.30 

±1SD 

27 

94 

1999 

0                      89 

531 

4.99 

±1SD 

2 

263 

Bremer  Bay 

1992 

12,000               19,280 

79,000 

— 

AS 

— 

25 

1993 

16,170               44,010 

63,608 

— 

AS 

— 

32 

1994 

15,700                28,458 

42,500 

— 

AS 

— 

102 

1999 

2161                  4156 

6150 

0.48 

±1  SD 

3 

256 

Esperance 

1993 

14.326               32,  252 

61,800 

— 

AS 

5 

50 

1994 

10,700               20,080 

40,100 

— 

AS 

— 

150 

1995 

10,900               31,900 

45,647 

— 

AS 

6 

105 

1999 

3454                17,396 

31,793 

0.80 

±1SD 

8 

257 

West  coast 

1993 

14,500                41,250 

78,000 

— 

AS 

— 

55 

1994 

3100                   8714 

29,000 

— 

AS 

— 

133 

1996 

43,300                60,228 

77,200 

0.28 

±1  SD 

4 

96 

1998 

9112                18,985 

28,951 

0.52 

±1SD 

28 

240 

1999 

3948                  5275 

6651 

0.25 

rtlSD 

4 

396 

which  accuracy  was  suspect.  However,  because  of  the 
small  size  of  the  fishery  and  the  difficulty  in  secur- 
ing additional  research  funds,  no  fishery-independent 
method  of  estimating  biomass  other  than  the  DEPM 
was  undertaken.  We  recognized  that  this  study  could 
not,  therefore,  unequivocally  determine  whether  or  not 
any  improvement  in  accuracy  had  been  achieved;  there- 
fore, we  focused  on  improving  the  consistency  between 
available  data  sets  (in  terms  of  the  broader  economic 
environment)  in  order  to  improve  the  decision-making 
process  for  what  is  a  small-scale  fishery. 

We  examine  the  relationship  between  relative  trends 
in  the  B 


SP-DEPM 


and  spawning  area  of  Sardinops  in  each 
of  four  regional  fisheries.  We  propose  a  method  of  using 
the  relationship  between  spawning  area  and  egg  pres- 
ence-absence data  as  an  indicator  of  Bsp  that  is  simpler 
than  the  methods  of  Mangel  and  Smith  (1990)  and 
Zenitani  and  Yamada  (2000).  We  specifically  chose  to 
keep  the  retrospective  analysis  simple  in  recognition  of 
data  limitations,  i.e.,  short  time  series  and  low  numbers 
of  samples  for  some  DEPM  surveys.  In  particular,  our 
analyses  did  not  rely  on  substantial  knowledge  of  vari- 
ous parameters  associated  with  estimating  BSPDEPM. 


This  investigation  (an  objective  re-assessment  of  Sar- 
dinops Bsp  DEPM)  was  undertaken  because  of  the  con- 
trast provided  by  the  significant  collapse  in  distribu- 
tion, coupled  with  the  sudden  and  substantial  decrease 
in  Bsp  DEPM  that  followed  mass  mortality  in  1998-99. 
Although  problems  in  obtaining  accurate  BSP  estimates 
with  the  DEPM  will  not  be  resolved  by  the  present 
study,  greater  consistency  between  indicators  of  the 
magnitude  of  Bsp  and  the  development  of  a  transparent 
and  objective  technique  to  identify  apparent  discrepan- 
cies between  the  two  will  facilitate  better  management 
of  this  key  pelagic  resource. 


Methods 

Estimates  of  spawning  biomass  with  the  DEPM 

The  estimates  of  Bsp  used  in  this  study  (Table  D  were 
obtained  by  using  the  DEPM,  which  relies  on  ichthyo- 
plankton  surveys  to  estimate  egg  production  and  tem- 
porally concurrent  samples  of  adult  fish  to  estimate  the 
adult  parameters  of  fecundity,  sex  ratio,  and  weight. 


620 


Fishery  Bulletin  102(4) 


The  egg  and  adult  data  were  subsequently  combined  in 
the  DEPM  model  (Parker,  1985),  as  follows,  to  estimate 
spawning  biomass: 


B 


SP  DEPM 


<APWk)/<SFR), 


where  A   =   spawning  area; 

P   =  egg  production  (numbers  of  eggs  before 

losses  due  to  mortality); 
W  =  weight  of  adult  fish; 
k    =   conversion  factor  to  bring  the  various  units 

to  a  value  in  metric  tons; 
S   =   spawning  fraction;  the  proportion  of  females 

that  spawn  per  day; 
F  =   fecundity;  number  of  eggs  produced  by  a 

female;  and 
R   =   ratio  of  females  to  males  by  weight. 

The  DEPM  provides  a  point  estimate  of  spawning 
biomass,  with  upper  and  lower  statistical  bounds.  In 
those  individual  surveys  where  all  parameters  could  be 
estimated,  estimates  of  coefficient  of  variation  (CV)  for 


B, 


were  undertaken  by  using  the  delta  method 


to  sum  the  CV  of  the  component  parameters  (Parker, 
1985).  In  turn,  the  CV  was  used  to  provide  an  estimate 
of  variability  around  the  point  estimate;  specifically,  we 
used  ±  1  standard  deviation  to  indicate  the  upper  and 
lower  bounds  around  the  point  estimates.  In  several 
surveys,  particularly  when  the  DEPM  was  initially  be- 
ing applied  in  WA,  few  adult  samples  meant  that  val- 
ues for  adult  parameters  (spawning  fraction,  sex  ratio, 
fecundity,  weight)  could  not  be  estimated  and  therefore 
a  CV  for  the  final  estimate  of  Bsp  DEPM  could  likewise 
not  be  estimated.  Although  sex  ratio  and  weight  could 
be  reasonably  estimated  from  the  regular  sampling  of 
commercial  catches  around  the  survey  period  and  fecun- 
dity could  be  estimated  from  a  relatively  small  sample 
(e.g.,  70-100  fish),  estimating  the  spawning  fraction  was 
more  difficult.  In  this  latter  case,  the  upper  and  lower 
bounds  for  the  Bsp  DEPM  estimate  were  not  based  on  a 
statistical  measure  but  rather  on  what  were  thought  to 
be  likely  low  and  high  values  of  spawning  fraction,  re- 
spectively, for  Sardinops  from  other  surveys  in  WA  and 
elsewhere  (e.g.,  Alheit,  1993,  Fletcher  et  al.  1996).  Prior 
knowledge  of  likely  Bsp  DEPM  values  when  applying  the 
DEPM,  specifically  for  the  purpose  of  providing  expert 
management  advice,  has  recently  been  used  successfully 
for  Sardinops  in  South  Australia  (Ward  et  al.  2001). 

Adult  samples 

Twenty  DEPM  surveys  were  conducted  between  1991 
and  1999  to  identify  stocks  and  to  estimate  spawn- 
ing biomass  of  Sardinops  of  southwestern  Australia 
(Fletcher  et  al.,  1996a,  1996b;  Fletcher  et  al.3;  senior 
author's  unpubl.  data).  The  surveys  were  performed 
during  the  peak  spawning  months  for  Sairlinops  off  the 
west  coast,  Albany,  Bremer  Bay,  and  Esperance  regions. 
The  timing  of  the  DEPM  survey  cruises  in  each  region 
was  based  on  gonadosomatic  indices  for  samples  obtained 


from  commercial  catches,  as  described  in  Gaughan  et  al. 
(2002).  The  aim  was  to  obtain  samples  from  35  catches 
of  adult  fish,  as  recommended  by  Alheit  (1993),  but 
this  number  was  never  achieved  and  in  some  cases  no 
samples  were  obtained  (Table  1).  For  each  catch  sampled, 
the  ovaries  from  15-50  females  were  immediately  placed 
in  10%  formalin  and  subsequently  prepared  histologi- 
cally for  microscopic  examination.  The  remainder  of  the 
subsample  was  processed  to  obtain  mean  female  weight 
and  sex  ratio  by  weight.  Mature  ovaries  were  retained 
for  estimation  of  fecundity. 

Plankton  sampling  and  estimation  of  egg  production 

Plankton  sampling  extended  from  nearshore  waters 
to  the  edge  of  the  continental  shelf  (Fig.  2).  Sampling 
stations  were  generally  spaced  uniformly,  typically  2-4 
nautical  miles  apart,  along  transects  perpendicular  to 
the  shore.  Analysis  of  Sardi/iops  egg  distribution  from 
surveys  conducted  in  the  early  1990s  indicated  that 
these  surveys  sufficiently  covered  the  distribution  of 
the  spawning  stock  (Fletcher  and  Tregonning,  1992; 
Fletcher  et  al.,  1994),  and  later  geostatistical  analyses 
of  Sardinops  egg  distribution  patterns  confirmed  that 
the  spacing  of  transects  and  stations  were  adequate  to 
effectively  represent  the  spatial  distribution  (Fletcher 
and  Sumner,  1999).  The  earlier  surveys  were  used  to 
refine  the  spatial  range  of  subsequent  surveys.  The 
number  of  plankton  samples  taken  in  each  survey  has 
generally  increased  since  the  early  1990s  (Table  1). 

Sardinops  eggs  were  collected  by  using  vertical  tows 
that  allowed  the  water  column  to  be  sampled  from  a 
maximum  depth  of  70  m  to  the  surface;  Fletcher  (1999) 
showed  that  Sardinops  eggs  off  southern  Australia  are 
typically  restricted  to  the  upper  70  m.  Bongo  nets  with 
diameters  of  either  60  or  26  cm  and  constructed  of 
either  500-  or  300-micron  mesh  were  used;  the  change 
to  smaller  nets  was  made  to  reduce  sample  volume  and 
hence  sorting  time,  whereas  the  change  to  smaller  mesh 
was  made  to  increase  efficiency  in  capturing  yolksac 
larvae;  these  changes  did  not  affect  the  sampling  ef- 
ficiency for  Sardinops  eggs.  Tow  speed  was  standard- 
ized at  1  m/s.  All  samples  were  collected  between  0630 
and  1800  hours  and  immediately  preserved  in  5-10% 
formalin  and  seawater. 

Plankton  samples  were  examined  under  a  dissecting 
microscope.  Sardinops  eggs  were  identified,  classified 
into  12  developmental  stages  (White  and  Fletcher4),  and 


1  Fletcher  W.  J.,  B.  Jones,  A.  F.  Pearce,  and  W.  Hosja.  1997.  En- 
vironmental and  biological  aspects  of  the  mass  mortality  of 
pilchards  (Autumn  1995 1  in  Western  Australia.  Fisheries 
Research  Report,  Fisheries  Department  Western  Australia 
106,  115  p.  Department  of  Fisheries,  Government  of  West- 
ern Australia,  168-170  St.  Georges  Tee.  Perth,  WA  6000. 
Australia. 

4  White.  K.  V..  and  W.  J.  Fletcher.  1998.  Identifying  the 
developmental  stages  for  eggs  of  the  Australian  pilchard. 
Sardinops  sagax.  Fisheries  Research  Division  WA.  Fisher- 
ies Research  Report  103,  21  p.  Department  of  Fisheries, 
Government  of  Western  Australia.  168-170  St.  Georges  Tee, 
Perth,  WA  6000.  Australia. 


Gaughan  et  al.:  Distribution  of  Sard/nops  sagax  off  southwestern  Australia 


621 


Table  2 

Correlations  between 

1)  proportion  of  positive  stations  (PPS)  and  pr 

jportional  spawning  areE 

(PSA) and  2) 

PPS 

and  estimated 

spawning  area  (km2) 

result 

ng  from  surveys  of  Sai 

•dinops  sagax  eggs 

at  four  regions  in 

southwestern  Austra 

lia. 

PPS 

is  the  pro- 

portion  of  the  total  number 

jf  plan 

kton  sampling  stations  that  contai 

ned  at  least  one  S. 

sagax 

egg  that  had  been 

spawned  on  the 

previous  night.   PSA 

is  the 

propoi 

tion  of  the  total 

survey  area  that 

consisted  of  spawning  area.   The  values  for 

spav 

ming  area 

are  also  provided. 

Survey 

PPS 

PSA(%) 

Correlation  I 

Area  (km-) 

Coi 

relation  II 

Albany 

Jul 
Jul 
Jul 
Jul 
Jul 
Jul 

91 
92 
93 
94 
95 
97 

0.46 
0.58 
0.52 
0.60 
0.33 
0.19 

37 
51 
36 
60 
28 
21 

1806 
2686 
2391 
6672 
1977 
2224 

Jul 

99 

0.15 

1 

0.94 

107 

0.68 

Bremer  Bay 

Jul 
Jul 
Jul 

92 
93 
94 

0.61 
0.72 
0.70 

64 
67 
71 

2807 
2809 
4474 

Jun 

99 

0.12 

12 

0.99 

908 

0.86 

Esperance 

Jul 
Jul 
Apr 

93 
94 
95 

0.50 

0.57 
0.61 

44 
73 
36 

5715 
9796 
5277 

May  99 

0.09 

3 

0.81 

7840 

0.83 

West  coast 

Jul 
Jul 
Aug 

93 
94 
96 

0.53 
0.30 
0.23 

62 
38 
16 

8012 
5199 
2202 

Aug  98 

0.10 

7 

1835 

Aug 

99 

0.12 

10 

0.98 

1836 

0.96 

counted.  Estimation  of  egg  production  was  undertaken 
by  fitting  a  negative  exponential  model  (Picquelle  and 
Stauffer,  1985)  and  was  derived  from  the  y-axis  inter- 
cept of  the  regression  model,  representing  time  0.  The 
number  of  stages  used  to  fit  the  model  depended  on  the 
egg  abundance  for  each  stage;  the  best  fitting  model 
was  chosen  visually  from  an  iterative  sequence  of  fits. 
The  best  fit  was  not  necessarily  that  with  the  smallest 
CV  but  rather  that  which  intuitively  did  not  violate  our 
understanding  of  natural  mortality  rates  as  determined 
from  the  literature.  For  example,  the  slope  of  the  regres- 
sion model  must  be  negative  and  egg  mortality  rates 
should  fall  within  the  broad  range  of  0.9-3.9/d  (e.g., 
Smith  et  al.,  1989). 

Estimation  of  spawning  area 

According  to  water  temperatures  during  each  survey  and 
the  stage  of  egg  development,  Sardinops  eggs  were  deter- 
mined to  have  been  spawned  either  the  previous  night 
("day-1")  or  two  nights  previous  ("day-2")  as  described 
by  Fletcher  et  al.  (1996).  The  total  survey  area  was  esti- 
mated by  constructing  a  polygon  around  all  stations.  The 
spawning  area  was  defined  as  the  area  in  which  day-1 
Sardinops  eggs  were  found  (Fletcher  et  al.,  1996a).  The 
areas  of  the  polygons  around  stations  that  had  day-1 
eggs,  referred  to  as  positive  stations,  were  summed  to 
estimate  the  spawning  area  for  each  zone.  When  positive 


stations  occurred  on  the  margin  of  the  sampling  area, 
polygons  for  these  positive  stations  were  drawn  as  for  the 
embedded  positive  stations,  but  the  areas  of  these  poly- 
gons were  extended  by  a  standardized  amount  beyond 
the  sampling  areas  (Wolf  and  Smith,  1986). 

The  proportion  of  positive  stations  (PPS)  was  calcu- 
lated for  each  survey.  The  proportion  of  the  survey  area 
(PSA)  that  consisted  of  spawning  area  was  also  evalu- 
ated in  each  case.  PPS  and  PSA  were  positively  corre- 
lated at  each  region  (Table  2);  this  result  was  expected 
and  indicated  that  PPS  provides  a  realistic  representa- 
tion of  changes  in  spawning  area.  The  relationships 
between  PPS  and  the  areal  estimates  of  spawning  area 
were  not  as  strong,  but  these  latter  estimates  suffered 
as  potential  predictors  of  biomass  in  our  study  because 
of  the  large  differences  in  numbers  of  plankton  samples 
collected  between  surveys  (Table  1).  PPS  is  thus  not 
only  an  objective  measure  but  can  also  be  considered 
as  an  index  of  spawning  area. 

Modeling  of  spawning  biomass 

The  collapse  in  distribution  of  Sardinops  at  each  of  four 
locations  in  southern  Western  Australia  in  1999  is  shown 
by  the  decline  in  spawning  area  (Fig.  2).  The  importance 
of  this  collapse  in  providing  contrast  for  model  fitting 
in  otherwise  poor  data  sets  (few  points  with  either  flat 
or  clumped  distributions)  is  evident  from  linear  fits  of 


622 


Fishery  Bulletin  102(4) 


Albany 


t 


1991 


1992 


-..'  ■       **"'.'• 

1993 

1994 

^»^ 

.    ;/: 

1? 

1995 


1997 


••-'  .■'■■ 


■■:■: 


1999 


Figure  2 

Sardinops  sagax  spawning  area  (distribution  of  eggs  spawned  on  the  night  previous 
to  sampling)  from  ichthyoplankton  surveys  conducted  between  1991  and  1999  for 
Albany,  Bremer  Bay.  Esperance,  and  west  coast  regions. 


Gaughan  et  al.:  Distribution  of  Sardinops  sagax  off  southwestern  Australia 


.    623 


Bremer  Bay 


1184 

1189 

119  4 

119.9 

120  4 

34 

t 

N 

34  5 

35 

1992 

1993 


1994 


— ■'•■■ 


1999 

1204 


Figure  2  (continued) 


BSP.DEPM  against  PPS  for  the  south  coast  locations 
(Fig.  3).  Because  we  wished  to  examine  the  relationship 
between  trends  in  DEPM-based  estimates  of  Bsp  and 
PPS  with  the  aim  of  improving  estimates  of  Bsp,  the 
development  of  an  appropriate  model  is  described  here 
from  first  principles,  followed  by  a  selectivity  analysis 


of  error  variance  to  choose  the  optimal  estimator  of  Bsp. 
Given  that  we  did  not  have  a  means  of  assessing  the  level 
of  accuracy  of  the  "adjusted"  estimates,  and  the  aim  was 
therefore  to  improve  consistency  between  data  sets  for 
the  purpose  of  enhancing  the  decision-making  process, 
our  criteria  in  choosing  an  optimal  estimator  was  to 


624 


Fishery  Bulletin  102(4) 


Esperance 


120  5         121         1215         122         122  5         123         123  5         124 


t 
N 


1993 


• 


1994 


1995 

.     ;% 

1999 

Figure  2  (continued) 


minimize  variance.  In  terms  of  improving  management 
of  the  Sardinops  fisheries  in  southwestern  Australia, 
we  considered  this  approach  appropriate  because  of  the 
relatively  conservative  exploitation  rates  that  have  been 
adopted  by  the  Management  Advisory  Committee. 


considered  the  following  general  relationship  between 
Bsp  DEPM  and  PPS  holding  over  time: 


(B 


SP-DEPM 


i,  =  cx0  (PPS,)'-  +  a,  +  er 


Hi 


Model  selection 

The  procedure  used  in  the  present  study  was  to  first 
invoke  a  general  model  and  then  use  the  data  to  drive 
a  simplification  process  in  order  to  avoid  a  specification 
error.  Thus,  for  each  of  the  fishery  regions,  we  first 


where  A  is  common  for  all  areas,  aQ  and  a2  are  area 
specific  constants,  and  the  error  term  e,  is  independent, 
homoscedastic  and  normal  with  a  mean  of  zero.  This 
model  was  chosen  specifically  in  order  to  be  amenable 
to  Taylor  series  expansion  during  the  simplification 
process.  In  satisfying  dimensional  and  conservational 


Gaughan  et  al.:  Distribution  of  Sardinops  sagax  off  southwestern  Australia 


625 


West  Coast 


114.5  115  1155  116 


N 

^    ,                              1993 

1994 

1996 


1998 


... 


1999 


1145  115  1155 


Figure  2  (continued) 


arguments,  a,  must  be  zero;  therefore  the  above  family 
of  models  reduces  to 


(Bspdepm\  =  a0(PPS,)^  +  e, 


(2) 


An  attempt  to  fit  the  linear  regression  model,  composed 
of  the  natural  logarithms  of  either  side  of  the  above  rela- 


tion, to  the  observed  data  was  unsuccessful  because  of 
heterogeneity  of  error  variance. 

Direct  estimation  of  model  II  with  a  nonlinear  regres- 
sion procedure  gave  estimates  of  A  that  were  near  1  and 
with  large  standard  errors  on  account  of  the  small  size 
of  the  data  sets.  However,  residual  diagnostics  were  sat- 
isfactory. Because  the  observed  PPS  values  fell  between 


626 


Fishery  Bulletin  102(4) 


1990  1992  1994  1996 


O         04 

Q. 

O 

a. 

0.2 


O         04 

CL 

O 


1998  2000 


Bremer  Bay 


1990       1992       1994       1996       1998       2000 


Esperance 


1990        1992        1994        1996        1998       2000 


West  coast 


1994  1996 

Year 


PPS 


Figure  3 

Pints  of  the  proportions  of  positive  stations  (-•-)  and  BSP.DEPM  estimates  (--■--)  for  the  four  Sardinops 
sagax  fisheries  in  southwestern  Australia.  The  right-hand  panel  shows  linear  fits  of  the  relationship 
between  proportion  of  positive  stations  and  BSP.DEPM  estimates. 


Gaughan  et  al.:  Distribution  of  Sardinops  sagax  off  southwestern  Australia 


627 


Table  3 

Parameter  estimates  for  two  models  (III  and  IV,  see  text  for  details)  of  Sardinops  sagax  spawning  biomass,  including  tests  for 
zero  coefficients,  at  each  of  four  regions  in  southwestern  Australia. 


Model  III 

Model  IV 

Albany 

SE  of  estimate:  6011 
Bsp 

SE  of  Bsp 

((22) 

P-level 

Albany 

SE  of  estimate:  6343 

BSP 

SEo{Bsp 

((23) 

P-level 

Intercept          -4597.15 
PPS                  56.780.19 

2419.94 
6638.90 

-1.90 
8.55 

0.070661 
1.91E-08 

PPS            45,910.11 

3552.36 

12.92 

4.96E-12 

Bremer  Bay 

SE  of  estimate:  4016 

BSP 

SE  of  Bsp 

r(  14) 

P-level 

Bremer  Bay 

SE  of  estimate:  3994 

BSP 

SE  of  BSP 

((18) 

P-level 

Intercept          -1436.88 
PPS                 45,341.85 

1573.94 
3638.2 

-0.91 
12.46 

0.37674 
5.75E-09 

PPS            42,784 

2308.30 

18.53 

9.47E-12 

Esperance 

SE  of  estimate:  4639 

BSP 

SE  ofBsp 

r(20) 

P-level 

Esperance 

SE  of  estimate:  10,214 

BSP 

SEo{Bsp 

((21) 

P-level 

Intercept          15,840.52 
PPS                   17,790.27 

1751.38 
4097.73 

9.045 
4.34 

0.000317 
1.66E-08 

PPS            48,377.13 

5095.09 

9.49 

4.76E-09 

West  coast 

SE  of  estimate:  15,280 

Bsp 
Intercept  5316.10 

PPS  63,192.69 


West  coast 

SE  of  estimate:  15,330 


SEofBgp        ((32) 

4826.58  1.10 

23,161.29  2.73 


P-level 
0.278929 
0.010251 


PPS 


BSP 

84,615.52 


12,615.76 


r(33l 
6.71 


P-level 
1.22E-07 


zero  and  one,  and  A  was  also  close  to  1,  model  II  was 
able  to  be  recast  in  a  more  tractable  form  by  using  the 
Taylor  series  expansion  of  the  RHS  of  model  II  about 
PPS  =  1,  leading  to  the  relationship 


iB.sp.pps»/=  <x0PPS,  +  S+et, 


(3) 


where  the  expected  value  of  8  is  approximately 
-0.25a„(A-l).  Details  of  the  derivation  are  provided  in 
Appendix  1. 

Fitting  the  regression  model  III  to  the  DEPM-based 
estimates  of  Bsp  gave  the  estimated  coefficients  shown 
in  the  left  hand  column  of  Table  3.  Residual  diagnostics 
showed  that  model  III  was  satisfactory.  Because  none 
of  the  intercept  terms  were  significantly  different  from 
zero,  the  parsimonious  model 


{BSP-pps)t=  a0(PPS)t  +  et 


(4l 


was  fitted,  giving  the  results  in  the  right  hand  column 
of  Table  3.  Residual  diagnostics  were  also  satisfactory 
for  these  models. 

Optimal  estimation  of  spawning  biomass 

We  now  have  available  two  unbiased  estimates  of  Bsp: 
estimator  1  (i.e.,  Bsp _DEPM)  with  associated  error  e'. 


which  has  an  expected  value  of  0  and  variance  Var(e') 
=  rjj2;  and  estimator  2  (i.e.,  Bsppps)  which  model  IV 
of  the  previous  section  fitted  to  the  values  of  Bsp _DEPM 
with  error  e,  which  had  an  expected  value  of  0  and 
variance  Var(e)  =  o~2.  Thus  estimator  2  can  be  seen  to 
be  unbiased  and  with  full  error  term  (e+e).  In  order  to 
obtain  an  optimal  predictor,  i.e.,  with  minimum  vari- 
ance, of  spawning  biomass  *>Bsp  0 „„„,„/),  we  considered 
the  weighted  average  of  the  two  estimators  above: 

Bsp-Optimai  =  (("''  estimator  1  +  (1-wO  estimator  2), 
with  weight  w:  0<«'<1. 

We  must  choose  the  weight  w  of  estimator  1  in  order 
to  minimize  the  variance  ( Var(Bsp  0 „,„„„/))  of  the  esti- 
mator B.Sp.0p,„„o/. 

Var(Ssp.oP,„„n/»  =  Var(B»e+(l-w)(e+e')) 
=  Varie  +{l-w)e  I 
=  Var(e)  +  ( l-u;)2Var(e')  +  2(  1-w  i 
covariance(e.e') 

=  a2+il-w)'2al2+2(l-wtc><71p 

where  p  =  correlation  between  e  and  e'.  For  Var  (Bsp 
Optimal  t°  be  a  minimum,  the  w  derivative  must  be  zero, 
yielding 


628 


Fishery  Bulletin  102(4) 


0.4  0.6 

DEPM  weighting 

Figure  4 

The  relationship  between  error  variance  between  the  two  estimators  for  spawn- 


the  optimally  weighted  estimate  of  Bsp  (i.e.,  B 


SPOPTIUAL 


).  Error  variance  is 


shown  for  a  range  of  error  correlations  from  r  =  -0.9  to  -0.2. 


0  =  ( 1-w )  <JX  +  ap, 
which  requires  iv  =  1  +  opl  o~v 

In  the  event  of  p  =  -axla,  Bsp .0pllmal  will  have  w  = 
0,  i.e.,  the  optimal  estimator  will  just  be  estimator  2 
alone. 

The  limited  sample  information  available  indicates 
that  c7j  is  approximately  equal  to  ex,  which  we  therefore 
assume  in  order  to  simplify  the  next  analysis.  Because 
BSp_ppS  is  based  on  estimates  of  PPS,  which  can  be 
estimated  with  more  confidence  than  Bsp  DEPM,  it  must 
be  expected  that  often,  if  not  always,  Variance! BSPmPPS) 
<  Variance(Bsp.D£PW). 

i.e.,     Var(e+e')  <  Var(<?)  +  f  ,  for  small  £  >0; 
i.e.,      a2  +  ct,2  +  2p  ct,ct<  a-  +  e; 
i.e.,     2p  <  (-rjj-  +  e)laxa. 

This  requires  p  <  -0.5  when  crl  =  a,  and  e  is  small.  This 
relation  has  an  important  role  in  our  decision  of  what 
is  the  best  estimator. 

In  Figure  4  the  error  variance  for  the  estimator  Bsp 
optimal  's  shown  for  various  DEPM  weightings  and  a 
theoretical  range  of  error  correlations  (i.e.,  between  e 
and  e)  from  r  =  -0.9  to  -0.2.  Our  aim  was  to  choose  a 
DEPM  weighting  that  provides  minimal  error  variance 
along  the  most  stable  regions  of  the  suite  of  error  cor- 
relation curves,  i.e.,  where  the  error  correlation  curves 
are  flattest.  The  error  correlation  curves  from  -0.4  to 
-0.7  were  the  most  stable  and  across  these  the  DEPM 
weightings  from  0.3  to  0.7  had  the  smallest  error  vari- 
ance. Therefore  we  choose  0.5  as  our  preferred  DEPM 
weighting,  which  lies  centrally  within  a  stable  part  of 
the  range  of  theoretical  error  correlations. 


Results 

The  decline  in  spawning  area  in  each  region  (Fig.  2) 
corresponded  to  declines  in  Bsp  DEm  ( Table  1),  which  in 
turn  were  reflected  by  the  Bsp .optimal  estimates  (Fig.  5). 
We  recognize  that  imbalance  in  the  intensity  of  samples 
between  years  poses  a  problem  for  the  interpolation  of 
data  between  sampling  stations  but  we  contend  that  the 
collapse  in  distribution  observed  is  of  sufficient  contrast 
to  be  a  reliable  reflection  of  the  estimated  709f  decrease 
in  Sardinops  biomass  that  resulted  from  the  1998-99 
epidemic  (Gaughan  et  al.,  2000;  Ward  et  al.,  2001).  Note 
that  we  have  used  Albany  (Fig.  2A)  as  the  primary  sup- 
port for  this  contention  because  of  the  larger  data  set. 
The  same  pattern  was  observed  at  all  regions,  although 
it  was  not  so  marked  for  the  west  coast  Sardinops  (Fig. 
2D)  because  estimated  Bsp  (this  term  hereafter  is  used 
generically)  had  already  declined  substantially  between 
1996  and  1998. 

Despite  sometimes  large  intervals  between  consecutive 
surveys,  there  were  two  broad  patterns  in  the  trends  for 
Sardinops  Bsp  during  the  1990s  (Fig.  5).  Within  each 
region  on  the  south  coast  (Albany,  Bremer  Bay,  and 
Esperance),  Bs 


P  DEPM 


remained  relatively  high  in  the 
early  to  mid  1990s  before  decreasing  substantially  by 
1999.  In  contrast  to  the  results  from  south  coast  DEPM 
surveys,  the  west  coast  estimated  BSP_DEPM  fluctuated 
widely  (Table  1).  This  fluctuation  resulted  in  a  rela- 
tively poor  fit  of  the  optimal  model  and  correspondingly 
wide  CIs.  Since  1996,  when  substantially  more  samples 
were  routinely  collected  during  each  survey  on  the  west 
coast,  there  has  also  been  a  decrease  in  Bsp  consistent 
with  that  observed  on  the  south  coast. 

Inconsistency  in  the  determination  of  variability  es- 
timates around  some  B^PI)EPM  estimates  precludes  any 
definitive  statements  about  the  relative  precision  of  the 


Gaughan  et  al.:  Distribution  of  Sardinops  sagax  off  southwestern  Australia 


629 


60000  ■ 

Albany 

50000  • 

30000  ■ 

20000  ■ 

■ 

. 

I 

, 

■ 

10000  ■ 

0  ■ 

■ 

1994  1996  1998  2000 


80000  - 

3remer  Bay 

60000  - 

40000  - 

■     \ 

.p 

20000  - 

"      / 

0  - 

"\  £ 

Year 


Figure  5 

Sardinops  sagax  BSP_DEPM  estimates  Icircles  with  error  bars) 
and  Bsp  qptimai  estimates  (squares  with  accompanying  95rr 
confidence  interval  boundaries!  at  four  regions  of  southwestern 
Australia.  In  each  case  the  confidence  intervals  that  extend 
below  the  .v-axis  have  not  been  shown. 


Bspoptimal  estimates.  Notwithstanding  this,  the  CIs  for 
the  optimal  estimates  always  encompassed  the  DEPM 
point  estimates.  Because  the  CIs  were  so  broad  in  rela- 
tion to  the  point  estimates,  only  the  point  estimates  for 
BSP_DEPM  and  BSP  0PTIMAL  are  further  compared. 

The  estimated  Bsp .optimal  indicates  that  for  Albany 
the  spawning  biomasses  were  underestimated  in  1992 
and  1999  and  overestimated  in  1997  and  that  the  differ- 
ence between  estimates  in  each  case  was  greater  than 
25%  (Fig.  5,  Table  4).  Although  the  DEPM  estimated 
that  the  Albany  Bsp  remained  steady  between  1995 
(17,544  t)  and  1997"  (18,597  t),  the  PPS  almost  halved 
from  0.33  to  0.19  for  these  same  surveys  (Tables  1  and 
2).  For  Bremer  Bay  the  estimates  for  Bsp  DEPM  and  the 
Bsp-optimal  were  within  20%  (Fig.  5,  Table  4).  In  Esper- 


ance  the  B 


SP  OPTIMAL 


estimate  indicates  that  the  DEPM 


underestimated  Bsp  by  199r  in  1994,  but  overestimated 
Bsp  by  37%  in  1999. 

The  DEPM  estimates  of  Sardinops  Bsp  on  the  west 
coast  had  the  poorest  fit  against  PPS.  Thus,  the  optimal 
estimates  of  BSP  differed  by  >30%  in  four  of  the  five 
DEPM-based  Bsp  estimates.  In  particular,  the  1994  and 
1999  DEPM  estimates  were  too  low.  and  those  for  1996 
and  1998  were  too  high. 


Discussion 

Egg  presence-absence  analysis,  i.e.,  proportion  of  positive 
stations  (PPS),  was  used  to  objectively  assess  changes 
in  the  spawning  area  of  Sardinops  along  the  south  and 
lower  west  coasts  of  WA  between  1991  and  1999.  The 


630 


Fishery  Bulletin  102(4) 


60000  - 

Espe 

ranee 

50000  - 

40000  ■ 

30000  ■ 

. 

■. 

• 

20000  ■ 

10000  ■ 
0  ■ 

■ 

1992        1994        1996        1998        2000 


aoooo  - 

. 

West  Coas 

60000  - 

40000  - 

■     V 

■ 

20000  - 

\ 

y\ 

■ 

0  - 

/     \ 

i 

1994        1996 

Year 


Figure  5  (continued) 


collapse  in  distribution  was  evident  in  1999  for  three  of 
the  four  regions  examined  and  has  been  attributed  to  a 
combination  of  fishing  mortality,  several  years  of  poor 
recruitment,  and  two  mass  mortality  events  (Murray 
and  Gaughan,  2003).  The  spawning  stock  in  Albany  and 
Bremer  Bay  decreased  to  a  point  where  the  annual  total 
allowable  catch  (TAC)  in  these  fisheries  was  reduced  to 
zero.  The  concurrent  decreases  in  Bsp _DEPM  and  PPS  at 
the  south  coast  regions  in  1999,  estimated  shortly  after 
the  progression  of  an  epidemic  mass  mortality  (Gaughan 
et  al.,  2000),  indicates  a  positive  relationship  between 
Bsp-depm  and  PPS.  This  widespread  response  provides 
support  for  the  concept  of  using  the  PPS-BSP  DEPM  rela- 
tionship to  objectively  detect,  albeit  retrospectively, 
particularly  suspect  estimates  of  Bsp_DEPM. 

The  marked  decline  in  Bsp  in  1999  to  a  very  low 
level  at  Albany  provided  sufficient  contrast  in  the  time 
series  of  data  to  allow  detection  of  an  overestimation  of 
spawning  biomass  in  1997.  Although  the  difference  may 
not  appear  to  be  overly  large,  the  critical  factor  in  this 
particular  case  is  that  the  Bsp  of  18,597  metric  tons  (t) 


was  seen  to  be  healthy,  whereas  an  estimate  of  13,660  t 
would  have  clearly  indicated  to  the  Management  Advi- 
sory Committee  a  downward  trend  in  Sar-dmops  Bsp.  In 
turn,  such  a  result  would  have  supported  the  contention 
that  the  stock  was  in  decline,  which  was  expected  be- 
cause of  several  years  of  poor  recruitment,  as  evidenced 
by  catch-at-age  data  (Gaughan  et  al.,  2002).  Further- 
more, in  1998,  during  the  6  months  prior  to  the  mass 
mortality,  the  purse-seine  fleet  in  Albany  experienced 
significant  difficulties  in  meeting  catch  expectations, 
which  also  indicated  that  the  stock  was  at  a  low  level. 
Although  we  cannot  assess  precision  of  the  revised  es- 
timates of  Bsp,  it  is  likely  that  the  Bsp-OPTIMAI  ^01"  1997 
still  overestimates  the  actual  stock  size. 

The  evidence  for  a  decline  in  Bsl,  at  Bremer  Bay  from 
1994  to  1999,  as  suggested  by  the  decline  in  PPS,  was 
supported  by  trends  in  catch  curves  for  that  period, 
which  showed  very  low  levels  of  recruitment  (Gaughan 
et  al.,  2002).  The  recruitment  trends  ensured  that  the 
annual  TACs  for  Bremer  Bay  after  the  mid  1990s  did 
not  increase  but  instead  were  gradually  reduced.  The 


Gaughan  et  al.:  Distribution  of  Sardmops  sagax  off  southwestern  Australia 


631 


Table  4 

Comparison 

of  estimates  of  spawning 

biomass  for  Sardi 

nops  sagax  from  four  regions 

in  southwest 

era 

Aust 

ralia.  Estimates 

obtained  by 

using 

the  daily  egg  produc 

ion  method  (DEPM)  were  re-estimated  by 

using 

a  model  that 

considered  the  proportion 

of  positive  stations 

iPPS.  see  text  and  Table  2)  during  each  of  the  DEPM  surveys. 

In  turn,  a  weighted 

or 

optimal,  estimate  was 

derived  from 

the  p 

•evious  two  estimates.  The  difference  and  ratio  between  the  optimal  and  the  DEPM  esti 

mates  are  provided  for 

comparison. 

Optimal  estimates  that  fall  outside  the  95ri  confidence  intervals  for  the  DEPM  estimates 

and  optimal:DEPM  ratios 

that  differ  from  1  by  greater  than  0.25 

are  shown  in  bold  type. 

DEPM  estimate 

PPS  estimate 

Optimal  estimate 

DEPM-PPS  estimate 

OptimahDEPM 

Albany 

1991 

19,300 

20,190 

20,209 

-890 

1.05 

1992 

16,994 

25,456 

21,811 

-8462 

1.28 

1993 

23,432 

22,823 

23,653 

609 

1.01 

1994 

31.330 

26,334 

29.438 

4996 

0.94 

1995 

17,544 

14.484 

16,347 

-1197 

0.93 

1997 

18,597 

8339 

13.660 

10,258 

0.73 

1999 

89 

6584 

3488 

-6495 

39.19 

Bremer  Bay 

1992 

19.280 

21,346 

22,689 

-2066 

1.18 

1993 

44,010 

25,195 

37.407 

-6603 

0.85 

1994 

28,458 

24.495 

29,204 

3963 

1.03 

1999 

4156 

4199 

4645 

-43 

1.12 

Esperance 

1993 

32,252 

23,542 

28,220 

-4032 

0.88 

1994 

20,080 

26,838 

23,827 

-6758 

1.19 

1995 

31.900 

28,721 

30,705 

3179 

0.96 

1999 

17.396 

4238 

10.875 

13,158 

0.63 

West  coast 

1993 

41,250 

48,318 

43.048 

-7068 

1.04 

1994 

8714 

27,350 

17,049 

-18,636 

1.96 

1996 

60,228 

20,968 

39,845 

39,260 

0.66 

1998 

18,985 

9117 

13,723 

9868 

0.72 

1999 

5725 

10.940 

7714 

-5665 

1.35 

very  poor  fit  for  Esperance  may  reflect  the  low  sample 
size  or  may  be  indicative  of  a  certain  level  of  decoupling 
of  BSP  and  PPS  not  evident  in  the  other  south  coast 
regions. 

The  1996  estimate  for  the  west  coast  was  hampered 
by  poor  estimation  of  adult  parameters  resulting  from 
a  low  number  of  adult  samples  obtained;  the  Bsp  DEPM 
estimate  for  that  year  appeared  to  be  much  too  high 
and,  intuitively,  was  not  used  as  the  basis  for  making 
management  decisions  at  that  time.  The  precautionary 
decision  to  use  the  lower  bound  rather  than  the  "best" 
estimate  from  the  1996  west  coast  DEPM  survey  was 
therefore  justified.  In  contrast,  the  estimate  of  B^pDEP!il 
of  8714  t  in  1994  for  the  west  coast  Sardmops  stock 
appears  to  have  been  too  low.  The  lack  of  an  obvious 
collapse  in  distribution  off  the  west  coast  was  partly 
due  to  the  marked  changes  in  the  intensity  and  dis- 
tribution of  sampling  after  1996.  Another  contributing 
factor  may  have  been  a  change  in  the  distribution  of 
the  spawning  adults  because  of  the  anomalously  warm 
water  in  the  Indian  Ocean  in  the  late  1990s  (Webster 
et  al.,  1999)  during  the  last  major  La  Nina.  The  PPS 


of  only  0.10  in  1998,  before  the  epidemic  mortality,  may 
therefore  have  been  the  result  of  behaviorally  mediated 
changes  in  the  distribution  of  Scuxlmops  in  response  to 
the  warmer  than  average  water  temperatures  (Gaughan 
et  al.,  2000).  We  recognize  that  other  factors  may  also 
have  influenced  the  distribution  of  Sardinops  off  the 
west  coast  but  our  relatively  short  time  series  of  data 
precluded  development  of  more  definitive,  alternative 
hypotheses  at  this  time.  The  potential  for  unusual  en- 
vironmental conditions  to  influence  spawning  behavior 
applies  equally  to  the  south  coast  Sardinops;  interpreta- 
tion of  PPS  data  therefore  also  requires  consideration 
of  environmental  conditions  in  each  case.  As  our  time 
series  of  biomass  estimates  is  extended  through  further 
DEPM  surveys,  hypotheses  regarding  the  influence  of 
the  environment  will  be  further  developed.  Prelimi- 
nary hypotheses  have  already  been  presented  to  the 
Management  Advisory  Committee  and  thus  form  part 
of  current  management  deliberations. 

The  results  from  this  retrospective  analysis  will  im- 
mediately be  used  to  reassess  the  Bsp  estimates  ob- 
tained for  Sardinops  in  WA  during  the  1990s  before 


632 


Fishery  Bulletin  102(4) 


refitting  them  to  Hall's  (2000)  integrated  models  for 
the  three  adult  assemblages  on  the  south  coast  of  WA. 
The  integrated  model  is  tuned  with  BSPDEPM  estimates. 
Therefore,  replacing  BSPDEPM  estimates  with  BSP_0PTI_ 
MAL  estimates  will  result  in  a  model  that  better  simu- 
lates the  size  of  the  Sardinops  stocks  off  southern  WA. 
Although  the  changes  may  appear  trivial,  it  is  impor- 
tant that  re-estimating  the  most  deviant  estimates  of 
Bspdepm  can  De  undertaken  in  a  manner  that  satisfies 
demands  by  stakeholders,  including  industry,  for  open- 
ness and  clarity  in  the  provision  of  scientific  advice. 

As  further  DEPM  surveys  are  conducted  to  assess  the 
status  of  the  Sardinops  stocks  in  the  five  to  six  years 
following  the  1998-99  mortality  event,  more  reliable  re- 
lationships between  PPS  and  Bsp  DEPM  will  be  developed. 
To  assist  this  process,  the  relative  merits  of  the  data 
for  individual  DEPM  surveys  can  also  be  re-examined, 
particularly  those  data  that  this  study  has  indicated  to 
have  resulted  in  poor  estimates  of  BSPDEPM.  An  ongoing 
iterative  approach  that  employs  retrospective  analyses 
will  be  undertaken  in  an  attempt  to  continuously  re- 
duce the  variance  of  the  PPS-BSP _DEPM  relationship. 
This  approach  will  permit  further  refinement  of  Hall's 
(2000)  integrated  model,  a  process  already  in  prog- 
ress (Stephenson  et  al.5)  and  will  therefore  contribute 
to  increased  confidence  in  the  scientific  advice  that  is 
provided  for  management  of  the  Sardinops  fisheries  in 
WA.  Eventually,  PPS  alone  may  be  sufficient  to  provide 
an  indication  of  spawning  biomass  with  an  acceptable 
level  of  precision. 

Besides  contributing  to  the  integrated  model,  the  BSP 
optimal  point  estimates  obtained  over  nearly  a  decade  in 
each  of  four  management  regions  now  provide  a  clearer 
indication  of  potential  maximum  biomass  levels  against 
which  industry  members  can  plan  their  businesses.  Be- 
cause of  the  highly  variable  recruitment  of  many  small 
pelagic  fish,  purse-seine  businesses  that  target  fish  such 
as  Sardinops  should  not  invest  at  levels  that  require  an 
economic  return  based  on  maximum  biomass  sizes.  For 
the  purse-seine  fishing  zones  in  southern  WA  the  maxi- 
mum spawning  biomass  from  which  purse-seine  indus- 
try members  can  expect  their  TAGS  to  be  determined 
are  as  follows:  west  coast  40,000  t,  Albany  29,000  t, 
Bremer  Bay  37,000  t,  and  Esperance  30,000  t.  Although 
these  values  provide  an  upper  limit  to  business  plan- 
ning, maximum  biomasses  should  not  represent  invest- 
ment targets.  These  values  provide  an  indication  of 
the  maximum  size  for  the  industry  but,  because  of  the 
"natural  and  social  disarray"  that  can  result  "from  har- 
vesting marine  fish  species  at  the  crest  of  their  produc- 
tion" (Smith  2000),  the  industry  should  be  structured 
at  a  level  that  focuses  on  longer-term  average  biomass 
and  that  includes  industry's  ability  to  survive  during 
periods  of  low  stock  size.  Maximum  and  average  Bsp 
for  Sardinops  at  each  of  the  four  management  regions 


Stephenson,  P.,  N.  Hall,  and  D.  Gaughan.  2004.  Unpubl. 
data.  Department  of  Fisheries.  Western  Australian  Ma- 
rine Laboratories,  North  Beach,  Western  Australia  6920, 
Australia. 


in  southwestern  Australia  will  be  further  investigated 
during  ongoing  development  of  the  integrated  model  and 
as  more  information  becomes  available. 


Conclusion 

Even  large  numbers  of  plankton  samples  can  result  in 
imprecise  estimates  of  egg  production  for  use  in  DEPM 
calculations  (e.g.,  Mangel  and  Smith,  1990).  Relative 
trends  in  spawning  area  that  can  be  obtained  from  the 
same  survey  by  using  egg  presence-absence  analysis 
provide  a  secondary  means  of  assessing  trends  in  the 
status  of  stocks.  This  egg  presence-absence  analysis 
will  be  particularly  useful  for  stocks  already  assessed 
by  using  DEPM  surveys  and  more  so  for  those  that  do 
not  have  large  amounts  of  ancillary  information,  such 
as  long  time-series  of  catch-at-age  data,  or  meaningful 
effort  data. 

Detection  of  either  upwards  or  downwards  bias  in 
estimates  of  Bsp  will  be  considered  in  the  integrated 
model  and  also  communicated  to  industry  members  to 
increase  their  understanding  of  the  stock  in  each  region. 
Although  this  review  of  biomass  trends  of  Sardinops 
during  the  1990s  cannot  change  how  Sardinops  were 
managed  during  that  period,  an  increased  understand- 
ing of  both  the  stock  sizes  and  the  science  behind  the 
biomass  assessments  will  facilitate  ongoing  manage- 
ment processes. 


Acknowledgments 

The  authors  sincerely  thank  all  Department  of  Fish- 
eries staff  involved  with  the  collection  and  analysis 
of  historical  ichthyoplankton-survey  data,  in  particu- 
lar Stuart  Blight,  Gary  Buckenara,  Cameron  Dawes- 
Smith,  Rick  Fletcher,  Kieren  Gosden,  Matt  Robinson, 
Rob  Tregonning,  Ken  White,  Bruce  Webber,  and  other 
personnel  on  PV  Baudin  and  PV  McLaughlan.  We  also 
thank  Kevern  Cochrane  (FAO)  for  a  detailed  review  of 
an  earlier  version  of  this  manuscript.  We  are  grateful 
to  the  Fisheries  Research  and  Development  Corporation 
(Canberra)  that  provided  funding  through  Project  92/25 
for  the  earlier  DEPM  surveys. 


Literature  cited 

Alheit.  J. 

1993.     Use  of  the  daily  egg  production  method  for  estimat- 
ing biomass  of  clupeoid  fishes:  a  review  and  evaluation. 
Bull.  Mar.  Sci.  53:750-767. 
Fletcher,  W.  J. 

1991.  A  test  of  the  relationship  between  otolith  weight 
and  age  fin-  the  pi Ich a rd  Sa rilin o/j.s ■  nctipili-hnrdua  Can 
J.  Fish.  Aquat.  Sci.  48:35-38. 

1995.  Application  of  the  otolith  weight-age  relationship 
for  the  pilchard,  Sardinops  sagax  neopilchardus.  Can. 
J.  Fish.  Aquat.  Sci.  52:657-664. 

1999.     Vertical  distribution  of  pilchard  tSardinops  sagax) 


Gaughan  et  al.:  Distribution  of  Sardinops  sagax  off  southwestern  Australia 


633 


eggs  and  larvae  off  southern  Australia.     Mar.  Freshw. 
Res.  50:117-122. 
Fletcher,  W.  J„  N.  C.  H.  Lo,  E.  A.  Hayes,  R.  J.  Tregonning,  and 
S.  J.  Blight. 

1996.  Use  of  the  daily  egg  production  method  to  estimate 
the  stock  size  of  Western  Australian  sardines  (Sardinops 
sagax).     Mar.  Freshw.  Res.  47:819-825. 

Fletcher  W.  J.,  and  N.  R.  Sumner. 

1999.  Spatial  distribution  of  sardine  i Sardinops  sagax) 
eggs  and  larvae:  an  application  of  geostatistics  and 
resampling  to  survey  data.  Can.  J.  Fish.  Aquat.  Sci. 
56:907-914 

Fletcher.  W.  J.,  and  R.  J.  Tregonning. 

1992.     The  distribution  and  timing  of  spawning  by  the 

Australian  pilchard  (Sardinops  sagax  neopilchardus) 

off  Albany,  Western  Australia.     Aust.  J.  Mar.  Freshw. 

Res.  46:1437-1449. 
Fletcher,  W.  J.,  R.  J.  Tregonning,  and  G.  J.  Sant. 

1994.     Interseasonal  variation  in  the  transport  of  pilchard 

eggs  and  larvae  off  southern  Western  Australia.     Mar. 

Ecol.  Prog.  Ser.  111:209-224. 
Gaughan,  D.  J„  W.  J.  Fletcher,  and  J.  P.  McKinlay. 

2002.  Functionally  distinct  adult  assemblages  within  a 
single  breeding  stock  of  the  sardine,  Sardinops  sagax: 
management  units  within  a  management  unit.  Fish 
Res.  59:217-231. 

Gaughan,  D.  J..  R.  W.  Mitchell,  and  S.  J.  Blight. 

2000.  Impact  of  mortality,  possibly  due  to  herpes  virus, 
on  pilchard  Sardinops  sagax  stocks  along  the  south 
coast  of  Western  Australia  in  1998-99.  Mar.  Freshw. 
Res.  51:601-612. 

Hall,  N.  G. 

2000.     Modelling  for  fishery  management,   utilising 

data  for  selected  species  in  Western  Australia.     Ph.D. 

diss. ,199  p.     Murdoch  Univ.,  Perth,  Australia. 
Hunter,  J.  R„  and  N.  C.  H.  Lo. 

1997.  The  daily  egg  production  method  of  biomass  estima- 
tion: some  problems  and  potential  improvements.  Oz- 
eanografika  2:41-69. 

Mangel,  M„  and  P.  E.  Smith. 

1990.     Presence-absence  sampling  for  fisheries  man- 
agement.    Can.  J.  Fish.  Aquat.  Sci.  47:1875-1887. 
Murray,  A.  G.,  and  D.  J.  Gaughan. 

2003.  Using  an  age-structured  model  to  simulate  the 
recovery  of  the  Australian  pilchard  (Sardinops  sagax) 


population  following  epidemic  mass  mortality.     Fish. 

Res.  60:415-426. 
Parker,  K. 

1985.     Biomass  model  for  the  egg  production  model.     In  An 

egg  production  method  for  estimating  spawning  biomass 

of  pelagic  fish:  application  to  the  northern  anchovy. 

Engraulis  mordax  (R.  M.  Lasker,  ed.i,  p.  5-6.     NOAA 

Tech.  Rep.  NMFS  36. 
Picquelle.  S..  and  G.  Stauffer. 

1985.  Parameter  estimation  for  an  egg  production  method 
of  anchovy  biomass  assessment.  In  An  egg  production 
method  for  estimating  spawning  biomass  of  pelagic  fish: 
application  to  the  northern  anchovy,  Engraulis  mordax 
(R.  M.  Lasker,  ed.i,  p.  7-15.  NOAA  Tech.  Rep.  NMFS 
36. 

Smith,  P.  E. 

2000.  Pelagic  fish  early  life  history:  CalCOFI  overview.  In 
Fisheries  oceanography:  an  integrated  approach  to  fish- 
eries ecology  and  management  (P.  J.  Harrison  and  T.  R. 
Parsons,  eds.l,  p.  8-28.     Blackwell  Science.  Oxford,  UK. 

Smith.  P.  E„  H.  Santander.  and  J.  Alheit. 

1989.     Comparison  of  the  mortality  rates  of  pacific  sar- 
dine, Sardinop  sagax.  and  Peruvian  anchovy,  Engraulis 
ringens,  eggs  off  Peru.     Fish.  Bull.  87:497-508. 
Uphoff,  J.  H. 

1993.     Determining  striped  bass  spawning  stock  status 
from  the  presence  or  absence  of  eggs  in  ichthyoplankton 
survey  data.     N.  Am.  J.  Fish.  Manag.  13:645-657. 
Ward  T.  M.,  F.  Hoedt,  L.  McLeay,  W.  F.  Dimmlich,  M.  Kinloch, 
G.  Jackson,  R.  McGarvey.  P.  J.  Rogers,  and  K.  Jones. 

2001.  Effects  of  the  1995  and  1998  mass  mortality 
events  on  the  spawning  biomass  of  sardine,  Sardinops 
sagax,  in  South  Australian  waters.  ICES  J.  Mar.  Sci. 
58(4):865-875. 

Webster  P.  J.,  A.  M.  Moore,  J.  P.  Loschnigg,  and  R.  R.  Leben. 

1999.  Coupled  ocean-atmosphere  dynamics  in  the  Indian 
Ocean  during  1997-98.     Nature  401:356-360. 

Wolf,  P..  and  P.E.Smith. 

1986.  The  relative  magnitude  of  the  1985  Pacific  sardine 
spawning  biomass  off  southern  California.  CalCOFI 
Rep.  27:25-31. 

Zenitani.  H.,  and  S.  Yamada. 

2000.  The  relation  between  spawning  area  and  biomass 
of  Japanese  pilchard,  Sardinops  melanostictus,  along  the 
Pacific  coast  of  Japan.     Fish.  Bull.  98:842-848. 


Appendix  1— Derivation  of  model  III  from  model  II 

We  start  with  model  II 


"SP_DEPM  ~  K,  )""•->     +  er 


For  brevity,  y  is  used  to  denote  PPS.  We  first  note  that 
y'~=yy"'~1'-  Now  we  obtain  the  Taylor  series  expansion  of 
yU-D  about. v=l  giving 


Because  y  is  a  proportion,  it  satisfies  0<y<l  so  that 
the  higher  powers  of  {y— 1)  will  be  individually  and  col- 
lectively small.  If  y  is  close  to  0.5,  a  further  algebraic 
simplification  of  the  second  term  is  possible,  giving  the 
identity 

v(y-l)  =  -0.25  +  (v-0.5)2. 


vi;'"=  1  +  (A-1)(v-1)-mA-1)(A-2i/2'<y-I): 
+(A-l)(A-2)(A-3)/3!(.v-l)-'  +  ...  . 

Multiplying  this  expression  through  by  y  gives 

yl  =  y  +  _y(_v-l)(A-l) 
+  ytterms  involving  second  and  higher  powers  of  t  v— 1 )]. 


When  v  is  in  the  range  0.25<v<0.75,  the  right-hand  side 
of  this  identity  remains  close  to  -0.25.  Thus  model  II 
may  be  simplified  to 

Bsp  pps  =  a„(PPS)  -  0.25a„(  A-l )  +  <?, 

which  is  the  form  of  model  III. 


634 


Abstract— Understanding  recoloni- 
zation  processes  of  intertidal  fish 
assemblages  is  integral  for  predict- 
ing the  consequences  of  significant 
natural  or  anthropogenic  impacts  on 
the  intertidal  zone.  Recolonization  of 
experimentally  defaunated  intertidal 
rockpools  by  fishes  at  Bass  Point,  New 
South  Wales  I  NSW),  Australia,  was 
assessed  quantitatively  by  using  one 
long-term  and  two  short-term  studies. 
Rockpools  of  similar  size  and  position 
at  four  sites  within  the  intertidal  zone 
were  repeatedly  defaunated  of  their 
fish  fauna  after  one  week,  one  month, 
and  three  months  during  two  short- 
term  studies  in  spring  and  autumn  i  5 
months  each),  and  every  six  months 
for  the  long-term  study  (12  months). 
Fish  assemblages  were  highly  resil- 
ient to  experimental  perturbations — 
recolonizing  to  initial  fish  assemblage 
structure  within  1-3  months.  This 
recolonization  was  primarily  due  to 
subadults  (30-40  mm  TL)  and  adults 
i >40  mm  TL)  moving  in  from  adjacent 
rockpools  and  presumably  to  abun- 
dant species  competing  for  access  to 
vacant  habitat.  The  main  recolonizers 
were  those  species  found  in  highest 
numbers  in  initial  samples,  such  as 
Bathygobius  cocosensis,  Enneaptery- 
gius  rufopileus,  and  Girella  elevata. 
Defaunation  did  not  affect  the  size 
composition  of  fishes,  except  during 
autumn  and  winter  when  juveniles 
(<30  mm  TL)  recruited  to  rockpools. 
It  appears  that  Bass  Point  rockpool 
fish  assemblages  are  largely  con- 
trolled by  postrecruitment  density- 
dependent  mechanisms  that  indicate 
that  recolonization  may  be  driven  by 
deterministic  mechanisms. 


Fish  recolonization  in  temperate 

Australian  rockpools: 

a  quantitative  experimental  approach 

Shane  P.  Griffiths 

Environmental  Science 

and 

Institute  for  Conservation  Biology 
University  of  Wollongong 
Wollongong,  New  South  Wales,  Australia 
Present  address:  CSIRO  Marine  Research 

233  Middle  Street 

Cleveland,  Queensland  4163  Australia 
Email  address:  shanegnffiIhsacsiro.au 

Ron  J.  West 

Environmental  Science 

University  of  Wollongong 

Wollongong,  New  South  Wales,  Australia 

Andy  R.  Davis 

Institute  for  Conservation  Biology 

University  of  Wollongong 

Wollongong,  New  South  Wales,  Australia 

Ken  G.  Russell 

School  of  Mathematics  and  Applied  Statistics 

University  of  Wollongong 

Wollongong,  New  South  Wales,  Australia 


Manuscript  submitted  28  April  2003 
to  thr  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
5  May  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:634-647(2004). 


Rocky  intertidal  fishes  are  faced  with 
many  biotic  (competition  and  food 
availability)  and  abiotic  (temperature 
and  salinity)  factors  that  can  influ- 
ence their  distribution  and  abundance 
(Gibson.  1982).  Despite  occupying  a 
dynamic  environment,  the  fish  assem- 
blages in  intertidal  rockpools  have 
been  widely  shown  to  remain  persis- 
tent through  time  (Grossman.  1982. 
1986;  Collette.  1986).  These  commu- 
nities can  also  rapidly  return  to  their 
original  state  after  major  or  even 
catastrophic  perturbations  (Moring, 
1996).  Such  resilience  is  less  common 
among  assemblages  of  invertebrates 
(Connell,  1972;  Astles,  1993)  because 
recolonization  of  substrata  is  normally 
dependent  upon  successful  larval 
settlement  (Paine  and  Levin,  1981). 
In  contrast,  fish  can  rapidly  colonize 
available  habitat  by  larval  recruit- 


ment from  the  plankton  (Willis  and 
Roberts,  1996.  Beckley,  2000;  Griffiths 
2003a)  but  also  by  the  relocation  of 
subadults  and  adults  from  adjacent 
rockpools  (Beckley,  1985a;  Griffiths, 
2003a).  Under  natural  conditions  rock- 
pools can  be  defaunated  by  events  such 
as  hurricanes  (Moring,  1996)  and,  in 
some  regions,  by  seasonal  freezing  of 
rockpool  water  (Thomson  and  Lehner, 
1976;  Moring,  1990).  These  events 
can  create  new  microhabitats  or  open 
existing  ones  for  fish  to  colonize,  and 
therefore  have  the  potential  to  change 
fish  assemblage  structure. 

Understanding  recolonization  pro- 
cesses of  intertidal  fish  assemblages 
is  integral  for  predicting  the  conse- 
quences of  natural  or  anthropogenic 
impacts  on  the  intertidal  community. 
The  role  of  disturbance  and  recoloni- 
zation processes  in  structuring  inter- 


Griffiths  et  al.:  Fish  recolonization  in  temperate  Australian  rockpools 


635 


Tasman  Sea 

•^Shellharbniir 
Bass  Point 


5  km 


Figure  1 

Map  illustrating  the  four  sampling  sites  at  Bass  Point  and  the  location  of  the  study  loca- 
tion in  the  Illawarra  region.  New  South  Wales,  Australia. 


tidal  rockpool  fish  assemblages  has  received  considerable 
attention  in  many  countries  of  the  world  (Bussing,  1972; 
Matson  et  al.,  1986;  Yoshiyama  et  al.,  1986;  Prochazka 
and  Griffiths,  1992;  Lardner  et  al.,  1993;  Prochazka, 
1996;  Faria  and  Almada,  1999;  Silberschneider  and 
Booth,  2001).  Such  studies  have  identified  patterns  in 
the  rates  of  recovery,  variation  in  species  and  size  com- 
position of  recolonizing  fish  assemblages  (Polivka  and 
Chotkowski,  1998;  Beckley,  2000),  and  homing  abilities 
of  many  intertidal  fishes  (Green  1971;  Yoshiyama  et  al., 
1992;  Griffiths.  2003b). 

Rockpools  can  be  regarded  as  "island'  habitats  (Under- 
wood and  Skilleter,  1996)  among  an  inhospitable  rocky 
landscape.  Therefore,  there  is  probably  a  balance  be- 
tween immigration  (recruitment  and  relocation)  and 
emigration  (mortality)  of  fishes  after  a  disturbance, 
sensu  the  equilibrium  theory  of  island  biogeography 
(MacArthur  and  Wilson,  1967).  After  a  period  of  time, 
the  number  of  species  and  individuals  in  a  defaunated 
rockpool  can  be  expected  to  reach  an  asymptote  when  a 
carrying  capacity  is  reached.  It  is  difficult  to  generalize 
about  recolonization  rates  of  rockpools  by  fishes  from 
the  current  literature  mainly  owing  to  the  diversity  of 
methods  used,  their  differing  effectiveness  in  sampling 
fish,  and  the  varying  intensity  of  the  sampling  regime. 
For  example,  most  studies  have  used  only  small  sample 
sizes  (<10  pools)  and  have  sampled  at  a  range  of  time 
intervals  from  days  (Mistry  et  al.,  1989;  Matson  et  al., 
1986;  Polivka  and  Chotkowski,  1998)  to  years  (Thomson 
and  Lehner,  1976;  Lardner  et  al.,  1993;  Mahon  and 
Mahon,  1994).  A  second  problem  in  measuring  and 
comparing  fish  recolonization  patterns  between  stud- 
ies is  that  many  researchers  have  sampled  fish  using 
an  anesthetic  (Mahon  and  Mahon.  1994;  Pfister,  1995, 
1997)  or  ichthyocide  (Beckley,  1985a,  1985b,  2000;  Wil- 


lis and  Roberts,  1996;  Silberschneider  and  Booth,  2001), 
which  may  affect  subsequent  catches  (Yoshiyama  et 
al.,  1986)  and  possibly  result  in  fish  assemblages  never 
reaching  preperturbation  conditions  (see  Mok  and  Wen, 
1985;  Lockett,  1998). 

Nonetheless,  recolonization  of  rockpools  by  fishes  is 
generally  a  rapid  process,  beginning  within  days,  or 
even  hours,  after  defaunation  (Collette,  1986).  and  com- 
plete recolonization  to  preperturbation  levels  can  take  a 
few  weeks  (Collette,  1986;  Faria  and  Almada.  1999)  to 
several  months  (Mok  and  Wen.  1985;  Willis  and  Rob- 
erts, 1996;  Polivka  and  Chotkowski,  1998). 

The  aims  of  this  study  were  to  quantitatively  deter- 
mine 1)  the  period  required  for  intertidal  rockpools  to 
recover  to  preperturbation  levels,  2)  the  fish  species 
(permanent  residents,  opportunist,  or  transients)  respon- 
sible for  recolonizing  rockpools,  3)  whether  recoloniza- 
tion patterns  differ  between  the  four  sites  at  Bass  Point 
and  between  the  times  of  year  when  defaunation  took 
place,  and  4)  whether  fish  comprise  different  life-history 
stages  before  and  after  a  disturbance  (sampling) — by 
examination  of  length-frequency  distributions. 


Methods 

Study  site  and  experimental  design 

Spatial  and  temporal  variation  in  fish  recolonization  pat- 
terns were  investigated  in  three  separate  studies  under- 
taken along  the  north-  and  south-facing  rocky  platforms 
at  Bass  Point  (34°58'S,  150°93'E),  New  South  Wales, 
Australia  (Fig.  1).  Bass  Point  is  a  large  rocky  headland 
that  extends  approximately  3  km  into  the  Tasman  Sea. 
Two  short-term  recolonization  studies  (each  around  5 


636 


Fishery  Bulletin  102(4) 


months  in  duration)  were  undertaken  in  spring-summer 
and  autumn-winter  (hereafter  referred  to  as  spring  and 
autumn  studies,  respectively),  and  a  long-term  recolo- 
nization  study  spanned  a  12-month  period.  Rockpools 
for  each  of  the  three  studies  were  selected  at  four  sites 
at  Bass  Point,  NSW,  which  are  named  Maloney's  Bay 
(MB),  The  Chair  (TO,  Gravel  Loader  (GL),  and  Beaky 
Bay  (BB)  (Fig.  1).  Each  of  the  four  sites  are  separated 
by  about  1  km.  Rockpools  were  selected  at  each  site 
(50-200  m  apart)  according  to  similar  physical  param- 
eters (i.e.,  volume,  surface  area,  and  substrate  type) 
and  particularly  according  to  their  vertical  elevation 
on  the  rock  platform.  Because  higher  pools  might  have 
less  chance  of  fish  recolonization  because  they  are  less 
frequently  inundated  by  seawater  (Griffiths  et  al.,  2003). 
every  effort  was  made  to  select  pools  located  in  the  mid- 
intertidal  zone  (1-1.5  m  above  MLLW  [mean  lower  low 
water])  and,  although  pools  were  visually  similar,  they 
varied  in  volume,  ranging  from  762  to  2160  liters  (or 
0.76-2.16  m3).  The  bottom  of  the  rockpools  consisted  of 
pebbles,  cobbles,  and  small  boulders. 

For  the  short-term  studies,  four  rockpools  were  sam- 
pled and  fish  removed  at  each  of  the  four  sites.  In  the 
spring  study  (beginning  7  September  1999),  they  were 
then  resampled  1  week,  1  month,  and  3  months  af- 
ter the  preceding  sampling  date  (referred  to  as  the 
"1-week,"  "1-month,"  and  "3-month"  samples  in  this 
article).  This  study  ended  on  8  February  2000,  after  a 
period  of  5  months.  After  this  date  a  period  of  at  least 
three  months  was  given  for  pools  to  re-establish  fish 
assemblages  before  beginning  the  autumn  study  on  15 
May  2000.  Rockpools  were  sampled  in  exactly  the  same 
manner  as  for  the  spring  study,  with  sampling  ending 
on  17  September  2000.  For  each  study,  64  samples  were 
taken  giving  a  total  of  128  samples  for  the  short-term 
studies.  It  is  important  to  note  that  although  every  ef- 
fort was  made  to  resample  pools  after  exactly  the  same 
time  intervals,  this  was  not  possible  because  of  daily 
time  and  height  of  tides  and  wave  heights.  For  example, 
for  the  "1  week"  samples,  the  number  of  days  between 
samples  was  actually  between  7  and  10  days. 

To  determine  whether  frequent  sampling  in  the  short- 
term  studies  affected  the  structure  of  rockpool  fish  as- 
semblages, a  long-term  study  was  undertaken  by  using 
four  different  rockpools  at  the  same  four  sites  that  were 
sampled  in  the  short-term  studies.  Four  rockpools  at 
each  site  were  considered  adequate  because  Griffiths 
(2003a)  was  able  to  detect  significant  differences  in  the 
numbers  of  fish  species  and  individuals  in  rockpools 
between  sites  and  months  using  four  rockpools  per  site 
in  the  same  region  that  was  surveyed  in  our  study. 
Rockpools  were  initially  sampled  on  22  September  1999 
and  then  resampled  twice  at  intervals  of  six  months 
(20  April  2000  and  11  September  2000).  A  total  of  48 
samples  were  taken  for  this  study. 

Data  collection 

Fish  were  collected  by  hand  after  completely  emptying 
each  rockpool  with  a  VMC  12V  battery-powered  bilge 


pump  of  9029  L/h  capacity  by  using  the  methods  of 
Griffiths  (2000).  A  thorough  search  of  each  pool  was 
conducted  by  overturning  all  rocks  and  boulders,  search- 
ing all  crevices  and  shaking  algal  fronds  until  all  fish 
that  could  be  seen  were  removed.  Fishes  were  identified 
and  total  lengths  (TL)  were  measured.  Fork  length  (FL) 
was  also  measured  for  economically  significant  species. 
Fish  were  categorized  as  being  juveniles  (<30  mm), 
subadults  (30-40  mm),  or  adults  (>40  mm).  Fish  were 
then  released  alive  into  rockpools  or  the  shallow  subtidal 
10-30  m  away  from  the  rockpool  being  sampled,  which 
was  considered  to  be  the  approximate  distance  that  fish 
may  be  displaced  by  waves  and  surge  during  significant 
natural  disturbances,  such  as  storms.  Each  species  was 
categorized  by  its  residential  status  in  rockpool  habitats 
according  to  the  definitions  of  Griffiths  (2003c)  in  order 
to  better  understand  the  types  of  fish  responsible  for 
recolonization.  These  categories  were  "permanent  resi- 
dents," "opportunists,"  and  "transients." 

Statistical  analyses 

A  repeated-measures  ANOVA  (RM-ANOVA)  was  used 
(SPSS  vers.  6.1;  SPSS,  Chicago,  IL)  to  test  for  sig- 
nificant differences  in  the  numbers  of  species  and  indi- 
viduals between  sampling  intervals  (within-subjects 
factor)  and  sites  (among-subjects  factor).  Short-  and 
long-term  experiments  were  analyzed  with  two  sepa- 
rate RM-ANOVAs.  For  the  short-term  study  a  third 
factor  of  season  (i.e.,  spring  or  autumn;  among-sub- 
jects factor)  was  added.  All  factors  were  considered 
fixed.  Assumption  of  sphericity  of  the  variance-covari- 
ance  matrix  was  tested  by  using  Mauchly's  criterion 
and,  if  violated,  F  tests  were  performed  with  Green- 
house-Geisser-adjusted  degrees  of  freedom.  Student- 
Newman-Keuls  iSNK)  tests  were  used  for  a  posteriori 
comparisons  among  means  (numbers  of  species  and 
individuals)  in  RM-ANOVAs. 

Nonmetric  multidimensional  scaling  (nMDS)  was 
used  to  examine  similarities  in  fish  assemblage  struc- 
ture between  sampling  intervals  and  sites.  Data  were 
fourth-root  transformed,  to  reduce  the  influence  of 
highly  abundant  taxa,  and  a  similarity  matrix  was 
constructed  by  using  the  Bray-Curtis  similarity  coef- 
ficient (Clarke,  1993).  Stress  values  are  given  for  all 
ordination  plots;  these  values  describe  the  quality  of 
the  representation  of  multidimensional  relationships 
of  the  data  in  a  two-dimensional  plane.  Stress  factors 
of  less  than  0.2  (<0.2  is  considered  to  give  a  good  rep- 
resentation of  sample  "relatedness"  and  to  prevent  the 
prospect  of  drawing  false  inferences)  were  obtained  for 
each  ordination  (Clarke.  1993). 

Analysis  of  similarities  (ANOSIM)  was  used  to  test 
whether  fish  assemblages  in  a  priori  groups  differed  sta- 
tistically (Clarke,  1993).  Abundance  data  for  each  spe- 
cies were  pooled  for  the  four  rockpools  at  each  site  and 
time.  Each  ANOSIM  comparison  involved  generating 
4999  random  permutations  of  the  data  to  calculate  the 
probability  that  observed  differences  in  the  structure  of 
the  fish  assemblages  among  a  priori  groups  could  arise 


Griffiths  et  al.:  Fish  recolonization  in  temperate  Australian  rockpools 


637 


by  chance.  Similarity  percentages  (SIMPER)  were  used 
to  determine  which  species  were  responsible  for  differ- 
ences between  selected  groups.  This  analysis  involved 
calculating  the  average  contribution  of  each  species  in 
each  pair  of  groups  and  comparing  this  contribution  to 
the  overall  dissimilarity  of  fish  assemblages  between 
the  groups.  All  multivariate  analyses  were  carried  out 
with  PRIMER  (Plymouth  routines  in  multivariate  eco- 
logical research)  software  (version  5.2.2,  PRIMER-E 
Ltd.,  Roborough,  Plymouth,  UK). 


Results 

Composition  of  rockpool  fish  assemblages 

A  total  of  3658  fish  representing  38  species  and  19 
families  was  caught  in  176  samples  from  32  rockpools  at 
Bass  Point  between  7  September  1999  and  22  September 
2000  (Table  1),  corresponding  to  densities  of  0.5  and  19 
species/m3  (mean  4.4  [±2.9]  /m3)  and  0.5  and  80  fish/m3 
(mean  15.6  | ±14.6]  /m3),  respectively.  The  most  numeri- 
cally abundant  taxa  were  permanent  rockpool  residents 
representing  the  families  Gobiidae  iBathygobius  cocosen- 
sis),  Tripterygiidae  iEnneapterygius  rufopileus),  Clinidae 
iHeteroclinus  whiteleggi  and  H.  fasciatus),  Blenniidae 
(Pa?-ablenrjius  intermedins),  and  Gobiesocidae  {Aspasmo- 
gaster  costatus),  although  the  temporary  resident  Girella 
elevata  was  the  third  most  abundant  species.  The  ten 
most  numerically  abundant  species  represented  92%  of 
the  catch  (Table  1).  Three  species,  G.  elevata,  Scorpis 
lineolatus,  and  Myxus  elongatus,  represented  by  504 
fish  were  considered  to  be  of  economic  significance.  All 
economically  important  fishes  were  caught  as  juveniles 
in  the  rockpools  and  89%  of  the  fish  measured  less  than 
100  mm  FL. 

Numbers  of  species  and  individuals 

For  the  short-term  studies,  the  mean  number  of  species 
differed  significantly  between  sampling  intervals  and 
sites  (RM-ANOVA,  Table  2).  With  respect  to  the  site 
factor,  there  were  significantly  more  species  caught  at 
BB  than  at  the  other  three  sites  and  the  latter  three 
sites  did  not  differ  from  each  other  (SNK  test).  Only 
the  "1-week"  samples  accounted  for  significantly  fewer 
species  than  the  initial  samples  (Fig.  2).  However,  the 
mean  number  of  species  caught  in  the  "1-month"  and 
"3-month"  samples  did  not  differ  significantly  from  the 
initial  samples  at  all  sites  (Fig.  2). 

The  mean  number  of  individuals  differed  significantly 
between  sampling  intervals  and  sites,  although  there 
was  also  a  significant  time x site  interaction  (RM-ANO- 
VA, Table  2).  A  close  investigation  of  the  significant 
interval xsite  interaction,  with  primary  interest  in  the 
interval  factor,  revealed  that  the  number  of  individuals 
in  the  initial  samples  did  not  differ  significantly  from 
samples  taken  after  three  months  at  the  exposed  sites 
(MB  and  TC),  but  they  did  differ  significantly  at  shel- 
tered locations  (GL  and  BB)  (Fig.  2).  It  appeared  that  the 


Number  of  species 


£     50 


-5    30  _ 


I     10 


Number  of  individuals 


;ni1^inW 


GL 


Site 


Figure  2 

Mean  i±SE)  numbers  of  species  and  individuals  (m~3) 
caught  in  rockpools  at  Bass  Point,  New  South  Wales 
between  7  September  1999  and  17  September  2000 
during  the  short-term  recolonization  studies  (combined 
for  spring  and  autumn)  between  sampling  intervals 
separated  by  1  week,  1  month,  and  3  months.  Key  to 
sites:  Maloney's  Bay  (MB).  The  Chair  (TC  I,  Gravel  Loader 
(GL),  Beaky  Bay  <BBi. 


Loader  and  Beaky  Bay  sites  initially  supported  unusu- 
ally high  numbers  of  individuals  and  these  high  numbers 
may  have  accounted  for  significantly  fewer  individuals 
caught  in  the  subsequent  samples  (Fig.  2). 

For  the  long-term  study,  the  number  of  individuals 
significantly  differed  among  sampling  times  but  did 
not  for  number  of  species  I  RM-ANOVA,  Table  3).  The 
significant  difference  in  the  mean  number  of  individuals 
was  due  to  fewer  individuals  caught  in  the  "12-month" 
samples  when  fish  numbers  were  pooled  for  all  sites 
(SNK  test,  Fig.  3). 

Variation  in  abundance  of  major  recolonizing  species 

The  rank  abundances  of  the  numerically  dominant 
species  were  consistent  for  B.  cocosensis  and  E.  rufopi- 
leus across  all  sampling  intervals  for  all  studies,  even 
though  their  relative  abundances  varied  considerably 
(Table  4).  In  contrast,  the  ranks  of  the  least  common 
of  the  six  species,  namely  H.  whiteleggi,  P.  intermedius. 


638 


Fishery  Bulletin  102(4) 


Table  1 

Numbers  of  fish  for  each  species 

caught  from  rockpools 

at  four 

sites  at  Bass  Point 

NSW, 

during  sho 

rt-term 

(spring  and 

autumn)  and  long-term  recolonization  studies  conducted  between 

7  September  1999 

and  22 

September 

2000.  * 

=  species  of 

commercial  or  recreational  significance  lor  bothi. 

Family  and  scientific  name 

Spring  s 

-udy 

Autumn  study 

Long-term  study 

Total 

Muraenidae 

Gymnothorax  prasinus 

41 

13 

16 

70 

Gymnothorax  cribroris 

1 

— 

— 

1 

Plotosidae 

Cnidoglanis  macrocephalus 

1 

— 

— 

1 

Gobiesocidae 

Alabes  dorsalis 

— 

2 

2 

Aspasmogaster  costatus 

134 

41 

65 

240 

Aspasmogaster  liorhyncha 

5 

20 

4 

29 

Syngnathidae 

Urocampus  carinirostris 

1 

— 

— 

1 

Scorpaenidae 

Scorpaena  cardinalis 

1 

— 

— 

1 

Serranidae 

Acanthistius  ocellatus 

33 

35 

16 

84 

Epinephelus  daemelii 

— 

2 

— 

2 

Plesiopidae 

Track inops  taeniatus 

1 

— 

— 

1 

Girellidae 

Girella  elevata* 

210 

93 

74 

377 

Scorpididae 

Microcanthus  strigatus 

1 

— 

— 

1 

Scorpis  lineolatus* 

94 

10 

16 

120 

Pomacentridae 

Abudefduf vaigiensis 

1 

— 

— 

1 

Parma  microlepis 

— 

1 

— 

1 

Chironemidae 

Chironemus  marmoratus 

30 

11 

10 

51 

Mugilidae 

Myxus  elongatus 

3 

4 

— 

7 

Labridae 

Halichoeres  nebulosus 

— 

1 

— 

1 

Notolabrus  gymnogenis 

— 

1 

— 

1 

Blennidae 

Parablennius  intermedins 

102 

78 

47 

227 

htiblcnnius  melvagris 

8 

in 

4 

22 

Tripterygiidae 

Lepidoblennius  haplodactylus 

22 

26 

22 

70 

Norfolkia  clarkei 

7 

4 

3 

14 

Enneapterygius  rufopileus 

354 

188 

157 

699 

( 'linidae 

Heteroclinus  fascial  us 

59 

48 

38 

145 

Heteroclinus  nasutus 

1 

— 

— 

1 

Heteroclinus  heptaeolus 

24 

— 

1 

25 

Heteroclinus  johnstoni 

— 

2 

— 

2 

Heteroclinus  whiteleggi 

138 

66 

39 

243 

Ophiclinus  gracilis 

15 

16 

6 

37 
continued 

Griffiths  et  al.:  Fish  recolonization  in  temperate  Australian  rockpools 


639 


Table  1 

(continued) 

Family  and  scientific  name 

Spring  s 

tudy 

Aut 

umn  study 

Long-term  study 

Total 

Gobiidae 

Bathygobius  cocosensis 

583 

293 

285 

1161 

Callogobius  depressus 

6 

2 

7 

15 

Callogobius  mucosus 

1 

— 

— 

1 

Priolepis  cincta 

1 

— 

— 

1 

Gobiidae  sp. 

1 

— 

— 

1 

Microdesmidae 

Gunnellichthys  monostigma 

1 

— 

— 

1 

Tetraodontidae 

Torquigener  pleurogramma 

— 

1 

— 

1 

Totals 

1880 

968 

810 

3658 

Table  2 

Results  of  repeated-measures  ANOVAs  for  significant 
differences  in  numbers  of  species  and  number  of  individ- 
uals (/m:!  I  caught  at  Bass  Point  during  two  short-term 
recolonization  studies  among  sampling  intervals  (time) 
(within-subjects  factor),  seasons  (spring  and  autumn) 
and  sites  (among-subjects  factors).  Both  numbers  of 
species  and  individuals  data  were  log1(1(.v+l)  trans- 
formed before  analysis,  which  removed  heteroscedas- 
ticity  in  the  data.  Mauchly's  criterion  for  sphericity  of 
variances  was  violated  for  number  of  species  (P=0.025); 
therefore  the  analysis  was  performed  with  Greenhouse- 
Geisser-adjusted  degrees  of  freedom.  Mean  squares 
(MS)  and  significance  levels  are  shown  and  significant 
results  are  given  in  boldface.  *  =  P<0.05;  **  =  P<0.01; 
***  =  P<0.001. 


Source 


Number  of 
species 

df  MS 


Number  of 
fish 


df        MS 


Among  subjects 
Season  tSe) 
Site(S) 
SxSe 
Residual 

Within  subjects 
TimeiT) 
TxS 
TxSe 
TxSxSe 
Residual 


1 

3 

3 

24 

2.16 
6.47 
2.16 
6.47 
51.75 


Mauchlv's  criterion  W 


14.94 
46.80* 

3.66 
9.38 

18.03** 

2.93 
2.54 
1.89 
1.41 

0.569* 


1  1292.24 

3  2782.52** 
3         63.53 

24  370.27 


3 
9 
3 
9 

72 


825.70** 
247.99 

94.86 
137.84 

79.71 

0.625 


and  A.  costatus,  varied  considerably  among  sampling 
intervals  for  each  study.  This  result  probably  reflects 
their  generally  low  abundances,  because  differences  in 


Table  3 

Results  of  repeated-measures  ANOVAs  for  significant 
differences  in  numbers  of  species  and  number  of  indi- 
viduals (/m3l  caught  at  Bass  Point  during  the  long-term 
recolonization  study  among  sampling  intervals  (within- 
subjects  factor)  and  sites  (among-subjects  factors).  Both 
numbers  of  species  and  individuals  data  were  log10(.v+l) 
transformed  before  analysis,  which  removed  heterosce- 
dasticity  in  the  data.  Mean  squares  ( MS )  and  significance 
levels  are  shown  and  significant  results  are  given  in  bold. 
**  =  P<0.01. 


Number  of 
species 


Source 


df 


MS 


Number  of 
individuals 

MS 


Among  subjects 

Site(S)                               3          86.93  176.18 

Residual  12           27.49  169.67 

Within  subjects 

TimeiT)                             2             3.26  451.38** 

TxS                                     6             4.44  111.82 

Residual  24             2.84  55.45 

Mauchly's  criterion  W                       0.622  0.705 


ranks  can  be  a  result  of  a  few  incidences  of  low  indi- 
vidual counts. 

The  mean  number  of  the  six  most  abundant  recolo- 
nizing  species  showed  considerable  variability  in  space 
and  time.  For  the  short-term  study,  densities  of  these 
species  differed  significantly  among  sites  and  among 
time  intervals  or  at  least  for  higher  order  interactions 
containing  these  effects  (Table  5).  No  definitive  conclu- 
sions could  be  made  regarding  the  effects  of  defaunation 
on  these  species  because  short-term  recolonization  pat- 
terns for  each  species  were  clearly  variable  within  and 
among  seasons  (Fig.  4).  However,  the  mean  number  of 
fish  was  generally  highest  in  initial  samples  and  low- 


640 


Fishery  Bulletin  102(4) 


Table  4 

Ranked  abundances  of  the 

six  most  abundant  species  overall  for  each  samp' 

ing  interval  in 

the  spring,  autumn,  and  long-term 

experiments.  Total  numbers  offish  caught  during  each  sampling 

occasion  from  16  rockpools 

from  four  sites  are 

shown  in 

paren- 

theses.  l=initial  samples; 

2=samples 

taken  after  1  week:  3=sa 

mples  taken 

after  1  month 

4=samp 

es  taken 

after  3  months. 

Species  having  equally  ran 

ked  abundances  are 

denoted  by  an  "= 

'  sign. 

Species 

Spring 

study 

Autumr 

study 

Long-term  study 

1 

2 

3 

4 

1 

2 

3 

4 

1 

2 

3 

Bathygobius  cocosensis 

1(153) 

K69l 

1(73) 

1(59) 

2(81i 

1  (20) 

1(58) 

1(41) 

1 (115i 

1  (114) 

1  ( 63 1 

Enneapterygius  rufopileus 

2  1 1091 

2(50) 

4(33) 

2(46) 

1  (84) 

2(12) 

2(29) 

3(25) 

2  ( 73  i 

2(43) 

2(41) 

Girella  elevata 

3(461 

3(36) 

3(40) 

3(38) 

3(341 

3  ( 10 1 

3  ( 16 1 

5H3) 

5(5) 

6(3) 

4(  12i 

Heteroclinus  whiteleggi 

6(11) 

=4(7) 

2(51i 

5(35) 

6(9)       = 

=  4(5) 

6(7l 

2(371 

6(4) 

-4  '  27  i 

6(8) 

Parablennius  intermedins 

4(27) 

=4(7) 

6(10) 

6(23) 

=  4(14)      = 

=4(5) 

4(15i 

4(23) 

4H4i 

5(14i 

3(19i 

Aspasmogaster  costatus 

5(21i 

=  4(7) 

5(13) 

4(36) 

=  4(14) 

6(3) 

5(8) 

6(7) 

3(22) 

3(34) 

5(9) 

est  in  the  1-week  samples  at  each  site  for  the  majority 
of  dominant  species.  For  the  long-term  study  only  the 
mean  number  of  B.  cocosensis  and  E.  rufopileus  differed 
significantly  among  sampling  times  (Table  5),  and  this 


E       30 


ro      25 


Number  of  species 


35  "I  Number  of  individuals 


Figure  3 

Mean  (±SE)  numbers  of  species  and  individuals  i  m'i 
caught  in  rockpools  at  Bass  Point,  New  South  Wales, 
between  22  September  1999  and  11  September  2000 
(pooled  for  all  four  sites)  during  the  long-term  recoloniza- 
t  ion  study.  Intervals  between  sampling  for  the  long-term 
study  were  six  months. 


difference  was  due  to  lower  numbers  being  caught  in 
the  12-month  samples  (Fig.  4). 

Fish  assemblage  structure 

No  clear  patterns  emerged  in  the  nMDS  ordination  plots, 
with  the  exception  of  separation  of  the  "1  week"  samples 
from  all  other  samples  at  BB  during  the  autumn  study 
(Fig.  5).  ANOSIM  supported  these  visual  interpretations 
of  ordination  plots  and  revealed  that  fish  assemblages 
did  not  differ  significantly  among  sampling  times  at  any 
of  the  four  sites  for  the  spring  and  long-term  studies 
(Table  6)  because  abundant  species  E.  rufopileus,  B.  coco- 
sensis, H.  fasciatus,  and  P.  intermedius  were  common  in 
all  samples  (SIMPER  analysis).  For  the  autumn  study, 
the  results  of  ANOSIM  complemented  those  of  RM- 
ANOVA  in  that  significant  differences  among  sampling 
intervals  were  detected  only  at  BB  (Table  6).  At  this 
site  the  initial  samples  and  "1-week"  samples  differed 
significantly  in  their  fish  assemblages,  which  was  due 
to  higher  numbers  of  G.  elevata  and  B.  cocosensis  in  the 
initial  samples  (SIMPER  analysis). 

Length-frequency  distributions 

Removal  of  fishes  from  rockpools  did  not  have  any 
apparent  effects  on  the  length-frequency  distributions 
for  at  least  two  species  (B.  cocosensis  and  E.  rufopi- 
leus) for  which  there  were  sufficient  data  to  construct 
length-frequency  histograms.  Unfortunately,  the  less 
abundant  recolonizing  species,  namely  P.  intermedius, 
A.  costatus,  and  H.  whiteleggi,  were  caught  in  too  few 
numbers  to  ascertain  the  impacts  of  defaunation  on 
their  size  compositions.  Rockpools  were  mainly  recolo- 
nized  by  subadults  and  adults  for  B.  cocosensis  and  E. 
rufopileus  in  all  three  studies  (Figs.  6  and  7).  However, 
cohorts  of  small  juveniles  (15-30  mm)  were  evident  in 
the  "3-month"  samples  during  spring  and  the  initial 
autumn  studies  (February  to  June),  which  could  then 
be  clearly  identified  in  subsequent  samples  (Figs.  6 
and  7). 


Griffiths  et  al.:  Fish  recolonization  in  temperate  Australian  rockpools 


641 


Bathygobius  cocosensis 


Enneapterygius  rulopileus 


Spring 


^-j*—  i*-!  ^r1^ 


Long-term 


Site 


I 


I 


i 


I 


Girella  elevata 


Spring 


I 


Hsf^Ft  fj 


ftA 


I 


ff^M 


\ 


5 
4 
3 

2- 
1  - 


Long-term 


[fitt 


I 


If 


MB 


I 


"Y~ 


- T- 
BB 


[~J  Initial 

H  1  week 

!\]  1  month 

H  3  months 


7J  Initial 
PI  6  months 
ffl  12  months 


Figure  4 

Mean  l±SEl  numbers  of  fish  (/m3)  for  the  six  most  abundant  species  caught  from  rockpools  at  four  sites 
(MB=Maloney's  Bay,  TC=The  Chair,  GL  =  Gravel  Loader,  BB  =  Beaky  Bay)  during  short-term  (spring  and  autumn) 
(four  sampling  events)  and  long-term  (three  sampling  events)  recolonization  studies  undertaken  between  7 
September  1999  and  22  September  2000. 


Table  5 

Results  of  repeated-measures  ANOVAs  for  significant  differences  in  numbers  of  individuals  (/m3 )  representing  the  six  most  abun- 
dant species  caught  at  Bass  Point  during  the  short-term  recolonization  studies  among  sampling  intervals  (within-subjects  factor) 
and  sites  and  seasons  lamong-subjects  factors).  Data  were  log1(1(.v+l)  transformed  before  analysis  to  remove  heteroscedasticity  in 
the  data.  Mauchly's  criterion  for  sphericity  of  variances  was  violated  (P<0.001)  for  species  denoted  by  '';  therefore  analysis  was 
performed  using  Greenhouse-Geisser-adjusted  degrees  of  freedom.  Greenhouse-Geisser  degrees  of  freedom  used  for  within-sub- 
jects factors  where  Mauchly's  criterion  for  sphericity  of  variances  was  violated  lP<0.001):  Time  (Ti=  1.55;  SexT  =  1.55;  SxT  = 
4.64;  SexSxT  =  4.64;  Residual  =  34.14.  Mean  squares  and  significance  levels  are  shown  and  significant  results  are  given  in  bold. 
Degrees  of  freedom  are  shown  in  parentheses.  *  =  P<0.05;  **  =  P<0.01;  ***  =  P<0.001. 


Among-subjects  factors 


Within-subjects  factors 


Season {Se) 

Site(S) 

SexS 

Residual 

Time  (T) 

SexT 

SxT 

SexSxT 

Residual 

Species 

(li 

(3) 

(3) 

(24) 

(3) 

(3) 

(91 

(9) 

(72) 

Bathygobius  cocosensis0 

5.20 

24.50*** 

3.62 

2.18 

43.99*** 

2.21 

10.68** 

3.22 

1.70 

Enneapterygius  rufopileusG 

0.66 

4.38*** 

0.01 

0.17 

0.70*** 

0.05 

0.08* 

0.07* 

0.04 

Girella  elevata 

0.22 

0.49 

0.03 

0.28 

0.12** 

0.07 

0.05* 

0.07** 

0.03 

Heteroclinus  whiteleggi0 

12.09 

20.77* 

1.65 

4.90 

18.99*** 

6.61* 

4.31* 

4.22 

1.72 

Parablennius  intermedius 

0.00 

4.85* 

0.20 

1.20 

7.88** 

0.20 

1.92 

0.29 

1.07 

Aspasmogaster  costatus 

0.37 

0.54** 

0.15 

0.10 

0.11* 

0.09 

0.03 

0.07* 

0.03 

Discussion 

This  study  has  shown  that  fish  assemblages  can  quickly 
return  to  preperturbation  levels  after  significant  distur- 
bance. This  resilience  appears  to  be  driven  mainly  by 
postsettlement  movements  of  fishes,  although  recruit- 


ment may  periodically  play  a  significant  role  in  popula- 
tion replenishment.  Not  only  does  this  provide  an  insight 
into  the  ecology  of  rockpool  fish  assemblages,  but  this 
information  may  also  provide  a  basis  for  future  sampling 
protocols  where  the  confounding  effects  of  sampling  may 
be  minimized. 


642 


Fishery  Bulletin  102(4) 


Parablennius  intermedius 


Aspasmogaster  costalus 

Spring  -r 


2_ 


Long-term 


;JkjL 


J-Jt 


i 


j 


ft 


TC 


Site 


Figure  4  (continued) 


Heteroclinus  whiteleggi 


B 


Maloney's  Bay 

o 

O       A 

■ 

.        °       A 

■  o 

o 

A 

Stress  =  0.19 

Maloney's  Bay 
°  o 


Stress  =  0  in 


Maloney's  Bay 


■'J 


■      ♦ 


The  Chair 


v> 


o 

■' 
o 


A 

Slits,  =  n  l  J 


The  Chair 
A 

A 

■ 

♦          * 

♦ 

Stress  =  0  1  l 

Gravel  Loader 

■ 

A    o    o*8 

A 

A 

Sires, 

A 

=  U.I6 

Gravel  Loader 

O 

■ 

»CAO 

°  4 

o     o 

Stress  = 

i  15 

Gravel  Loader 

■ 

*      * 

o 

O 

■ 

Stress  = 

0.03 

Beaky  Bay 


°    '   «"'»o 


Beaky  Bay 


Beaky  Bay 

♦  A* 


Short-term  study: 
■  Initial 
o   1  week 
o  1  month 
A  3  months 


Long-term  study: 
■   Initial 
«   6  months 
A    12  months 


Figure  5 

Nonmetrie  multidimensional  scaling  (MDSl  plots  for  comparison  of  fish  assemblages  from  four  sites  at  Bass  Point 
in  respect  to  initial  samples  and  those  taken  after  1  week,  1  month,  and  3  months  during  (A)  spring  and  iBi 
autumn  short- term  studies,  and  after  6  months  and  12  months  during  (C)  the  long-term  study.  Each  coordinate 
represents  a  single  mckpool  sample.  Stress  values  are  shown. 


Griffiths  et  al.:  Fish  recolonization  in  temperate  Australian  rockpools 


643 


f$  Autumn 

Initial  n=81 


C        Long-term 
Initial  1=115 


0    10  20  30  40  50  60  70  80 


0    10  20  30  40  50  60  70  80 
Total  length  (mm) 


0    10  20   30  40  50  60  70  80 


Figure  6 

Length-frequency  distributions  (in  2-mm  intervals)  for  Bathygobius  cocosensis  caught 
in  the  initial  samples  and  after  1  week,  1  month,  and  3  months  later  in  the  short- 
term  study  and  caught  in  the  initial  sampling  and  6  months  and  12  months  later 
in  the  long-term  study.  Samples  were  pooled  for  all  sites  during  the  (Al  spring  and 
(Bl  autumn  short-term  recolonization  studies  and  for  all  sites  during  (C)  the  long- 
term  recolonization  study.  Studies  were  conducted  between  7  September  1999  and 
22  September  2000.  Samples  sizes  are  shown. 


Results  of  ANOSIM  testing 
during  the  spring,  autumn. 

Table  6 

for  differences  in  fish  assemblage  structure  among  sampling  intervals  at  four  sites  at  Bass  Point 
and  long-term  recolonization  studies.  Significant  results  are  shown  in  bold. 

Site 

Spr 

ng 

Autumn 

Long 

term 

R 

P  value 

R                  P  value 

R 

P  value 

Maloney's  Bay 
The  Chair 
Gravel  Loader 
Beaky  Bay 

0.002 
0.185 
0.083 
0.025 

0.474 
0.057 
0.165 
0.383 

0.029               0.393 
0.014                 0.422 
0.027                 0.389 
0.726               0.000 

-0.101 
0.063 
0.190 

0.044 

0.752 
0.284 
0.051 
0.348 

For  rockpool  fish  assemblages  in  southeastern  Aus- 
tralia, a  period  of  one  week  appears  insufficient  for 
recolonization  of  all  species  if  fish  are  removed  during 
sampling,  whereas  intervals  of  one  to  three  months 
appear  sufficient  for  rockpool  fish  assemblages  at  most 
Bass  Point  locations  to  recolonize  to  preperturbation 
levels.  It  is  possible  that  recolonization  times  may  be 
decreased  if  all  fish  are  returned  to  rockpools  immedi- 
ately after  sampling.  However,  this  should  not  provide 
a  foundation  for  subsequent  studies  with  other  defauna- 
tion  methods,  such  as  anesthetics  or  ichthyocides.  The 
possible  residual  effects  of  these  other  sampling  meth- 
ods, such  as  the  mortality  of  mobile  and  sessile  inver- 


tebrates or  the  residues  from  chemical  anesthetics  and 
ichthyocides  are  possible  factors  that  may  complicate 
fish  recolonization  patterns  (see  Lockett,  1998)  and 
certainly  require  additional  investigation.  Nonethe- 
less, in  recolonization  studies  with  chemical  sampling 
methods  similar  recolonization  times  as  those  of  the 
present  study  were  found.  For  example,  recolonization 
of  rockpools  defaunated  by  ichthyocides  was  shown  to  be 
complete  within  1  month  (Grossman,  1982;  Prochazka, 
1996)  and  3  months  (Beckley,  1985a;  Willis  and  Rob- 
erts, 1996;  Polivka  and  Chotkowski,  1998). 

Spatial  variability  in  fish  recolonization  patterns 
was  not  definitive  with  regard  to  species  composition 


644 


Fishery  Bulletin  102(4) 


Spring 


Autumn 


/~>     Long-term 

Initial  n=73 


i — i     i     F1  i — I 

0      10     20     30     40     50     60 


Total  length  (mm) 


Figure  7 

Length-frequency  distributions  (in  2-mm  intervals)  for  Enneapterygius  rufopileus 
caught  in  the  initial  samples  and  after  1  week,  1  month,  and  3  months  later  in  the 
short-term  study  and  caught  in  the  initial  sampling  and  6  months  and  12  months 
later  in  the  long-term  study.  Samples  were  pooled  for  all  sites  during  the  (A)  spring 
and  (B)  autumn  short-term  recolonization  studies  and  for  all  sites  during  (C)  the 
long-term  recolonization  study.  Studies  were  conducted  between  7  September  1999 
and  22  September  2000.  Samples  sizes  are  shown. 


because  samples  were  generally  widely  dispersed  in 
nMDS  ordination  plots.  The  relatively  low  stress  values 
(<0.2)  indicate  that  high  variability  in  fish  assemblages 
at  the  level  of  individual  rockpools  is  probably  respon- 
sible for  the  patterns  observed.  However,  some  spatial 
variability  in  fish  recolonization  patterns  was  evident 
(Fig.  2)  and  appeared  to  be  dependent  to  some  extent 
on  exposure  of  sites  to  predominant  swell.  There  is 
evidence  to  suggest  that  wave  exposure  can  affect  the 
distribution  of  intertidal  fishes  (Gibson,  1972;  Ibanez  et 
al.,  1989),  although  there  is  apparently  no  study  that 
has  investigated  this  effect  in  relation  to  fish  recoloni- 
zation in  rockpools.  In  the  present  study,  recolonization 
appeared  more  rapid  at  wave-exposed  sites  (MB  and 
TC)  compared  to  more  sheltered  sites.  This  may  have 
been  the  result  of  the  close  distance  between  rockpools 
at  exposed  sites  (within  meters  of  each  other),  whereas 
at  both  sheltered  sites,  rockpools  were  significantly 
farther  apart.  Consequently,  defaunated  rockpools  at 
exposed  sites  may  recolonize  more  quickly  if  the  major 
recolonizers  are  derived  from  neighboring  rockpools 
as  has  been  documented  elsewhere  (Beckley,  1985a; 
Polivka  and  Chotkowski,  1998). 

Fish  recolonization  patterns  were  not  influenced  by 
the  time  of  year  that  rockpools  were  defaunated  in 
either  short-term  or  long-term  studies.  The  numbers 


of  species  and  individuals  consistently  returned  to  pre" 
perturbation  levels  within  a  few  weeks,  but  this  return 
to  previous  levels  may  partially  be  a  consequence  of 
the  relatively  small  number  of  species  that  are  nor- 
mally found  in  rockpools  at  any  given  time.  In  such 
situations  a  significant  differences  could  only  occur  if 
large-scale  changes  in  abundances  were  recorded.  The 
lack  of  temporal  variation  in  recolonization  rates  was 
surprising  because  recolonization  was  expected  to  be 
more  rapid  during  summer,  when  the  larvae  of  residents 
and  warm  water  transients  are  expected  to  be  avail- 
able for  settlement  (Beckley,  1985a;  Willis  and  Roberts, 
1996).  Recruitment  was  not  the  major  mechanism  driv- 
ing fish  recolonization  in  the  present  study  because  the 
majority  of  recolonizers  were  subadults  and  adults  that 
would  have  relocated  from  nearby  rockpools.  Although 
many  of  the  fish  captured  in  each  pool  were  tagged, 
the  vast  majority  offish  caught  in  the  same  rockpool  in 
subsequent  sampling  events  were  not  tagged.  Griffiths 
(2003b)  showed  that  the  common  recolonizing  species  in 
the  present  study  moved  between  a  few  rockpools  within 
a  limited  home  range.  Therefore,  postsettlement  fishes 
from  surrounding  rockpools  were  probably  moving  into 
the  study  rockpools  between  each  sampling  event. 

The  movement  of  postsettlement  fishes  from  adjacent 
rockpools  also  appears  to  control  the  resilience  of  rock- 


Griffiths  et  al.:  Fish  recolonization  in  temperate  Australian  rockpools 


.  645 


pool  fish  assemblages.  Therefore,  the  composition  of  spe- 
cies in  newly  recolonized  rockpools  is  probably  depen- 
dent upon  the  relative  abundances  of  species  in  nearby 
rockpools.  Species  having  the  highest  local  abundances, 
such  as  B.  cocosensis  and  E.  rufopileus,  are  therefore 
more  likely  to  be  the  primary  recolonizers  because  va- 
cant habitats  have  a  higher  probability  of  being  located 
by  these  species  during  high-tide  excursions  throughout 
the  intertidal  zone  (also  see  Polivka  and  Chotkowski, 
1998).  These  species  are  also  versatile  and  can  exploit 
a  range  of  microhabitats  and,  as  a  result,  can  occupy 
almost  any  rockpool  within  the  intertidal  zone  (Griffiths 
et  al.,  2003).  This  is  particularly  true  for  B.  cocosensis. 
In  contrast,  less  abundant  species  such  as  H.  whiteleggi 
often  occupy  more  specific,  and  perhaps  less  abundant, 
microhabitats  such  as  algal  cover  (see  Marsh  et  al.. 
1978;  Bennett  and  Griffiths,  1984)  that  may  require 
longer  periods  to  locate  than  more  abundant  habitats, 
such  as  cobble-covered  substratum. 

Processes  regulating  fish  assemblages 

The  structure  of  multispecies  assemblages  can  be 
regarded  as  being  regulated  by  either  deterministic  or 
stochastic  processes  (see  Grossman,  1982).  Assemblages 
regulated  by  deterministic  processes  generally  occur  in 
environments  where  conditions  are  constant  or  fluctuate 
consistently  over  time.  The  structure  of  these  assem- 
blages is  generally  predictable.  This  can  be  maintained 
through  a  number  of  factors  including  partitioning  of 
resources  in  finite  supply  (Schoener,  1974;  Behrents, 
1987)  and  interspecific  competition,  which  prevents  any 
single  species  being  competitively  dominant  (Buss  and 
Jackson,  1979). 

In  contrast,  assemblages  regulated  by  stochastic  pro- 
cesses generally  exist  in  unpredictable  environments. 
Here,  the  resources  are  available  on  a  random  or  pe- 
riodic basis,  which  prevents  superior  competitors  from 
dominating  the  assemblage  (Sale,  1977,  1978).  The  suc- 
cess of  particular  species  can  be  compared  to  winning 
a  "lottery"  for  living  space  (Sale,  1977,  1978,  1982). 
Consequently,  stochastically  regulated  assemblages  are 
generally  species  rich  (Sale,  1977). 

Rockpool  fish  assemblages  are  often  persistent  for 
lengthy  periods,  even  after  catastrophic  natural  dis- 
turbances, such  as  hurricanes  (Moring,  1996),  and  con- 
tinual experimental  eliminations  (Grossman,  1982; 
Collette,  1986).  For  example,  Collette  (1986)  found  two 
species — Pholis  gunnellus  and  Tautogolabrus  adsper- 
sus — to  be  dominant  over  19  years  of  study  in  two  New 
England  rockpools,  whereas  the  rank  of  dominant  spe- 
cies in  the  rockpools  of  Barbados  showed  no  evidence 
of  change  over  six  years  (Mahon  and  Mahon.  1994). 
Similar  stability  and  persistence  were  evident  in  the 
present  study,  where  B.  cocosensis,  E.  rufopileus  and 
G.  elevata  were  consistently  the  highest  ranked  species 
in  each  collection  for  all  three  studies,  regardless  of 
the  period  between  sampling.  This  finding  may  indi- 
cate that  deterministic  processes  probably  regulate  the 
Bass  Point  fish  assemblage.  If  this  is  the  case,  it  may 


seem  ironic  because  the  intertidal  zone  is  subjected  to 
a  high  frequency  of  stochastic  events.  It  would  be  easy 
to  assume  that  such  events  could  eliminate  fishes  from 
rockpools  and  thus  leave  microhabitats  for  other  species 
to  exploit.  This  kind  of  process  has  been  documented  for 
some  sessile  intertidal  invertebrate  assemblages  that 
rely  on  the  availability  of  vacant  substrata  for  success- 
ful recruitment  of  larvae  (see  examples  by  Raffaelli 
and  Hawkins.  1996).  However,  the  locomotory  capabili- 
ties and  morphological  and  physiological  adaptations  of 
resident  intertidal  fishes  allow  them  to  cope  with  such 
disturbances  by  being  able  to  cope  temporarily  with  ad- 
verse conditions  (Martin.  1995).  As  a  result,  the  abun- 
dance of  resident  species  may  be  little  affected  under 
normal  disturbance  regimes. 


Conclusions 

The  results  of  this  study  have  significantly  increased  an 
understanding  of  the  patterns  of  recolonization  of  rock- 
pools by  fishes  and  some  of  the  processes  that  underpin 
these  patterns.  Such  an  understanding  of  recoloniza- 
tion processes  may  improve  our  ability  to  predict  the 
consequences  of  significant  natural  and  anthropogenic 
disturbances  on  not  only  the  fish  assemblages  but  also 
on  other  intertidal  community  assemblages  that  may  be 
maintained  by  the  presence  offish  (see  Coull  and  Wells, 
1983;  Connell  and  Anderson,  1999). 

On  a  more  technical  note,  the  recolonization  rates 
observed  in  the  present  study  may  provide  insight  for 
other  researchers  aiming  to  stud}'  natural  temporal 
variation  of  rockpool  fish  assemblages  by  minimizing 
the  possibility  of  confounding  effects  of  sampling.  This 
may  be  particularly  important  for  long-term  monitoring 
programs,  such  as  for  marine  protected  areas  (MPAs). 
that  may  require  detection  of  changes  in  community 
structure  over  time.  Finding  sufficient  numbers  of  simi- 
lar-size pools  at  a  single  location  for  monitoring  can  be 
difficult:  therefore  repeated  visits  to  the  same  rockpools 
may  often  be  required.  For  southeastern  Australian 
rockpools.  we  feel  that  a  period  of  one  to  three  months 
is  required  before  resampling  the  same  rockpools  with 
the  methods  employed  in  this  study.  Although  fish  were 
not  returned  to  rockpools  immediately  after  sampling 
in  the  present  study,  we  feel  that  this  practice  may 
significantly  increase  recolonization  rates.  However, 
the  results  of  the  present  study  should  not  provide  a 
foundation  for  studies  using  other  defaunation  methods, 
such  as  anesthetics  or  ichthvocides.  because  other  fac- 
tors, such  as  chemical  residues  remaining  in  rockpools, 
may  complicate  fish  recolonization  patterns.  Further 
investigation  into  these  other  factors  will  be  necessary 
in  the  future. 


Acknowledgments 

We  sincerely  thank  Jade  Butler  and  Alan  Griffiths  for 
help  with  fieldwork.  This  paper  is  partly  based  upon 


646 


Fishery  Bulletin  102(4) 


research  included  in  a  Ph.D.  by  S.  P.  Griffiths  funded  by 
an  Australian  Postgraduate  Award.  Additional  funding 
was  granted  by  the  Institute  for  Conservation  Biology, 
University  of  Wollongong,  Shellharbour  City  Council, 
The  Ecology  Lab.  Pty.  Ltd.,  and  Ocean  Beach  Hotel 
Fishing  Club.  S.  P.  Griffiths  would  like  to  thank  CSIRO 
Marine  Research  for  support  during  the  preparation  of 
this  article. 


Literature  cited 

Astles,  K.  L. 

1993.     Patterns  of  abundance  and  distribution  of  spe- 
cies in  intertidal  rockpools.     J.  Mar.  Biol.  Assoc.  U.K. 
73:555-569. 
Beckley.  L.  E. 

1985a.     Tide-pool  fishes:  recolonization  after  experimental 

elimination.     J.  Exp.  Mar.  Biol.  Ecol.  85:287-295. 
1985b.     The  fish  community  of  East  Cape  tidal  pools  and 
an  assessment  of  the  nursery  function  of  this  habitat.     S. 
Afr.  J.  Zool.  20:21-27. 
2000.     Species  composition  and  recruitment  of  tidal 
pools  fishes  in  KwaZulu-Natal,  South  Africa.     Afr. 
Zool.  35:29-34. 
Behrents,  K.  C. 

1987.     The  influence  of  shelter  availability  on  recruit- 
ment and  early  juvenile  survivorship  of  Lythrypnus 
dalli  Gilbert  (Pisces:  Gobiidae).     J.  Exp.  Mar.  Biol. 
Ecol.  107:45-59. 
Bennett,  B.  A.,  and  C.  L.  Griffiths. 

1984.     Factors  affecting  the  distribution,  abundance  and 
diversity  of  rock-pool  fishes  on  the  Cape  Peninsula, 
South  Africa.     S.  Afr.  J.  Zool.  19:97-104. 
Buss.  L.  W.,  and  J.  B.  C.  Jackson. 

1979.     Competitive  networks:  nontransitive  competitive 
relationships  in  cryptic  coral  reef  environments.     Am. 
Nat.  113:223-234. 
Bussing,  W.  A. 

1972.     Recolonization  of  a  population  of  supratidal  fishes 
at  Eniwetok  Atoll,  Marshall  Islands.     Atoll  Res.  Bull. 
154:1-4. 
Clarke,  K.  R. 

1993.     Non-parametric  multivariate  analyses  of  changes 
in  community  structure.     Aust.  J.  Ecol.  18:117-143. 
Collette,  B.  B. 

1986.     Resilience  of  the  fish  assemblage  in  New  England 
tidepools.     Fish.  Bull.  84:200-204. 
Connell.J.  H. 

1972.     Community  interactions  on  marine  rock  intertidal 
shores.     Ann.  Rev.  Ecol.  Syst.  3:169-192. 
Connell,  S.  D.,  and  M.  J.  Anderson. 

1999.     Predation  by  fish  on  assemblages  of  intertidal 
epibiota:  effects  of  predator  size  and  patch  size.     J. 
Exp.  Mar.  Biol.  Ecol.  241:15-29. 
Coull,  B.  C,  and  J.  B.  J.,  Wells. 

1983.     Refuges    from    fish    predation:    experiments 
with  phytal  meiofauna  from  the  New  Zealand  rocky 
intertidal.     Ecology  64:1599-1609. 
Faria,  C,  and  V.  Almada. 

1999.     Variation   and   resilience   of  rocky   intertidal 
fish   in   western    Portugal.     Mar.    Ecol.    Prog.    Ser. 
184:197-203. 
Gibson,  R.  N. 

1972.     The  vertical  distribution  and  feeding  relationships 


of  intertidal  fish  on  the  Atlantic  coast  of  France.     J. 
Anim.  Ecol.  41:189-207. 
1982.     Recent  studies  on  the  biology  of  intertidal  fishes. 
Oceanogr.  Mar.  Biol.  Ann.  Rev.  20:363-414. 
Green,  J.M. 

1971.     High  tide  movements  and  homing  behaviour  of 
the  tidepool  sculpin  Oligocottus  maculosus.     J.  Fish. 
Res.  Can.  28:383-389. 
Griffiths,  S.  P. 

2000.     The  use  of  clove  oil  as  an  anaesthetic  and  method 
for  sampling  intertidal  rockpool  fishes.     J.  Fish  Biol. 
57:1453-1464. 
2003a.     Spatial  and  temporal  dynamics  of  temperate 
Australian  rockpool  ichthyofaunas.     Mar.  Freshw.  Res. 
54:163-176. 
2003b.     Homing  behaviour  of  intertidal  rockpool  fishes 
in  south-eastern  New  South  Wales,  Australia.     Aust. 
J.  Zool.  51:387-398. 
2003c.     Rockpool  ichthyofaunas  of  temperate  Austra- 
lia: species  composition,  residency  and  biogeographic 
patterns.     Estuar.  Coast  Shelf  Sci.  58:173-186. 
Griffiths,  S.  P.,  R.  J.  West,  and  A.  Davis. 

2003.     Effects  of  intertidal  elevation  on  the  rockpool 
ichthyofaunas  of  temperate  Australia.     Environ.  Biol. 
Fishes  68:197-204 
Grossman,  G.  D. 

1982.     Dynamics  and  organisation  of  a  rocky  intertidal 
fish  assemblage:  the  persistance  and  resilience  of  taxo- 
cene  structure.     Am.  Nat.  119:611-637. 
1986.     Long  term  persistence  in  a  rocky  intertidal  fish 
assemblage.     Environ.  Biol.  Fishes  15:315-317. 
Ibanez,  M.,  I.  Miguel,  D.  San  Millan,  and  I.  Ripa. 

1989.     Intertidal  ichthyofauna  of  the  Spanish  Atlantic 
coast.     Sci.  Mar.  53:451-455. 
Lardner,  R.,  W.  Ivantsoff,  and  L.  E.  L.  M.  Crowley. 

1993.  Recolonization  by  fishes  of  a  rocky  intertidal 
pool  following  repeated  defaunation.  Aust.  Zool.  29: 
85-92. 

Lockett,  M.  M. 

1998.     The  effect  of  Rotenone  on  fishes  and  its  use  as  a 
sampling  technique:  a  survey.     Z.  Fischkd.  5:13-45. 
MacArthur,  R.  H.,  and  E.  O.  Wilson. 

1967.     The  theory  of  island  biogeography,  224  p.     Princ- 
eton Univ.  Press,  Princeton,  NJ. 
Mahon,  R.,  and  S.  D.  Mahon. 

1994.  Structure  and  resilience  of  a  tidepool  fish  assem- 
blage at  Barbados.     Environ.  Biol.  Fishes  41:171-190. 

Marsh,  B.,  T.  M.  Crowe,  and  W.  W.  Siegfried. 

1978.     Species  richness  and  abundance  of  clinid  fish 
(Teleostei:  Clinidae)  in  intertidal  rock  pools.     S.  Afr. 
J.  Zool.  13:283-291. 
Martin.  K.  L.  M. 

1995.  Time  and  tide  wait  for  no  fish:  intertidal  fishes 
out  of  water.     Environ.  Biol.  Fishes  44:165-181. 

Matson,  R.  H.,  C.  B.  Crabtree,  and  T.  R.  Haglund. 

1986.     Ichthyofaunal  composition  and  recolonization  in 
a  central  California  tidepool.     Calif.  Fish  and  Game 
72:227-231. 
Mistry,  S.  D.,  E.  K.  Lizerbram,  and  E.  R.  Parton. 

1989.     Short-term  ichthyofaunal  recruitment  in  northern 
California  tidepools.     Copeia  1989:1081-1084. 
Mok,  H„  and  P.  Wen. 

1985.  Intertidal  fish  community  ecology  on  Lu  Tao  Island 
(Green  Island),  Taiwan.     J.  Taiwan  Mus.  38:81-118. 


Griffiths  et  al.:  Fish  recolonization  in  temperate  Australian  rockpools 


647 


Moring,  J.  R. 

1990.     Seasonal   absence   of  fishes   in   tidepools  of  a 
boreal  environment  (Maine,  USA).     Hydrobiologia  194: 
163-168. 
Moring.  J.  R. 

1996.  Short-term  changes  in  tidepools  following  two 
hurricanes.     Hydrobiologia  328:155-160. 

Paine.  R.  T.,  and  S.  A.  Levin. 

1981.     Intertidal  landscapes:  disturbance  and  the  dynam- 
ics of  pattern.     Ecol.  Monogr.  47:37-63. 
Pfister.  C.A. 

1995.  Estimating  competition  coefficients  from  census 
data:  a  test  with  field  manipulations  of  tidepool  fishes. 
Am.  Nat.  146:271-291. 

1997.  Demographic  consequences  of  within  year  variation 
in  recruitment.     Mar.  Ecol.  Prog.  Ser.  153:  229-238. 

Polivka,  K.  M.,  and  M.  A.Chotkowski. 

1998.  Recolonization  of  experimentally  defaunated 
tidepools  by  northeast  Pacific  intertidal  fishes.  Co- 
peia:456-462. 

Prochazka.  K. 

1996.  Seasonal  patterns  in  a  temperate  intertidal  fish 
community  on  the  west  coast  of  South  Africa.  Environ. 
Biol.  Fishes  45:133-140. 

Prochazka,  K.,  and  C.  L.  Griffiths. 

1992.     The  intertidal  fish  fauna  of  the  west  coast  of 
South  Africa — species,  community  and  biogeographic 
patterns.     S.  Afr.  J.  Zool.  27:115-120. 
Raffaelli,  D.,  and  S.  Hawkins. 

1996.     Intertidal  ecology,  372  p.     Chapman  and  Hall, 
London. 
Sale,  P.  F. 

1977.  Maintenance  of  high  diverity  in  coral  reef  fish 
communities.     Am.  Nat.  111:337-359. 


1978.     Coexistance  of  coral  reef  fishes — a  lottery  for  living 

space.     Environ.  Biol.  Fishes  3:  85-102. 
1982.     Stock-recruit  relationships  and  regional  coexistance 
in  a  lottery  competitive  system:  a  simulation  study.     Am. 
Nat.  120:139-159. 
Schoener,  T.  W. 

1974.     Resource  partitioning  in  ecological  communities. 
Science  185:27-39. 
Silberschneider.  V.,  and  D.  J.,  Booth. 

2001.     Resource  use  by  Enneapterygius  rufopileus  and 
other  rockpool  fishes.     Environ.  Biol.  Fishes  61:195- 
204. 
Thomson,  D.  A.,  and  C.  E.  Lehner. 

1976.     Resilience  of  a  rocky  intertidal  fish  community 
in  a  physically  unstable  environment.     J.  Exp.  Mar. 
Biol.  Ecol.  22:-29. 
Underwood,  A.  J.,  and  G.  A.  Skilleter. 

1996.     Effects  of  patch-size  on  the  structure  of  assemblages 
in  rock  pools.     J.  Exp.  Mar.  Biol.  Ecol.  197:  63-90. 
Willis,  T  J.,  and  C.  D.  Roberts. 

1996.     Recolonization  and  recruitment  of  fishes  to  inter- 
tidal rockpools  at  Wellington,  New  Zealand.     Environ. 
Biol.  Fishes  47:329-343. 
Yoshiyama,  R.  M.,  C.  Sassaman.  and  R.  N.  Lea. 

1986.     Rocky  intertidal  fish  communities  of  California: 
temporal  and  spatial  variation.     Environ.  Biol.  Fishes 
17:23-40. 
Yoshiyama,  R.  M.,  K.  B.  Gaylord,  M.  T  Philippart,  T.  R.  Moore. 
J.  R.  Jordan,  C.  C.  Coon,  L.  L.  Schalk,  C.  J.  Valpey,  and 
I.  Tosques. 

1992.  Homing  behaviour  and  site  fidelity  in  the  inter- 
tidal sculpins  I  Pisces:  Cottidael.  J.  Exp.  Mar.  Biol. 
Ecol.  160:115-130. 


648 


Abstract — We  have  studied  the  repro- 
ductive biology  of  the  goldlined  sea- 
bream  (Rhabdosargus  sarba)  in  the 
lower  Swan  River  Estuary  in  Western 
Australia,  focusing  particularly  on 
elucidating  the  factors  influencing 
the  duration,  timing,  and  frequency 
of  spawning  and  on  determining 
potential  annual  fecundity.  Our 
results  demonstrate  that  1)  Rhab- 
dosargus sarba  has  indeterminate 
fecundity,  2)  oocyte  hydration  com- 
mences soon  after  dusk  (ca.  18:30  hi 
and  is  complete  by  ca.  01:30-04:30  h 
and  3)  fish  with  ovaries  containing 
migratory  nucleus  oocytes,  hyd rated 
oocytes,  or  postovulatory  follicles  were 
caught  between  July  and  November. 
However,  in  July  and  August,  their 
prevalence  was  low,  whereas  that  of 
fish  with  ovaries  containing  substan- 
tial numbers  of  atretic  yolk  granule 
oocytes  was  high.  Thus,  spawning 
activity  did  not  start  to  peak  until 
September  (early  spring),  when  salini- 
ties were  rising  markedly  from  their 
winter  minima.  The  prevalence  of 
spawning  was  positively  correlated 
with  tidal  height  and  was  greatest 
on  days  when  the  tide  changed  from 
flood  to  ebb  at  ca.  06:00  h.  i.e.,  just 
after  spawning  had  ceased.  Because 
our  estimate  of  the  average  daily 
prevalence  of  spawning  by  females 
during  the  spawning  season  (July 
to  November)  was  36.5%,  individual 
females  were  estimated  to  spawn, 
on  average,  at  intervals  of  about  2.7 
days  and  thus  about  45  times  during 
that  period.  Therefore,  because  female 
R.  sarba  with  total  lengths  of  180, 
220,  and  260  mm  were  estimated  to 
have  batch  fecundities  of  about  4500, 
7700,  and  12,400  eggs,  respectively, 
they  had  potential  annual  fecundities 
of  about  204,300,  346,100  and  557,500 
eggs,  respectively.  Because  spawn- 
ing occurs  just  prior  to  strong  ebb 
tides,  the  eggs  of/?,  sarba  are  likely 
to  be  transported  out  of  the  estuary 
into  coastal  waters  where  salini- 
ties remain  at  ca.  359«.  Such  down- 
stream transport  would  account  for 
the  fact  that,  although  R.  sarba  exhib- 
its substantial  spawning  activity  in 
the  lower  Swan  River  Estuary,  few 
of  its  early  juveniles  are  recruited 
into  the  nearshore  shallow  waters  of 
this  estuary. 


Manuscript  submitted  9  June  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
28  April  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:648-660(2004). 


Factors  influencing  the  timing  and  frequency  of 
spawning  and  fecundity  of  the  goldlined  seabream 
(Rhabdosargus  sarba)  (Sparidae) 
in  the  lower  reaches  of  an  estuary 

S.  Alexander  Hesp 

Ian  C.  Potter 

Centre  for  Fish  and  Fisheries  Research 

School  of  Biological  Sciences  and  Biotechnology 

Murdoch  University 

South  Street 

Murdoch,  Western  Australia  6150,  Australia 

E-mail  address  (for  I  C  Potter,  contact  author):  i.pottera'murdoch  edu.au 

Sonja  R.  M.  Schubert 

Ernst-Moritz  Arndt  Universitaet,  Hansestadt  Greifswald 
F.-L.-Jahn  StraBe  15a 
17487  Greifswald,  Germany 


The  goldlined  seabream  {Rhabdosar- 
gus sarba)  is  an  important  recreational 
and  commercial  fish  species  in  numer- 
ous regions  throughout  the  Indo-west 
Pacific  (van  der  Elst,  1988;  El-Agamy, 
1989;  Kuiter,  1993).  Although  this 
species  is  a  protandrous  hermaphro- 
dite in  certain  regions,  e.g.,  the  waters 
of  Hong  Kong  and  South  Africa  (Yeung 
and  Chan,  1987;  Garratt,  1993).  it  is  a 
rudimentary  hermaphrodite  in  a  range 
of  environments  in  Western  Australia 
(Hesp  and  Potter,  2003).  Rudimentary 
hermaphrodites  are  those  species  in 
which  the  juveniles  possess  gonads 
consisting  of  both  immature  testicular 
and  ovarian  tissues  that,  in  adults, 
develop  permanently  into  either  func- 
tional testes  with  rudimentary  ovar- 
ian tissue  or  functional  ovaries  with 
rudimentary  testicular  tissue  (Buxton 
and  Garratt,  1990).  In  Western  Aus- 
tralia, R.  sarba  attains  similar  maxi- 
mum lengths,  i.e.,  346-370  mm,  in 
temperate  marine  coastal  waters  and 
the  lower  reaches  of  the  Swan  River 
Estuary  on  the  lower  west  coast  of 
Australia  and  in  a  large  subtropical 
embayment  ca.  800  km  farther  north 
(Hesp  et  al.,  2004).  However,  the  max- 
imum age  recorded  for  this  species  in 
the  estuary.  7  years,  was  far  less  than 
that  for  the  other  two  environments: 
temperate  marine  coastal  waters  (11 


years)  and  a  large  subtropical  embay- 
ment (13  years)  (Hesp  et  al.,  2004). 

Although  R.  sarba  is  typically  re- 
garded as  a  marine  species  that  fre- 
quently uses  estuaries  as  a  nursery 
area  (e.g.,  Wallace,  1975;  Potter  and 
Hyndes,  1999;  Smith  and  Suthers, 
2000),  it  spawns  in  the  lower  Swan 
River  Estuary  as  well  as  in  coastal 
waters  outside  this  estuary  (Hesp  and 
Potter,  2003).  However,  this  sparid 
attains  maturity  later  in  the  estuary 
than  in  those  nearby  coastal  marine 
waters.  If  this  indication  that  the  on- 
set of  spawning  for  R.  sarba  in  the 
Swan  River  Estuary  is  related  to  the 
attainment  of  higher  salinities  in  the 
spring,  it  would  parallel  the  situation 
recorded  for  the  spotted  seatrout  (Cy- 
noscion  nebulosus)  in  the  estuaries  of 
the  Gulf  of  Mexico  where  this  species 
completes  its  entire  life  cycle  ( Brown- 
Peterson  et  al.,  2002). 

Despite  the  importance  and  wide- 
spread occurrence  of  R.  sarba.  and 
the  great  value  of  fecundity  data  for 
stock  assessments  (Hunter  et  al., 
1992:  Nichol  and  Acuna,  2001),  on- 
ly one  attempt  has  apparently  been 
made  to  estimate  the  annual  fecun- 
dity of  wild  populations  of  this  sparid 
(El-Agamy,  1989).  Although  El-Aga- 
my (1989)  recognized  that  R.  sarba 
is  a  "fractional"  spawner  and  has  a 


Hesp  et  al.:  Timing  and  frequency  of  spawning  and  fecundity  of  Rhabdosargus  sarba 


649 


protracted  spawning  season,  he  recorded  the  fecundity 
of  this  species  as  the  number  of  larger  eggs  (diameter 
>180  fjm)  estimated  to  be  present  in  the  ovaries  of 
mature  females  just  prior  to  the  commencement  of  the 
spawning  period.  Thus,  the  very  strong  possibility  that 
some  eggs  with  diameters  <180  /jm  would  have  been 
destined  to  have  become  fully  mature  and  released  at 
some  stage  during  the  protracted  spawning  season,  i.e., 
the  species  has  indeterminate  fecundity,  was  not  taken 
into  account. 

In  species  with  indeterminate  fecundity,  the  distri- 
bution of  oocyte  diameters  essentially  forms  a  contin- 
uum, reflecting  the  continuous  maturation  of  oocytes 
throughout  the  spawning  season  and  thus  the  progres- 
sion through  to  maturity  of  some  of  the  small  and  pre- 
vitellogenic  oocytes  that  were  present  at  the  beginning 
of  the  spawning  period.  Consequently,  counts  of  the 
standing  stock  of  larger  oocytes  found  just  prior  to  the 
onset  of  spawning  will  result  in  an  underestimate  of  the 
potential  annual  fecundity  of  such  species  (Hunter  et  al., 
1985.  1992;  Lisovenko  and  Andrianov,  1991).  Estimation 
of  the  annual  fecundity  of  species  with  indeterminate 
fecundity  thus  requires  a  combination  of  data  on  batch 
fecundity  and  spawning  frequency  (Hunter  et  al..  19851. 
Batch  fecundity,  i.e.,  the  number  of  oocytes  released 
during  a  single  spawning  event,  can  be  estimated  by 
counting  the  number  of  hydrated  oocytes  present  in 
ovaries  immediately  prior  to  that  spawning  (Hunter 
et  al.,  1985).  The  frequency  with  which  a  fish  spawns 
during  the  spawning  period  can  be  determined  from  the 
frequency  of  mature  female  fish  possessing  ovaries  with 
either  hydrated  oocytes  or  postovulatory  follicles  (POFs) 
of  a  known  age  (Hunter  and  Macewicz,  1985). 

The  spawning  of  many  marine  species  of  teleosts  and 
invertebrates  is  correlated  with  lunar  periodicity  and  the 
associated  tidal  cycles  (e.g.,  Schwassmann,  1971;  Taylor. 
1984;  Greeley  et  al.,  1986;  Hoque  et  al.,  1999),  with  the 
spawning  of  such  fish  species  typically  peaking  around 
the  full  or  new  moon  (or  both)  (e.g.,  Johannes.  1978; 
Taylor  and  DiMichele.  1980;  Greeley  et  al.,  1986).  Many 
fish  and  invertebrates  with  pelagic  eggs  spawn  on  high 
or  ebb  tides  that  enable  eggs  and  the  subsequent  larval 
stages  to  be  transported  away  from  spawning  areas, 
in  which  planktivorous  predators  are  concentrated.  This 
process  thus  reduces  the  likelihood  of  those  early  life 
cycle  stages  being  subjected  to  predation  (Taylor,  1984; 
Johnson  et  al.,  1990;  Morgan,  1990).  The  fact  that  there 
is  very  little  recruitment  of  the  early  0+  individuals  of 
R.  sarba  into  the  lower  Swan  River  Estuary,  where  ex- 
tensive spawning  occurs,  indicates  that  tides  transport 
the  eggs  of  this  species  from  spawning  areas  in  the 
estuary  into  coastal  marine  waters  (Hesp  and  Potter, 
2003;  Hesp  et  al.,  2004). 

This  investigation,  which  involved  a  detailed  study  of 
the  females  of  R.  sarba  in  the  lower  Swan  River  Estuary, 
had  the  following  aims:  1)  to  test  the  hypothesis  that 
R.  sarba  has  indeterminate  fecundity;  2)  to  establish 
the  period  during  the  day  when  the  oocytes  of  R.  sarba 
become  hydrated  and  when  ovulation  and  spawning 
occur;  3)  to  establish  whether  R.  sarba  spawns  mainly 


when  salinities  are  high  and  thus  approach  those  of  the 
marine  waters  in  which  this  species  typically  breeds 
and  whether  spawning  is  correlated  with  the  strength 
and  type  (ebb  vs  flood)  of  tide  in  the  lower  reaches  of 
the  Swan  River  Estuary;  4)  to  estimate  the  average 
frequency  of  spawning  for  R.  sarba  during  the  spawning 
period;  5)  and  to  determine  the  relationship  between 
batch  fecundity  and  fish  length,  and  to  use  this  rela- 
tionship, in  combination  with  the  average  spawning 
frequency,  to  calculate  the  potential  annual  fecundity 
of  R.  sarba  of  different  sizes. 


Materials  and  methods 

Tide,  lunar  phase,  and  salinity 

The  maximum  daily  tidal  heights  at  the  mouth  of  the 
Swan  River  Estuary  were  calculated  by  using  the  tidal 
prediction  data  of  the  Coastal  Data  Centre  at  the  Depart- 
ment of  Planning  and  Infrastructure,  Government  of 
Western  Australia  (http://www.coastaldata. transport. 
wa.gov.au).  The  maximum  tidal  range  at  the  mouth  of 
the  Swan  River  Estuary  is  small,  i.e.,  <0.8  m,  and  tides 
can  be  diurnal  or  semidiurnal,  depending  on  the  time 
of  year  (Spencer,  1956).  Salinity  was  measured  on  each 
sampling  occasion  by  using  a  Yellow  Springs  Instru- 
ments salinity  meter  (YSI  model  number  30,  Yellow 
Springs  Instrument  Co.,  Inc.,  Yellow  Springs.  OH). 

Sampling 

During  2001  and  2002,  female  Rhabdosargus  sarba 
were  collected  by  seine  netting  in  nearshore  shallow 
waters  at  distances  of  ca.  2.5  to  5  km  from  the  mouth 
of  the  Swan  River  Estuary,  and  by  rod  and  line  fishing 
in  water  depths  of  10-12  m  at  a  distance  of  ca.  150  m 
from  the  shore  (for  details  of  sampling  region  and  seine 
net,  see  Hesp  and  Potter,  2003).  Sampling  was  under- 
taken at  least  once  weekly  between  July  and  November, 
the  period  when  R.  sarba  reach  maturity  in  the  lower 
Swan  River  Estuary  (Hesp  and  Potter,  2003).  It  was 
restricted  to  the  hours  between  dusk  (ca.  18:00  h)  and 
dawn  (ca  06:00  h)  because  extensive  seine  netting  and 
angling  during  the  day  in  our  earlier  study  failed  to  yield 
any  R.  sarba.  The  failure  to  capture  R.  sarba  by  these 
methods  during  daylight  reflected  the  offshore  movement 
of  this  species  from  the  shallows  prior  to  dawn  and  a 
far  stronger  targeting  of  bait  by  the  large  numbers  of 
the  banded  toadfish  {Torquigener  pleurogramma)  that 
feed  in  the  offshore  waters  of  the  lower  estuary  during 
the  day.  Because  the  lower  reaches  of  the  Swan  River 
Estuary  act  as  a  shipping  harbor,  alternative  sampling 
methods,  such  as  gill  netting  and  spearing,  could  not 
be  used  to  catch  R.  sarba  during  the  day.  The  data  for 
2000  and  2001  were  augmented  by  those  derived  from 
fish  collected  from  the  same  location  by  using  the  same 
methods  in  1998  and  1999  (Hesp  and  Potter,  2003).  In 
total,  the  results  of  the  present  study  are  based  on  an 
examination  of  over  2000  R.  sarba,  of  which  510  were 


650 


Fishery  Bulletin  102(4) 


Table  1 

Characteristics  of  each  macroscopic  stage  in  the  development  of  the  ovaries  of  Rhabdosargus  sarba,  and  its  corresponding  histo- 
logical characteristics.  Adapted  from  Laevastu  (19651.  Terminology  for  oocyte  stages  follows  Wallace  and  Selman  (1989). 


Stage 


Macroscopic  characteristics 


Histological  characteristics 


I     Virgin 


Ovary  is  very  small  and  strand-like. 


II     Immature 

and  resting 


Small  and  transparent.  Yellowish-orange  in 
color.  Oocytes  not  visible  through  ovarian 
wall. 


Ill     Developing     Slightly  larger  than  stage  II.  Reddish  color. 
Oocytes  visible  through  ovarian  wall. 


Rhabdosargus  sarba  is  a  rudimentary  hermaphrodite,  sensu 
Hesp  and  Potter  (2003).  Thus,  the  gonads  of  small  juveniles 
contain  only  connective  tissue.  Larger  juveniles  possess 
gonads  (ovotestes)  in  which  each  ovarian  lobe  consists  of 
an  immature  ovarian  and  testicular  zone,  separated  by 
connective  tissue.  The  ovotestes  develop  later  into  gonads 
containing  almost  entirely  ovarian  tissue  (functional  ovaries) 
or,  in  the  case  of  males,  gonads  containing  almost  entirely 
testicular  tissue  (functional  testes). 

Ovigerous  lamellae  highly  organized.  Chromatin  nucleolar 
and  perinucleolar  oocytes  dominate  the  complement  of 
oocytes.  Oogonia  sometimes  present.  Chromatin  nucleolar 
oocytes  present  in  all  subsequent  ovarian  stages. 

Chromatin  nucleolar,  perinucleolar  and  cortical  alveolar 
oocytes  present. 

Cortical  alveolar  and  yolk  granule  oocytes  abundant. 


IV  Maturing        Larger   than   stage    III.    Reddish-orange    in 

color.  Yolk  granule  oocytes  visible  through 
ovarian  wall. 

V  Mature  Larger  than  stage  IV  occupying  half  to  two      Yolk  granule  oocytes  predominant. 

thirds  of  body  cavity.  Extensive  capillaries 
visible  in  ovarian  wall. 

VI  Spawning       Hydrated    oocytes    visible    through    ovarian       Migratory  nucleus  oocytes,  hydrated  oocytes,  or  postovula- 

wall.  Note  that  fish  with  ovaries  in  "spawn-       tory  follicles  present, 
ing  condition"  can  only  be  detected  macro- 
scopically  when  caught  during  the  hydration 
period. 


VII     Spent 


Smaller  than  V  and  VI  and  flaccid.  Some  yolk       Some  remnant  yolk  granule  oocytes  present,  all  or  almost  al 
granule  oocytes  visible  through  ovarian  wall.       of  which  are  typically  undergoing  atresia. 


VIII     Spent  and    Small  and  dark  red. 
recovering 


Extensive  scar  tissue  present.  Ovarian  lamellae  becoming 
reorganized.  No  yolk  granule  oocytes  present. 


females  with  stage-V  (mature)  or  stage-VI  (spawning) 
ovaries  (see  Table  1  for  definitions  of  these  stages). 

During  the  above  sampling,  R.  sarba  was  collected 
for  up  to  2  hours  at  intervals  commencing  at  18:30, 
21:30,  00:30,  and  03:30  h  on  1-2  September  2001  and 
for  up  to  2  hours  at  intervals  commencing  at  18:30  and 
22:30  h  on  13  September  2001.  One  of  the  ovarian  lobes 
of  up  to  five  fish  caught  during  each  of  these  above 
time  intervals  was  cut  into  several  pieces,  preserved  in 
10''  neutrally  buffered  formalin  solution  and  used  for 
determining  the  distributions  of  oocyte  diameters  at 
the  above  different  times.  The  other  lobe  was  used  for 
histology  to  determine  the  oocyte  stages  present  in  that 
lobe,  and  thus,  by  extrapolation,  also  the  stages  of  the 
oocytes  in  the  lobe  that  had  been  preserved  in  formalin. 
The  resultant  comparisons  were  used,  in  conjunction 
with  data  from  other  times,  to  elucidate  the  pattern  of 


oocyte  development  during  hydration  and  the  duration 
of  hydration  and  timing  of  ovulation. 

Gonadal  staging  and  histology  of  ovaries 

The  sex,  total  length  (to  the  nearest  1  mm),  and  total 
weight  and  gonad  weight  (to  the  nearest  0.01  g)  of  each 
fish  were  recorded.  From  its  macroscopic  appearance, 
each  gonad  was  assigned  to  one  of  the  following  stages  in 
maturation,  based  on  the  scheme  of  Laevastu  (1965),  i.e., 
I  =  virgin,  II  =  immature  and  resting,  III  =  developing, 
IV  =  maturing,  V  =  mature,  VI  =  spawning,  VII  =  spent, 
VIII  =  spent  and  recovering.  The  corresponding  histolog- 
ical characteristics  of  each  macroscopic  stage  are  shown 
in  Table  1.  When  hydrated  oocytes  could  be  seen  through 
the  ovarian  wall  of  a  fish,  a  note  was  made  as  to  whether 
they  were  distributed  throughout  the  ovary  or  were  in 


Hesp  et  al.:  Timing  and  frequency  of  spawning  and  fecundity  of  Rhabdosargus  sarba 


651 


the  ovarian  duct  and  thus  whether  or  not  ovulation  had 
commenced  at  the  time  of  capture  of  that  fish. 

For  all  histological  studies  of  the  gonads,  part  of  the 
mid  region  of  one  of  the  ovarian  lobes  was  placed  in 
Bouin's  fixative  for  ca.  48  hours,  dehydrated  in  a  series 
of  ethanols,  embedded  in  paraffin  wax,  cut  into  6-um 
sections,  and  stained  with  Mallory's  trichrome.  The  ova- 
ries were  fixed  within  1-3  hours  of  capture  of  the  fish. 

To  test  the  hypothesis  that  R.  sarba  has  indetermi- 
nate fecundity,  the  diameters  of  100  oocytes  in  histo- 
logical sections  of  stage-VI  ovaries  of  two  fish  caught 
during  the  spawning  period  were  measured  to  the  near- 
est 10  /jm  by  using  an  eyepiece  graticule  in  a  compound 
microscope  and  the  stage  of  each  of  those  oocytes  was 
recorded.  Measurements  were  restricted  to  oocytes  in 
which  a  nucleus  was  visible  in  their  center  to  ensure 
that  the  oocytes  had  been  sectioned  through  their  center 
and  that  the  diameters  were  thus  measured  accurately. 
This  approach  could  not  be  used  to  measure  the  oocyte 
diameters  of  hydrated  oocytes  in  histological  sections 
because  the  nucleus  of  these  oocytes  undergoes  germi- 
nal vesicle  breakdown. 

Histological  sections  of  numerous  ovaries  were  used 
to  determine  the  timing  of  the  formation  and  degen- 
eration of  postovulatory  follicles  (POFs).  An  age  was 
assigned  to  the  POFs  found  in  ovaries  of  fish  caught 
at  different  times  of  the  day,  based  on  the  timing  of 
ovulation  and  the  degree  to  which  those  POFs  had 
degenerated  (Hunter  and  Goldberg,  1980;  Hunter  and 
Macewicz,  1985).  Histological  sections  were  also  used 
to  determine  the  relative  abundance  of  the  different 
stages  of  atresia  in  ovaries  at  different  times  during 
the  spawning  period. 

The  jars  containing  the  ovarian  lobes  that  had  been 
preserved  in  formalin  at  the  different  time  intervals  on 
1.  2,  and  13  September  2001  (see  earlier)  were  shaken 
until  the  oocytes  of  each  ovary  had  become  evenly  sus- 
pended in  the  solution.  The  resultant  solution  from 
each  ovary  was  then  passed  through  a  125-/jm  sieve 
to  remove  the  smallest  oocytes,  and  we  were  able  thus 
to  focus  our  study  more  specifically  on  the  vitellogenic 
oocytes.  Comparisons  of  the  appearance  of  the  larger 
oocytes  under  a  dissecting  microscope  with  those  of  the 
different  oocyte  stages  in  histological  sections  of  the 
other  ovarian  lobe  of  the  same  fish  were  used  to  allocate 
the  oocytes  observed  under  the  dissecting  microscope 
to  a  specific  stage  in  oocyte  development.  Each  oocyte 
in  a  representative  subsample  of  100  oocytes  from  each 
formalin-preserved  ovarian  lobe  was  measured  under  a 
dissecting  microscope  with  an  eyepiece  graticule.  This 
approach  enabled  the  diameters  of  hydrated  oocytes  to 
be  measured  accurately,  which  was  not  possible  with 
histological  sections  (see  earlier). 

Categorization  of  stages  in  atresia,  fecundity  estimates, 
and  spawning  frequency 

On  the  basis  of  their  histological  characteristics,  atretic 
oocytes  were  allocated  to  either  the  a  or  /3  stages,  by 
using  the  criteria  of  Hunter  and  Macewicz  (1985).  Mature 


ovaries  were  categorized  according  to  the  proportions 
of  their  a  and  /3  atretic  oocytes  (Hunter  and  Macewicz, 
1985).  Thus,  atretic  state  0  =  ovaries  with  yolked  oocytes 
but  no  a  atretic  oocytes;  atretic  state  1  =  ovaries  in  which 
less  than  509c  of  the  yolked  oocytes  are  in  the  a  stage 
of  atresia;  atretic  state  2  =  ovaries  in  which  less  than 
509c  of  the  yolked  oocytes  are  a  atretic  and  atretic  state 
3  =  ovaries  which  contain  no  yolked  oocytes  but  do  pos- 
sess |3  atretic  oocytes.  During  the  present  study,  atretic 
state  1  ovaries  were  further  divided  into  three  categories 
on  the  basis  of  the  percentage  of  a  atretic  yolk  granule 
oocytes  in  histological  sections,  namely  early  (<10%),  mid 
(10-359?  I  and  late  (36-50%)  atretic  state  1,  an  approach 
similar  to  that  adopted  by  Farley  and  Davis  (1998). 

The  batch  fecundities  of  31  R.  sarba  were  estimated 
from  the  number  of  hydrated  oocytes  in  one  of  the  ovar- 
ian lobes  of  fish  that  had  been  preserved  in  109c  neu- 
trally buffered  formalin.  These  fish  were  chosen  because 
histological  examination  of  their  other  ovarian  lobe  dem- 
onstrated that  the  ovaries  were  in  atretic  state  0  or  early 
state  1,  i.e.,  less  than  10%  of  their  yolk  granule  oocytes 
were  atretic  and  newly  formed  POFs  were  not  present 
(Hunter  et  al.,  1992;  Nichol  and  Acuna,  2001).  The  for- 
malin-preserved ovarian  lobe  was  dried  with  blotting  pa- 
per and  ca.  180-200  mg  of  tissue  was  removed  from  each 
of  its  anterior,  middle,  and  posterior  regions  and  weighed 
to  the  nearest  1  mg.  These  pieces  of  tissue  were  placed  on 
separate  slides,  covered  with  309c  glycerol  and  examined 
under  a  dissecting  microscope.  The  oocytes  were  then 
teased  apart  and  the  number  of  hydrated  oocytes  record- 
ed. The  number  of  hydrated  oocytes  in  each  of  the  three 
pieces  of  ovarian  tissue  of  known  weight  were  then  used, 
in  conjunction  with  the  weight  of  both  ovarian  lobes,  to 
estimate  the  total  number  of  hydrated  oocytes  (=batch 
fecundity)  that  would  have  been  present  in  the  pair  of 
ovarian  lobes  of  each  fish.  The  prevalence  of  spawning  on 
any  given  night  is  expressed  as  the  percentage  of  female 
fish  with  hydrated  eggs  (ovarian  stage  VD  among  all  fe- 
male fish  with  stage-V  I  mature)  and  stage-VI  (spawning) 
ovaries.  These  estimates  were  based  on  an  examination 
of  samples  collected  between  22:00  and  01:30  h,  when  it 
was  possible  to  determine  which  female  fish  were  going 
to  spawn  in  the  ensuing  few  hours  (see  Hunter  et  al., 
1985,  for  further  details  of  this  method). 


Results 

Although  mean  monthly  salinities  in  the  lower  Swan 
River  Estuary  in  late  spring  to  early  winter  were  close 
to  that  of  full  strength  sea  water  (359cc),  they  fell  pre- 
cipitously to  a  minimum  of  23%c  (minimum  individual 
value=14'?c)  in  August,  and  then  rose  sharply  in  early 
to  mid-spring  (Fig.  1). 

Staging  of  the  ovaries  and  confirmation  of 
indeterminate  fecundity 

The  characteristics  of  each  macroscopic  stage  of  the 
ovaries  of  R.  sarba  and  the  corresponding  histologi- 


652 


Fishery  Bulletin  102(4) 


40 


35 


30 


25 


20 


15  L 


80 


60       S 


40 


20 


J  0 


Month 

Figure  1 

Mean  monthly  salinities  I±1SE)  at  the  bottom  of  the  water  column 
in  the  lower  Swan  River  Estuary  throughout  the  year  and  the  preva- 
lences of  atresia  in  mature  ovaries  of  Rhabdosargus  sarba  between 
July  and  November,  which  are  shown  as  histograms,  together  with 
the  number  of  fish  examined.  Closed  rectangles  on  the  horizontal 
axis  refer  to  summer  and  winter  months,  and  the  open  rectangles 
to  autumn  and  spring  months. 


cal  characteristics  are  presented  in  Table  1.  Because 
stages  V  and  VI  could  be  distinguished  macroscopi- 
cally  only  during  the  period  of  oocyte  hydration,  the 
macroscopic  data  for  these  two  stages  had  to  be  com- 
bined for  other  times.  The  diameters  of  the  oocytes  in 
histological  sections  of  an  ovarian  lobe  from  each  of  two 
mature  female  R.  sarba  caught  during  the  spawning 
season — oocyte  diameters  that  were  typical  of  those 
from  mature  R.  sarba  during  this  period — formed  an 
essentially  continuous  distribution  (Fig.  2).  This  distri- 
bution reflected  the  presence  of  oocytes  at  all  stages  in 
development  from  chromatin  nucleolar  oocytes  to  yolk 
granule  oocytes  and  demonstrated  that  R.  sarba  has 
indeterminate  fecundity  sensu  Hunter  et  al.  (1985). 
Thus,  the  potential  annual  fecundity  is  not  fixed  prior 
to  the  commencement  of  the  spawning  period  and  conse- 
quently the  potential  annual  fecundity  of  R.  sarba  has  to 
be  estimated  by  using  a  combination  of  batch  fecundity 
and  spawning  frequency. 

Period  of  hydration  and  spawning 

The  diameters  of  oocytes  in  ovaries  of  fish  collected  at 
intervals  on  1-2  September  2001  and  13  September 
2001  and  which  had  been  retained  on  the  125-^im  sieve, 
produced  a  modal  class  that,  for  each  time  interval, 
fell  between  420  and  600  fim  (Fig.  3).  At  ca.  18:30  h  on 
1  September  2001,  the  oocyte  diameters  formed  a  single 
mode,  and  the  vast  majority  of  oocytes  were  less  than 
720  ;im  and  produced  a  modal  class  at  420-539  um 
(Fig.  3).  However,  by  ca.  21:30  h  on  the  same  evening. 


the  maximum  diameter  of  the  oocytes  had  increased 
markedly  and  the  distribution  of  the  oocyte  diameters 
was  beginning  to  become  bimodal.  with  modal  classes  at 
480-539  and  780-839  ^m.  By  00:30  h  on  2  September, 
the  oocyte  diameter  distributions  had  become  markedly 
bimodal,  and  the  modal  diameter  class  of  the  largest 
oocytes  at  this  time,  and  also  at  03:30  h,  lay  between 
840  and  959  ^(m  (Fig.  3).  The  oocyte  diameter  frequen- 
cies on  13  September  were  essentially  the  same  as  those 
at  similar  times  on  1  September;  the  distributions  were 
unimodal  at  18:30  h  and  bimodal  at  22:30  h  (Fig.  3). 
The  oocyte  diameters  of  each  fish  within  a  given  time 
slot  on  1,  2,  and  13  September  exhibited  essentially  the 
same  distribution. 

Histological  sections  showed  that,  at  18:30  h  on 
1  September  2001,  most  of  the  mature  ovaries  contained 
migratory  nucleus  stage  oocytes,  i.e.,  oocytes  in  which 
the  nucleus  was  migrating  towards  the  edge  of  the  cy- 
toplasm and  a  conspicuous  lipid  droplet  was  present  in 
the  cytoplasm  (Fig.  4A).  However,  it  was  difficult  at  this 
time  to  distinguish  migratory  nucleus  oocytes  from  yolk 
granule  oocytes  under  a  dissecting  microscope  (Fig.  4B). 
By  21:30  h,  the  yolk  and  lipid  of  the  larger  oocytes  had 
begun  to  coalesce  and  the  nucleus  could  sometimes  be 
seen  near  the  edge  of  the  cytoplasm  (Fig.  4Cl.  Their 
relatively  larger  size,  translucent  appearance,  and  one's 
ability  to  detect  their  lipid  droplet  enabled  these  hydrat- 
ing  oocytes  to  be  far  more  readily  distinguished  from 
yolk  granule  oocytes  under  a  dissecting  microscope 
than  was  the  case  earlier  in  the  evening  (cf.  Fig.  4,  B 
and  D).  By  00:30  h.  the  largest  oocytes  had  increased 


Hesp  et  al.:  Timing  and  frequency  of  spawning  and  fecundity  of  Rhabdosargus  sarba 


653 


further  in  size  and  all  of  their  lipid  and  yolk  material 
had  coalesced  (Fig.  4E).  Under  the  dissecting  micro- 
scope, these  hydrated  oocytes  were  of  similar  appear- 
ance to  the  corresponding  oocytes  at  21:30  h  (Fig.  4F). 
Although  mature  fish  with  ovaries  containing  the  above 
stages  in  oocyte  hydration  were  frequently  found  in 
nearshore  shallow  waters,  the  numbers  of  such  fish  in 
these  waters  declined  markedly  after  about  00:30  h 
and  none  of  the  few  fish  caught  there  after  this  time 
contained  recently  formed  POFs.  However,  fish  with 
ovaries  containing  newly  formed  POFs  were  caught  in 
offshore  deeper  waters. 

Histological  examination  demonstrated  that,  when 
hydrated  oocytes  were  present  in  the  ovarian  duct, 
the  ovary  contained  recently  formed  POFs,  which  are 
formed  by  the  thecal  and  granulosa  layers  of  the  oocytes 
that  surround  the  zona  radiata  externa  (Fig.  5A).  Newly 
formed  POFs  (0-6  h  old)  possess  a  conspicuous  lumen 
and  their  granulosa  cells  contain  prominent  darkly 
stained  nuclei  (Fig.  5B).  These  newly  formed  postovula- 
tory  follicles  were  first  observed  in  the  ovaries  of  females 
caught  at  ca.  01:30  h  and  were  present  in  the  ovaries  of 
several  fish  caught  in  the  ensuing  four  hours.  In  con- 
trast, no  newly  formed  POFs  were  found  in  the  ovaries 
of  R .  sarba  at  dusk,  i.e.,  ca.  18:30  h.  At  this  time,  the 
POFs  comprised  one  of  two  morphological  forms.  The 
first  and  least  degenerate  form  was  less  well  organized 
than  newly  formed  POFs  and  its  nuclei  were  becoming 
pycnotic  (Fig.  5C);  the  second  form  was  smaller  and 
highly  degenerate  and  its  nuclei  had  become  far  less 
visible  or  undetectable  (Fig.  5D).  The  least  degenerate 
of  the  two  forms  of  POFs  in  ovaries  of  fish  caught  at  ca. 
22:00  and  01:00  h  (Fig.  5,  D  and  E)  represents  stages  in 
degeneracy  that  are  intermediate  between  those  of  the 
two  different  forms  described  above  for  the  ovaries  of 
fish  caught  at  18:30  h.  These  POFs  were  thus  compact 
and,  although  some  of  their  nuclei  were  still  detectable, 
they  were  markedly  pycnotic. 


Chromatin  nucleolar  oocytes 
Perinucleolar  oocytes 
D  Cortical  alveolar  oocytes 
Yolk  granule  oocytes 


30 


20 


10 


■  Chromatin  nucleolar  oocytes 

E3  Perinucleolar  oocytes 

fJJ  Cortical  alveolar  oocytes 

B  Yolk  granule  oocytes 


0  100  200  300  400  500 

Oocyte  diameter  (urn) 

Figure  2 

Percent  frequency  distributions  for  the  oocyte  diameters 
of  different  oocyte  stages  in  histological  sections  of  stage- 
VI  ovaries  of  two  female  Rhabdosargus  sarba. 


Influence  of  salinity  and  tides  on  spawning 

Both  a  and  /5  atretic  oocytes  were  frequently  observed 
in  the  ovaries  of  R.  sarba.  The  chorion  (zona  radiata) 
of  the  early  a  atretic  vitellogenic  oocyte  was  distorted, 
fragmented,  and  had  moved  inwards  (Fig  6A).  By  the  /3 
atretic  stage,  the  yolk  and  lipid  had  been  resorbed  and 
a  large  proportion  of  the  oocyte  volume  was  occupied  by 
vacuoles  (Fig.  6B). 

Sixty-two  percent  and  72%  of  the  stage-V  and  stage- 
VI  ovaries  sectioned  in  July  and  August,  respectively, 
were  at  mid  or  late  atretic  state  1,  i.e.,  11-50%  of  their 
yolk  granule  oocytes  were  a  atretic  (Fig.  1).  However, 
the  prevalence  of  these  mid-late  state-1  ovaries  declined 
precipitously  to  28%  in  September,  as  salinities  rose 
markedly,  and  remained  at  a  similar  level  until  the  end 
of  spawning  in  late  November. 

Histological  sections  showed  that,  in  July  and  August, 
only  39r/c  of  the  57  pairs  of  ovarian  lobes  of  R.  sarba 
that  were  macroscopically  assigned  as  stage  V  and 
stage  VI  contained  migratory  nucleus  oocytes,  hydrated 


oocytes,  or  POFs,  i.e.,  were  at  stage  VI.  However,  in  the 
following  two  months,  76%  of  the  88  pairs  of  ovarian 
lobes  of  R.  sarba,  that  were  macroscopically  assigned 
as  stage  V  or  stage  VI,  were  shown  by  histology  to  be 
at  stage  VI. 

During  September,  when  spawning  activity  was  great- 
est, the  prevalence  of  spawning  (PS)  was  positively 
correlated  (P<0.05)  with  maximum  daily  tidal  height 
(T).  PS  =  91.72T  +  20.73  (/-2  =  0.46,  number  of  sampling 
occasions=10)  (Fig.  7A). 

Data  for  the  same  days  as  those  used  to  provide  the 
points  shown  in  Fig.  7A  demonstrated  that  the  preva- 
lence of  spawning  (PS)  is  inversely  correlated  with  the 
difference  in  hours  between  the  time  when  spawning  is 
believed  to  cease  (ca.  06:00  h,  see  later)  and  the  time 
of  high  tide.  PS  =  -8.26(T)  +  78.22  (r2=0.49,  number  of 
sampling  occasions=10)  (Fig.  7B).  Thus,  the  prevalence 
of  these  "spawning"  females  was  greatest  on  those  days 
when  the  time  that  the  tide  was  about  to  change  from 
flood  to  ebb  coincided  with  the  time  when  R.  sarba  is 
considered  to  cease  spawning. 


654 


Fishery  Bulletin  102(4) 


1  and  2  September 


1  and  2  September 


30  r 
20 

10 

0 
30 

20 

10 


g-     30 
lL 

20 

10 

0 
30 

20 

10 


0  L 


18:30  h 
n  =  5 


Ql 


i i i — i — i — i — i — t — i — i — i 


21:30h 
n  =  5 


tn= 


00:30  h 
rt  =  5 


hJ 


i i — i — 1_ 


_j — i — i — i 


03:30  h 
n  =  2 


r 


*<P^<&<&4PoPjP<&<& 


■<?  $><W 


13  September 
13  September 


18:30  h 
n  =  5 


Hliliri 


_i i i — i — i — i — i — i 


22:30  h 
n=5 


M 


0= 


\?  it  nr  t»°  or  A1'  <tr  qP^^Tr 


'.c?^ 


Oocyte  diameter  (jim) 


Figure  3 

Frequency  distributions  for  the  oocyte  diameters  in  mature  ovaries  of 
Rhabdosargus  sarba  caught  on  1-2  and  13  September  2001.  The  ovaries  had 
been  preserved  in  formalin  and  their  oocytes  had  been  passed  through  a 
125-^m  sieve.  The  time  of  commencement  of  each  2-hour  sampling  interval 
is  shown,  n  =  number  offish  used  for  oocyte  diameter  measurements. 


Batch  fecundity,  spawning  frequency, 
and  potential  annual  fecundity 

The  relationship  between  batch  fecundity  iBF)  and 
total  length  iTL)  shown  in  Figure  8,  and  between  batch 
fecundity  and  somatic  weight  (W)  are  described  by  the 
following  equations: 

InBF  =  5.00251nrL-17.557 

(P<0.001;  r2=0.52,  n  =  30), 
BF  =  1997e00105W 

(P<0.001;  r2=0.55,  n=30). 


The  batch  fecundities  of  R.  sarba.  predicted  by  the  above 
equations  for  fish  with  lengths  of  180,  220.  and  260  mm, 
were  ca.  4500,  7700.  and  12,400,  eggs,  respectively,  and 
for  fish  with  somatic  weights  of  100,  150,  and  200  g  were 
ca.  5700,  9600,  and  16,300  eggs,  respectively.  The  average 
daily  prevalence  of  spawning  during  the  spawning  period 
was  36.5^.  Thus,  during  this  period,  individual  females 
spawned,  on  average,  once  every  2.7  days  and  therefore 
about  45  times  during  the  spawning  season.  The  poten- 
tial annual  fecundities  of  female  R.  sarba  with  lengths 
of  180,  220.  and  260  mm  were  thus  estimated  to  be  ca. 
204,300,  346,100,  and  557,500  eggs,  respectively. 


Hesp  et  al.:  Timing  and  frequency  of  spawning  and  fecundity  of  Rhabdosargus  sarba 


655 


Figure  4 

Histological  sections  of  ovaries  of  individual  Rhabdosargus  sarba  caught  on 
1  and  2September  2001  at  ca.  18:30  h  (A),  21:30  h  (Cl  and  00:30  h  (E)  and 
photographs  of  the  oocytes  from  the  other  lobe  of  the  ovary  of  the  same  three 
fish  (B,  D  ,Fi.  c  =  coalescing  yolk  and  lipid;  ho=hydrating  oocyte;  l  =  lipid 
droplet;  mn=migratory  nucleus  oocyte;  n  =  nucleus;  yg=yolk  granule  oocyte. 
Scale  bars  =  200  fim  in  A,  C,  and  D  and  250  jim  in  B,  D,  and  F. 


Discussion 

Oocyte  hydration,  ovulation,  and  spawning  periods 

Because  histological  studies  showed  that  the  ovaries  of 
numerous  fish  caught  on  different  occasions  between 
18:30  and  20:30  h  did  not  contain  recently  formed  POFs, 
we  deduced  that  these  fish  had  not  spawned  in  the 
previous  few  hours.  However,  at  this  time,  the  ovaries 
of  many  fish,  that  were  designated  macroscopically  as 
at  stage  V  and  stage  VI,  often  contained  numerous 
migratory  nucleus-stage  oocytes  and,  towards  the  end 
of  this  period,  often  a  few  oocytes  in  the  early  stages  of 
hydration.  Although  the  frequency  distributions  of  the 
oocyte  diameters  offish  examined  on  both  1  and  13  Sep- 


tember were  still  unimodal  at  18:30  to  20:30  h,  they  had 
become  bimodal  by  21:30  to  23:30  h  (Fig.  3),  reflecting 
the  fact  that,  by  this  time,  numerous  oocytes  had  become 
markedly  enlarged  through  hydration.  The  above  data 
demonstrate  that  hydration  typically  commences  soon 
after  dusk.  Furthermore,  because  a  number  of  J?,  sarba 
caught  between  01:30  and  04:30  h,  and  particularly 
towards  the  end  of  this  time  interval,  contained  ovaries 
undergoing  ovulation  and  had  newly  formed  POFs,  the 
period  between  the  onset  of  hydration  and  commence- 
ment of  ovulation  typically  lasts  about  7-10  hours,  which 
is  very  similar  to  the  duration  estimated  for  species 
such  as  the  black  sea  anchovy  [Engraulis  encrasicholus) 
(Lisovenko  and  Andrianov,  1991)  and  ballyhoo  (Hemir- 
amphus  brasiliensis)  (McBride  et  al.,  2003).  Although  we 


656 


Fishery  Bulletin  102(4) 


Figure  5 

Histological  sections  through  ovaries  of  Rhabdosargus  sarba  showing 
lAl  the  outer  layers  of  a  yolk  granule  oocyte,  and  postovulatory  follicles  in 
the  ovaries  offish  collected  at  (B)  ca.  02:00,  (C  and  D)  ca.  18:30,  (E)  ca. 
22:00  and  ca.  (F)  01:00  h.  g=granulosa  layer;  lu  =  lumen;  t  =  thecal  layer; 
yg=yolk  granule;  yv=yolk  vesicle;  zre  =  zona  radiata  externa.  Scale  bars  = 
25  ftm  in  A,  and  50  pra  in  B-F. 


caught  several  R.  sarba  with  new  POFs  in  their  ovaries 
and  hydrated  oocytes  in  their  oviducts,  we  were  able  to 
catch  only  one  individual  of  this  species  in  which  the 
ovaries  possessed  new  POFs  and  no  hydrated  oocytes. 
The  latter  fish,  which  had  clearly  just  completed  spawn- 
ing, was  caught  between  05:00  and  06:00  h. 

Several  species  are  known  typically  to  complete 
spawning  in  the  10-14  hours  after  the  time  when  their 
oocytes  commence  hydration,  e.g.,  the  northern  anchovy 
(Engraulis  mordax)  (Hunter  and  Macewicz,  1985),  the 
spotted  seatrout  (Cynoscion  nebulosus)  (Brown-Peterson 
et  al.,  1988)  and  the  horse  mackerel  (Trachurus  trachu- 
rus)  (Karlou-Riga  and  Economidis,  1997).  Furthermore, 
spawning  is  typically  completed  2-5  hours  after  the 
commencement  of  ovulation,  e.g.,  the  spotted  seatrout 


(Cynoscion  nebulosus)  (Brown-Peterson  et  al.,  1988), 
the  Black  Sea  anchovy  [Engraulis  encrasicholus)  (Lisov- 
enko  and  Andrianov,  1991)  and  the  weakfish  {Cynoscion 
regalis)  (Taylor  and  Villoso,  1994).  These  consistent 
data,  when  considered  in  conjunction  with  the  similar 
duration  of  hydration  of  R.  sarba,  and  the  capture  of  a 
very  recently  spawned  fish  between  05:00  and  06:00  h. 
provide  very  strong  circumstantial  evidence  that,  in 
the  lower  Swan  River  Estuary,  R.  sarba  spawns  mainly 
between  02:00  and  06:00  h. 

The  newest  POFs  in  the  ovaries  of  R.  sarba  caught 
at  dusk,  i.e.,  at  18:30  h,  had  degenerated  to  an  extent 
similar  to  those  of  ca.  12-h-old  POFs  in  the  ovaries  of 
other  species,  e.g.  the  skipjack  tuna  (Katsuwonus  pela- 
mis)  (Hunter  et  al.,  1986)  and  the  whitemouth  croaker 


Hesp  et  al.:  Timing  and  frequency  of  spawning  and  fecundity  of  Rhabdosargus  sarba 


657 


^m^ 


Figure  6 

Histological  sections  of  ovaries  of  Rhabdosargus  sarba  showing  an  oocyte  at 
the  (A)  a  and  (B)  /3  stages  of  atresia.  c  =  chorion;  t= thecal  layer;  v= vacuole. 
Scale  bars  =  100  pm  in  A  and  50  j.tm  in  B. 


iMicropogonias  furnieri)  (Macchi  et  al.,  2003).  This 
finding  provides  further  evidence  that  R.  sarba  spawns 
close  to  dawn.  Certainly,  the  state  of  degeneration  of  the 
newest  POFs  in  mature  ovaries  of  R.  sarba  at  dusk  pro- 
vides very  strong  circumstantial  evidence  that  spawn- 
ing could  not  have  occurred  during  at  least  most  of  the 
previous  daylight  hours. 

Our  results  demonstrate  that  the  prevalence  of  fish 
with  ovaries  at  mid  to  late  atretic  state  1  declined 
precipitously  as  salinities  increased  from  their  winter 
minima  in  July  and  August  and  that  this  decrease 
was  accompanied  by  an  increase  in  the  prevalence  of 
migratory  nucleus  oocytes,  hydrated  oocytes  or  POFs. 
The  implication  that  the  oocytes  of  R.  sarba  are  often 
inhibited  from  undergoing  final  oocyte  maturation  when 
salinities  are  low  parallels  the  conclusions  drawn  for 
the  influence  of  salinity  on  the  gonadal  development  of 
Cynoscion  nebulosus  in  estuaries  entering  the  Gulf  of 
Mexico  (Brown-Peterson  et  al.,  2002).  The  resorption  of 
yolk  granule  oocytes  by  the  ovaries  of  R.  sarba  in  July 
and  August  would  help  conserve  energy  at  a  time  when, 
if  those  oocytes  progressed  through  to  final  maturation 
and  were  released,  they  would  be  exposed  to  salinities 
that  are  known  to  be  lower  than  those  required  for  op- 
timal development  (Mihelkakis  and  Kitajima,  1994). 

Frequency  of  spawning 

Because  ovulation  lasts  for  ca.  2-5  hours,  the  POFs  that 
were  present  in  ovulating  ovaries  and  that  showed  no 
detectable  signs  of  degeneration  were  presumably  <3 
hours  old.  It  then  follows  that,  when  POFs  at  intermedi- 
ate and  advanced  stages  of  degeneration  were  present 
in  those  same  ovaries,  those  POFs  were  presumably  ca. 
24  and  ca.  48  hours  old,  respectively.  Thus,  the  pres- 
ence of  these  three  very  distinct  forms  of  POFs  in  the 
same  ovary  of  a  fish  implies  that  individual  R.  sarba  are 
capable  of  spawning  on  at  least  three  successive  days 
during  that  part  of  the  month  when  spawning  activity 
is  greatest. 


The  estimated  average  frequency  of  spawning  by  R. 
sarba,  i.e.,  once  every  2.7  days,  is  essentially  the  same 
as  that  recorded  for  several  other  species,  including  e.g., 
spotted  seatrout  (Cynoscion  nebulosus)  (Brown-Peter- 
son et  al.,  1988.  2002),  red  drum  iSciaenops  ocellatus) 
(Wilson  and  Nieland,  1994),  and  common  snook  (Centro- 
pomus  undecimalis)  (Taylor  et  al.,  1998).  The  resultant 
conclusion  that  R.  sarba  spawns  about  45  times  during 
a  spawning  period  is  comparable  to  that  estimated  for 
black  sea  anchovy  (Engraulis  encrasicholus)  (Lisovenko 
and  Andrionov,  1991)  and  cobia  (Rachycentron  canadum) 
(Brown-Peterson  et  al.,  2001).  However,  spawning  fre- 
quency does  vary  markedly  among  species. 

Relationship  between  spawning  time  and  tidal  cycle 

Although  seine  netting  between  00:30  and  05:30  h  on  a 
number  of  days  yielded  no  female  R.  sarba  with  newly 
formed  POFs,  rod-and-line  angling  in  deeper  water 
between  01:30  and  04:30  h  yielded  several  females  in 
which  the  ovaries  contained  both  newly  formed  POFs 
and  concentrations  of  hydrated  oocytes  in  their  oviducts, 
and  also  running  ripe  males.  This  finding  provides 
strong  circumstantial  evidence  that,  just  prior  to  ovula- 
tion, R.  sarba  moves  from  nearshore  shallows  to  offshore 
deeper  waters. 

Because  R.  sarba  typically  spawns  just  prior  to  the 
commencement  of  a  relatively  strong  ebb  tide,  the  fer- 
tilized eggs  would  likely  be  transported  downstream 
and  out  of  the  estuary.  The  conclusion  that  eggs  are 
swept  out  of  the  estuary  is  supported  by  the  fact  that 
only  15  larvae  of  R.  sarba  were  caught  during  extensive 
sampling  of  the  lower  Swan  River  Estuary  and  that 
virtually  all  of  these  larvae  were  caught  at  its  mouth 
(Gaughan  et  al.,  1990).  A  downstream  movement  of 
eggs  would  be  further  facilitated  by  R.  sarba  spawn- 
ing in  deeper  waters,  where  the  current  is  greatest. 
Emigration  of  eggs  from  the  estuary  would  enhance 
the  chances  of  survival  of  the  eggs  of  this  essentially 
marine  species  by  ensuring  that  they  would  develop  in 


658 


Fishery  Bulletin  102(4) 


60 


40 


I  ° 


0  0.1  0  2  0  3  0  4  0  5  0  6  0  7 

Predicted  maximum  change  in  daily  tidal  level  (m) 


B 


Number  of  hours  between  high  tide  and 
estimated  time  of  spawning  completion  (06:00  h) 

Figure  7 

Relationship  between  prevalence  of  spawning  and 
(A)  the  predicted  maximum  change  in  daily  tidal  level 
and  (Bi  the  difference  between  the  time  of  high  tide 
and  the  time  at  which  spawning  in  Rhabdosargus  sarba 
is  estimated  to  be  completed  in  the  lower  Swan  River 
Estuary  I  06:00  h).  Each  point  represents  the  data  for 
a  separate  sampling  occasion. 


11.5 

11.0 

m 

>,    105 

§     100 

o 

CD 

~      9.5 
o 

1      9.0 

9 

■   :  \^^  •  * 

c 

^^-""» 

8.5 

•     • 

• 

8.0 

• 

7.5 

S.2                         5.3                        54                         55                        5.6 

In  total  length  (mm) 

Figure  8 

Relationship  between  batch  fecundity  (  =  number  of 

hydrated  oocytes)  and  total  length  Imml  for  Rhabdosar- 

gus  sarba. 

Spratelloicles  robustus  was  particularly  numerous  in 
some  of  our  seine-net  catches,  a  movement  of  the  eggs 
ofi?.  sarba  out  of  the  estuary  would  also  enhance  their 
chances  of  avoiding  predation  by  that  species. 

A  downstream  transport  of  eggs  would  account  for 
the  relatively  few  young  0+  juveniles  that  are  recruited 
into  the  nearshore  shallow  waters  of  the  estuary  ( Hesp 
et  al.,  2004).  Indeed,  substantial  recruitment  into  these 
nearshore  waters,  presumably  as  a  result  of  immigra- 
tion from  coastal  marine  waters,  does  not  occur  until 
R.  sarba  is  about  one  year  old  and  about  140  mm  in 
length  (Hesp  et  al..  2004).  Because  R.  sarba  settles  at  a 
length  of  ca.  12  mm  (Hesp  et  al.,  2004)  and  ca.  30  days 
of  age  (Neira1),  this  immigration  back  into  the  estuary 
does  not  occur  until  11  months  after  settlement.  In 
contrast  to  the  situation  in  the  Swan  River  Estuary,  R. 
sarba  elsewhere  typically  spawns  in  marine  waters  and 
their  larvae  often  enter  estuaries  on  flood  tides  (e.g.. 
Miskiewicz,  1986;  Neira  and  Potter,  1992). 


a  marine  environment  in  which  salinity  remained  con- 
stantly at  ca.  35'</,  rather  than  in  one  in  which  sudden 
rainfall  could  result  in  sudden  marked  declines  in  salin- 
ity. However,  the  possession  of  spawning  cycles  linked 
to  lunar  and  tidal  periodicities  can  reduce  the  likeli- 
hood of  predation  (Taylor,  1984).  For  example,  Johannes 
(1978)  pointed  out  that,  because  the  spawning  of  many 
reef-dwelling  fishes  is  synchronized  with  the  lunar  cycle 
and  occurs  on  high  or  ebbing  tides,  their  eggs  would  be 
transported  away  from  reefs,  where  the  concentration 
of  predators  is  high,  and  consequently  the  likelihood 
of  predation  during  the  early  stages  of  life  would  be 
reduced.  Because  planktivorous  fishes  are  abundant  in 
estuaries  (Johnson  et  al.,  1990;  Morgan,  1990),  includ- 
ing the  Swan  River  Estuary  where  the  planktivorous 


Potential  annual  fecundity 

The  estimates  of  potential  annual  fecundity  derived  for 
R.  sarba  during  the  present  study,  which  ranged  from 
109,000  to  2,417,000  eggs  for  fishes  of  188  and  266  mm 
total  length,  respectively,  greatly  exceed  those  of  El- 
Agamy  (1989),  which  ranged  from  23,000  to  99,000  eggs 
for  fishes  of  170  and  260  mm  total  length,  respectively. 
However,  because  El-Agamy  ( 1989 )  based  his  estimates 
on  the  number  of  large  oocytes  present  in  the  ovaries  of 
individual  R.  sarba,  he  did  not  take  into  account  the  fact 


1  Neira.  F.  J.  2004.  Personal  commun.  Australian  Maritime 
College.  Faculty  of  Fisheries  and  Marine  Environment. 
PO  Box  21,  Beaconsfield.  Tasmania  7270.  Australia. 


Hesp  et  al.:  Timing  and  frequency  of  spawning  and  fecundity  of  Rhabdosargus  sarba 


659 


that  this  species  has  indeterminate  fecundity.  Thus,  the 
values  recorded  for  the  annual  fecundity  of  R.  sarba  in 
the  Arabian  Gulf  almost  certainly  represent  a  marked 
underestimate  of  the  true  annual  fecundity  of  this  spe- 
cies in  that  region. 


Acknowledgments 

Our  gratitude  is  expressed  to  colleagues  and  friends 
for  help  with  sampling — to  G.  Thomson  for  preparing 
the  histological  sections,  to  F.  J.  Neira  for  information 
on  the  larval  phase  of  R.  sarba,  and  to  D.  Fairclough 
and  two  anonymous  reviewers  for  invaluable  comments 
on  the  manuscript.  Financial  support  was  provided  by 
Murdoch  University. 


Literature  cited 

Brown-Peterson,  N.  J.,  R.  M.  Overstreet,  J.  M.  Lotz, 
J.  S.  Francis,  and  K.  M.  Burns. 

2001.  Reproductive  biology  of  cobia,  Rachycentron 
canadum.  from  coastal  waters  of  the  southern  United 
States.     Fish.  Bull.  99:15-28. 

Brown-Peterson,  N.  J.,  M.  S.  Peterson,  D.  L.  Nieland, 
M.  D.  Murphy,  R.  D.  Taylor,  and  J.  R.  Warren. 

2002.  Reproductive  biology  of  female  spotted  seatrout, 
Cynoscion  nebulosus,  in  the  Gulf  of  Mexico:  differences 
among  estuaries?     Environ.  Biol.  Fishes  63:405-415. 

Brown-Peterson,  N.,  P.  Thomas,  and  C.  R.  Arnold. 

1988.  Reproductive  biology  of  the  spotted  seatrout, 
Cynoscion  nebulosus,  in  south  Texas.  Fish.  Bull.  86:373- 
388. 

Buxton,  C.  D..  and  P.  A.  Garratt 

1990.     Alternative  reproductive  styles  in  seabreams  ( Pisces: 
Sparidae).     Environ.  Biol.  Fishes  28:113-124. 
El-Agamy,  A.  E. 

1989.  Biology  of  Sparus  sarba  Forskal  from  the  Qatari 
water,  Arabian  Gulf.  J.  Mar.  Biol.  Assoc.  India.  31:129- 
137. 

Farley,  J.  H.,  and  T.  L.  O.  Davis. 

1998.     Reproductive  dynamics  of  southern  bluefin  tuna, 
Thunnus  maccoyii.     Fish.  Bull.  96:223-236. 
Gaughan,  D.  J.,  F.  J.  Neira,  L.  E.  Beckley  and  I.  C.  Potter. 

1990.  Composition,  seasonality  and  distribution  of  the  ich- 
thyoplankton  in  the  lower  Swan  Estuary,  south-western 
Australia.     Aust.  J.  Mar.  Freshw.  Res.  41:529-543. 

Garratt.  P.  A. 

1993.     Comparative  aspects  of  the  reproductive  biology  of 
seabreams  (Pisces:  Sparidae).     Ph.D.  thesis,  vol.  1,  175 
p.     Rhodes  Univ..  Grahamstown,  South  Africa. 
Greeley,  M.  S„  D.  R.  Marion,  and  R.  MacGregor  III. 

1986.     Semilunar  spawning  cycles  of  Fundulus  similes 
( Cyprinodontidae).     Environ.   Biol.   Fishes    17:125- 
131. 
Hesp,  S.  A.,  N.  G.  Hall  and  I.  C.  Potter. 

2004.     Size-related  movements  of  Rhabdosargus  sarba 
in  three  different  environments  and  their  influence  on 
estimates  of  von  Bertalanffy  growth  parameters.     Mar. 
Biol. 
Hesp,  S.  A.,  and  I.  C.  Potter. 

2003.  Reproductive  biology  of  the  tarwhine  Rhabdos- 
argus sarba  (Sparidae)  in  Western  Australian  waters, 


in  which  it  is  a  rudimentary  hermaphrodite      J.  Mar. 
Biol.  Assoc.  U.K.  83:1333-1346. 
Hoque,  M.  M.,  A.  Takemura,  M.  Matsuyama,  S.  Matsuura, 
and  K.  Takano. 

1999.     Lunar  spawning  in  Siganus  eanaliculatus.     J.  Fish 
Biol.  55:  1213-1222. 
Hunter  J.  R.,  and  S.  R.  Goldberg. 

1980.     Spawning  incidence  and  batch  fecundity  in  northern 
anchovy,  Engraulis  mordax.     Fish.  Bull.  77:641-652. 
Hunter,  J.  R.,  N.  C.  H.  Lo,  and  R.  J.  H.  Leong. 

1985.  Batch  fecundity  in  multiple  spawning  fishes.  In 
An  egg  production  method  for  estimating  spawning  bio- 
mass  of  pelagic  fish:  application  to  the  northern  anchovy 
(Engraulis  mordax)  (R.  Lasker,  ed.l,  p.  67-77.  NOAA 
Tech.  Rep.  NMFS  36. 
Hunter,  J.  R.,  and  B.  J.  Macewicz. 

1985.  Measurement  of  spawning  frequency  in  multiple 
spawning  fishes.  In  An  egg  production  method  for  esti- 
mating spawning  biomass  of  pelagic  fish:  application  to 
the  northern  anchovy  (Engraulis  mordax)  (R.  Lasker, 
ed.i,  p.  79-94.     NOAA  Tech.  Rep.  NMFS  36. 

Hunter,  J.  R.,  B.  J.  Macewicz,  L.  N.  Chyan-huei,  and 
C.  A.  Kimbrell. 

1992.  Fecundity,  spawning,  and  maturity  of  female  Dover 
sole  Mierostomus  paeificus,  with  an  evaluation  of  assump- 
tions and  precision.     Fish.  Bull.  99:101-128. 

Hunter  J.  R.,  B.  J.  Macewicz,  and  J.  R.  Sibert. 

1986.  The  spawning  frequency  of  skipjack  tuna  Katsu- 
wonus  pelamis,  from  the  South  Pacific.  Fish.  Bull. 
84:895-903. 

Johannes,  R.  E. 

1978.     Reproductive  strategies  of  coastal  marine  fishes 
in  the  tropics.     Environ.  Biol.  Fishes  3:65-84. 
Johnson,  W.  S.,  D.  M.  Allen,  M.  V.  Ogburn,  and  S.  E.  Stancyk. 

1990.  Short-term  predation  responses  of  adult  bay 
anchovies  Anehoa  mitchilli  to  estuarine  zooplankton 
availability.     Mar.  Ecol.  Prog.  Ser.  64:55-68. 

Karlou-Riga  C,  and  P.  S.  Economidis. 

1997.  Spawning  frequency  and  batch  fecundity  of  horse 
mackerel,  Trachurus  trachurus  (L.),  in  the  Saronikos 
Gulf  (Greece).     J.  Appl.  Ichthyol.  13:97-104. 

Kuiter,  R.  H. 

1993.  The  complete  diver's  and  fishermen's  guide  to 
coastal  fishes  of  south-eastern  Australia.  437  p.  Craw- 
ford House  Press,  Bathurst,  Australia. 

Laevastu,  T. 

1965.     Manual  of  methods  in  fisheries  biology.     FAO, 
Rome. 
Lisovenko.  L.  A.,  and  D.  P.  Andrianov. 

1991.  Determination  of  absolute  fecundity  of  intermit- 
tently spawning  fishes.     Vop.  Ikhtiol.  31:631-641. 

Macchi,  G.  J.,  E.  M.  Acha  and  M.  I.  Militelli. 

2003.     Seasonal  egg  production  of  whitemouth  croaker 
(Micropogonias  furnieri)  in  the  Rio  de  la  Plata  estuary. 
Argentina-Uruguay.     Fish.  Bull.  101:332-342. 
McBride,  R.  S.,  J.  R.  Styer  and  R.  Hudson. 

2003.     Spawning  cycles  and  habitats  for  ballyhoo  (Hemir- 
amphus  brasiliensis)  and  balao  (H.  balao)  in  south 
Florida.     Fish.  Bull.  101:583-589. 
Mihelkakis,  A.,  and  C.  Kitajima. 

1994.  Effects  of  salinity  and  temperature  on  incubation 
period,  hatching  rate,  and  morphogenesis  of  the  silver 
sea  bream,  Sparus  sarba  (Forskal,  1775).  Aquaculture 
126:361-371. 

Miskiewicz,  A.  G. 

1986.     The  season  and  length  at  entry  into  a  temper- 


660 


Fishery  Bulletin  102(4) 


ate  Australian  estuary  of  the  larvae  of  Acanthopagrus 
australis,  Rhabdosargus  sarba  and  Chrysophrys  aura- 
tus  iTeleostei:  Sparidaei.  In  Indo-Pacific  fish  biology: 
proceedings  of  the  second  international  conference  on 
Indo-Pacific  fishes  (T.  Uyeno,  A.  R.  Taniuchi  and  K. 
Matsuura,  eds.),  p.  740-747.  Ichthyological  Society  of 
Japan,  Tokyo. 

Morgan,  S.  G. 

1990.  Impact  of  planktivorous  fishes  on  dispersal,  hatch- 
ing, and  morphology  of  estuarine  crab  larvae.  Ecology 
71:1639-1652. 

Neira,  F.  J.,  and  I.  C.  Potter. 

1992.  Movement  of  larval  fishes  through  the  entrance  chan- 
nel of  a  seasonally  open  estuary  in  Western  Australia. 
Estuar.  Coast.  Shelf  Sci.  35:213-224. 

Nichol,  D.  G.,  and  E.  I.  Acuna. 

2001.  Annual  and  batch  fecundities  of  yellowfin  sole, 
Limanda  aspera,  in  the  eastern  Bering  Sea.  Fish. 
Bull.  99:108-122. 

Potter,  I.  C.,  and  G.  A.  Hyndes. 

1999.  Characteristics  of  the  ichthyofauna  of  southwest- 
ern Australian  estuaries,  including  comparisons  with 
holarctic  estuaries  and  estuaries  elsewhere  in  temperate 
Australia:  areview.     Aust.  J.  Ecol.  24:395-421. 

Schwassmann,  H.  O. 

1971.     Biological  rhythms.     In  Fish  physiology,  vol.  VI  (W. 

S.  Hoar  and  D.  J.  Randall,  eds. ),  p  371-428.     Academic 

Press  Inc.,  New  York,  NY. 
Smith,  K.  A.,  and  I.  M.  Suthers. 

2000.  Consistent  timing  of  juvenile  fish  recruitment 
to  seagrass  beds  within  two  Sydney  estuaries.  Mar. 
Freshw.  Res.  51:765-776. 

Spencer,  R.  S. 

1956.  Studies  in  Australian  estuarine  hydrology.  2.  The 
Swan  River.     Aust.  J.  Mar.  Freshw.  Res.  7:193-253. 


Taylor,  M.  H. 

1984.     Lunar   synchronisation   of  fish   reproduction. 
Trans.  Am.  Fish.  Soc.  113:484-493. 
Taylor,  M.  H„  and  L.  DiMichele. 

1980.     Ovarian  changes  during  the  lunar  spawning  cycle 
of  Fundulus  heteroclitus.     Copeia  1980:118-125. 
Taylor,  M.  H.,  and  E.  P.  Villoso. 

1994.     Daily  ovarian  cycles  in  weakfish.     Trans.  Am. 
Fish.  Soc.  123:9-14. 
Taylor,  R.  G.,  H.  J.  Grier,  and  J.  A.  Whittington. 

1998.     Spawning  rhythms  in  common  snook  in  Florida.     J. 
Fish  Biol.  53:502-520. 
van  der  Elst,  R. 

1988.  A  guide  to  the  common  sea  fishes  of  Southern 
Africa,  2nd  ed.,  398  p.  Struik  Publishers,  Cape  Town, 
South  Africa. 

Wallace.  J.  H. 

1975.  The  estuarine  fishes  of  the  east  coast  of  South 
Africa.  III.  Reproduction.  South  African  Association  for 
Marine  Biological  Research.  Oceanographic  Research 
Institute,  Investigational  Report  41.  The  Oceanographic 
Research  Inst.,  Durban,  South  Africa. 

Wallace,  R.  A.,  and  K.  Selman. 

1989.  Cellular  and  dynamic  aspects  of  oocyte  growth  in 
teleosts.     Am.  Zool.  21:325-343. 

Wilson,  C.  A.,  and  D.  L.  Nieland. 

1994.     Reproductive  biology  of  red  drum,  Sciaenops  ocel- 
latus,  from  the  neritic  waters  of  the  northern  Gulf  of 
Mexico.     Fish  Bull.  92:841-850. 
Yeung,  W.  S.  B..  and  S.  T.  H.  Chan. 

1987.  The  gonadal  anatomy  and  sexual  pattern  of  the 
protandrous  sex-reversing  fish,  Rhabdosargus  sarba 
(Teleostei:  Sparidae).     J.  Zool.  Lond.  212:521-532. 


661 


Abstract — Novel  data  on  the  spatial 
and  temporal  distribution  of  fishing 
effort  and  population  abundance  are 
presented  for  the  market  squid  fishery 
lLoligo  opaleseens)  in  the  Southern 
California  Bight,  1992-2000.  Fishing 
effort  was  measured  by  the  detection 
of  boat  lights  by  the  Defense  Meteo- 
rological Satellite  Program  (DMSPi 
Operational  Linescan  System  (OLS). 
Visual  confirmation  of  fishing  vessels 
by  nocturnal  aerial  surveys  indicated 
that  lights  detected  by  satellites  are 
reliable  indicators  of  fishing  effort. 
Overall,  fishing  activity  was  con- 
centrated off  the  following  Channel 
Islands:  Santa  Rosa,  Santa  Cruz, 
Anacapa,  and  Santa  Catalina.  Fishing 
activity  occurred  at  depths  of  100  m  or 
less.  Landings,  effort,  and  squid  abun- 
dance (measured  as  landings  per  unit 
of  effort.  LPUE)  markedly  declined 
during  the  1997-98  El  Nino;  land- 
ings and  LPUE  increased  afterwards. 
Within  a  fishing  season,  the  location 
of  fishing  activity  shifted  from  the 
northern  shores  of  Santa  Rosa  and 
Santa  Cruz  Islands  in  October,  the 
typical  starting  date  for  squid  fishing 
in  the  Bight,  to  the  southern  shores 
by  March,  the  typical  end  of  the  squid 
season.  Light  detection  by  satellites 
offers  a  source  of  fine-scale  spatial 
and  temporal  data  on  fishing  effort  for 
the  market  squid  fishery  off  Califor- 
nia, and  these  data  can  be  integrated 
with  environmental  data  and  fishing 
logbook  data  in  the  development  of  a 
management  plan. 


Fishery  dynamics  of  the  California 
market  squid  (Loligo  opaleseens), 
as  measured  by  satellite  remote  sensing 


Michael  R   Maxwell 

University  of  California 

c/o  Southwest  Fisheries  Science  Center 

8604  La  Jolla  Shores  Drive 

La  Jolla,  California  92037 

Present  address:    Department  of  Biology 
University  of  San  Diego 
5998  Alcala  Park 
San  Diego,  California  92110 

E-mail  address.  maxwellm@sandiegoedu 


Annette  Henry 

California  Department  of  Fish  and  Game 
8604  La  Jolla  Shores  Drive 
La  Jolla,  California  92037 


Christopher  D.  Elvidge 

NOAA  National  Geophysical  Data  Center 

325  Broadway 

Boulder,  Colorado  80305 


Jeffrey  Safran 

Vinita  R.  Hobson 

Ingrid  Nelson 

Benjamin  T.  Tuttle 

Cooperative  Institute  for  Research  in 
Environmental  Sciences 
University  of  Colorado 
Boulder,  Colorado  80303 

John  B.  Dietz 

Cooperative  Institute  for  Research  on 

Atmosphere 

Colorado  State  University 

Fort  Collins,  Colorado  80523 

John  R.  Hunter 

Southwest  Fisheries  Science  Center 
NOAA  National  Marine  Fisheries  Service 
8604  La  Jolla  Shores  Drive 
La  Jolla,  California  92037 


Manuscript  submitted  24  February  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
17  June  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:661-670  (2004). 


The  market  squid  I  Loligo  opalesce?is) 
(also  known  as  the  opalescent  inshore 
squid,  FAO  [Roper  et  al„  1984])  is  cur- 
rently the  largest  revenue  fishery  for 
California  (Vojkovich,  1998;  CDFG, 
2000).  The  fishery's  importance  rose 
steadily  in  the  1980s  and  1990s,  in 
response  to  increased  demand  in  Asia 
coupled  with  declines  in  other  fisheries 
off  the  U.S.  West  Coast.  Market  squid 
is  a  short-lived  species  (Jackson,  1994; 
Butler  et  al.,  1999)  whose  abundance 
appears  to  be  readily  impacted  by 
environmental  variability.  For  exam- 
ple, squid  landings  plummeted  during 
the  1997-98  El  Nino  but  reached  a 
record  high  in  the  following  year 
(CDFG,  2000).  Considered  an  inte- 
gral component  of  California's  pelagic 
fishery,  the  market  squid  was  included 
in  the  Coastal  Pelagic  Species  Fishery 
Management  Plan  as  approved  by  the 
Pacific  Fisheries  Management  Council 
in  1998.  In  this  plan,  federal  authority 
is  invoked  to  monitor  the  fishery  to 
ensure  the  provisions  of  the  Magnu- 
son-Stevens  Fishery  Conservation  and 
Management  Act  of  1996.  If  a  resource 


is  estimated  as  overfished,  the  Coun- 
cil is  to  consider  implementing  active 
management  measures. 

The  lack  of  records  of  fishing  effort, 
such  as  vessel  logbooks  or  observer 
data,  hampered  initial  attempts  to 
formulate  a  management  plan  for  the 
market  squid.  The  nature  of  the  fish- 
ery, however,  suggested  an  alternative 
measure  of  fishing  effort:  the  detec- 
tion of  boat  lights  by  satellites.  The 
market  squid  is  typically  harvested  on 
shallow  nearshore  spawning  grounds 
in  the  Southern  California  Bight  and 
Monterey  Bay  (Vojkovich,  1998).  At 
night,  specialized  lightboats  shine 
high  intensity  (c.  30,000  watt)  lights 
on  the  water,  which  attract  and  con- 
gregate the  squid  near  the  surface. 
Seiner  boats  then  capture  the  con- 
centrated squid  with  purse-seine  nets 
(Vojkovich,  1998).  The  lights  of  the 
fishing  boats  are  detected  and  record- 
ed by  the  U.S.  Air  Force  Defense  Me- 
teorological Satellite  Program  (DMSP) 
Operational  Linescan  System  (OLS). 

DMSP-OLS  satellites  continuously 
orbit  the  planet,  acquiring  data  on 


662 


Fishery  Bulletin  102(4) 


meteorology  and,  incidentally,  nighttime  light  sources 
(Croft,  1978;  Elvidge  et  al.,  1997,  2001b).  Nighttime 
light  detection  by  satellites  has  proven  useful  for  vari- 
ous environmental  questions,  such  as  identifying  the 
extent  of  forest  fires  (Elvidge  et  al.,  2001a)  and  the 
effects  of  urban  lighting  on  sea  turtle  nest  selection 
and  hatchling  survivorship  (Salmon  et  al.,  2000).  Com- 
pilation of  light  data  for  the  global  light-fishing  squid 
fleet  has  contributed  to  examinations  of  the  fishery's 
ecosystem  impacts  (Rodhouse  et  al.,  2001).  For  local 
squid  fisheries,  the  locations  of  boat  lights  over  space 
and  time  are  particularly  valuable  in  cases  where  na- 
tional boundaries  pose  constraints  on  the  collection  of 
effort  data  (e.g.,  Illex  argentinus  in  the  southwestern 
Atlantic,  Waluda  et  al.,  2002). 

In  this  study,  we  used  boat  lights  to  quantify  the  spa- 
tial and  temporal  patterns  of  market  squid  fishing  ac- 
tivity in  the  Southern  California  Bight  over  the  period 
1992-2000.  The  bight  has  come  to  represent  the  great 
majority  of  squid  landings  off  California  (Vojkovich, 
1998;  Butler  et  al.,  1999;  CDFG,  2000).  An  important 
component  of  our  study  is  ground-truthing  work  that 
validates  the  feasibility  of  using  light  data  as  a  measure 
of  fishing  effort.  This  estimate  for  fishing  effort  enables 
us  to  present  novel  landings-per-unit-of-effort  (LPUE) 
data  for  the  market  squid.  A  companion  paper  analyzes 
the  light  detection  properties  of  the  DMSP-OLS  satellites 
over  the  Southern  California  Bight  (Elvidge  et  al.1). 


Materials  and  methods 

Light  detection  by  satellites 

The  DMSP  is  a  polar  orbiting  satellite  system  that 
acquires  daytime  and  nighttime  data  during  each  orbit. 
The  OLS  is  an  oscillating  scan  radiometer  designed  for 
cloud  imaging.  A  full  technical  description  of  image 
acquisition  by  the  DMSP-OLS  system,  and  the  subse- 
quent processing  of  images,  appears  in  a  companion 
paper  (Elvidge  et  al.1).  Briefly,  the  DMSP-OLS  acquired 
nighttime  data  for  over  2200  satellite  orbits  over  the 
Southern  California  Bight  (i.e.,  117°  to  122°  W,  32°30' 
to  34°30'N)  between  26  April  1992  and  4  April  2001. 
Four  different  satellites  were  employed  during  this  time. 
Three  overlapped  in  operation  dates,  producing  mul- 
tiple images  for  some  dates.  On  all  dates,  images  were 
acquired  between  18:30  and  22:00  Pacific  Standard  Time 
(PST),  with  20:21  PST  being  the  average  time.  The  satel- 
lite images  were  processed  into  geo-referenced  images  of 
boat  lights  and  clouds.  This  process  involved  superimpos- 
ing a  field  of  grid  cells  onto  the  satellite  image,  which 
quantified  the  satellite's  "field  of  view,"  the  extent  of 
detected  clouds,  and  the  area  available  for  light  detec- 
tion. Image  pixels  of  lights  were  taken  directly  from  the 


1  Elvidge,  C.  D.,  J.  Safran,  M.  R.  Maxwell,  K.  E.  Baugh,  A. 
Henry,  and  J.  R.  Hunter.  Unpubl.  data.  Satellite  based 
indices  of  lightboat  fishing  effort. 


satellite  image.  Pixels  were  identified  as  lights  by  their 
visible  band  digital  number.  The  images  were  subjected 
to  quality-control  procedures  to  correct  for  atmospheric 
noise  and  to  eliminate  images  overly  contaminated  by 
solar  glare,  sunlight,  heavy  lunar  illumination,  or  those 
containing  missing  data.  Fixed  sources  of  lights,  such  as 
city  lights  along  the  southern  California  coast,  the  city  of 
Avalon  (Santa  Catalina  Island),  off-shore  oil  platforms, 
and  naval  installations,  were  masked  from  the  light 
detection  algorithm. 

Data  deliveries  were  irregular  during  1992,  result- 
ing in  gaps  in  the  early  part  of  the  time  series.  For 
1992-98,  only  data  collected  during  the  dark  half  of 
the  lunar  cycle  were  available.  To  control  for  lunar  il- 
lumination throughout  the  time  series,  we  restricted 
analysis  of  fishery  data  to  images  for  which  lunar  il- 
lumination was  less  than  0.02  lux  (lumens  per  square 
meter).  Images  for  analysis  were  evaluated  against  ad- 
ditional criteria.  For  a  given  image,  we  calculated  the 
number  of  total  grid  cells  that  were  not  used  for  light 
detection  because  of  glare,  missing  data,  or  the  mask- 
ing of  known  nonboat  lights.  If  the  resulting  number  of 
grid  cells  left  available  for  light  detection  was  at  least 
50%  of  the  original  number  of  cells,  we  retained  the 
image  for  analysis.  Cloud  coverage  can  obscure  light 
sources1;  therefore  we  used  only  images  from  nights 
when  clouds  covered  less  than  25%  of  the  grid  cells 
available  for  light  detection.  For  nights  with  multiple 
acceptable  images,  we  averaged  the  percent  cloud  cover- 
age and  the  number  of  detected  light  pixels. 

Ground-truthing:  aerial  observations  of  boat  activity 

To  determine  the  relationship  between  detected  light 
pixels  and  the  number  of  squid  fishing  vessels  on  the 
water,  35  aerial  surveys  were  conducted  from  10  June 
1999  to  18  May  2000.  Each  survey  took  place  in  a  Cessna 
337  Skymaster  flown  at  an  average  altitude  of  1160  m 
above  sea  level.  The  path  of  each  survey  covered  the 
main  areas  of  squid  fishing  activity  within  the  South- 
ern California  Bight  (Fig.  1):  from  San  Diego,  over  the 
Channel  Islands,  to  Point  Conception,  and  back  down  the 
coastline  to  San  Diego.  Each  survey  took  approximately 
four  hours  to  complete,  occurring  between  18:00  h  and 
midnight  PST.  These  times  encompassed  the  time  that 
the  DMSP-OLS  satellites  were  over  the  bight.  The  35 
surveys  produced  26  nights  of  usable  data.  Survey  data 
were  discarded  if  satellite  images  were  unavailable,  if 
flights  were  aborted  because  of  weather,  or  if  heavy  fog 
obscured  boat  visibility.  We  note  that,  for  this  ground- 
truthing  work,  we  did  not  restrict  our  analysis  to  nights 
with  lunar  illumination  of  less  than  0.02  lux.  Rather, 
we  used  all  of  the  acceptable  26  nights,  and  quantified 
lunar  illumination  as  a  proportion  of  the  moon's  phase, 
where  0.00  denoted  a  new  moon  and  1.00  denoted  a  full 
moon. 

All  vessels  on  the  water  were  identified  by  using  Fuji- 
non  10x50  gyroscopic  binoculars,  and  the  GPS  positions 
of  all  vessels  were  recorded.  Vessel  type  was  identified 
as  either  a  nonsquid  vessel  or  as  a  squid  fishing  vessel. 


Maxwell  et  al.:  Fishery  dynamics  of  Lo/igo  opalescens 


663 


34°N 


33°N 


Pt.  Conception 


saota      Barbara 


San  Miguel  Island . 


Santa  Rosa  Island-f 
B 


Los  Angeles 


A 

N 


Anacapa 
Santa  Cruz      island 
Island 


Santa  Barbara  Island 
;** ,. 


Santa  Catahna  Island 


San  Nicolas  Island 


Coast    Lines 

1  00m 

depth    contour 


1 

120°W. 


San  Clemente 
Island 


— r 
118°W 


1  2  1  °W. 


119°W 


B 


12 

2'0'0'W 

1 

row 

TO'W 

l 

641 

f.iri 

638  ( 

si 

CMS 

347 

we 

646 

644 

as 

69? 

663 

i,i.; 

661 

660 

ess 

658 

657 

655 

6SJ~ 

■ssy 

"65TBM. 

t„ 

67? 

e  ,>, 

675 

674 

672 

671 

670 

669 

668 

667 

666    665 

'    i 

34WN- 

i 

i. 

696 

B9! 

694 

693 

692 

691 

-k£ 

l8" 

1034 

r 

776 

717 

716 

715 

714 

713 

^ 

^ 

70S      •    " 

706 

705 

704 

703 

70? 

l 

1 

738 

735 

'- 

11 

730 

- 

775 

724 

723 

72? 

7?1 

741 

4ft| 

715%, 

i 
1033 

1 

i 
f 

777 

755 

754 

753 

75? 

751 

749 

748 

746 

745 

743 

74? 

740 
760 

759 

758 

775 

774 

773 

772 

771 

770 

769 

768 

767 

766 

- 
765 

764 

763 

V 

75V, 

32 

i 

895 

820 

819 

818 

817 

816 

815 

314C 

^13 

812 

811 

810 

809 

IB 

805 

804 

803 

802 

' 

WOWN- 

10 

841 

4 

838 

837 

836 

835 

834 

633 

832 

831 

830 

828 

827 

826 

825 
846 

824 

823 

822 

h 

858 

857 

856 

855 

854 

853 

851 

85o\ 

848 

847 

845 

844 

843 

ft 

896 

876 

875 

874 

873 

872 

871 

870 

869 

868 

867 

866 

865 

864 

863 

862 

861 

*V 

894 

893 

892 

891 

890 

889 

888 

887 

888 

885 

884 

883 

882 

881 

880 

879 

878 

^ 

897 

916 

n 

1 

I 

Figure  1 

Fishing  activity  for  the  market  squid  in  the  Southern  California  Bight,  from  26  April  1992  to  28  May  2000. 
(At  Composite  satellite  image  of  squid  fishing  vessel  lights  (black  marks).  Permanent  sources  of  lights  (e.g., 
city  lights,  offshore  oil  platforms,  naval  installations)  are  removed.  CalCOFI  stations  83.42  ("A";  34.18°N, 
119.51°W)  and  83.51  ("B";  33.88°N,  120.13°W)  are  indicated.  (B)  Squid  landings  as  reported  by  California 
Dept.  Fish  and  Game  fishing  blocks.  Gray:  blocks  that  account  for  6.8  million  kg  (2%)  or  more  of  the  land- 
ings from  blocks  651-896.  Black:  blocks  that  account  for  20.5  million  kg  (6%)  or  more.  Latitudinal  blocks 
1032-1035  are  indicated.  Santa  Cruz  Island  is  marked  "999"  to  aid  correspondence  with  A. 


664 


Fishery  Bulletin  102(4) 


Squid  fishing  vessels  could  not  always  be  distinguished 
as  light  boats  or  seiners  and  therefore  were  recorded  as 
"squid  fishing  vessels." 

The  numbers  of  squid  fishing  vessels  showed  large 
skew  in  their  frequency  distribution.  These  data  were 
transformed  by  x'  =  log10(x+l).  Similarly,  proportion 
lunar  phase  was  transformed  by  x'  =  arcsin(V.r),  and 
detected  light  pixels  were  transformed  by  x'  =  log]0(.v+l) 
to  correct  for  skew  (Zar,  1984).  These  transformations 
produced  normally  distributed  data  acceptable  for  re- 
gression analysis.  With  these  transformed  variables, 
multiple  stepwise  regression  (forward  selection)  was 
performed  with  the  software  S-Plus  2000  (MathSoft 
Inc.,  Cambridge,  MA)  to  examine  the  effects  of  squid 
fishing  vessels  and  the  proportion  lunar  phase  on  de- 
tected light  pixels.  Squid  fishing  vessels  and  proportion 
lunar  phase  showed  very  little  correlation  (r=-0.09). 

Fishery  characteristics,  1992-2000 

For  quantitative  analysis  of  the  fishery  data,  we  aggre- 
gated the  nightly  satellite  data  (i.e.,  light  pixels  detected 
on  the  water)  into  calendar  quarters,  as  suggested  by 
the  within-year  distribution  of  squid  landings  in  the 
bight  (Butler  et  al.,  1999).  To  standardize  conditions  of 
light  detection,  we  excluded  all  data  after  28  May  2000, 
because  this  was  the  starting  date  of  mandatory  shield- 
ing of  the  high  intensity  lights  of  the  lightboats.  This 
regulation  was  enforced  by  California's  Department  of 
Fish  and  Game  to  reduce  light  pollution  by  the  light- 
boats.  The  shields  did  not  totally  obscure  the  lightboats 
from  detection  by  the  satellites  (authors'  pers.  obs.)  but 
made  the  emitted  light  less  bright,  and,  hence,  less 
detectable  by  the  satellites.  Thus,  our  data  for  fishing 
effort  spanned  calendar  quarters  from  Jul-Sep  1992  to 
Jan-Mar  2000.  We  included  a  quarter  for  analysis  if  it 
contained  10  or  more  nights  of  acceptable  images.  By 
these  criteria,  we  described  effort  for  24  of  the  31  cal- 
endar quarters.  The  mean  number  of  nights  per  quarter 
was  26  (range=10-72  nights). 

The  quantity  (kg)  and  location  of  landed  market  squid 
were  recorded  by  California  Department  of  Fish  and 
Game  (CDFG)  throughout  the  1992-2000  study  pe- 
riod and  were  made  available  to  the  authors.  During 
this  study  period,  squid  fishing  in  the  bight  occurred 
exclusively  at  night  (Vojkovich,  1998).  The  squid  were 
landed  at  port  within  several  hours  after  being  caught; 
therefore  the  landings  for  a  given  day  corresponded  to 
the  previous  night's  effort.  Squid  fishermen  reported 
the  locations  of  their  hauls  by  CDFG  fishing  blocks.  We 
defined  catch  taken  from  the  Southern  California  Bight 
as  that  from  blocks  651-896  and  1032-1035  (Fig.  1). 
Blocks  651-896  are  typically  10'  latitude  x  10'  longitude 
and  can  be  used  to  locate  regions  of  high  catch.  Blocks 
1032-1035  are  large  latitudinal  bands,  generally  30' 
wide,  that  encompass  blocks  651-896.  We  used  blocks 
1032-1035  in  calculating  the  total  catch  in  the  bight, 
but  not  in  depicting  the  location  of  the  catch. 

To  construct  the  abundance  index  of  landings  per 
unit  of  effort  (LPUE).  we  first  estimated  the  number  of 


squid  fishing  vessels  for  each  night  of  satellite  data,  us- 
ing the  regression  results  of  the  ground-truthing  work 
(see  "Results"  section).  We  then  summed  the  nightly 
estimated  number  of  vessels  for  each  calendar  quarter. 
For  those  nights  for  which  we  had  estimated  numbers  of 
vessels,  we  also  summed  the  landed  catch  within  each 
calendar  quarter.  To  arrive  at  LPUE  for  the  quarter, 
we  divided  the  summed  landings  by  the  corresponding 
summed  effort. 

Environmental  data 

We  used  the  multivariate  ENSO  index  (MEI)  to  indicate 
overall  environmental  conditions  over  the  course  of  the 
1992-2000  study  period.  The  MEI  is  a  multivariate 
index  that  incorporates  sea  level  pressure,  surface  zonal 
and  meridional  wind  components,  sea  surface  tempera- 
ture, surface  air  temperature,  and  cloudiness  (Wolter 
and  Timlin,  1998).  The  MEI  index  is  calculated  for  the 
tropical  Pacific  (i.e.,  between  10°N  and  10°S,  from  Asia 
to  the  Americas),  and  its  monthly  values  appear  on 
the  website  http://www.cdc.noaa.gov/~kew/MEI/table. 
html.2 

Analysis  of  the  location  of  fishing  effort  over  the 
course  of  the  traditional  squid  fishing  season  in  the 
bight  led  to  an  investigation  of  oceanographic  data  for 
waters  surrounding  Santa  Cruz  Island  in  March.  Spe- 
cifically, we  examined  sea  temperature  from  two  sourc- 
es. First,  we  obtained  sea  surface  temperature  for  all 
satellite  nights  in  March  1993-2000  from  the  Physical 
Oceanography  Distributed  Active  Archive  Center  (PO. 
DAAC)  at  California  Institute  of  Technology  (Pasadena, 
CA).  These  data  were  reported  for  18x18  km  grids, 
which  were  approximately  the  size  of  the  10'xlO'  fish- 
ing blocks.  We  selected  the  grid  that  covered  block  686 
to  represent  the  northern  shore  of  the  island,  and  that 
which  covered  block  708  to  represent  the  southern  shore 
(Fig.  IB).  For  each  year  in  the  1993-2000  period,  we 
calculated  mean  March  temperature  for  both  blocks. 

The  second  source  of  sea  temperature  was  the  da- 
tabase maintained  by  the  California  Cooperative  Oce- 
anic Fisheries  Investigations  (CalCOFI).  Since  1950. 
the  CalCOFI  program  has  conducted  quarterly  survey 
cruises  along  transects  perpendicular  to  the  southern 
California  coast.  This  system  of  transects  incorporates 
66  geographically  fixed  stations.  At  each  station,  a 
conductivity-temperature-depth  (CTD)  instrument  is 
deployed.  Details  on  survey  methods  appear  on  the  web- 
site http://www-mlrg.ucsd.edu/calcofi. html. :1  along  with 
the  publicly  accessible  database.  For  April  1993-2000, 
we  obtained  temperatures  at  sea  surface  and  at  75  me- 
ters depth  at  two  stations  (Fig.  1A):  83.42  (northeast 
of  Santa  Cruz  Island:  34.18°N,  119.51°W)  and  83.51 
(southwest  of  Santa  Cruz  Island;  33.88DN,  120.13  W). 


-  NOAA-CIRES  Climate  Diagnostics  Center  website.  I  Ac- 
cessed 3  November  200.3.1 

3  California  Cooperative  Oceanic  Fisheries  Investigations 
website.     [Accessed  3  November  200.3.1 


Maxwell  et  al.:  Fishery  dynamics  of  Loligo  opalescens 


665 


One  measurement  was  made  at  each  station 
at  sea  surface  and  at  75  meters  depth  during 
April  (n  =  8  for  both  depths). 


Results 

Ground-truthing:  aerial  observations 
of  boat  activity 

Nonsquid  vessels  used  weak  lights  (i.e.,  much 
less  than  30,000  watts),  which  did  not  show 
in  the  satellite  images.  On  average,  23  squid 
fishing  vessels  were  observed  each  night  by 
the  aerial  surveys  (range  =  0-64  vessels,  n=26 
nights).  The  20:00-midnight  observation 
period  was  the  peak  time  for  attraction  of 
squid  by  the  light  boats.  Although  the  squid 
vessels  did  change  location  during  this  time, 
they  typically  left  their  lights  running  to 
continue  searching  for  squid.  The  number  of 
squid  vessels  explained  much  of  the  varia- 
tion in  detected  light  pixels;  proportion  lunar 
phase  failed  to  enter  the  analysis  as  a  signifi- 
cant variable  (Table  1).  Detected  light  pixels  increased 
with  the  number  of  squid  vessels  (Fig.  2). 

The  regression  analysis  yields  the  following  simpli- 
fied equation: 

log10(p, +l)  =  1.25xlog10(x, +1),  (1) 

where  xt  =  observed  number  of  squid  vessels;  and 
pt  =  detected  light  pixels  for  night  t. 

We  used  inverse  prediction  to  estimate  the  number 
of  squid  vessels  for  each  satellite  night  (Et\  in  the 
1992-2000  period  (Zar,  1984).  The  estimated  number 
of  squid  vessels  was  found  by  the  equation 


3.0- 

♦  ♦ 

♦ 

^. 

♦     * 

+ 

♦           ,' 

«>       2.0- 

♦            ♦     *- ' '  ♦ 

"cB 

*      *'<•      ♦ 

5. 

„'            ♦ 

S 

'    ♦ 

.D) 

'    ♦,-.*     • 

o 

ra       10- 

S  *" 

o 

s  "* 

<> 

*** 

0.0                                                         1.0                                                         2.0 

Log10  (observed  squid  vessels  +  1 ) 

Figure  2 

Plot  of  log10-transformed  number  of  squid  vessels  and  detected  light 

pixels.  Regression  line  taken  from  the  statistics  in  Table  1. 

Table  1 

Multiple  stepwise  (forward  selection)  regression  of  de- 
tected light  pixels  on  squid  fishing  vessels  (transformed: 
x'=log10(x+l)  and  proportion  lunar  phase  (transformed: 
.r'=arcsin(  Vr).  r2=0.64;  ANOVA:  Fl  24  =  42.66,  P<0.0001. 


Variable 


Coefficient  ±SE 


Squid  fishing  vessels 
INTERCEPT 
Proportion  lunar  phase 


1.25  ±0.19 
0.07  ±0.24 
not  entered 


<0.0001 
>0.75 
not  entered 


4=io"«-(ft+imB-i=1J»/ft+i-i- 


(2) 


The  ground  sample  distance  of  the  satellite  data  is 
2.7  km,  which  means  that  multiple  squid  vessels  may 
potentially  fit  into  one  pixel  of  detected  light.  This  could 
result  in  an  underestimation  of  effort.  The  severity  of 
this  problem  can  be  assessed  by  examining  the  coeffi- 
cient of  the  simple  linear  regression  of  log-transformed 
variables  represented  by  Equation  1.  One  of  four  sce- 
narios is  possible:  1)  boats  are  not  aggregated  (coef- 
ficient^), 2)  boats  are  aggregated  regardless  of  the 
number  of  boats  on  the  water  (coefficients),  3)  boats 
are  aggregated  only  when  many  boats  are  on  the  water 
(coefficient<l),  or  4)  boats  are  aggregated  only  when  few 
boats  are  on  the  water  (coefficient>l).  The  coefficient 
in  Equation  1  is  1.25,  which  fails  to  significantly  differ 
from  1.00  (f-test  for  regression  coefficient:  t  =  1.305, 
j30  =  1,  df  =  24,  P  >  0.2,  two-tailed;  power  <  0.5,  retrospec- 
tively calculated;  Zar,  1984).  This  result  suggests  that 
very  little  clumping  of  the  boats  occurred  (scenario  1), 


or  that  the  degree  of  clumping  was  independent  of  the 
number  of  boats  on  the  water  (scenario  2).  Although  the 
statistical  power  of  this  ^-test  is  not  high  (power<0.5), 
we  conclude  that  the  data  provide  more  support  for  sce- 
narios 1  and  2  over  scenarios  3  and  4.  Either  scenario, 
1  or  2,  allows  for  a  comparison  of  the  relative  values  of 
estimated  effort  and  LPUE  within  a  time  series. 

Fishery  characteristics,  1992-2000 

A  composite  satellite  image  of  all  squid  fishing  activity 
in  the  Southern  California  Bight  during  the  1992-2000 
study  period  revealed  major  concentrations  of  effort  off 
the  Channel  Islands,  especially  Santa  Rosa,  Santa  Cruz, 
Anacapa,  and  Santa  Catalina  (Fig.  1A).  Squid  fishing 
occurs  close  to  the  island  shores  and  is  bounded  by  the 
100-m  contour.  During  the  study  period,  379.2  billion 
kg  of  squid  were  landed  in  the  bight:  341.2  billion  from 
blocks  651-896  (Fig.  IB),  and  the  remainder  from  the 
large  blocks  1032-1035.  The  main  areas  of  fishing  activ- 
ity, as  indicated  by  satellite,  are  consistent  with  the 
blocks  of  high  catch  (Fig.  IB).  We  note  that  blocks  682 


666 


Fishery  Bulletin  102(4) 


and  720,  although  areas  of  high  catch,  do  not  appear  on 
the  satellite  composite  because  the  mainland  shore  was 
excluded  from  light  detection.  Further,  much  activity 
was  evident  around  Santa  Barbara  Island  (block  765). 
Although  this  block  represented  4.0  million  kg  (18th 
out  of  the  127  blocks),  it  did  not  rank  highly  enough  for 
inclusion  in  Fig.  IB. 

Analysis  of  temporal  trends  in  the  fishery  showed 
peaks  in  landed  catch  for  the  bight  in  the  fall  and 
winter  quarters  (Oct-Dec  and  Jan-Mar,  respectively; 
Fig.  3A).  There  was  a  near  absence  of  catch  during 


c    > 

J=   o 
<n    ^ 

■*-  a> 
>-.o 

P 

L±J~g 

si 

<5  to 

5   o> 


B 


.-c  m 

o  <fl 

o  o> 

o  > 

OS 

*-  ZJ 

x  cr 

" —  to 

HI  i- 

D  <° 

a.  °- 

_i  Ol 


o  S      30- 


rs.tt 


az 


1992    1993    1994    1995    1996    1997    1998    1999    2000 

Quarter 

Figure  3 

Time  series  of  market  squid  fishery  data  in  Southern  California 
Bight,  by  calendar  quarter  (Jul-Sep  1992  to  Jan-Mar  2000). 
The  Jan-Mar  quarters  are  marked  by  dashed  vertical  lines. 
(A)  Landings  are  in  kg  (blocks  651-896,  1032-1035).  (B)  Mean 
±SE  nightly  fishing  effort,  in  estimated  number  of  squid  vessels. 
(C)  Landings  per  unit  of  effort  (LPUEl:  summed  landings  (kg) 
on  satellite  nights  were  divided  by  summed  effort  (estimated 
number  of  squid  vessels)  on  the  corresponding  nights. 


most  of  1997-98  (Fig.  3A),  which  corresponded  to  the 
strong  El  Nino  event  during  this  period  (Fig.  4).  Effort 
data  revealed  surges  in  the  Oct-Dec  quarters  before 
the  1997-98  El  Nino  (Fig.  3B).  The  Oct-Dec  quarter 
of  1998  signalled  a  resumption  of  fishing  effort  follow- 
ing El  Nino,  but  effort  levels  for  1999  and  early  2000 
were  lower  than  pre-El  Nino  levels.  Interestingly,  squid 
abundance,  as  measured  by  landings  per  unit  of  effort 
(LPUE),  showed  a  rapid  increase  from  the  El  Nino  lows, 
and  squid  abundance  for  1999-2000  reached  the  high- 
est values  of  the  time  series  (Fig.  3C). 

Analysis  of  boat  locations  along  the  Channel 
Islands  revealed  a  shift  over  the  course  of  the 
fishing  season.  Compiling  the  satellite  data  to 
yield  composite  images  in  multiyear  sets,  we 
found  that  fishing  activity  in  October  consis- 
tently included  the  north  shore  of  Santa  Cruz 
Island  (Fig.  5,  A,C,E).  In  contrast,  fishing  ac- 
tivity in  March  showed  considerable  reduction 
along  the  north  side  of  Santa  Cruz  Is.,  but  activ- 
ity continued  along  the  island's  southern  shore 
(Fig.  5,  B,D,F).  Composite  images  for  December 
and  January  were  also  examined  for  all  of  the 
multiyear  sets.  December  marked  a  transitional 
stage  from  the  activity  in  October  to  reduction 
of  fishing  in  March  along  the  northern  shores. 
In  all  multiyear  sets,  the  December  lights  along 
northern  Santa  Cruz  Island  were  more  scat- 
tered and  less  dense  than  those  in  October. 
January  images  were  very  similar  to  those  for 
March.  Although  data  from  March  1993-95  in- 
dicated little  fishing  activity,  a  composite  image 
for  January  1993-95  was  very  similar  to  that 
for  March  1999-2000:  light  banks  occurred  off 
southern  Santa  Cruz,  southeastern  Santa  Rosa, 
and  around  Anacapa,  but  were  virtually  absent 
from  northern  Santa  Cruz  and  Santa  Rosa. 

Water  temperatures  around  Santa  Cruz  Is- 
land did  not  consistently  differ  between  north- 
ern and  southern  waters.  March  sea  surface 
temperatures,  measured  by  satellite,  were  very 
similar  for  the  island's  northern  and  south- 
ern shores  (Table  2).  April  sea  surface  tem- 
peratures, measured  at  CalCOFI  stations,  were 
slightly  warmer  to  the  northeast  of  the  island 
(Table  2).  Temperatures  at  75  meters,  however, 
were  nearly  identical  for  the  two  CalCOFI  sta- 
tions (Table  2). 


Discussion 

The  satellite  images  and  landings  data  corrobo- 
rated spatial  and  temporal  patterns  of  fishing 
activity  for  the  market  squid.  For  the  period 
1992-2000,  both  data  sets  indicated  intense 
harvesting  along  the  Channel  Islands  of  Santa 
Rosa,  Santa  Cruz,  Anacapa,  and  Santa  Cata- 
lina.  The  satellite  images  captured  additional 
information,  such  as  fishing  activity  being 


Maxwell  et  al.:  Fishery  dynamics  of  Loligo  opalescens 


667 


Table  2 

Water  temperature  (  C  I  for  the  northern  and  southern  waters  around  Santa  Cruz  Island,  March  and  April,  1993- 

-2000. 

Northern  waters                                                                               Southern  waters 

Depth  (m)                   Location                    Mean              Min             Max                        Location                   Mean             Min 

Max 

March  sea  surface  temperature,  as  measured  by  satellite  (PO.DAAC  datai' 

0                            Block  686                   14.5              12.8             15.7                       Block  708                  14.7              13.2 

15.9 

April  temperature,  measured  at  CalCOFI  stations2 

0                            Station  83.42             13.6              11.6             16.7                       Station  83.51            12.8             11.2 

14.5 

75                            Station  83.42               9.9                9.3             11.2                       Station  83.51            10.3               9.3 

11.2 

1  Measurements  made  on  multiple  nights  per  month  of  March  (range  of  measured  nights  per  month  of  March:  6-26 1.  "Mean"  is  the  overall  average 

of  the  mean  March  temperatures;  "Min"  is  the  minimum  of  the  mean  values,  "Max  is  the  maximum,  of  the  mean  values. 

2  One  measurement  made  at  each  station  at  each  depth  per  month  of  April  (n  =  8  for  both  depths). 

clearly  delimited  by  the  100-m  contour.  The  landings 
data,  reported  by  fishing  blocks,  were  much  cruder  in 
geographic  scale  and  failed  to  catch  this  subtlety. 

The  ground-truthing  work  conducted  by  aerial  sur- 
veys indicated  that  detected  light  pixels  are  useful  in 
estimating  the  number  of  squid  vessels  in  operation. 
This  result  is  consistent  with  examination  of  the  fish- 
ery for  the  squid  Illex  argentinus  in  the  southwestern 
Atlantic,  where  vessels  use  powerful  lamps  to  attract 
the  squid  to  lures  (Waluda  et  al.,  2002).  In  the  latter 
fishery,  analysis  of  images  acquired  by  the  DMSP-OLS 
satellites  revealed  a  good  fit  between  the  recorded  num- 
ber of  vessels  in  operation  on  a  given  night  and  the 
number  of  light  pixels  detected  (Waluda  et  al.,  2002). 

In  the  present  study,  the  fishery  data  showed  a 
strong  response  to  the  1997-98  El  Nino  event,  which 
was  one  of  the  strongest  events  on  record  (Wolter  and 
Timlin,  1998).  Fishing  effort  and  landings  tended  to 
peak  in  the  Oct-Dec  and  Jan-Mar  quarters  before 
the  1997-98  El  Nino.  Both  data  series  dramatically 
dropped  during  the  1997-98  El  Nino  and  showed  recov- 
ery afterwards.  Squid  abundance,  measured  as  LPUE, 
also  showed  a  pronounced  drop  and  rapid  increase  in 
response  to  the  El  Nino.  It  is  interesting  to  note  that 
another  index  of  market  squid  abundance,  the  occur- 
rence of  squid  beaks  in  the  scat  of  sea  lions,  showed 
similar  responses  to  earlier  El  Nino  events  (Lowry  and 
Carretta,  1999).  Squid  beak  occurrence  dropped  steeply 
during  the  strong  1983-84  El  Nino,  and  increased 
afterwards.  Beak  occurrence  also  dipped  and  rose  in 
response  to  a  milder  El  Nino  in  1992-93.  Significantly, 
Lowry  and  Carretta  (1999)  examined  southern  Chan- 
nel Islands:  Santa  Barbara,  San  Clemente,  and  San 
Nicolas.  Our  present  study  reflects  squid  abundance 
primarily  around  northern  Channel  Islands  (e.g.,  Santa 
Rosa,  Santa  Cruz,  Anacapa).  Taken  together,  these 
studies  may  indicate  that  El  Nino  exerts  a  bight-wide 
influence  on  squid  abundance. 

We  suggest  that  a  strong  El  Nino  event  changes  the 
reproductive  conditions  for  the  market  squid  in  the 


-2.0-1 1 1 1 1 1 1 1 h 

1992  1993  1994  1995  1996  1997  1998  1999  2000 
Year 

Figure  4 

Multivariate  ENSO  index  iMEI)  for  the  tropical  Pacific 
(between  10CN  and  10°Si,  by  month.  Data  were  obtained 
from  http://www.cdc.noaa.gov/~kew/MEI/table.html. 


Southern  California  Bight.  With  regard  to  spawning, 
the  spawning  population  becomes  less  abundant  on  the 
traditional  shallow-water  spawning  grounds.  Research 
on  a  congener,  the  South  African  chokka  squid  (Loligo 
vulgaris  reynaudii),  points  to  possible  environmental 
influences  on  spawning  for  loliginid  squid  (Roberts  and 
Sauer,  1994).  Off  South  Africa,  a  strong  El  Nino  can 
lead  to  reduced  upwelling  and  increased  turbidity.  In 
normal  years,  upwelling,  presumably  detected  by  the 
squid  as  an  influx  of  cold  water,  may  trigger  spawning 
behavior  (Roberts  and  Sauer.  1994).  In  El  Nino  years 
off  South  Africa,  reduced  upwelling  and  increased  tur- 
bidity on  the  inshore  spawning  grounds  are  thought  to 
force  the  spawners  into  deeper  water,  beyond  the  reach 
of  the  fishery  (Roberts  and  Sauer,  1994).  In  a  recent 
study,  catch  for  the  chokka  squid  increased  with  strong 
easterly  winds,  which  caused  upwelling,  and  decreased 
with  increased  turbidity  (Schon  et  al.,  2002).  In  the 
California  Current  System,  upwelling  decreases  during 
strong  El  Nino  events  (Schwing  et  al.,  2000).  Upwelling 


668 


Fishery  Bulletin  102(4) 


October  1992, 

1993.  1994 

V 

-O*^- 

Los 
Angeles 

A 

^ 

March  1993,  1994,  1995 


120VJ  n9*w 

October  1995,  1996,  1997 


March  1996, 

1997,  1998 

'"^. 

Los 

34'"  Nh 

.         -     ~ 

Angeles 

D 

•                         *C>^_ 

120°W  119°W 

October  1998,  1999,  2000 


120°W  119°W 

March  1999,  2000 


119°W 


Figure  5 

Location  of  fishing  activity,  as  indicated  by  black  areas,  for  the  early  ( October  i  and 
late  (March  l  parts  of  the  traditional  squid  fishing  season  in  the  Southern  California 
Bight.  1992-2000.  For  each  month,  a  multiyear  composite  image  is  shown.  (A)  October 
1992,  1993.  1994.  (Bi  March  1993.  1994,  1995.  (Cl  October  1995,  1996,  1997.  (Dl 
March  1996,  1997,  1998.  (E)  October  1998,  1999.  (F)  March  1999,  2000. 


in  the  Southern  California  Bight  was  reduced  during 
the  1997-98  El  Nino  (Hay ward,  2000).  It  is  not  known 
how  market  squid  adults  respond  to  changes  in  water 
temperature  or  turbidity,  or  whether  spawning  fish  shift 
to  other  habitats  during  El  Nino  events. 

A  strong  El  Nino  event  can  also  alter  feeding  and 
developmental  conditions  for  squid.  During  the  1997-98 
El  Nino,  macrozooplankton  abundance  substantially 
decreased  in  the  Southern  California  Bight  and  off 
Baja  California  (Lynn  et  al.,  1998;  Hayward,  2000;  La- 
vaniegos  et  al.,  2002).  Food  availability  affects  growth 
rates  of  loligind  squid  (Jackson  and  Moltschaniwskyj, 
2001).  Recently,  Jackson  and  Domeier  (2003)  indicated 
lower  growth  rates  for  the  market  squid  in  the  Southern 
California  Bight  during  the  1997-98  El  Nino. 

In  the  present  study,  fishing  effort  following  the 
1997-98  El  Nino  was  generally  below  pre-El  Nino  lev- 
els. The  subsequent  high  levels  of  catch  in  late   1999 


and  early  2000  may  indicate  that  squid  were  in  great 
abundance,  thereby  requiring  less  overall  catch  effort 
to  meet  market  demand.  A  strong  La  Nina  succeeded 
the  1997-98  El  Nino  (Lynn  and  Bograd.  2002;  Schwing 
et  al.,  2002),  with  strong  upwelling  and  high  macrozoo- 
plankton abundance  in  the  Southern  California  Bight 
by  spring  1999  (Schwing  et  al.,  2000;  Hayward,  2000). 
Indeed,  the  high  LPUE  in  the  present  study  in  late  1999 
and  early  2000  points  to  increased  squid  abundance  in 
response  to  a  more  productive  environment.  Alterna- 
tively, one  could  argue  that  increased  fishing  efficiency. 
not  increased  squid  abundance,  resulted  in  high  LPUE. 
One  manifestation  of  higher  fishing  efficiency  could  be 
a  contracted  fishing  range,  where  especially  productive 
pockets  are  identified  and  targeted.  An  overall  com- 
parison of  fishing  location  in  October  and  March  before 
and  after  El  Nino  did  not  support  this  explanation:  the 
total  spatial  extent  of  fishing  activity  was  not  greatly 


Maxwell  et  al.:  Fishery  dynamics  of  Loligo  opalescens 


.    669 


reduced  in  post-El  Nino  October  or  March.  A  noticeable 
concentration  of  fishing  effort  off  the  southern  shore 
of  Santa  Cruz  Island  was  evident  in  the  post-El  Nino 
period,  however.  The  landings  data  may  indicate  that 
this  southern  shore,  represented  by  blocks  708  and 
709,  was  indeed  productive.  In  the  pre-El  Nino  period 
(1992-96),  blocks  708  and  709  represented  3%  of  the 
landings  in  the  bight.  In  the  post-El  Nino  period  (1999 
to  early  2000),  these  two  blocks  came  to  represent  12% 
of  the  landings. 

The  spatial  distribution  of  fishing  activity  appears  to 
shift  over  the  course  of  the  squid  fishing  season.  In  the 
Southern  California  Bight,  October  and  March  mark 
the  traditional  beginning  and  end  of  the  squid  fishing 
season,  respectively  (Butler  et  al.,  1999).  In  the  present 
study,  fishing  activity  along  the  Santa  Rosa  and  Santa 
Cruz  Islands  moved  largely  to  the  southern  shores  by 
March,  leaving  the  northern  shores  relatively  unfished. 
This  spatial  shift  may  reflect  change  in  local  squid 
habitat  or  changes  in  the  fishermen's  behavior.  As  a 
rough  indicator  of  habitat  quality,  water  temperature 
did  not  consistently  differ  between  the  northern  and 
southern  waters  around  Santa  Cruz  Island  in  March 
and  April,  both  at  sea  surface  and  at  75  meters  depth. 
Wind  conditions,  on  the  other  hand,  change  consider- 
ably from  October  to  March.  The  northern  shores  of 
Santa  Rosa  and  Santa  Cruz  lie  on  the  rim  of  the  San- 
ta Barbara  Channel.  Wind  speed  and  wind  stress  are 
relatively  low  through  the  channel  in  the  fall  and  early 
winter  but  increase  significantly  in  March  to  remain 
high  throughout  the  spring  and  summer  (Winant  and 
Dorman,  1997;  Harms  and  Winant,  1998;  Dorman  and 
Winant,  2000).  It  remains  unresolved  whether  the  high 
winds  in  the  Channel  in  March  and  April  create  ocean- 
floor  turbulence  and  turbidity  that  discourage  squid 
spawning  (cf,  Roberts  and  Sauer,  1994),  or  whether 
fishermen  simply  eschew  the  rocky  Channel  in  favor  of 
the  southern  shores  of  the  islands. 

Although  satellite  remote  sensing  can  generate  a 
"neutral  party"  record  of  fishing  effort,  we  note  three 
caveats  associated  with  satellite  data.  First,  large  sta- 
tionary sources  of  light,  such  as  coastal  cities,  must  be 
excluded  when  quantifying  fishing  vessel  activity.  The 
exclusion  of  urban  light  sources  can  result  in  under- 
estimating effort,  because  boats  that  work  near  large 
light  sources  can  be  excluded  from  analysis.  We  were 
concerned  that  an  underestimation  of  effort  along  the 
mainland  coast  would  explain  this  study's  post-El  Nino 
increase  in  LPUE.  Landings  data,  however,  may  indi- 
cate that  effort  in  coastal  blocks  actually  declined  after 
the  1997-98  El  Nino.  Coastal  blocks  accounted  for  19% 
of  the  landings  in  the  pre-El  Nino  years  (1992-96), 
dropping  to  11%  of  landings  in  the  post-El  Nino  years 
(1999  to  early  2000). 

Second,  the  spatial  resolution  of  the  satellite  imag- 
es may  be  large  enough  to  allow  multiple  boats  to  fit 
into  one  "pixel"  of  detected  light.  Thus,  effort  may  be 
underestimated.  Analysis  of  the  ground-truthing  fly- 
overs, however,  did  not  indicate  a  strong  interaction 
between  boat  aggregation  and  nightly  fleet  size.  Boats 


may  have  indeed  aggregated  over  the  course  of  our 
study,  but  our  analysis  indicates  that  such  aggregation 
was  independent  of  nightly  fleet  size.  In  this  case,  the 
absolute  values  of  estimated  effort  and  LPUE  would  be 
underestimated  across  all  dates.  The  relative  values  of 
effort  and  LPUE,  however,  will  be  only  slightly  affected 
within  a  time  series;  therefore  we  place  confidence  in 
our  examinations  of  the  temporal  patterns  of  the  ef- 
fort-based data.  A  third  caveat  is  specific  to  the  present 
study.  The  ground-truthing  work  occurred  during  a  pe- 
riod of  relatively  low  fishing  effort  (1999-2000).  Future 
fly-overs  during  periods  of  greater  effort  will  be  useful 
in  corroborating  our  observed  relationship  between  fly- 
over and  satellite  data. 

The  present  study  demonstrates  that  light  detection 
by  satellite  remote  sensing  is  useful  for  examining  tem- 
poral and  spatial  patterns  of  fishing  effort  and  popula- 
tion abundance,  as  measured  by  LPUE.  Light  detection 
by  satellite  has  certain  drawbacks,  but  these  are  not 
insurmountable.  Importantly,  geo-referenced  satellite 
images  provide  an  independent  source  of  fishing  effort, 
which  can  be  feasibly  integrated  with  environmental 
data  through  GIS  analysis.  With  regard  to  market  squid 
off  California,  satellite  data  can  help  provide  fine-scale 
data  on  fishing  location  for  this  fishery's  ongoing  man- 
agement efforts4'5  (see  also  Mangel  et  al.,  2002).  Al- 
though mandatory  shielding  of  the  boat  lights  went  into 
effect  in  May  2000,  these  lights  are  still  detectable  by 
the  satellites  (authors'  pers.  obs.).  Recently,  effort  log- 
books have  become  mandatory  for  squid  fishermen  off 
California.  This  requirement  points  to  a  unique  oppor- 
tunity to  collect  and  corroborate  fishery-dependent  and 
independent  measures  of  fishing  effort. 


Acknowledgments 

We  owe  much  gratitude  to  personnel  of  California's 
Department  of  Fish  and  Game  for  their  assistance  in 
the  ground-truthing  work.  In  particular,  we  thank  the 
pilots  Jeff  Veal  and  Tom  Evans,  and  the  following  aerial 
observers:  D.  Bergen,  T.  Bishop,  S.  Carner,  D.  Hanan, 
C.  Kong,  J.  Kraus,  A.  Lohse,  S.  MacWilliams,  D.  Ono, 
M.  Songer,  J.  Wagner,  and  E.  Wilson.  We  also  thank 
Paul  Crone  for  collaboration  on  this  project,  Chris  Reiss 
for  extracting  CalCOFI  water  temperature  data.  Rich 
Cosgrove  for  assistance  with  mapping,  Kevin  Hill  for 
information  about  the  Pacific  Fisheries  Management 
Council,  and  George  Watters  and  anonymous  review- 
ers for  constructive  comments.  This  project  was  funded 
by  the  California  Department  of  Fish  and  Game  and 
U.S.  Department  of  Commerce  (NOAA  NESDIS  Ocean 
Remote  Sensing  Program). 


4  California  Department  of  Fish  and  Game.  2003.  Draft: 
Market  squid  fishery  management  plan.  [Available  from: 
Calif.  Dept.  Fish  Game,  4949  Viewridge  Avenue,  San  Diego, 
CA  92123.] 

5  Maxwell,  M.  R.,  L.  D.  Jacobson,  and  R.  Conser.  Manuscript  in 
review.  Eggs-per-recruit  model  for  management  of  the  Cali- 
fornia market  squid  (Loligo  opalescens)  fishery. 


670 


Fishery  Bulletin  102(4) 


Literature  cited 

Butler,  J.,  D.  Fuller,  and  M.  Yaremko. 

1999.  Age  and  growth  of  market  squid  (Loligo  opalescens  i 
off  California  during  1998.  Calif.  Coop.  Oceanic  Fish. 
Invest.  Rep.  40:191-195. 

CDFG  (California  Department  of  Fish  and  Gamei. 

2000.  Review  of  some  California  fisheries  for  1999:  market 
squid.  Dungeness  crab,  sea  urchin,  prawn,  abalone, 
groundfish,  swordfish  and  shark,  ocean  salmon,  nearshore 
finfish,  Pacific  sardine,  Pacific  herring.  Pacific  mackerel, 
reduction,  white  seabass,  and  recreational.  Calif.  Coop. 
Oceanic  Fish.  Invest.  Rep.  41:8-25. 

Croft,  T.  A. 

1978.     Nighttime  images  of  the  earth  from  space.     Sci. 
Am.  239:68-79. 
Dorman,  C.  E.,  and  C.  D.  Winant. 

2000.     The  structure  and  variability  of  the  marine  atmo- 
sphere around  the  Santa  Barbara  Channel.     Month. 
Weather  Rev.  128:261-282. 
Elvidge,  C.  D„  K.  E.  Baugh,  E.  A.  Kihn,  H.  W.  Kroehl,  and 
E.  R.  Davis. 

1997.  Mapping  city  lights  with  nighttime  data  from 
the  DMSP  Operational  Linescan  System.  Photogram. 
Engin.  Remote  Sens.  63:  727-734. 

Elvidge,  C.  D.,  V.  R.  Hobson,  K.  E.  Baugh,  J.  Dietz, 
Y.  E.  Shimabukuro,  Y.  E.  Krug,  E.  M.  L.  M.  Novo, 
and  F.  R.  Echavarria. 
2001b.     DMSP-OLS  estimation  of  rainforest  area  impacted 
by   ground   fires   in   Roraima,   Brazil:    1995   versus 
1998.     Int.  J.  Remote  Sens.  22:2661-2673. 
Elvidge.  C.  D.,  M.  L.  Imhoff,  K.  E.  Baugh,  V.  R.  Hobson, 
I.  Nelson,  J.  Safran,  ,  J.  B.  Dietz,  and  B.  T.  Tuttle. 

2001a.     Night-time  lights  of  the  world:  1994-1995.     ISPRS 
J.  Photogram.  Remote  Sens.  56:81-99. 
Harms,  S.,  and  C.  D.  Winant. 

1998.  Characteristic  patterns  of  circulation  in  the  Santa 
Barbara  Channel.     J.  Geophys.  Res.  103:3041-3065. 

Hayward,  T.  L. 

2000.  El  Nino  1997-98  in  the  coastal  waters  of  southern 
California:  a  timeline  of  events.  Calif.  Coop.  Oceanic 
Fish.  Invest.  Rep.  41:98-116. 

Jackson,  G.  D. 

1994.     Statolith  age  estimates  of  the  loliginid  squid  Loligo 
opalescens  (Mollusca:  Cephalopoda):  corroboration  with 
culture  data.     Bull.  Mar.  Sci.  54:554-557. 
Jackson,  G.  D.,  and  M.  L.  Domeier. 

2003.     The  effects  of  an  extraordinary  El  Nino  /  La  Nina 
event  on  the  size  and  growth  of  the  squid  Loligo  opalescens 
off  Southern  California.     Mar.  Biol.  142:925-935. 
Jackson,  G  D.,  and  N.  A.  Moltschaniwskyj. 

2001.  The  influence  of  ration  level  on  growth  and  stato- 
lith increment  width  of  the  tropical  squid  Sepioteuthis 
lessoniana  (Cephalopoda:  Loliginidae):  an  experimental 
approach.     Mar.  Biol.  138:819-825. 

Lavaniegos,  B.  E.,  L.  C. -Perez,  and  G.  Gaxiola-Castro. 

2002.  Plankton  response  to  El  Nino  1997-1998  and 
La  Nina  1999  in  the  southern  region  of  the  California 
Current.     Progress  Oceanog.  54:33-58. 

Lowry.  M.  S.,  and  J.  V.  Carretta. 

1999.  Market  squid  I  Loligo  opalescens)  in  the  diet  of 
California  sea  lions  (Zalophus  californianus)  in  South- 
ern California  (1981-1995).  Calif.  Coop.  Oceanic  Fish. 
Invest.  Rep.  40:196-207. 


Lynn,  R.  J.,  T.  Baumgartner,  J.  Garcia,  C.  A.  Collins. 
T  L.  Hayward,  K.  D.  Hyrenbach,  A.  W.  Mantyla, 
T.  Murphree,  A.  Shankle,  F.  B.  Schwing.  K.  M.  Sakuma, 
and  M.  J.  Tegner. 

1998.     The  state  of  the  California  Current,  1997-1998: 
transition  to  El  Nino  conditions.     Calif.  Coop.  Oceanic 
Fish.  Invest.  Rep.  39:25-49. 
Lynn.  R.  J.,  and  S.  J.  Bograd. 

2002.     Dynamic  evolution  of  the  1997-1999  El  Nino- 
La  Nina  cycle  in  the  southern  California  Current 
System.     Progress  Oceanog.  54:59-75. 
Mangel,  M..  B.  Marinovic,  C.  Pomeroy,  and  D.  Croll. 

2002.     Requiem  for  Ricker:  unpacking  MSY.     Bull.  Mar. 
Sci.  70:763-781. 
Roberts,  M.  J.,  and  W.  H.  H.  Sauer. 

1994.     Environment:  the  key  to  understanding  the  South 
African  chokka  squid  (Loligo  vulgaris  reynaudii)  life 
cycle  and  fishery?     Antarctic  Sci.  6:249-258. 
Rodhouse,  P.  G.,  C.  D.  Elvidge,  and  P.  N.  Trathan. 

2001.  Remote  sensing  of  the  global  light-fishing  fleet: 
an  analysis  of  interactions  with  oceanography,  other 
fisheries  and  predators.     Adv.  Mar.  Biol.  39:261-303. 

Roper,  C.  F.  E.,  M.  J.  Sweeney,  and  C.  E.  Nauen. 

1984.  Cephalopods  of  the  world.  An  annotated  and  illus- 
trated catalogue  of  species  of  interest  to  fisheries.  FAO 
species  catalogue,  vol.  3,  277  p.     FAO  Fish  Synop.  125. 

Salmon,  M..  B.  E.  Witherington,  and  C.  D.  Elvidge. 

2000.  Artificial  lighting  and  the  recovery  of  sea  turtles. 
In  Sea  turtles  of  the  Indo-Pacific:  research  management 
and  conservation  (N.  Pilcher  and  G.  Ismail,  eds.).  p. 
25-34.     ASEAN  Academic  Press,  London. 

Schbn,  P-J.,  W.  H.  H.  Sauer,  and  M.  J.  Roberts. 

2002.  Environmental  influences  on  spawning  aggrega- 
tions and  jig  catches  of  chokka  squid  Loligo  vulgaris 
reynaudii:  a  "black  box"  approach.  Bull.  Mar.  Sci. 
71:783-800. 

Schwing,  F.  B.,  C.  S.  Moore,  S.  Ralston,  and  K.  M.  Sakuma. 

2000.     Record  coastal  upwelling  in  the  California  Cur- 
rent in  1999.     Calif.  Coop.  Oceanic  Fish.  Invest.  Rep. 
41:148-160. 
Schwing.  F  B.,  T.  Murphree.  L.  deWitt.  and  P.  M.  Green. 

2002.     The  evolution  of  oceanic  and  atmospheric  anoma- 
lies in  the  northeast  Pacific  during  the  El  Nino  and  La 
Nina  events  of  1995-2001.     Progress  Oceanog.  54:459- 
491. 
Vojkovich,  M. 

1998.     The  California  fishery  for  market  squid  (Loligo 
opalescens).     Calif.  Coop.  Oceanic  Fish.  Invest.  Rep.  39:55- 
60. 
Waluda,  C.  M.,  P.  N.  Trathan,  C.  D.  Elvidge.  V.  R.  Hobson.  and 
P.  G.  Rodhouse. 

2002.     Throwing  light   on  straddling  stocks  of  Illex 
argentinus:  assessing  fishing  intensity  with  satellite 
imagery.     Can.  J.  Fish.  Aquat.  Sci.  59:592-596. 
Winant,  C.  D.,  and  C.  E.  Dorman. 

1997.  Seasonal  patterns  of  surface  wind  stress  and  heat 
flux  over  the  Southern  California  Bight.  J.  Geophys. 
Res.  102:5641-5653. 

Wolter,  K„  and  M.  S.  Timlin. 

1998.  Measuring  the  strength  of  ENSO  events:  how  does 
1997/98  rank?     Weather  53:315-324. 

Zar,  J.  H. 

1984.  Biostatistical  analysis.  2nd  edition,  718  p.  Prentice- 
Hall.  Englewood  Cliffs,  NJ. 


671 


Abstract-In  May  2001,  the  National 
Marine  Fisheries  Service  (NMFS) 
opened  two  areas  in  the  northwest- 
ern Atlantic  Ocean  that  had  been 
previously  closed  to  the  U.S.  sea 
scallop  I Placopecten  magellanicus) 
dredge  fishery.  Upon  reopening  these 
areas,  termed  the  "Hudson  Canyon 
Controlled  Access  Area"  and  the  "Vir- 
ginia Beach  Controlled  Access  Area." 
NMFS  observers  found  that  marine 
turtles  were  being  caught  inciden- 
tally in  scallop  dredges.  This  study 
uses  the  generalized  linear  model  and 
the  generalized  additive  model  fitting 
techniques  to  identify  environmen- 
tal factors  and  gear  characteristics 
that  influence  bycatch  rates,  and  to 
predict  total  bycatch  in  these  two 
areas  during  May-December  2001  and 
2002  by  incorporating  environmental 
factors  into  the  models.  Significant 
factors  affecting  sea  turtle  bycatch 
were  season,  time-of-day,  sea  sur- 
face temperature,  and  depth  zone.  In 
estimating  total  bycatch,  rates  were 
stratified  according  to  a  combination 
of  all  these  factors  except  time-of- 
day  which  was  not  available  in  fish- 
ing logbooks.  Highest  bycatch  rates 
occurred  during  the  summer  season, 
in  temperatures  greater  than  19°C, 
and  in  water  depths  from  49  to  57  m. 
Total  estimated  bycatch  of  sea  turtles 
during  May-December  in  2001  and 
2002  in  both  areas  combined  was  169 
animals  ( CV=  55.3 ),  of  which  164  ( 97% ) 
animals  were  caught  in  the  Hudson 
Canyon  area.  From  these  findings,  it 
may  be  possible  to  predict  hot  spots 
for  sea  turtle  bycatch  in  future  years 
in  the  controlled  access  areas. 


Magnitude  and  distribution  of  sea  turtle  bycatch 
in  the  sea  scallop  (Placopecten  magellanicus) 
dredge  fishery  in  two  areas  of  the 
northwestern  Atlantic  Ocean,  2001-2002 


Kimberly  T.  Murray 

Northeast  Fisheries  Science  Center 
National  Marine  Fisheries  Service 
166  Water  Street 

Woods  Hole,  Massachusetts  02543 
E-mail  address.  Kimberly  Murray<S'noaa  gov 


Manuscript  submitted  1  December  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
27  May  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:671-681  12004). 


Five  species  of  sea  turtles  in  the 
northwestern  Atlantic  Ocean  are 
protected  under  the  U.S.  Endangered 
Species  Act  of  1973.  The  loggerhead 
turtle  (Caretta  caretta)  is  listed  as  a 
threatened  species,  and  the  leather- 
back  (Dermochelys  coriacea),  hawks- 
bill  (Eretmochelys  imbricata),  Kemp's 
ridley  iLepidochelys  kempii),  and 
certain  populations  of  the  green  sea 
turtle  (Chelonia  mydas)  are  listed  as 
endangered.  Populations  of  each  of 
these  species  have  declined  principally 
as  a  result  of  human  activities  (NRC, 
1990). 

The  incidental  capture,  or  bycatch, 
of  sea  turtles  in  commercial  fisheries 
is  a  major  source  of  mortality  (NRC, 
1990;  Turtle  Expert  Working  Group, 
2000).  These  turtles  are  captured  in- 
cidentally in  pelagic  longlines  (Lewi- 
son  et  al.,  2004),  trawls  (Epperly, 
2003),  gill  nets  (Julian  and  Beeson. 
1998),  pound  nets,  weirs,  pots,  and 
traps  (NMFS  and  USFWS,  1991;  Al- 
len, 2000).  Such  threats  occur  at  vari- 
ous life  stages  of  a  population  and  at 
different  intensities,  and  consequent- 
ly have  implications  for  management 
policy  (Heppell  et  al.,  2003). 

The  U.S.  National  Marine  Fisher- 
ies Service  (NMFS)  has  implemented 
management  measures  in  both  the 
Atlantic  and  the  Pacific  in  the  form 
of  gear  modifications  or  time  and  area 
closures  to  reduce  sea  turtle  bycatch. 
For  example,  since  the  early  1990s, 
turtle  excluder  devices  (TEDs)  have 
been  required  in  all  inshore  and  off- 
shore shrimp  trawl  nets  in  southeast- 
ern U.S.  waters  (Epperly,  2003)  to  re- 
duce sea  turtle  mortality  (Henwood 


and  Stuntz,  1987).  Bycatches  of  sea 
turtles  in  the  U.S.  pelagic  longline 
fisheries  for  swordfish  and  tuna  (Wit- 
zell,  1999)  led  to  a  year-round  clo- 
sure of  a  2.6  million  nmi2  area  in  the 
northwestern  Atlantic  Ocean  to  these 
fisheries  beginning  in  2002. 

In  recent  years,  documented  inter- 
actions have  occurred  between  sea 
turtles  and  sea  scallop  dredges,  a  pre- 
viously unidentified  threat  in  recovery 
planning  efforts  (NMFS,  1991).  Dur- 
ing 2001  and  2002,  fisheries  observers 
aboard  commercial  sea  scallop  vessels 
documented  the  bycatch  of  sea  turtles 
in  two  small  regions  of  the  Mid-Atlan- 
tic Bight  (MAB).  These  areas,  termed 
the  "Hudson  Canyon  Controlled  Ac- 
cess Area"  (approximately  3150  km2i 
and  the  "Virginia  Beach  Controlled 
Access  Area"  (approximately  900  km2) 
were  closed  to  scallop  fishing  in  April 
1998  but  reopened  in  May  2001  on  a 
conditional  basis  (Fig.  1).  This  study 
uses  the  generalized  linear  and  gen- 
eralized additive  models  to  identify 
environmental  factors  and  gear  char- 
acteristics affecting  the  bycatch  rate 
of  sea  turtles  in  these  two  areas  and 
to  predict  total  bycatch  by  sea  scallop 
dredge  vessels  in  these  two  areas  in 
2001  and  2002. 


Methods 

The  fishery 

In  2001  and  2002.  137  and  93  com- 
mercial vessels,  respectively,  partici- 
pated in  the  Controlled  Area  Access 
Program  sea  scallop  fishery.  Although 


672 


Fishery  Bulletin  102(4) 


40"0'0"N 


74=0'0"W 

1 


A 


'  Hudson  Canyon 
Controlled  Access  Area 


Virginia  Beach 
Controlled  Access  Area 


Mid-Atlantic  Controlled  Access  Areas 
for  the  sea  scallop  fishery  2001  and  2002 

50  Fathom  Isobath 

100  Fathom  Isobath 


Figure  1 

Mid-Atlantic  controlled  access  areas  for  the  sea  scallop  fishery  2001 
and  2002. 


the  U.S.  commercial  scallop  fishery  operates  year-round, 
the  area  access  program  in  2001  began  on  1  May,  and 
in  2002  on  1  March,  and  ended  on  28  February  follow- 
ing the  respective  fishing  year  (1  March-28  February). 
Vessels  in  the  controlled  access  areas  fished  around 
the  clock  for  approximately  5-12  days,  accomplishing 
between  40  and  160  hauls  per  trip.  Dredges  in  the  con- 
trolled areas  were  generally  fished  at  depths  between 
45  and  75  m.  The  average  haul  duration  was  about  1 
hour.  Most  vessels  fished  two  dredges  simultaneously 
(one  from  each  side  of  the  vessel),  which  were  generally 
either  3.9  or  4.5  m  (13  or  15  ft)  wide. 

Vessels  in  the  Mid-Atlantic  typically  fish  with  a  New 
Bedford  style  scallop  dredge  equipped  for  soft-bottom 
substrates.  In  this  configuration,  tickler  chains  strung 
from  the  sweep  chain  run  horizontally  between  the 


dredge  frame  and  the  ring  bag  and  are  designed  to 
raise  scallops  off  the  bottom  and  into  the  bag.  Turtles 
become  entrapped  in  the  ring  bag  or  on  the  dredge 
frame.  For  dredging  on  hard  bottom,  such  as  in  New 
England,  vertical  up  and  down  chains  hang  over  the 
tickler  chains,  preventing  boulders  from  entering  the 
ring  bag  (Smolowitz,  1998).  New  Bedford  style  scallop 
dredges  have  also  been  used  in  U.S.  fisheries  in  the 
Pacific  (PSMFC1). 


1  Pacific  States  Marine  Fisheries  Commission  (PSMFC  I. 
2003.  Description  of  fishing  gears  used  on  the  Pacific 
Coast,  http://pcouncil.org/habitat/geardesc.pdf.  [Accessed 
6  April  2004.] 


Murray:  Magnitude  and  distribution  of  sea  turtle  bycatch  in  the  sea  scallop  dredge  fishery 


673 


Data  sources 

Observer  data  Observers  were  placed  on  randomly 
selected  vessels  fishing  in  the  controlled  areas  to  record 
the  bycatch  of  turtles  and  other  protected  species.  From 
May  to  December  in  2001  and  2002,  observers  sampled 
11%  of  the  commercial  fishing  effort  in  the  Hudson 
Canyon  region,  and  in  October  2001,  16%  of  the  effort  in 
Virginia  Beach.  No  trips  were  observed  in  the  Virginia 
Beach  region  during  2002  because  of  low  commercial 
fishing  effort  in  the  area.  Observers  were  on-  and  off- 
watch  on  an  irregular  schedule  throughout  a  24-hour 
period,  observing  on  average  65%  of  the  hauls  on  a  trip. 
When  a  dredge  was  hauled  on  board,  observers  recorded 
the  haul  location,  time,  depth,  tow  speed,  tow  duration, 
number  of  dredges  observed,  and  the  presence  or  absence 
of  turtle  bycatch.  In  2001,  observers  identified  20%  of  the 
turtles  that  came  aboard  as  loggerhead  sea  turtles  but 
were  unable  to  identify  the  remaining  80%.  As  a  result 
of  improved  observer  training  (NMFS  200.3),  observers 
identified  88%  of  the  turtles  as  loggerhead  sea  turtles 
in  2002,  but  they  were  unable  to  identify  the  remain- 
ing 12%.  Given  that  observers  document  the  loggerhead 
species  most  commonly  in  the  Mid-Atlantic  area,  and 
that  all  sea  turtles  positively  identified  were  loggerhead 
sea  turtles,  bycatch  estimates  in  this  analysis  are  con- 
sidered to  be  those  of  loggerhead  sea  turtles.  Although 
some  turtles  may  have  been  released  alive  or  injured, 
this  analysis  does  not  differentiate  between  live,  dead, 
and  injured  animals. 

Fishing  effort  data  Under  the  1982  Atlantic  Sea  Scallop 
Fishery  Management  Plan,  all  vessels  targeting  scallops 
must  complete  a  vessel  trip  report  ( VTR)  log  (as  of  1994) 
indicating  area  fished,  kept  and  discarded  catch,  and 
fishing  effort.  These  data  were  used  to  estimate  the  total 
fishing  effort  of  the  fleet.  In  calculating  fishing  effort, 
one  unit  of  effort  equals  a  single  dredge  haul  because 
vessels  may  fish  one  or  two  dredges  simultaneously  on 
each  haul.  Because  a  preliminary  analysis  showed  that 
tow  duration  or  dredge  length  does  not  significantly 
affect  the  probability  of  turtle  capture,  dredge  haul  effort 
was  not  standardized  for  these  two  variables.  All  VTR 
trips  from  May  to  December  in  the  controlled  areas  were 
used  in  the  analysis.  Because  completion  of  vessel  trip 
reports  is  mandatory  and  trips  to  the  controlled  areas 
were  closely  monitored,  it  was  assumed  that  the  VTR 
data  represented  100%  of  total  fishing  effort. 

Sea  surface  temperature  Sea  surface  temperature  at 
each  position  reported  in  the  observer  and  VTR  data- 
bases was  extracted  from  NOAA  AVHRR  (advanced 
very  high  resolution  radiometer)  coastwatch  satellite 
images.  A  Visual  Basic  (Microsoft  Corp.,  Redmond,  WA) 
routine  was  used  to  extract  temperatures  from  7-day 
composite  images  (3  days  forward  and  backward  from 
the  haul  date),  by  using  a  3x3  cell  window  at  1-km 
resolution.  Therefore,  a  9-km2  area  of  coverage  around 
each  coordinate  position  was  used  to  extract  sea  surface 
temperature.  Within  the  3x3  cell  search  radius,  the  pixel 


representing  the  warmest  temperature  was  used  to  avoid 
temperatures  affected  by  cloud  coverage. 

Data  analysis 

Missing  temperature  data  Sea  surface  temperature 
values  could  not  be  obtained  for  33%  of  the  VTR  data 
and  10%  of  the  observer  data  because  of  either  missing 
coordinate  positions  on  the  VTR  logs  or  bad  satellite 
images.  For  these  fishing  events,  sea  surface  tempera- 
ture was  predicted  by  using  a  linear  regression  based 
on  year,  month,  and  area.  For  the  observer  data,  area 
was  defined  as  either  Hudson  Canyon  or  Virginia  Beach 
access  areas  (r2=0.88).  For  the  VTR  data,  the  vessel's 
home  state  served  as  a  proxy  for  area  fished  because 
most  of  the  missing  temperature  values  were  due  to 
missing  coordinate  positions  (r2=0.86). 

Modeling  approach  Generalized  linear  model  (GLM) 
and  generalized  additive  model  (GAM)  fitting  techniques 
were  used  to  understand  and  predict  bycatch  rates  of  sea 
turtles  in  relation  to  environmental  variables,  fishing 
practices,  and  gear  characteristics  in  the  commercial 
sea  scallop  fishery.  Unlike  classic  linear  regression 
models,  GLMs  and  GAMs  allow  for  nonlinearity  and 
nonconstant  variance  structures  in  the  data  (Guisan 
et  al.,  2002).  GAMs  differ  from  GLMs  in  that  smooth 
functions  replace  the  linear  predictors  in  GLMs  (Hastie 
and  Tibshirani,  1990).  Smooth  functions,  or  "smoothers," 
summarize  the  trend  of  a  response  measurement  as  a 
function  of  multiple  predictors  (Hastie  and  Tibshirani, 
1990)  and  therefore  some  form  of  parametric  relation- 
ship between  the  response  and  explanatory  variables 
is  not  assumed  (Guisan  et  al.  2002).  Both  frameworks 
have  been  used  to  model  abundance  or  probability  events 
as  a  function  of  environmental  variables  (Frost  et  al., 
1999;  Denis  et  al.,  2002;  Guisan  et  al.,  2002;  Hamazaki, 
2002). 

A  modeling  approach  to  estimate  bycatch  of  sea  tur- 
tles in  the  sea  scallop  dredge  fishery  was  preferred  over 
the  ratio  method  (Cochran,  1977)  that  has  been  used 
to  estimate  bycatch  of  marine  mammals  and  turtles 
in  other  fisheries  (Epperly  et  al.,  1995;  Rossman  and 
Merrick,  1999).  With  the  ratio  method,  the  observed 
number  of  sea  turtles  divided  by  the  observed  effort  is 
used  to  calculate  a  bycatch  rate,  and  this  rate  is  then 
multiplied  by  total  commercial  fishing  effort  to  derive 
a  bycatch  estimate.  Bycatch  data  in  the  sea  scallop 
dredge  fishery  violate  the  underlying  assumptions  of  the 
ratio  method  (Cochran,  1977),  largely  because  sea  turtle 
bycatch  is  binomially  distributed  with  a  nonconstant 
variance.  An  analyis  of  binary  response  data  derived 
from  a  statistical  model  allows  bycatch  rates  to  be  pre- 
dicted by  using  factors  that  account  for  variability  in 
bycatch.  Moreover,  stratifying  bycatch  rates  according 
to  these  factors  will  reduce  variability  in  total  bycatch 
estimates.  For  the  sea  turtle  data  analyzed  in  the  pres- 
ent study,  the  GLM  approach  provided  a  more  accurate 
and  less  biased  mortality  estimate  than  that  derived 
using  the  ratio  method. 


674 


Fishery  Bulletin  102(4) 


GAM  smoothers  Before  a  GLM  was  constructed,  a 
GAM  helped  group  continuous  variables  into  catego- 
ries. Fitting  the  GLM  model  with  categorized  variables 
was  necessary  to  extrapolate  bycatch  rates  in  order  to 
derive  a  total  estimate  of  the  bycatch  of  sea  turtles  in 
scallop  dredges  in  the  controlled  access  areas.  All  of  the 
variables  tested  in  the  GLM  model  were  first  fitted  to 
a  GAM,  in  which  the  parameters  of  the  continuous  pre- 
diction variables  were  estimated  by  a  smoothing  spline. 
Variable  values  were  grouped  according  to  whether  they 
had  a  positive  or  negative  influence  on  the  bycatch  rate 
(i.e.,  the  group  explained  more  or  less  of  the  bycatch 
rate). 

Development  of  a  GLM  bycatch  model  Because  bycatch 
events  were  counts  ranging  from  zero  or  one,  a  logistic 
regression  was  used  to  model  the  probability  of  sea 
turtle  bycatch  (GLM  function,  SPLUS  6.1,  Seattle,  WA). 
Each  dredge  haul  is  a  data  point  and  the  response  was 
whether  turtle  bycatch  was  zero  or  one.  Probability  of 
sea  turtle  bycatch  (p)  was  calculated  as 


p  =  ey  1 1  +  ey 

y  =  P0+P1x1+P2x2+...+  pixi, 

where  pt  is  a  parameter  coefficient; 
xl   is  a  predictor  variable;  and 
y    is  a  sea  turtle  bycatch  event. 

Dredge  hauls  are  assumed  to  be  independent  because 
turtles  were  never  simultaneously  caught  in  both  dredges 
operating  from  a  vessel  during  a  single  haul. 

A  forward  stepwise  selection  method  was  used  to  de- 
termine the  best  fitting  model.  Model  parameters  were 
estimated  by  maximizing  the  log-likelihood  function. 
The  null  model  was  the  first  model  in  the  stepwise 
process  and  was  specified  with  a  single  intercept  term 
as 

H0:  logiturtle  bycatch)  =  1. 

At  each  step,  a  new  variable  was  added  to  the  null  model 
(Appendix  1)  and  tested  against  the  former  model  formu- 
lation (ANOVA  function,  chi-square  test)  to  determine 
the  better  fitting  model.  A  preliminary  assessment  of  a 
broad  suite  of  gear  characteristics  and  environmental 
factors  indicated  that  10  variables  could  significantly 
affect  bycatch  rates.  The  main  effects  of  each  variable 
were  tested  in  the  stepwise  selection  process  as  well 
as  the  interaction  between  season  and  temperature. 
Because  the  order  of  the  predictor  variables  affects  their 
significance,  main  effects  were  entered  in  various  orders. 
If  a  P-value  was  less  than  0.05,  then  the  additional  vari- 
able was  considered  to  explain  more  of  the  variability  in 
bycatch  than  a  model  without  that  variable.  Each  new 
model  was  also  compared  against  the  former  model  by 
using  the  Akaike  information  criterion  (AIC),  which  is 
defined  as 


A/C  =  -21og(L(0ly))  +  2.K', 

where  log(L(6H  y ))  =  the  numerical  value  of  the  log-likeli- 
hood at  its  maximum  point;  and 
K  =  the  number  of  estimable  parameters 
(Burnham  and  Anderson,  20021. 

The  AIC  is  a  measure  of  the  level  of  parsimony,  defined 
as  a  model  that  fits  the  data  well  and  includes  as  few 
parameters  as  necessary  (Palka  and  Rossman,  2001).  If 
the  AIC  value  decreases,  the  new  combination  of  vari- 
ables in  the  model  fit  the  data  better. 

To  investigate  whether  the  bycatch  data  are  over- 
dispersed,  that  is.  where  the  sampling  variance  exceeds 
the  theoretical  variance,  the  GLM  model  was  refitted 
by  using  a  quasi-likelihood  function.  When  data  are 
over-dispersed,  the  estimated  over-dispersion  parameter 
is  generally  between  1  and  4  (Burnham  and  Ander- 
son, 2002).  The  over-dispersion  parameter  fitted  to  the 
global  model  was  0.61,  indicating  these  data  were  not 
over-dispersed  and  error  assumptions  of  the  binomial 
model  were  appropriate  for  analyzing  these  data. 

Alias  patterns  in  the  final  model  were  examined  to 
assess  correlation  among  the  explanatory  variables. 
The  fit  of  the  final  model  was  assessed  by  plotting  the 
observed  turtle  bycatch  against  the  predicted  turtle 
bycatch.  The  r2  value  indicated  how  well  predictions 
from  the  linear  model  fit  the  actual  data. 

Bycatch  rate  estimates  The  spatial  and  temporal  strati- 
fication of  bycatch  rates  in  each  of  the  controlled  access 
areas  was  determined  by  the  explanatory  variables  in 
the  best-fitting  GLM.  Parameter  estimates  from  the 
model  were  used  to  predict  the  bycatch  rate  for  each 
stratum. 

The  coefficient  of  variation  (CV)  for  each  bycatch 
rate  was  estimated  by  bootstrap  resampling  (Efron  and 
Tibshirani,  199.3).  The  resampling  unit  was  a  scallop 
dredge  haul.  Replicate  bycatch  rates  were  generated 
with  the  best-fitting  GLM  model,  by  sampling  with 
replacement  1000  times  from  the  original  data  set. 
The  CV  was  defined  as  the  standard  deviation  of  the 
bootstrap  replicate  bycatch  rates  in  a  stratum  divided 
by  the  bycatch  rate  for  that  stratum  estimated  from  the 
original  data.  Variances  and  CVs  of  combined  estimates 
were  based  on  means  weighted  by  their  respective  vari- 
ances (Wade  and  Angliss,  1997). 

Total  bycatch  The  total  estimated  turtle  bycatch  in 
each  stratum  was  calculated  as  the  product  of  predicted 
bycatch  per  dredge  haul  (i.e.,  the  predicted  bycatch 
rate)  for  that  stratum  and  the  total  number  of  dredge 
hauls  accomplished  by  the  commercial  fishery  in  that 
stratum: 


^  Predicted  bycatch 
V  Dredge  hauls t 

where  ;  =  stratum 


■  [Total  dredge  hauls),. 


Murray:  Magnitude  and  distribution  of  sea  turtle  bycatch  in  the  sea  scallop  dredge  fishery 


675 


Table  1 

Analysis  of  deviance  for  significant  factors  a 
to  contruct  a  model  to  predict  total  bycatch 

ffectir 
AIC  = 

igsea  turtle  bycatch.  Significant  factors 
Akaike  information  criterion. 

were  used  to  stratify  bycatch  rates  and 

Model 

df 

Deviance 

Residual  df 

Residual  deviance 

Pfchil 

AIC 

null  model  only 

18,071 

405.29 

407.2989 

n  ui 'I  +  year 

1 

-2.33 

18,070 

402.96 

0.12626 

406.9611 

null  +  season 

2 

19.81 

18,069 

385.48 

0.00004 

391.4807 

null  +  season  +  temp 

1 

9.34 

18,068 

376.13 

0.00223 

384.1319 

null  +  season  +  temp  +  depth 

2 

17.23 

18,066 

358.89 

0.00018 

370.8983 

null  +  season  +  temp  +  depth  +  time  of  day 

1 

7.86 

18,065 

351.03 

0.00503 

365.0318 

null  +  depth  +  time  of  day  +  season  (temp) 

1 

1.64 

18,064 

349.39 

0.20011 

365.3903 

null  +  season  +  temp  +  depth  +  time  of  day 
+  state 

4 

3.77 

18,061 

347.25 

0.43746 

369.2579 

null  +  season  +  temp  +  depth  +  time  of  day 
+  dredge  frame  width 

2 

3.27 

18.063 

347.76 

0.19487 

365.7611 

null  +  season  +  temp  +  depth  +  time  of  day 
+  number  of  up  and  down  chains 

1 

0.54 

18,064 

350.48 

0.45955 

366.4849 

null  +  season  +  temp  +  depth  +  time  of  day 
+  number  of  tickler  chains 

1 

3.18 

18,064 

347.84 

0.07436 

363.8480 

Annual  bycatch  was  the  sum  of  the  stratified  bycatch 
estimates.  The  finite  population  correction  factor  (Co- 
chran, 1977)  was  applied  to  bycatch  estimates  in  stratas 
where  the  observer  coverage  was  greater  than  10%. 

Number  of  dredge  hauls  in  the  VTR  database  without 
coordinate  positions  (32%)  were  prorated  between  the 
stratified  areas  according  to  the  percentage  of  dredge 
hauls  with  known  coordinates  from  the  same  year, 
state,  and  stratified  areas. 


Results 

Observed  bycatch 

Nine  and  16  turtle  bycatch  were  observed  in  2001  and 
2002,  respectively,  in  the  Hudson  Canyon  controlled 
access  area.  Of  the  25  turtles  taken  in  the  Hudson 
Canyon  area  across  both  years,  21  (84%)  were  taken 
during  summer  months.  Two  turtle  bycatch  were 
observed  in  the  Virginia  Beach  access  area  during  fall 
2001 — the  only  time  when  there  was  observer  coverage 
in  this  area  across  both  years. 

GAM  smoothers 

Plots  of  the  smoothed  functions  in  the  GAM  revealed 
whether  the  continuous  variable  in  the  model  explained 
any  error  in  the  bycatch  rate  estimates.  For  example, 
a  plot  of  the  smooth  function  for  depth  as  a  covariate 
revealed  that  bycatch  rates  may  be  higher  between  49 
m  (27  fm)  and  57  m  (31  fm)  and  lower  around  this  zone 
(Fig.  2).  Likewise,  a  plot  of  the  smooth  function  for  tem- 
perature as  a  covariate  revealed  that  bycatch  rates  may 


be  higher  above  19°C.  These  plots  helped  bin  the  continu- 
ous variables  into  categories  (Appendix  1)  which  could 
then  be  tested  in  the  GLM.  All  continuous  variables  in 
the  GAM  were  categorized  in  a  similar  manner. 

GLM  bycatch  model 

Significant  factors  affecting  sea  turtle  bycatch  were 
season,  sea  surface  temperature,  depth  zone,  and  time- 
of-day  (Table  1).  These  variables  were  significant  despite 
the  order  in  which  they  were  tested  in  the  model.  The 
model  with  the  lowest  AIC  value  was  considered  the 
"best"  model,  although  time-of-day  could  not  be  included 
in  the  final  model  to  predict  bycatch  rates.  This  level  of 
information  is  not  recorded  in  commercial  fisheries  log- 
books; therefore  bycatch  rates  based  on  time-of-day  could 
not  be  extrapolated  to  total  bycatch.  Width  of  the  scallop 
dredge  frame,  number  of  tickler  chains,  and  number  of 
up  and  down  chains  were  not  significant  variables. 

Model  fit 

The  number  of  predicted  sea  turtle  bycatch  closely 
matched  the  observed  bycatch  in  both  years  in  all 
bycatch  strata  (Table  2).  Strata  were  defined  according 
to  variables  identified  in  the  GLM  as  having  a  significant 
effect  on  bycatch  rates.  The  relationship  between  actual 
and  observed  takes  was  strong  (r2=0.93),  indicating  that 
the  predictions  from  the  model  fitted  the  data  well. 

Bycatch  rate  estimates 

Bycatch  rates  were  stratified  by  season,  temperature  inter- 
val, and  depth  zone  (Table  3).  Because  year  was  not  a 


676 


Fishery  Bulletin  102(4) 


10  - 

y' 

0  - 
§•  -10- 

-20 

20                    25                   30                    35                   40                    45 

Depth  (fm) 

10  - 

'*■* 

0 

'"'''-•-•--•.-._f^^^ 

s  (temperature) 
o                   o 

^/ 

-30  _ 

0                    5                   10                 15                 20                  25                  30 

Temperature  (C) 

Figure  2 

Partial  fits  for  the  general  additive  model  (GAM)  of  sea  turtle  bycatch  with 

depth  and  temperature  as  covariates,  showing  the  relationship  estimated 

by  a  smoothing  spline.  Depths  between  27  fm  (49  ml  and  31  fm  (57  m).  and 

temperatures  above  19°C,  have  a  positive  influence  on  the  bycatch  rate. 

95^  confidence  bands  are  also  shown.  All  continuous  variables  in  the  GAM 

were  categorized  in  a  similar  manner.  The  "s"  on  the  y-axis  represents  a 

smoothed  function  for  each  variable  and  explains  the  effect  of  each  variable 

on  sea  turtle  bycatch  per  haul. 

significant  factor  in  the  final  model,  predicted  bycatch 
rates  were  the  same  for  2001  and  2002.  Highest  sea  turtle 
bycatch  rates  occurred  during  the  summer  season  (Aug- 
Sep),  in  temperatures  warmer  than  19C,  in  water  depths 
from  49  to  57  m.  Lowest  bycatch  rates  occurred  during  the 
fall  (Oct-Dec)  and  spring  (May-June),  in  temperatures 
cooler  than  19  C,  and  in  water  depths  less  than  49  m. 


Total  bycatch 

The  total  estimated  bycatch  of  sea  turtles  in  the  Mid- 
Atlantic  controlled  access  areas  in  2001  and  2002  com- 
bined was  169  animals  (CV=55.3)  (Table  4).  Of  this  total, 
164  animals  (97'*  )  were  caught  in  the  Hudson  Canyon 
area:  69  (42',  i  in  2001  and  95  (58%)  in  2002.  Total  esti- 


Murray:  Magnitude  and  distribution  of  sea  turtle  bycatch  in  the  sea  scallop  dredge  fishery 


.  677 


Table  2 

Observec 

versus 

predicted  number  of  turtle  byeatch,  by  stratum.  2001  and  2002.  Obs.= 

observed;  Pred. 

=predicted. 

Spr 

ng 

Summer 

] 

Fall 

Number  of 

Number  of 

Number  of 

Number  of 

Number  of 

Number  of 

obs. 

pred. 

obs. 

pred. 

obs. 

pred. 

Water  depth 

Temp. 

turtle  bycatch 

turtle  bycatch 

turtle  bycatch 

turtle  bycatch 

turtle  bycatch 

turtle  bycatch 

Shallow 

High 

0 

0 

0 

0 

0 

0 

Low 

0 

0 

0 

0 

0 

0 

Mid-depth 

High 

2 

1 

17 

16 

1 

2 

Low 

0 

0 

0 

0 

1 

0 

Deep 

High 

1 

1 

4 

5 

1 

0 

Low 

0 

0 

0 

0 

0 

0 

Table  3 

Stratification  of  turtle  bycatch 

rates  with  associated  CVs. 

N.C.E.=no  commercial  effort. 

Water  depth 

Temperature 

Spring  (May- June l 

Summer  (Aug-Sepi 

Fall(Oct-Dec) 

Shallow  (<49  ml 

High(>19°C) 

0.0000027(82.5) 

0.0000052(62.1) 

0.0000030(87.71 

Low(<19°C) 

0.0000002(99.5) 

N.C.E. 

0.0000002(106.6) 

Mid-depth  (49-57 

ml                High(>19°C) 

0.0032018(64.9) 

0.0061179(25.4) 

0.0035838(57.2) 

Low(<19°C) 

0.0002117(95.4) 

N.C.E. 

0.0002371(98.3) 

Deep(>57m) 

High(>19°C) 

0.0007578(73.8) 

0.0014512(41.9) 

0.0008485(80.5) 

Low(<19°C) 

0.0000500(92.91 

N.C.E. 

0.0000560(103.8) 

Table  4 

Total  bycatch  estimates  by 

year  and 

season 

with 

weighted  CVs  (%) 

N.C.E. 

=no  commercial  effort. 

Spring 

Summer 

Fall 

Total 

Hudson  Canyon 

2001 

10(89.2) 

50(61.5) 

9(105.8) 

69 

2002 

13(89.2) 

78(61.5) 

4(105.8) 

95 

Virginia  Beach 

2001 

N.C.E. 

N.C.E. 

5(105.8) 

5 

2002 

0 

0 

N.C.E. 

0 

Totals 

23 

128 

18 

169(55.31 

mated  bycatch  of  turtles  in  the  Virginia  Beach  area  was 
five  animals  in  2001  and  zero  animals  in  2002. 

Across  both  areas,  the  highest  bycatches  occurred  in 
summer  (128  turtles;  76%),  followed  by  spring  (23  tur- 
tles; U7( )  and  fall  (18  turtles;  10%)  (Table  5).  One  hun- 
dred thirty-two  (78%)  (CV=49.6)  sea  turtles  were  caught 
in  the  mid-depth  zone  from  49  to  57  m,  whereas  37  (22%) 
(CV=59.6)  sea  turtles  were  caught  in  waters  deeper 
than  57  m.  One-hundred  fifty-eight  (93%)  (CV=51.2)  sea 
turtles  were  caught  in  waters  warmer  than  19°C,  and  11 
(7%)  (CV=74.9)  in  waters  cooler  than  19°C. 


Discussion 

Use  of  bycatch  models 

Generalized  linear  and  generalized  additive  models 
help  to  identify  environmental  variables  or  fishing 
practices  that  influence  the  probability  of  sea  turtle 
bycatch.  In  estimating  total  mortality,  bycatch  rates  can 
then  be  stratified  according  to  these  factors,  reducing 
unexplained  variability  in  the  total  estimate.  More- 
over, understanding  factors  that  lead  to  a  high  or  low 


678 


Fishery  Bulletin  102(4) 


Table  5 

Total  bycatch  estimates  by  season,  depth. 

and  temperat 

jre  strata  in 

Hudson  Canyon  and  Virginia  Beach  controlled  access  areas 

in  2001  and  2002  with  95<7r  confidence  intervals.  Spring=May-Jun; 

Summer=Jul 

-Sep;  Fall= 

=Oct-Dec.  N.E.C 

=  no  commercial 

effort;  N.O.=no  observer 

coverage. 

Water  depth 

Temperature 

2001 

2002 

Total 

Spring 

Summer 

Fall 

Spring 

Summer 

Fall 

Shallow  l<49  ml 

High(>19°C) 

0 

0 

0 

0 

0 

0 

0 

Low«19°Ci 

0 

N.O. 

0 

0 

N.C.E. 

0 

0 

Mid-Depth  (49-57  mi 

High(>19°C) 

6(0-13) 

37(21-59) 

2(0-5) 

8(0-211 

65(34-961 

5(0-9) 

123 

Low(<19°C) 

2(0-8) 

N.C.E. 

4(0-191 

1(0-2) 

N.C.E. 

2i0-10i 

9 

Deep  (>57  ml 

High(>19°C) 

2(0-51 

13(4-25i 

2(0-51 

4i0-lll 

13(4-24) 

1  (0-1) 

35 

Low«19°C) 

0(0-11 

N.C.E. 

1  (0-6) 

0(0-1) 

N.C.E. 

1(0-2) 

2 

Total 

10 

50 

9 

13 

78 

9 

169 

probability  of  bycatch  can  motivate  bycatch  mitigation 
research.  Finally,  the  ability  to  predict  bycatch  on  the 
basis  of  explanatory  variables  allows  one  to  examine  the 
relative  effectiveness  of  different  management  measures 
designed  to  reduce  bycatch  (Kobayashi  and  Polovina2). 
Ultimately  this  framework  can  improve  the  assessment 
of  threats  to  turtles  and  broaden  conservation  options. 

Magnitude  of  bycatch 

During  May-December  in  2001  and  2002,  an  estimated 
169  animals  were  captured  incidentally  by  commercial 
sea  scallop  dredge  vessels  in  two  areas  of  the  Mid- Atlan- 
tic Bight.  Throughout  the  entire  Mid-Atlantic  Bight,  the 
magnitude  of  bycatch  was  probably  larger,  particularly 
because  the  factors  associated  with  the  high  bycatch 
rates  were  not  specific  to  the  controlled  access  areas.  Of 
the  11  observed  turtles  measured  for  size,  9  (82%)  were 
between  70-80  cm  straight  carapace  length  (the  large 
juvenile  stage).  Stage  class  models  indicate  that  the  long- 
term  survivability  of  loggerhead  sea  turtles  is  sensitive 
to  mortality  at  this  life  stage  (Crouse  et  al.,  1987). 

Factors  influencing  bycatch 

The  incidental  capture  of  turtles  occurs  where  there 
is  overlap  between  fishing  effort  and  turtle  habitat. 
The  elevated  probability  of  turtle  bycatch  occurring 
in  warm  waters,  during  summer,  at  depths  between 
50  and  60  m  is  consistent  with  the  habitat  regime  of 
loggerhead  sea  turtles  in  the  Mid-Atlantic  (Shoop  and 
Kenney,  1992;  Epperly  et  al.,  1995;  Coles  and  Musick, 
2000).  During  the  oceanic  phase  of  their  life  cycle,  sea 
turtles  occupy  habitats  at  specific  temperatures  or  with 


bathymetric  features  that  concentrate  prey  and  other 
areas  of  enhanced  productivity  (Polovina  et  al..  2000). 
In  Mid-Atlantic  waters,  high  aggregations  of  loggerhead 
sea  turtes  have  been  observed  in  the  summer,  in  waters 
22-49  m  deep,  at  temperatures  from  20°  to  24°C  (Shoop 
and  Kenney,  1992).  In  the  Hudson  Canyon  and  Virginia 
Beach  controlled  access  creas,  the  bycatch  of  sea  turtles 
was  associated  with  habitat  conditions  rather  than  gear 
characteristics.  From  these  findings,  it  may  be  possible 
to  predict  future  hotspots  for  sea  turtle  bycatch  in  the 
controlled  access  areas  where  fishing  effort  and  sea 
turtles  overlap  in  time  and  space.  These  hotspots  may 
be  centered  over  the  portion  of  the  Hudson  Canyon 
where  depths  are  between  50  and  60  m.  after  waters 
warm  to  19  C. 

Because  of  the  low  amount  of  observer  data  in  the 
Virginia  Beach  area,  predicted  bycatch  rates  for  this 
area  were  based  largely  on  conditions  within  the  Hud- 
son Canyon  area.  Sea  scallop  fishing  effort  occurs  year- 
round  both  north  and  south  of  the  Hudson  Canyon, 
and  high  concentrations  of  loggerhead  sea  turtles  (de- 
termined from  migratory  patterns)  exist  in  spring  and 
fall  from  North  Carolina  to  northern  Maryland  ( Shoop 
and  Kenney,  1992).  It  is  probable  that  the  distribution 
of  turtles  and  scallop  fishing  effort  co-occur  in  other 
regions  of  the  Mid-Atlantic,  particularly  south  of  the 
Hudson  Canyon.  The  scallop  dredge  fishery  in  the  Mid- 
Atlantic  is  a  complex,  dynamic  system;  there  may  be 
other  factors  influencing  the  bycatch  of  sea  turtles  in 
the  fishery  south  of  the  Hudson  Canyon  that  were  not 
observed.  However,  without  additional  data  on  turtle 
interactions  in  these  areas,  it  is  unwise  to  extrapolate 
bycatch  estimates  beyond  the  scope  of  the  data  in  this 
analysis. 


2  Kobayashi.  D.  R.,  and  J.  J.  Polovina.  2000.  Time/area 
closure  analysis  for  turtle  take  reductions.  Appendix  C, 
Environmental  Impact  Statement,  FMP  for  Pelagic  Fisher- 
ies of  the  Western  Pacific,  44  p.  NMFS  Honolulu,  Hawaii, 
96822. 


Conservation  management  options 

Time  and  area  closures  Models  of  turtle  migrations  can 
be  used  to  predict  interactions  with  fisheries  in  time 
and  space  to  maximize  the  efficiency  of  time  and  area 


Murray:  Magnitude  and  distribution  of  sea  turtle  bycatch  in  the  sea  scallop  dredge  fishery 


679 


closures  (Morreale,  1996).  The  results  of  this  analy- 
sis indicate  that  bycatch  rates  are  affected  by  season, 
depth,  and  sea  surface  temperature.  Within  certain 
months  and  depth  zones,  therefore,  the  time  when  sea 
surface  temperature  reaches  a  threshold  level  may  be 
the  time  to  trigger  an  area  closure.  For  example,  this 
type  of  management  approach  has  been  taken  in  the 
southeastern  United  States  to  regulate  turtle  bycatch 
in  the  large-mesh  gill-net  fishery.3  The  timing  of  sea- 
sonally adjusted  area  closures  is  based  upon  analyzing 
sea  surface  temperatures  in  relation  to  the  presence 
or  absence  of  sea  turtles  throughout  the  area  (Epperly 
et  al.,  1995;  Epperly  and  Braun-McNeill4).  In  addition, 
temperature  thresholds  currently  trigger  area  closures 
in  the  southern  California  driftnet  fishery  during  El 
Nino  conditions  to  prevent  the  incidental  capture  of  log- 
gerhead sea  turtles.5 

Results  from  the  present  study  can  be  used  to  help 
evaluate  potential  bycatch  reduction  under  different 
management  scenarios,  given  certain  assumptions.  For 
example,  had  the  portion  of  the  Hudson  Canyon  con- 
trolled access  area  between  depths  of  49  and  57  m  been 
closed  after  surface  waters  reached  19°C  in  the  summer 
(the  stratum  with  highest  bycatch),  the  closure  would 
have  reduced  bycatch  by  39%.  For  this  estimate,  it  is 
assumed  that  surface  temperatures  remain  above  19°C 
throughout  the  summer  and  drop  below  19°C  thereafter. 
Further,  this  bycatch  reduction  scenario  also  assumes 
that  fishing  effort  shifts  proportionately  to  the  fall  and 
spring  season  within  the  same  depth  zone  and  that 
bycatch  rates  remain  the  same  as  those  that  are  cal- 
culated. Alternatively,  fishing  effort  could  shift  within 
a  season  to  shallow  and  deep  depth  zones  if  scallop 
catch-per-unit-of-effort  were  not  affected.  Under  this 
assumption,  bycatch  would  be  reduced  by  60'"<  under 
the  same  time  and  area  closure.  However,  unless  there 
are  concurrent  reductions  in  fishing  effort,  bycatch 
reductions  achieved  by  these  measures  could  well  be 
offset  by  increases  in  bycatch  in  other  depth  strata 
and  seasons. 

Gear  or  fishing  modifications  Management  actions  to 
modify  gear  or  fishing  practices  can  be  evaluated  in  a 
similar  manner.  For  instance,  this  analysis  indicates 
that  bycatch  rates  are  influenced  by  the  time-of-day 
when  dredges  are  in  the  water.  Time-of-day  was  not  used 
to  stratify  bycatch  rates  or  to  extrapolate  total  bycatch 
estimates  because  of  limitations  in  the  fishing  effort 
data  ( VTR  records).  If  time-of-day  had  been  incorporated 
into  the  bycatch  model,  the  model  would  have  predicted 
higher  bycatch  rates  when  dredges  were  set  between  4 
am  and  4  pm  (day  tows).  If  the  stratum  with  the  highest 
bycatch  rate  (summer,  high  surface  temperatures,  and 


depths  between  49  and  57  m),  had  been  further  strati- 
fied by  time-of-day,  the  model  would  have  predicted  a 
bycatch  rate  of  0.008  sea  turtles/dredge  hauls  during  the 
day,  and  0.002  turtles/dredge  hauls  during  the  night.  If 
all  the  commercial  vessels  had  been  fishing  during  the 
day  in  this  stratum  (;?  =  6352  dredges  in  2001),  the  esti- 
mated bycatch  would  have  been  51  turtles.  If  the  vessels 
had  been  fishing  during  the  night,  the  total  estimated 
bycatch  would  have  been  13  turtles.  According  to  these 
rates  and  effort,  restricting  vessels  to  night-time  tows 
between  the  hours  of  4  pm  and  4  am  has  the  potential  to 
reduce  bycatch  by  75%  in  this  particular  stratum. 

Although  specific  gear  characteristics  did  not  show 
a  strong  relationship  to  sea  turtle  bycatch  in  this 
analysis,  further  work  should  be  conducted  to  evaluate 
whether  specific  gear  characteristics  could  be  modified 
to  decrease  bycatch.  For  example,  the  near  significance 
with  the  model  incorporating  number  of  tickler  chains 
(P=0.07)  warrants  further  testing  of  this  gear  charac- 
teristic. Tickler  chains  cover  the  mouth  of  the  dredge  in 
a  grid-like  configuration  with  the  vertical  up  and  down 
chains.  The  number  of  chains  on  the  bag  and  distance 
between  the  chains  may  help  to  prevent  sea  turtles  from 
entering  the  dredge  bag.  This  dredge  configuration  is 
currently  being  tested  for  sea  turtle  bycatch  reduction 
in  the  Hudson  Canyon  area  (DuPaul  and  Smolowitz6). 
Further  research  should  also  examine  the  behavior 
of  sea  turtles  in  relation  to  dredge  gear  for  a  more 
complete  understanding  of  how  and  when  turtles  are 
entrapped. 

Sea  turtles  and  scallop  dredge  interactions  cannot  be 
viewed  in  isolation  from  other  gear  types  and  conserva- 
tion measures.  Some  fisheries  that  co-occur  with  sea 
turtles  may  have  an  equal,  if  not  greater,  impact  on 
turtles  than  do  scallop  dredges  (e.g.,  the  shrimp  trawl 
fishery  in  the  Gulf  of  Mexico  (Henwood  and  Stuntz, 
1987]).  Changes  in  sea  turtle  abundance,  or  shifts  in 
fishing  effort,  may  increase  the  likelihood  of  encounters 
in  both  net  and  dredge  fisheries.  If  environmental  condi- 
tions associated  with  high  bycatch  rates  in  the  Hudson 
Canyon  and  Virginia  Beach  areas  are  consistent  across 
years,  it  may  be  possible  to  anticipate  and  deter  future 
interactions  from  occurring. 


Acknowledgments 

I  would  like  to  thank  Debra  Palka  and  Marjorie  Ross- 
man  for  help  with  analytical  and  statistical  approaches 
to  bycatch  estimation.  Andy  Solow  and  Andy  Beet  at  the 
Marine  Policy  Center,  Woods  Hole  Oceanographic  Insti- 
tute, also  provided  guidance  in  the  statistical  analysis. 
David  Mountain  provided  invaluable  help  in  acquiring 


3  Final  Rule.  FR  67:  71895-71900.  3  December  2002. 

4  Epperly,  S.  P.  and  J.  Braun-McNeill.  2002.  Unpubl.  data. 
The  use  of  AVHRR  Imagery  and  the  management  of  sea  turtle 
interactions  in  the  Mid-Atlantic  Bight.  NMFS  Southeast 
Fisheries  Science  Center,  Miami,  Florida,  33149. 

5  Final  Rule,  FR  68:  69962-69967,  16  December  2003. 


6  DuPaul.  W.  P.,  and  R.  Smolowitz.  2003.  Unpubl.  data. 
Industry  trials  of  a  modified  sea  scallop  dredge  to  minimize 
the  catch  of  sea  turtles.  Virginia  Institute  of  Marine  Sci- 
ence, Gloucester  Point,  Virginia,  23062,  and  Coonamessett 
Farm,  East  Falmouth,  Massachusetts,  02536. 


680 


Fishery  Bulletin  102(4) 


sea  surface  temperature  for 
data.  Frederic  Serchuk,  Richa 
Marjorie  Rossman,  and  Pau 
Fisheries  Science  Center  all 
of  the  manuscript.  Jeffrey  S 
mous  reviewers  provided  val 
peer  review.  Finally,  I  wish  to 
collected  data  on  interactions 
sea  scallop  dredge  fishery. 


the  Observer  and  VTR 
rd  Merrick,  Debra  Palka, 
Rago  at  the  Northeast 
provided  initial  reviews 
eminoff  and  two  anony- 
uable  comments  during 
thank  the  observers  who 
between  turtles  and  the 


Literature  cited 

Allen,  L.  K. 

2000.     Protected  species  and  New  England  fisheries:  an 
overview  of  the  problem  and  conservation  strategies. 
Northeastern  Naturalist  7(41:411-418. 
Burnham,  K.  P.  and  D.  R.  Anderson. 

2002.     Model  selection  and  multimodel  inference:  a  prac- 
tical information-theoretic  approach,  2nd  ed.,  488  p. 
Springer-Verlag,  New  York,  NY. 
Cochran,  W.G. 

1977.     Sampling  techniques,  3rd  ed.,  428  p.     John  Wiley 
&  Sons,  New  York,  NY. 
Coles,  W.  C.  and  J.  A.  Musick. 

2000.     Satellite  sea  surface  temperature  analysis  and 
correlation   with   sea   turtle   distribution   off  North 
Carolina.     Copeia  2:551-554. 
Crouse,  D.  T.,  L.  B.  Crowder,  and  H.  Caswell. 

1987.     A  stage-based  population  model  for  loggerhead 
sea  turtles  and  implications  for  conservation.     Ecology 
68(51:1412-1423. 
Denis.  V.,  J.  Lejeune,  and  J.  P.  Robin. 

2002.  Spatio-temporal  analysis  of  commercial  trawler 
data  using  General  Additive  models:  patterns  of  Loliginid 
(sic)  squid  abundance  in  the  north-east  Atlantic.  ICES 
J.  Mar.  Sci.,  59:633-648. 

Efron,  B.  and  R.  J.  Tibshirani. 

1993.     An  introduction  to  the  boostrap,  436  p.     Chapman 
&  Hall,  New  York,  NY. 
Epperly,  S.  P. 

2003.  Fisheries-related  mortality  and  turtle  excluder 
devices  (TEDs).  In  The  biology  of  sea  turtles,  vol.  II 
(P.  L.  Lutz,  J.  A.  Musick.  and  J.  Wyneken,  eds.),  p.  339- 
353.     CRC  Press,  Boca  Raton,  FL. 

Epperly,  S.  P.,  J.  Braun,  A.  J.  Chester,  F.A.  Cross, 
J.  V.  Merriner,  and  P.  A.  Tester. 

1995.     Winter  distribution  of  sea  turtles  in  the  vicin- 
ity of  Cape  Hatteras  and  their  interactions  with  the 
summer  flounder  trawl  fishery.     Bull.  Mar.  Sci.  56(2): 
547-568. 
Frost,  K.  J..  L.  F.  Lowry,  and  J.  M.  Ver  Hoef. 

1999.     Monitoring  the  trend  of  harbor  seals  in  Prince 
William  Sound,  Alaska,  after  the  Exxon  Valdez  oil 
spill.     Mar.  Mamm.  Sci.  15(21:494-506. 
Guisan,  A.,  T.  C.  Edwards,  and  T.  Hastie. 

2002.     Generalized   linear   and  generalized   additive 
models  in  studies  of  species  distributions:  setting  the 
scene.     Ecol.  Model.  157:89-100. 
Hamazaki,  T. 

2002.  Spatiotemporal  prediction  models  of  cetacean 
habitats  in  the  mid-western  North  Atlantic  ocean  (from 
Cape  Hatteras,  North  Carolina,  USA  to  Nova  Scotia. 
Canada).     Mar.  Mamm.  Sci.  18<4):134-151. 


Hastie,  T  J.,  and  R.  J.  Tibshirani. 

1990.  Generalized  additive  models,  329  p.  Chapman 
and  Hall,  London. 

Henwood,  T.  A„  and  W.  E.  Stuntz. 

1987.     Analysis  of  sea  turtle  captures  and  mortalities 
during  commercial  shrimp  trawling.     Fish.  Bull.  85: 
813-817. 
Heppell.  S.  S.,  L.  B.  Crowder,  D.  T.  Crouse.  S.  P.  Epperly,  and 
N.  B.  Frazer. 

2003.  Population  models  for  Atlantic  loggerheads.  In 
Loggerhead  sea  turtles  (A.  Bolten  and  B.  Witherington. 
eds.),  pp.  255-273.  Smithsonian  Books,  Washington, 
D.C.. 

Julian.  F.,  and  M.  Beeson. 

1998.  Estimates  of  marine  mammal,  turtle,  and  seabird 
mortality  for  two  California  gillnet  fisheries:  1990- 
1995.     Fish.  Bull.  96:271-284. 

Lewison,  R.  L.,  S.  A.  Freeman,  and  L.  B.  Crowder. 

2004.  Quantifying  the  effects  of  fisheries  on  threatened 
species:  the  impact  of  pelagic  longlines  on  loggerhead  and 
leatherback  sea  turtles.     Ecology  Letters  7:221-231. 

Morreale,  S.  J.,  E.  A.  Standora,  J.  R.  Spotila,  and  F.  V.Paladino. 
1996.     Migration  corridor  for  sea  turtles.     Nature  384: 
319-320. 
NMFS  (National  Marine  Fisheries  Service  I. 

2003.     Fisheries  observer  program  manual,  217  p.     North- 
east Fisheries  Observer  Program,  Northeast  Fisheries 
Science  Center,  Woods  Hole.  MA. 
NMFS  and  USFWS  (National  Marine  Fisheries  Service  and 
U.S.  Fish  and  Wildlife  Service). 

1991.  Recovery  plan  for  U.S.  population  of  loggerhead 
turtle,  64  p.  National  Marine  Fisheries  Service,  Wash- 
ington, D.C. 

NCR  (National  Research  Council). 

1990.     Decline  of  the  sea  turtles:  causes  and  prevention, 
259  p.     National  Academy  Press.  Washington,  D.C. 
Palka.  D.  L..  and  M.  C.  Rossman. 

2001.     Bycatch  estimates  of  coastal  bottlenose  dolphin 
(Tursiops  truncatus)  in  U.S.  Mid-Atlantic  gillnet  fisher- 
ies for  1996-2000.     Northeast  Fisheries  Science  Center 
Reference  Document  01-15,  77  p. 
Polovina.  J.  J.,  D.  R.  Kobayashi,  D.  M.  Parker.  M.  P.  Seki,  and 
G.  H.  Balazs. 

2000.     Turtles  on  the  edge:  movement  of  loggerhead  tur- 
tles (Caretta  caretta)  along  oceanic  fronts,  spanning 
longline  fishing  grounds  in  the  central  North  Pacific, 
1997-1998.     Fish.  Oceanogr.  9:71-82. 
Rossman,  M.  C.  and  R.  L.  Merrick. 

1999.  Harbor  porpoise  bycatch  in  the  Northeast  mul- 
tispecies  sink  gillnet  fishery  and  the  Mid-Atlantic 
coastal  gillnet  fishery  in  1998  and  during  January-May 
1999.     Northeast  Fish.  Sci.  Cent.  Ref.  Doc.  99-17.  36  p. 

Shoop,  R.  C.  and  R.  D.  Kenney. 

1992.  Seasonal  distributions  and  abundances  of  logger- 
head and  leatherback  sea  turtles  in  waters  of  the  north- 
eastern United  States.     Herpetol.  Monogr.  6:43-67. 

Smolowitz,  R. 

1998.  Bottom  tending  gear  used  in  New  England.  In 
Effects  of  fishing  gear  on  the  sea  floor  of  New  England 
(E.  Dorsey  and  J.  Pederson.  eds.),  p.  46-52.  Conserva- 
tion Law  Foundation,  Boston,  MA. 

Turtle  Expert  Working  Group. 

2000.  Assessment  update  for  the  Kemp's  ridley  and  log- 
gerhead sea  turtle  populations  in  the  western  North 
Atlantic.  U.S.  Dep.  Commer.  NOAA  Tech.  Memo. 
NMFS-SEFSC-444,  115  p. 


Murray:  Magnitude  and  distribution  of  sea  turtle  bycatch  in  the  sea  scallop  dredge  fishery 


681 


Wade,  P.  R.,  and  R.  P.  Angliss. 

1997.  Guidelines  for  assessing  marine  mammal  stocks: 
report  of  the  GAMMS  workshop  April  3-5,  1996,  Seattle, 
Washington.  U.S.  Dep.  Commer.,  NOAA  Tech.  Memo. 
NMFS-OPR-12.  93  p. 


Witzell,  W.  N. 

1999.  Distribution  and  relative  abundance  of  sea  turtles 
caught  incidentally  by  the  U.S.  pelagic  longline  fleet  in 
the  western  North  Atlantic  Ocean,  1992-1995.  Fish. 
Bull.  97:200-211. 


Appendix  1 

Categorical  variables 

examined  in  an  analysis  of  factors  affecting  sea  turtle  bycatch  in  the  sea  scallop  dredge  fishery.  Frequency 

of  observed  dredges  ir 

each  category  is  also  shown. 

Number  of 

Number  of 

Number  of 

Number  of 

observed 

observed 

observed 

observed 

dredges  in 

dredges  in 

dredges  in 

dredges  in 

Hudson 

Virginia 

Hudson 

Virginia 

Variable 

Category 

Canyon  2001 

Beach  2001 

Canyon  2002 

Beach  2002 

Year 

2001  or  2002 

9493 

520 

8059 

0 

Season 

Spring  =  May  and  June 

3919 

0 

1987 

0 

Summer  =  July,  August,  September 

2719 

0 

3764 

0 

Fall  =  October,  November,  December 

2855 

520 

2308 

0 

State  in  which 

Connecticut 

199 

0 

595 

0 

scallops  were  landed 

Massachusetts 

4925 

0 

5628 

0 

New  Jersey 

2849 

0 

740 

0 

Rhode  Island 

112 

0 

474 

0 

Virginia 

1408 

520 

622 

0 

Frame  width' 

Small  =  3.0-3.9  m  ( 10-13  ft) 

560 

0 

443 

0 

category 

Medium  =  >3.9  m  and  <4.5  m  (15  ft) 

3987 

122 

3013 

0 

Large  =  <4.5  m-4.8  m  ( 15-16  ft ) 

4946 

398 

4603 

0 

Number  of  up  and 

Code  1  =  0  chains 

4256 

520 

2171 

0 

down  chains  used- 

Code  2  =  1-4  chains 

4089 

0 

5378 

0 

Code  3  =  >4  chains 

1148 

0 

510 

0 

Number  of  tickler 

Code  1  =  <2  chains 

6890 

520 

4469 

0 

chains  used3 

Code  2  =  >2  chains 

2603 

0 

3590 

0 

Time-of-day 

Day  =  4  am-4  pm 

5514 

346 

4854 

0 

Night  =  4  pm-4  am 

3979 

174 

3205 

0 

Sea  surface 

Hi  =  >19'C 

3910 

518 

4883 

0 

temperature 

Low  =  <19°C 

5583 

2 

3176 

0 

Depth 

Shallow  =  40-<49  m  (22-27  fmi 

1089 

42 

782 

0 

Mid-Depth  =  49-57  m  (27-31  fin) 

3371 

280 

3642 

0 

Deep  =  >57-88  m  (31-48  fm) 

5033 

198 

3635 

0 

;  Width  of  the  dredge  fra 

me. 

2  Vertical  chains  attache 

i  to  the  sweep  on  the  bottom  of  the  dredge  that 

prevent  rocks  from  entering  the  chain 

bag.  Number  of  up 

and  down  chains 

were  influenced  by  bottom  type. 

3  Horizontal  chains  attac 

hed  to  the  sweep  on  the  bottom  of  the  dredge  that  help  stir  up  contents  of  the  sea  bottom.  Number  of  ti< 

kler  chains  were 

influenced  by  bottom  t\ 

pe. 

682 


Abstract  —  Numerous  studies  have 
applied  skeletochronology  to  sea  turtle 
species.  Because  many  of  the  studies 
have  lacked  validation,  the  applica- 
tion of  this  technique  to  sea  turtle 
age  estimation  has  been  called  into 
question.  To  address  this  concern,  we 
obtained  humeri  from  13  known-age 
Kemp's  ridley  (Lepidochelys  kempii) 
and  two  loggerhead  (Caretta  caretta) 
sea  turtles  for  the  purposes  of  examin- 
ing the  growth  marks  and  comparing 
growth  mark  counts  to  actual  age.  We 
found  evidence  for  annual  deposition 
of  growth  marks  in  both  these  spe- 
cies. Corroborative  results  were  found 
in  Kemp's  ridley  sea  turtles  from  a 
comparison  of  death  date  and  amount 
of  bone  growth  following  the  comple- 
tion of  the  last  growth  mark  (n=76). 
Formation  of  the  lines  of  arrested 
growth  in  Kemp's  ridley  sea  turtles 
consistently  occurred  in  the  spring  for 
animals  that  strand  dead  along  the 
mid-  and  south  U.S.  Atlantic  coast. 
For  both  Kemp's  ridley  and  loggerhead 
sea  turtles,  we  also  found  a  propor- 
tional allometry  between  bone  growth 
•  humerus  dimensions)  and  somatic 
growth  (straight  carapace  length  i, 
indicating  that  size-at-age  and  growth 
rates  can  be  estimated  from  dimen- 
sions of  early  growth  marks.  These 
results  validate  skeletochronology  as 
a  method  for  estimating  age  in  Kemp's 
ridley  and  loggerhead  sea  turtles  from 
the  southeast  United  States. 


Validation  and  interpretation  of  annual 
skeletal  marks  in  loggerhead  (Caretta  caretta) 
and  Kemp's  ridley  (Lepidochelys  kempii)  sea  turtles 


Melissa  L.  Snover 

Duke  University  Marine  Laboratory 

135  Duke  Marine  Lab  Road 

Beaufort,  North  Carolina  28516 

Present  addresss:  Pacific  Fisheries  Environmental  Laboratory 

1352  Lighthouse  Ave. 

Pacific  Grove,  California  93950 
E-mail  address:  melissa  snover  g  noaa  gov 

Aleta  A.  Hohn 

Center  for  Coastal  Fisheries  and  Habitat  Research 
National  Marine  Fisheries  Service,  NOAA 
101  Pivers  Island  Road 
Beaufort,  North  Carolina  28516 


Manuscript  submitted  1 5  August  200,3 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
ii  June  2004  by  the  Scientific  Editor 

Fish.  Bull.  102:682-692  (2004  I. 


The  basic  tenet  of  skeletochronology  is 
that  bone  growth  is  cyclic  and  has  an 
annual  periodicity  in  which  bone  for- 
mation ceases  or  slows  before  new,  rel- 
atively rapid  bone  formation  resumes 
(Simmons,  1992;  Castanet  et  al„  1993; 
Klevezal,  1996).  This  interruption  of 
bone  formation  is  evidenced  within 
the  primary  periosteal  compacta  by 
histological  features,  which  take  two 
forms  in  decalcified  and  stained  thin- 
sections.  The  most  common  form  is  a 
thin  line  that  appears  darker  than  the 
surrounding  tissue,  termed  the  "line 
of  arrested  growth"  (LAG)  (Castanet 
et  al.,  1977).  The  second,  less-common 
form  is  a  broader  and  less  distinct  line 
that  also  stains  darker,  referred  to  as 
an  annulus  (Castanet  et  al.,  1977). 
Alternating  with  LAGs  or  annuli  are 
broad  zones  that  stain  homogeneously 
light,  and  represent  areas  of  active 
bone  formation.  Together,  a  broad 
zone  followed  by  either  a  LAG  or  an 
annulus  represents  a  skeletal  growth 
mark  (GM)  (Castanet  et  al.,  1993).  To 
apply  skeletochronology  to  a  species, 
the  annual  periodicity  of  the  GM  must 
be  validated. 

Validation  studies  are  necessary 
not  only  to  confirm  the  annual  nature1 
of  the  GM  but  also  to  identify  and  in- 
terpret anomalous  LAGs.  Anomalous 
LAGs  that  are  a  common  problem  in 
skeletochronology  studies  of  reptiles 


and  amphibians  include  double  (Chin- 
samy  et  al.,  1995;  El  Mouden  et  al., 
1997;  Guarino  et  al.,  1998),  splitting 
(Guarino  et  al.,  1995;  1998;  Coles  et 
al.,  2001),  and  supplemental  (Guarino 
et  al,  1995;  Lima  et  al.,  2000;  Tren- 
ham  et  al.,  2000)  lines.  In  addition  to 
anomalous  LAGs,  there  are  two  other 
difficulties  typical  in  skeletochronol- 
ogy studies;  compression  of  LAGs  at 
the  periphery  of  the  bone  and  resorp- 
tion of  the  innermost  LAGs.  In  older 
animals  the  GMs  are  compressed  at 
the  outer  periphery  of  the  bone  as  a 
result  of  decreased  growth.  Francil- 
lon-Vieillot  et  al.  (1990)  term  this 
phenomenon  "rapprochement"  and  it 
is  a  problem  when  the  LAGs  become 
too  close  together  to  be  differentiated 
— usually  in  the  small  phalangeal 
bones  used  in  amphibian  studies  lEg- 
gert  and  Guyetant.  1999;  Lima  et  al., 
2000;  Leclair  et  al..  2000). 

In  addition  to  anomalous  and  com- 
pressed LAGs,  the  loss  of  early  GMs 
through  endosteal  resorption  is  an- 
other problem  with  skeletochronol- 
ogy. Although  this  does  not  present  a 
problem  with  most  amphibian  species 
(Kusano  et  al.,  1995;  Castanet  et  al., 
1996;  Sagor  et  al.,  1998),  the  prob- 
lem is  extreme  in  skeletochronology 
studies  of  loggerhead  (Caretta  caretta; 
Klinger  and  Musick.  1995;  Zug  et  al., 
1995;  Parham  and  Zug,  1997),  green 


Snover  and  Hohn:  Validation  and  interpretation  of  skeletal  marks  in  Caretta  caretta  and  Lepidochelys  kempu 


683 


Table  1 

Species  and  history  of  known 

-age  sea  turtles  analyzed  in  this  study. 

Sample 

Species 

History  during  captivity 

Age  (yr) 

LK-1 

Lepidochelys  kempu 

Captive  for  first  year,  then  released 

5.0 

LK-2 

L.  kempii 

Captive  for  first  year,  then  released 

6.5 

LK-3 

L.  kempii 

Captive  for  first  year,  then  released 

4.5 

LK-4 

L.  kempii 

Tagged  and  released  after  hatching 

1.27 

LK-5 

L.  kempii 

Tagged  and  released  after  hatching 

1.70 

LK-6 

L.  kempii 

Tagged  and  released  after  hatching 

1.72 

LK-7 

L.  kempii 

Tagged  and  released  after  hatching 

2.37 

LK-8 

L.  kempii 

Tagged  and  released  after  hatching 

2.37 

LK-9 

L.  kempii 

Tagged  and  released  after  hatching 

3.25 

LK-10 

L.  kempii 

Tagged  and  released  after  hatching 

2.0 

LK-11 

L.  kempii 

Tagged  and  released  after  hatching 

2.75 

LK-12 

L.  kempii 

Tagged  and  released  after  hatching 

3.0 

LK-13 

L.  kempii 

Tagged  and  released  after  hatching 

4.25 

CC-1 

Caretta  caretta 

Captive  during  entire  life 

29.4 

CC-2 

C.  caretta 

Captive  for  first  two  years,  then  released 

8.0 

(Chelonia  mydas;  Zug  and  Glor,  1998;  Zug  et  al.,  2002) 
and  Kemp's  ridley  (Lepidochelys  kempii;  Zug  et  al., 
1997)  sea  turtles.  In  each  of  these  studies,  the  authors 
used  various  protocols  to  estimate  the  number  of  lay- 
ers lost.  Any  protocol  estimating  the  number  of  layers 
lost  to  resorption  relies  on  the  concept  that  the  spatial 
pattern  of  the  LAGs  is  representative  of  the  growth  of 
the  animal.  To  confirm  this  assumption,  researchers 
must  establish  a  correlation  between  bone  dimensions 
and  body  size  (Hutton,  1986;  Klinger  and  Musick,  1992; 
Leclair  and  Laurin,  1996). 

Two  of  the  studies  that  have  applied  skeletochronology 
to  sea  turtles  have  demonstrated  annual  GMs  in  both 
juvenile  (Klinger  and  Musick,  1992)  and  adult  (Coles  et 
al.,  2001)  loggerhead  sea  turtles.  Numerous  additional 
studies  have  applied  skeletochronology  to  sea  turtles. 
To  date,  the  technique  has  been  applied  to  loggerhead 
(Zug  et  al.,  1986;  Zug  et  al,  1995;  Bjorndal  et  al.,  2003), 
green  (Bjorndal  et  al.,  1998;  Zug  and  Glor,  1998;  Zug 
et  al.,  2002),  Kemp's  ridleys  (Zug  et  al.,  1997),  and 
leatherback  [Dermochelys  coriacea)  (Zug  and  Parham, 
1996)  sea  turtles.  What  is  needed  for  the  appropriate 
application  of  skeletochronology  to  sea  turtle  species 
is  additional  validation  of  annual  GMs  and  a  guide  to 
their  interpretation.  Furthermore,  because  resorption  is 
a  problem  in  sea  turtle  bones,  the  validation  of  a  pro- 
portional allometry  between  bone  and  somatic  growth 
is  necessary  to  enable  back-calculation. 

In  this  study,  we  address  each  of  these  issues  for 
Kemp's  ridley  and  loggerhead  sea  turtles  by  examining 
humeri  from  known-age  animals.  We  analyzed  each 
humerus  without  prior  knowledge  of  the  animal's  age 
and  we  present  the  results  of  our  analyses,  including 
reinterpretations  of  bones  for  which  we  were  incorrect 
in  our  age  assessments.  The  purpose  of  this  study  was 


to  use  known-age  samples  both  to  validate  the  likeli- 
hood that  GMs  are  annual  and  as  a  learning  tool  for  the 
best  guide  to  interpreting  GM  in  wild  animals. 


Materials  and  methods 

We  obtained  samples  from  two  known-age  loggerhead 
and  13  known-age  Kemp's  ridley  sea  turtles  (Table  1). 
In  addition,  we  collected  samples  from  240  wild  logger- 
head and  262  wild  Kemp's  ridley  sea  turtles.  With  the 
exception  of  one  loggerhead,  CC-1,  all  of  the  sea  turtles 
died  in  the  wild  and  samples  were  retrieved  from  the 
carcasses.  Sample  CC-1  died  in  captivity. 

Sample  preparation 

Zug  et  al.  (1986)  analyzed  skeletal  elements  of  the  cra- 
nium and  right  forelimb  of  loggerhead  sea  turtles  and 
determined  that  the  humerus  was  most  suited  to  skeleto- 
chronology studies.  Therefore,  we  also  used  the  humerus. 
Specimens  arrived  as  either  dried  bones  or  whole  flippers. 
For  flippers,  we  dissected  out  the  humerus,  which  v/as 
then  flensed,  boiled,  and  air-dried  for  at  least  two  weeks. 
We  cross-sectioned  each  humerus  at  a  site  just  distal  to 
the  deltopectoral  crest.  At  this  site,  the  ratio  of  cortical 
to  cancellous  bone  is  highest  (Zug  et  al.,  1986),  and  the 
region  immediately  distal  to  the  insertion  scar  of  the 
deltopectoral  muscle  on  the  ventral  side  of  the  bone  maxi- 
mizes that  ratio  (see  Zug  et  al.,  1986  for  diagrams  of  the 
loggerhead  sea  turtle  humerus).  This  site  also  provided 
a  landmark  that  allowed  us  to  section  at  equivalent  sites 
on  every  humerus. 

We  removed  2-3  mm  thick  sections  at  that  site  us- 
ing a  Buehler®  isomet  low  speed  saw.  This  section  was 


684 


Fishery  Bulletin  102(4) 


fixed  in  109c  formalin  then  decalcified  by  using  a  com- 
mercial decalcifying  agent  (RDO,  Apex  Engineering 
Products  Corporation,  Calvert  City,  Kentucky).  Time 
to  decalcification  varied  with  the  size  of  the  bone  and 
the  strength  of  the  solution,  usually  between  12  and  36 
hours.  Following  decalcification,  25-f.im  thick  cross-sec- 
tions were  made  by  using  a  freezing-stage  microtome. 
Sections  were  stained  in  Erlich's  hematoxylin  diluted 
1:1  with  distilled  water  (Klevezal,  1996)  and  mounted 
on  slides  in  100%  glycerin. 

Known-age  sea  turtles 

We  received  the  humeri  from  each  of  two  captive,  known- 
age  loggerhead  sea  turtles  after  they  died  (Table  1).  The 
first  specimen,  CC-1,  was  held  in  an  outdoor  tank  during 
the  summer  months  and  inside  a  greenhouse  during  the 
winter  months  (this  turtle  was  the  same  captive  female 
noted  in  Swartz,  1997).  The  second,  CC-2,  was  raised  in 
captivity  for  two  years  then  released  from  Panama  City, 
Florida,  into  the  Gulf  of  Mexico. 

For  the  Kemp's  ridley  sea  turtles,  we  received  humeri 
from  13  dead  known-age  animals  (Table  1).  The  head- 
start  Kemp's  ridleys  were  raised  in  captivity  for  one 
year,  then  released  as  part  of  a  binational  program  oper- 
ated jointly  by  state  and  federal  U.S.  agencies  and  the 
Instituto  Nacional  de  la  Pesca  (INP)  of  Mexico  (Klima 
and  McVey,  1995).  The  coded-wire-tagged  (CWT)  Kemp's 
ridley  sea  turtles  were  tagged  and  released  as  hatch- 
lings.  This  tagging  program  is  operated  jointly  by  the 
U.S.  National  Marine  Fisheries  Service  (NMFS)  Galves- 
ton Laboratory  and  the  INP  of  Mexico  as  a  means  of 
gaining  a  better  understanding  of  the  early  life  history 
of  the  Kemp's  ridley  sea  turtle  (Caillouet  et  al.,  1997). 

Using  the  methods  described  previously,  we  prepared 
stained  thin-sections  from  the  humeri.  Without  prior 
knowledge  of  the  animal's  history,  the  number  of  visible 
LAGs  was  quantified  for  each  bone  and  a  minimum  age 
estimated.  Our  age  estimates  were  then  compared  to 
the  age  information  available  for  each  animal. 

Indirect  validation  of  annual  growth  marks 

Peabody  (1961)  and  Castanet  et  al.  (1993)  suggested  that 
the  correlation  between  the  width  of  the  last  zone  formed 
and  the  date  of  death  provided  an  indirect  means  of  vali- 
dating that  deposition  of  the  LAG  occurs  annually  and  at 
the  same  time  of  year  for  an  individual  population.  We 
applied  this  method  to  76  wild  Kemp's  ridley  sea  turtles 
for  which  humeri  displayed  between  one  and  five  LAGs. 
Each  of  these  animals  had  stranded  dead  along  the 
Atlantic  coast  between  Maryland  and  North  Carolina. 
Thin-sections  were  prepared  of  the  humeri  as  described 
above.  We  quantified  the  width  of  the  last  zone  formed 
by  measuring  the  outside  diameter  of  the  whole  section 
(D0)  and  the  diameter  of  the  last  competed  LAG  (DL), 
between  the  lateral  edges  of  the  bone  on  an  axis  paral- 
lel to  the  dorsal  edge.  The  amount  of  bone  growth  after 
the  last  LAG  (D0-DL)  was  plotted  against  the  Julian 
stranding  date,  with  the  assumption  that  stranding 


date  approximated  date  of  death.  Least-squares  linear 
regressions  were  fitted  to  the  data. 

Validation  of  the  relationship  between 
LAG  diameter  and  body  size 

In  order  to  relate  GM  diameters  to  somatic  growth  rates, 
there  must  be  a  constant  proportionality  between  bone 
growth  and  somatic  growth  (Chaloupka  and  Musick, 
1997).  To  address  this  proportionality,  we  took  eight 
morphometric  measurements  of  240  wild  loggerhead  and 
262  wild  Kemp's  ridley  humeri,  using  digital  calipers 
or  a  tape  measure  when  dimensions  were  beyond  the 
range  of  the  calipers.  Measurements  of  maximum  length, 
longitudinal  length,  proximal  width,  distal  width,  delto- 
pectoral  crest  width,  lateral  diameter  at  sectioning  site, 
ventral  to  dorsal  thickness  at  sectioning  site,  and  mass 
were  recorded.  We  compared  these  measurements  with 
the  carapace  length,  measured  as  standard  straight-line 
length  (SCL)  from  the  nuchal  notch  to  the  posterior  end 
of  the  posterior  marginal,  using  a  least-squares  linear 
regression.  For  mass,  the  data  were  natural-log  trans- 
formed to  form  a  linear  regression. 


Results 

Known-age  Kemp's  ridley  sea  turtles 

Three  Kemp's  ridley  sea  turtles  captive  for  one  year 
and  then  released  were  recovered  4.5  to  6.5  years  after 
hatching  (Table  1).  Sample  LK-1  had  minimal  resorp- 
tion and  four  complete  GMs,  each  comprising  one  zone 
followed  by  a  LAG.  An  additional  zone  was  seen  at  the 
periphery  and  the  LAG  that  would  complete  this  last 
GM  was  not  yet  visible  at  the  outer  edge  of  the  humerus 
cross-section.  From  GM  counts  and  death  date,  we  esti- 
mated the  age  of  this  animal  accurately  at  five  years 
(Fig.  1).  Sample  LK-2  retained  five  completed  and  one 
incomplete  GM;  however,  we  observed  a  large  area  of 
resorption  in  the  interior  region  of  the  cross-section 
that  potentially  obscured  additional  GMs.  We  aged  this 
animal  at  a  minimum  of  5.5  years,  the  actual  age  being 
6.5  years.  Sample  LK-3  displayed  four  completed  GMs 
and  one  incomplete  mark.  Without  prior  knowledge  of 
this  animal's  age,  we  estimated  the  age  accurately  at 
4.5  years  based  on  layer  count  and  time  of  death. 

Ten  of  the  Kemp's  ridley  sea  turtle  samples  were 
tagged  and  released  after  hatching,  and  no  time  was 
spent  in  captivity  (Table  1).  Results  from  these  ten  re- 
covered animals  allowed  us  the  opportunity  to  study 
and  interpret  the  early  GM  patterns  in  noncaptive  ani- 
mals. The  first  year  mark  for  Kemp's  ridley  sea  turtles 
appeared  to  be  a  poorly  defined  annulus,  as  evidenced 
by  LK-4  (Fig.  2A).  In  turtles  greater  than  two  years 
old,  similar  first  year  marks  also  appeared  more  or  less 
distinctly  (Figs.  2B  and  3).  Additional  marks,  which  can 
only  be  interpreted  as  supplemental  lines  given  the  age 
of  the  animal,  appeared  between  GM  one  and  the  outer 
edge  of  the  bone  in  LK-6  (Fig.  2B)  and  LK-10.  Specimens 


Snover  and  Hohn:  Validation  and  interpretation  of  skeletal  marks  in  Caretta  caretta  and  Lepidochelys  kempu 


685 


LK-7  and  LK-8  were  difficult  to  inter- 
pret and  in  our  initial  assessment  we 
underestimated  age  by  one  year.  In 
both  of  these  samples,  the  LAG  rep- 
resenting the  end  of  the  second  GM 
was  very  close  to  the  outer  edge  of  the 
bone  cross-section  and  was  difficult 
to  differentiate  from  the  edge.  Hence 
these  samples  were  not  counted  in  the 
initial  assessment.  Because  both  of 
these  animals  died  in  the  fall,  there 
would  have  been  a  full  growing  sea- 
son, and  hence  a  growth  zone,  follow- 
ing the  completion  of  the  second  GM. 
Both  of  these  animals  were  recovered 
dead  in  Cape  Cod,  Massachusetts, 
during  the  fall  of  1999  when  record 
numbers  of  cold-stunned  sea  turtles 
stranded  in  that  region. 

Humerus  cross-sections  from  LK-9 
through  LK-13  (Fig.  3)  showed  poorly 
defined  annuli  at  the  end  of  the  first 
GM  —  annuli  similar  to  the  poorly 
defined  annulus  in  LK-4  (Fig.2A). 
Subsequent  GMs  in  these  humerus 
cross-sections  contained  well-defined 
LAGs.  Without  prior  knowledge  of 
these  animals'  history  we  accurately 
aged  each  of  them  from  GM  counts 
and  stranding  date.  Specimens  LK-9 
through  LK-13  demonstrated  clearly 
that  well-defined  LAGs  were  depos- 
ited at  the  end  of  year  two  and  in 
subsequent  years,  providing  evidence 
that  any  lines  between  the  year-one 
annulus  and  the  year-two  LAGs  were 
supplemental. 

Known-age  loggerhead  sea  turtles 

The  first  known-age  loggerhead  sea 
turtle,  CC-1,  was  29.4  years  old. 
Eleven  LAGs  were  discernible  around 
the  circumference  of  the  bone  cross- 
section  (Fig.  4A),  although  the  LAGs 
become  too  compressed  on  the  lateral 
edges  of  the  bone  to  be  differentiated; 
hence  counts  were  made  on  the  ven- 
tral and  dorsal  edges.! Fig.  4).  Trac- 
ing the  LAGs  from  the  lateral  to  the 
ventral  edge  of  the  bone,  we  observed 
that  these  LAGs  at  some  point  became 
bifurcating  and  splitting  LAGs  and  we 
interpreted  each  branch  as  a  separate 
LAG.  An  additional  nine  LAGs  can 
still  be  seen  within  the  resorption  zone 
in  most  areas  of  the  bone  (Fig.  4B).  On 
the  dorsal  side  of  the  cross-section,  at 
least  four  less-distinct  LAGs  or  annuli 
could  still  be  observed;  these  had  been 


LAG-4  CAG"-3- 


E 
GM-' 


Figure  1 

Image  of  a  humerus  cross-section  from  a  headstart  Kemp's  ridley  {Lepido- 
chelys kempii,  LK-1)  sea  turtle.  GM-1  refers  to  growth  mark  one;  LAG-2, 
LAG-3,  and  LAG-4  refer  to  the  lines  of  arrested  growth  ending  growth  marks 
two,  three,  and  four.  Curved  dashed  lines  highlight  GM-1  and  the  LAG. 
Black  bar  represents  1  mm  in  length.  This  specimen  was  5.0  years  old. 


Annulus  ending 
GM-1 


B 


/? 

Supplementa 


lines 


Annulus  ending 
GM-1 


Figure  2 

Images  of  humeri  cross-sections  of  two  coded-wire-tagged  Kemp's  ridley 
sea  turtles  (L.  kempii).  GM-1  refers  to  growth  mark  one.  Black  bar  repre- 
sents 1  mm  for  both  images.  (A)  Specimen  LK-4  was  1.27  years  old.  (B) 
Specimen  LK-6  was  1.72  years  old. 


686 


Fishery  Bulletin  102(4) 


Annulus  ending 
GM-1 


3  "t  4 

LAG-2  '        \ 
LAG-3LAG-4 


Figure  3 

Image  of  humerus  cross-section  from  a  coded-wire-tagged  Kemp's  ridleys  (L. 
kempii).  Black  bar  represent  1  mm  in  length.  GM-1  refers  to  growth  mark 
one;  LAG-2,  LAG-3,  and  LAG-4  refer  to  the  lines  of  arrested  growth  ending 
growth  marks  two,  three,  and  four.  Curved  black  lines  highlight  LAGs  or 
annuli.  This  specimen,  LK-13,  was  4.75  years  old. 


resorbed  in  all  other  parts  of  the  bone  (Fig.  4C).  There 
had  been  a  great  deal  of  remodeling  within  the  bone  and 
much  of  the  inner  portion  of  the  bone  had  been  resorbed. 
Summing  all  of  these  GMs,  we  gave  a  minimum  age  esti- 
mate of  24  years  without  prior  knowledge  of  the  history 
of  the  animal.  The  outermost  20  GMs  contained  well- 
defined  LAGs  that  were  spaced  close  together,  whereas 
the  four  interior-most  visible  GMs  contained  LAGs  or 
annuli  that  were  spaced  farther  apart  (Fig.  4).  The 
number  of  layers  completely  resorbed  was  five. 

A  second  known-age  loggerhead  sea  turtle,  CC-2,  was 
eight  years  old.  We  assigned  a  minimum  age  estimate 
of  five  years.  Just  outside  of  the  resorption  area  was 
a  series  of  three  LAGs  that  were  very  close  together 
(Fig.  5).  In  our  initial  analysis,  we  assumed  that  three 
LAGs  so  close  together  could  not  each  be  deposited  an- 
nually and  we  interpreted  the  triple  LAGs  as  a  single 
LAG  with  an  anomalous  appearance.  We  re-evaluated 
this  assumption  after  learning  its  history.  The  animal 
was  in  captivity  for  two  years  and  then  released  at 
42.7  cm  SCL  in  October  1994.  Counting  back  from  the 
outside  of  the  bone,  the  outermost  of  the  triplet  LAGs 
would  represent  spring  1996.  Given  this  evidence,  our 
best  interpretation  of  this  bone  section  was  that  the 
innermost  of  the  triplets  of  LAGs  indicated  release  and 
was  therefore  not  an  annual  mark.  The  next  LAG  was 
likely  deposited  the  following  spring  (1995)  and  was 
likely  an  annual  mark.  The  third  of  the  closely  spaced 
LAGs  likely  represented  spring  1996,  indicating  that 
the  animal  did  not  grow  significantly  in  its  first  year 
in  the  wild  (Fig.  5).  Following  the  three  closely  spaced 
LAGs.  there  were  four  additional  indistinct  LAGs  or 


annuli  that  represented  the  remaining  years  at  large. 
The  outermost  of  these  was  very  close  to  the  edge  of  the 
bone,  indicating  that  the  animal  did  not  grow  much,  if 
at  all,  during  the  last  summer  of  its  life. 

Indirect  validation  of  annual  growth  marks 

For  Kemp's  ridley  sea  turtles,  there  was  a  significant 
increase  in  the  amount  of  bone  deposited  after  the  last 
LAG  from  20  June  to  30  November  (Fig.  6).  The  LAGs 
near  the  outer  edges  of  the  bones  were  fully  visible  in 
strandings  that  occurred  after  20  June.  Earlier  detec- 
tion of  the  outer  LAGs  was  unlikely  because  a  certain 
amount  of  bone  formation  must  occur  following  the  LAG 
before  it  can  be  discerned  from  the  edge.  There  was  not 
a  significant  relationship  between  bone  growth  and  date 
from  1  December  to  19  June.  The  slope  of  this  regression 
was  very  close  to  zero  (6  =  -0.003).  indicating  no  trend, 
either  increasing  or  decreasing,  in  the  amount  of  bone 
deposited  during  this  time  (Fig.  6). 

Validation  of  the  relationship  between  LAG  diameter 
and  body  size 

The  regressions  of  the  eight  morphometric  measure- 
ments of  loggerhead  and  Kemp's  ridley  sea  turtle  humeri 
against  SCL  revealed  high  correlations  between  bone 
dimension  and  body  size  (Table  2).  Most  importantly  for 
purposes  of  back-calculation,  the  lateral  diameter  at  the 
sectioning  site  of  the  humerus  i  distal  to  the  insertion 
scar  of  the  deltopectoral  muscle)  and  the  body  length  of 
the  animal  was  highly  correlated. 


Snover  and  Hohn:  Validation  and  interpretation  of  skeletal  marks  in  Caretta  caretta  and  Lepidochetys  kempii 


687 


Discussion 

Validation  of  the  annual  nature  of  growth  marks 

Our  results  supported  annual  deposition  of  GMs  in  log- 
gerhead and  Kemp's  ridley  sea  turtles.  The  headstarted 
and  older  CWT  Kemp's  ridley  sea  turtles  in  particular 
highlighted  the  likelihood  of  annual  marks.  These  ani- 
mals displayed  sharp  and  regularly  spaced  LAGs  that 
were  consistent  with  the  actual  ages  of  the  animals. 
The  results  from  the  CWT  Kemp's  ridley  sea  turtles  also 
emphasized  the  difficulties  in  interpreting  early  GMs. 
From  these  animals  we  concluded  that  in  general  Kemp's 
ridley  sea  turtles  deposit  a  poorly  defined  annulus  in 
their  first  year  and  well-defined  LAGs  starting  with  the 
end  of  the  second  year  and  in  following  years. 

For  loggerhead  sea  turtles,  only  CC-2  spent  any  time 
in  the  wild.  The  number  of  GMs  deposited  after  the 
animal  was  released  (determined  from  the  appearance 
of  the  anomalous  triplet  of  LAGs)  was  consistent  with 
the  number  of  years  for  which  the  animal  was  at  large, 
considering  that  the  first  mark  was  deposited  at  release. 
This  indicated  that  not  less  than  one  GM  was  deposited 
per  year,  and  that  additional  or  supplemental  LAGs 
or  annuli  indistinguishable  from  annual  lines  may  be 
deposited  under  extreme  conditions,  such  as  at  the  time 
of  release  into  the  wild.  Fortunately,  in  this  case,  these 
extreme  conditions  were  not  frequent  enough  to  have  a 
serious  impact  on  age  estimates.  For  the  life-time  cap- 
tive animal,  CC-1,  our  estimated  minimum  age  was 
five  years  shorter  than  the  actual  age  of  29.4  years  and 
clearly  demonstrated  that  not  more  than  one  GM  was 
deposited  each  year.  Because  of  the  relatively  large  size 
of  the  sea  turtle  humerus,  in  comparison  to  phalanges 
of  amphibians,  rapprochement  did  not  appear  to  be  a 
problem  in  our  attempts  to  discern  LAGs.  This  bone 
was  similar  in  appearance  to  adult  wild  loggerhead  and 
Kemp's  ridleys  sea  turtles  with  rapprochement  of  the 
peripheral  LAGs  and  resorption  of  most  of  the  interior 
GMs.  Although  accurate  age  estimates  cannot  be  made 
of  these  bones  through  skeletochronology,  if  rapproche- 
ment correlates  to  the  timing  of  sexual  maturity,  counts 
of  the  compressed  GMs  can  provide  valuable  informa- 
tion on  postreproductive  longevity  and  adult  survival. 
This  information  can  be  combined  with  average  age  at 
reproductive  maturation  for  piecing  together  the  life 
history  of  sea  turtles.  Although  our  sample  size  for 
loggerhead  sea  turtles  was  very  small  (two),  the  size 
complements  a  tetracycline-injection  study  that  previ- 
ously validated  annual  GMs  for  juvenile  loggerhead 
sea  turtles  from  Chesapeake  Bay  (Klinger  and  Musick, 
1992).  In  addition,  an  adult  loggerhead  sea  turtle  from 
that  same  study  stranded  dead  8.25  years  after  in- 
jection and  provided  evidence  of  annual  deposition  of 
growth  marks  in  adults  (Coles  et  al.,  2001). 

The  indirect  validation  results  for  Kemp's  ridley  sea 
turtles  highlighted  the  cyclic  nature  of  bone  growth; 
bone  deposition  increases  from  late  spring  through  early 
summer  to  fall  and  no  bone  deposition  occurs  from  De- 
cember to  spring.  From  this  information  we  inferred 


10 


11 


B 


-9- 
-10- 


11 


=14/15= 

16 


12 


13 


17 


19, 


20 


c 


-24- 


"./-23- 


.-22- 


J2.Y 


J20- 


Figure  4 

Images  of  different  portions  of  the  humerus  cross-sec- 
tions of  CC-1  (Caretta  caretta).  Black  bar  represents  1 
mm  in  length  for  all  views.  (A  and  Bl  The  outer  edge 
of  the  bone  is  at  the  top  of  the  photo.  (C)  The  outer 
edge  of  the  bone  is  towards  the  bottom  of  the  photo.  For 
all  views,  lines  of  arrested  growth  (LAGs)  are  labeled 
with  numbers;  low  numbers  represent  the  most  recently 
deposited  LAGs  (near  the  outer  edge  of  the  bone)  and 
higher  numbers  represent  the  earlier  LAGs. 


that  LAGs  form  annually  in  the  spring  for  Kemp's  rid- 
ley sea  turtles  that  strand  along  the  mid-  to  southeast 
U.S.  Atlantic  coast  and  that  these  LAGs  are  visible  at 
the  edges  of  the  bones  by  late  spring  to  early  summer. 


688 


Fishery  Bulletin  102(4) 


triple  LAG 


^> . 


>-'.c  •*    a  C  o 
y'  ?    f    ^    5  >    Q 

J3   -j    -  ^    * 

a       *•»  '•  ■•"  h"  ■  ■     ' 


\  ' 


■-■> 


'¥> 


o  " 


Figure  5 

Image  of  a  section  of  the  humerus  cross-section  of 
CC-2  (C.  caretta).  Outer  edge  of  bone  is  towards  the 
bottom  of  the  photo.  Solid  lines  (upper  left  I  highlight  a 
series  of  triple  lines  of  arrested  growth  (LAGs);  curved 
dashed  lines  highlight  the  three  diffuse  LAGs.  Black 
bar  represents  1  mm  in  length. 


Most  studied  species  of  reptiles  and  amphibians  deposit 
GMs  within  their  bones  (Castanet  et  al.,  1993;  Smirina, 
1994).  For  some  of  these  species,  the  annual  nature  of  the 
GM  has  been  validated  (e.g..  Tucker,  1997;  de  Buffrenil 
and  Castanet,  2000;  Trenham  et  al.,  2000).  For  others, 
it  is  consistent  with  their  ecology  that  the  marks  must 
represent  annual  events  (Castanet  et  al.,  1993).  Growth 
marks  observed  in  loggerhead  (Zug  et  al.,  1986;  Zug  et 
al.,  1995;  Coles  et  al.,  2001),  Kemp's  ridley  (Zug  et  al., 
1997),  and  green  (Zug  and  Glor,  1998;  Zug  et  al.,  2002) 
sea  turtles  are  similar  in  structure  to  those  observed  in 
other  species  of  reptiles  and  amphibians.  Drawing  on 
previous  studies  of  reptiles  and  amphibians,  validation 
studies  on  sea  turtles,  and  the  evidence  presented  in  this 
article,  we  assert  that  GM  in  bones  of  sea  turtles  are 
likely  deposited  primarily  with  an  annual  periodicity. 

Given  these  results,  on  the  surface  it  seems  contradic- 
tory that  in  two  validation  studies  annual  GMs  could 
not  be  confirmed.  For  serpentine  species,  Collins  and 
Rodda  (1994)  injected  brown  snakes  with  a  fluorescent 
marker  and  kept  them  in  captivity  for  one  year  under 
two  different  feeding  regimes.  Five  or  six  GMs  vary- 
ing in  distinctness  were  identified  beyond  the  fluores- 
cent marks  in  bone  cross-sections.  Statistical  analyses 
showed  that  these  marks  may  relate  to  shedding  events. 
It  is  unclear  if  the  GM  pattern  prior  to  captivity  was 
similar  to  what  was  seen  after  the  fluorescent  mark. 
The  forced  feeding  component  of  that  study  may  have 
induced  higher  growth  rates  than  would  be  found  in 
nature,  causing  the  shedding  events  to  appear  as  his- 
tological marks  in  the  bone. 

In  a  sea  turtle  study,  Bjorndal  et  al.  (1998)  did  not 
find  GMs  in  the  humeri  of  green  sea  turtle  bones.  They 
suggested  that  the  tropical  marine  habitat  of  the  study 


5  -i 


LU        3  - 


150 


250 


350 
Julian  date 


450 


550 


Figure  6 

Julian  date  of  stranding  plotted  against  the  amount  of 
bone  deposited  peripherally  to  the  last  LAG  in  Kemp's 
ridley  sea  turtles  (L.  kempii:  n  =  76).  D0  represents 
the  outside  diameter  of  the  humerus,  DL  represents 
the  diameter  of  the  last  LAG.  Julian  dates  on  x-axis 
equate  to  20  June  through  19  June;  therefore  num- 
bers that  are  greater  than  365  represent  the  Julian 
date  plus  365.  Solid  lines  represent  linear  regressions 
that  were  run  separately  for  6  months.  20  June  to  31 
November  I  filled  squares)  and  1  December  to  19  June 
(open  squares).  The  regression  for  the  first  six  months 
was  significant  (P<0.006i  and  the  regression  for  the 
second  six  months  was  not  significant  lP=  0.27 1. 


population  (approximately  21°07'N)  allowed  for  con- 
tinual activity  and  growth  and  inhibited  GM  forma- 
tion. However,  GMs  have  been  clearly  demonstrated 
in  green  sea  turtles  from  the  coastal  waters  of  Florida 
(approximately  29°N)  (Zug  and  Glor,  1998)  and  Hawaii 
(approximately  22°N)  (Zug  et  al.,  2002).  Other  studies 
of  reptiles  and  amphibians  in  tropical  and  warm  tem- 
porate  climates  have  reported  distinct  GMs  in  species 
that  remain  active  year-round  (i.e.,  do  not  hibernate 
or  estivate)  (Patnaik  and  Behera,  1981;  Estaban  et  al., 
1996;  Guarino  et  al..  1998). 

Interpretation  of  anomalous  LAGs 

Although  our  sample  sizes  were  small,  especially  for  log- 
gerhead sea  turtles,  several  characteristics  were  noted 
in  the  analyses  of  the  samples  that  would  affect  how 
anomalous  LAGs  are  interpreted.  Three  interpretations 
of  double  and  bifurcating  LAGs  are  provided.  The  first 
interpretation  is  that  if  double  LAGs  appear  frequently 
in  individual  bones  and  throughout  the  sample,  they 
likely  indicate  an  ecology  that  has  two  growth  cycles 
per  year  (Castanet  et  al.,  1993).  In  this  case  the  two 
LAGs  are  distinct  from  each  other  over  the  entire  bone 
cross-section.  This  pattern  was  observed  in  the  newt 
Triturus  marmoratus  living  at  a  high  altitude  where 
the  animals  had  both  winter  and  summer  dormancy 
periods  (Castanet  and  Smirina.  1990;  Caetano  et  al., 
1985;  Caetano  and  Castanet,  1993).  The  second  inter- 
pretation of  double  LAGs  is  that  they  result  from  a  brief 


Snover  and  Hohn:  Validation  and  interpretation  of  skeletal  marks  in  Caretta  caretta  and  Lepidoche/ys  kempn 


689 


Table  2 

Regressions  equations  and  statistics  from  correlations  between  dimensions  of  the  humerus  and  notch-to-tip  straight  carapace 
length  (SCL,  cm)  in  loggerhead  and  Kemp's  ridley  sea  turtles.  All  F  statistics  are  significant  at  P<0.005. 


Humeral  measurement 


Model  equation 


SE  slope 


Loggerhead  sea  turtles  In  =243) 
Maximal  length  (ML,  mm) 
Longitudinal  length  ILL,  mm) 
Proximal  width  (PW.  mm) 
Deltopectoral  crest  width  (DCW,  mm) 
Site  of  sectioning  width  ISW,  mm) 
Site  of  sectioning  thickness  ( ST,  mm) 
Distal  width  (DW,  mm) 
Mass(M,  g) 

Kemp's  ridley  sea  turtles  (rc=262) 
Maximal  length  (ML,  mm) 
Longitudinal  length  (LL,  mm) 
Proximal  width  iPW,  mm) 
Deltopectoral  crest  width  (DCW,  mm) 
Site  of  sectioning  width  ( S  W,  mm ) 
Site  of  sectioning  thickness  (ST.  mm) 
Distal  width  ( DW,  mm) 
Mass  (M,  g) 


SCL  =  0.44xA/L  +  5.97 
SCL  =  0.47xLL  +  4.85 
SCL  =  1.06xPW  +  7.31 
SCL=  1.69xDCW  +  6.04 
SCL  =  2.38xSW  +  5.48 
SCL  =  4.13xST  +  11.62 
SCL  =  1.28xZW+5.43 
ln(SCL)  =  0.30xln(M)  +  2.94 

SCL  =  0.43xML  +  4.69 
SCL  =  0.47xLL  +  3.11 
SCL  =  1.12xPW+4.39 
SCL  =  1.69xDCW  +  3.35 
SCL  =  2.48xSW  +  2.74 
SCL  =  4.16xST+  4.79 
SCL=  1.36xDW+  0.227 
LNiSCL)  =  0.30xLN(M)  +  2.89 


0.0064 

4814 

0.95 

0.0064 

5381 

0.96 

0.015 

4857 

0.95 

0.026 

4069 

0.94 

0.037 

4110 

0.94 

0.080 

2682 

0.92 

0.021 

3684 

0.94 

0.0022 

18905 

0.99 

0.0040 

10970 

0.98 

0.0039 

14772 

0.98 

0.010 

12390 

0.98 

0.017 

10200 

0.98 

0.033 

5715 

0.96 

0.072 

3306 

0.93 

0.013 

11435 

0.98 

0.0023 

16305 

0.98 

interruption  of  hibernation  (Hemelaar  and  van  Gelder, 
1980).  In  this  instance  little  bone  deposition  would 
occur  and  the  layers  would  not  be  distinct  from  each 
other  over  the  entire  bone,  thus  giving  the  appearance 
of  a  bifurcating  LAG  (Hemelaar  and  van  Gelder,  1980). 
The  third  interpretation  of  double  or  bifurcating  LAGs 
is  that  they  result  from  extreme  decreased  growth  over 
the  active  period,  which  places  annual  LAGs  very  close 
to  each  other  and  in  some  cases  they  appear  to  merge 
(de  Buffrenil  and  Castanet,  2000). 

With  the  first  two  interpretations,  a  double  or  bifur- 
cating LAG  would  be  counted  as  one  for  the  purposes  of 
age  estimation,  whereas  the  third  interpretation  would 
necessitate  counting  each  LAG  or  bifurcating  branch 
separately.  Coles  et  al.  (2001)  interpreted  a  bifurcat- 
ing LAG  as  one  LAG  in  an  adult  loggerhead  sea  turtle 
that  was  recovered  8.25  years  after  it  had  been  injected 
with  oxytetracycline.  In  cross-sections  of  the  humerus, 
Coles  et  al.  (2001)  reported  seven  LAGs  following  the 
tetracycline  mark,  six  plus  the  bifurcating  LAG.  The 
animal  was  marked  on  20  June  1989  and  recovered 
dead  on  22  September  1997.  It  is  reasonable  to  assume 
that,  as  with  Kemp's  ridley  sea  turtles  from  the  same 
region,  the  LAGs  form  in  the  spring,  and  Coles  et  al. 
(2001)  showed  that  the  oxytetracycline  mark  overlaid 
one  of  the  LAGs — likely  the  LAG  deposited  in  spring 
of  1989.  Therefore,  there  should  have  been  eight  LAGs 
deposited  after  the  tetracycline  mark,  not  seven,  each 
representing  the  spring  of  years  1990  through  1997. 
In  this  case,  then,  the  bifurcating  mark  in  this  bone 
should  be  counted  as  two  LAGs. 


Similarly,  for  splitting  LAGs,  where  numerous  thin- 
ner LAGs  branch  out  from  what  appears  to  be  one  thick 
LAG,  Francillon-Vieillot  et  al.  (1990)  examined  different 
bones  from  the  same  animal  and  determined  whether 
each  thin  LAG  comprising  splitting  LAGs  should  be 
counted  as  one  LAG.  In  our  analysis  of  the  adult  log- 
gerhead sea  turtle,  CC-1,  we  observed  several  bifurcat- 
ing and  splitting  LAGs.  each  of  which  eventually  split 
into  two  or  more  thinner  LAGs.  We  counted  each  of  the 
thin  LAGs  as  one.  Because  the  LAG  count  was  close  to 
the  actual  age  of  the  animal,  this  interpretation  ap- 
pears to  have  been  appropriate  for  compressed  LAGs 
in  adult  humeri. 

The  question  remains  as  to  whether  this  is  the  ap- 
propriate interpretation  for  double  or  bifurcating  LAGs 
in  juveniles.  Wild  loggerhead  growth  rates  have  been 
monitored  in  an  ongoing  mark-recapture  study  in  Pam- 
lico and  Core  Sounds  in  North  Carolina  (Epperly  et  al., 
1995).  Epperly  et  al.  (1995)  currently  have  65  growth 
rates  for  49  juvenile  loggerhead  sea  turtles  between  45.1 
and  81.0  cm  SCL  at  initial  capture  that  were  at-large 
for  one  year  (±0.1  year).  The  mean  annual  growth  rate 
for  all  of  the  animals  is  2.09  cm/yr.  However,  of  the 
65  growth  records,  11  of  them  displayed  an  annual  in- 
crease of  0.3  cm  or  less  in  SCL  (Braun-McNeill1).  Hence 
it  is  not  uncommon  for  juvenile  loggerhead  sea  turtles  to 


Braun-McNeill,  J.  2004.  Personal  commun.  Center  for 
Coastal  Fisheries  and  Habitat  Research,  National  Marine 
Fisheries  Service,  NOAA,  101  Pivers  Island  Rd.,  Beaufort, 
NC  28516 


690 


Fishery  Bulletin  102(4) 


grow  little  or  not  at  all  over  the  course  of  a  year.  Using 
the  equation  for  width  at  sectioning  site  from  Table  2, 
we  found  that  the  increase  in  bone  diameter  for  these  11 
animals  was  =0.13  mm  or  less,  which  places  the  LAGs 
very  close  together.  Because  it  not  uncommon  for  sea 
turtle  to  exhibit  little  or  no  growth  over  a  year,  LAGs 
spaced  closely  together  very  likely  represent  distinct 
years  as  also  determined  by  de  Buffrenil  and  Castanet 
(2000).  Although  the  sample  sizes  are  still  small  for  a 
definitive  answer,  our  results  indicate  that  counting 
the  LAGs  individually  is  the  correct  interpretation  of 
double  or  bifurcating  LAGs  in  juvenile  as  well  as  adult 
loggerhead  sea  turtles. 

Similarly,  our  results  indicate  the  same  interpretation 
for  double  or  bifurcating  LAGs  in  juvenile  Kemp's  ridley 
sea  turtles.  The  CWT  Kemp's  ridley  sea  turtles,  samples 
LK-7  and  LK-8,  displayed  LAGs  near  the  outer  edge 
of  the  bone  and  a  small  amount  of  bone  was  deposited 
after  the  LAGs.  These  animals  were  each  2.25  years 
old  and  had  one-year  marks  visible  in  the  humeri  but 
no  LAGs  or  annuli  other  than  those  at  the  periphery. 
Other  CWT  samples  clearly  indicated  that  LAGs  are 
deposited  at  the  end  of  the  second  GM.  The  indirect 
validation  results  demonstrated  that  LAGs  were  visible 
in  bone  tissue  by  late  spring  or  early  summer.  It  seemed 
that  the  LAGs  at  the  outer  edge  of  the  LK-7  and  LK-8 
bones  were  the  LAGs  ending  the  second  GM  and  that 
very  little  growth  occurred  over  the  subsequent  growing 
season.  Both  of  these  animals  were  recovered  as  dead 
strandings  resulting  from  a  major  cold  stun  event  in 
Cape  Cod,  Massachusetts,  in  1999;  hence  their  growth 
rates  may  have  been  anomalous  in  their  last  year  of 
life.  Had  these  animals  survived  the  cold  stun  event, 
they  would  have  deposited  a  year-three  LAG  very  close 
to  year  two,  giving  the  appearance  of  a  double  or  bi- 
furcating LAG. 

Another  anomaly  in  skeletochronology,  supplemen- 
tal lines,  may  form  as  a  result  of  temporary  stressful 
environmental  events  such  as  droughts.  In  support  of 
this,  Rogers  and  Harvey  (1994)  noted  a  supplemental 
line  in  11  of  43  specimens  of  the  toad  Bufo  cognatus, 
and  in  10  of  these  animals  the  supplemental  line  was 
within  a  growth  zone  that  corresponded  to  a  drought 
year.  Most  skeletochronology  studies  that  have  noted 
the  presence  of  supplemental  lines  have  indicated  that 
supplemental  lines  are  easily  identified  as  such  because 
they  are  less  distinct  and  do  not  appear  around  the 
entire  circumference  of  the  bone.  In  general,  the  same 
has  been  observed  in  sea  turtles.  Supplemental  lines 
do  appear  but  are  generally  easily  differentiated  from 
LAGs  by  appearance.  An  exception  to  this  was  the 
presence  of  supplemental  marks  in  one-  to  two-year-old 
Kemp's  ridley  sea  turtles.  These  marks  were  similar 
in  appearance  to  the  first  year  annuli.  We  were  able 
to  identify  these  marks  as  supplemental  only  by  the 
observation  of  known-age  animals.  In  addition,  there 
appeared  to  be  a  supplemental  line  in  CC-2  that  rep- 
resented when  the  animal  was  released;  hence,  highly 
stressful  events  may  cause  the  deposition  of  nonannual 
lines,  but  these  events  are  likely  to  be  relatively  rare 


in  wild  turtles  and  not  likely  to  interfere  significantly 
with  age  estimations. 

Resorption  of  early  growth  marks 

The  loss  of  the  early  GMs  due  to  endosteal  resorption 
and  remodeling  of  the  interior  region  of  the  bone  is  a  lim- 
iting factor  in  the  application  of  skeletochronology  to  sea 
turtles.  From  our  findings,  it  was  possible  to  accurately 
age  juvenile  Kemp's  ridley  sea  turtles  up  to  at  least  5 
years  from  GM  counts  and  this  may  be  true  for  other 
sea  turtle  species  (e.g.,  Bjorndal  et  al.,  2003),  with  the 
possible  exception  of  the  leatherback  sea  turtle  (Zug  and 
Parham,  1996).  Because  sea  turtles  have  distinct  life- 
cycle  stages,  we  suggest  that  in  order  to  age  a  population 
of  sea  turtles,  one  must  acquire  an  ontogenetic  series  of 
samples  spanning  all  sizes  and  stages.  Average  duration 
can  be  determined  for  each  ontogenetic  stage  and  the 
approximate  age  of  older  animals  with  extreme  resorp- 
tion can  be  estimated.  Because  GM  patterns  appear  to 
mimic  somatic  growth  rates,  once  growth  through  each 
life-cycle  stage  is  understood,  backcalculation  techniques 
can  be  used  to  estimate  the  number  of  layers  resorbed. 


Conclusions 

For  many  species,  skeletochronology  is  not  a  perfect 
method  for  age  estimation.  As  GMs  are  histological 
expressions  of  variation  in  rates  of  osteogenesis  (Casta- 
net et  al.,  1993).  external  factors  and  individual  varia- 
tion will  affect  the  appearance  of  the  marks  (Castanet 
et  al.,  1993,  Esteban  et  al.,  1996,  Wave  and  Gregory. 
1998).  Endosteal  resorption  also  serves  to  confound  this 
technique  and  is  the  primary  difficulty  in  the  application 
of  the  technique  to  sea  turtles.  However,  the  evidence 
presented  in  the  present  study  gives  strong  support 
to  the  concept  that  GMs  are  deposited  on  an  annual 
basis  in  sea  turtles  and  that  the  spatial  pattern  of  the 
GMs  correspond  to  the  growth  rates  of  the  animal.  The 
GMs  therefore  provide  invaluable  information  on  age 
and  growth  that  cannot  otherwise  be  easily  obtained, 
and  age  determination  by  skeletocronology  is  valid  and 
appropriate  for  the  study  of  sea  turtles. 


Acknowledgments 

We  thank  L.  Crowder,  S.  Heppell,  A.  Read,  and  D. 
Rittschof  for  their  valuable  comments  on  earlier  ver- 
sions of  this  manuscript.  A.  Gorgone.  B.  Brown  and  J. 
Weaver  provided  assistance  with  the  preparation  of  the 
humeri.  Most  of  the  humeri  were  received  through  the 
Sea  Turtle  Stranding  and  Salvage  Network,  a  coopera- 
tive endeavor  between  the  National  Marine  Fisheries 
Service,  other  federal  and  state  agencies,  many  academic 
and  private  entities,  and  innumerable  volunteers.  We 
especially  thank  R.  Boettcher  and  W.  Teas.  In  addition, 
humeri  were  received  from  F.  Swartz  at  the  Univer- 
sity of  North  Carolina-Chapel  Hill  Institute  of  Marine 


Snover  and  Hohn:  Validation  and  interpretation  of  skeletal  marks  in  Caretta  caretta  and  Lepidochelys  kempu 


691 


Science,  B.  Higgins  at  the  National  Marine  Fisheries 
Service-Galveston  Lab,  the  Virginia  Marine  Science 
Museum  Stranding  Program,  the  Maryland  Department 
of  Natural  Resources,  and  the  Massachusetts  Audu- 
bon Society  in  Wellfleet.  Funding  was  provided  by  the 
National  Marine  Fisheries  Service.  All  work  was  done 
under  and  complied  with  the  provisions  of  the  Sea  Turtle 
Research  Permit  TE-676379-2  issued  by  the  U.S.  Fish 
and  Wildlife  Service. 


Literature  cited 

Bjorndal,  K.  A.,  A.  B.  Bolten,  R.  A.  Bennet,  E.  R.  Jacobson, 
T.  J.  Wronski,  J.  J.  Valeski,  and  R  J.  Eliazar. 

1998.     Age  and  growth  in  sea  turtles:  limitations  of  skel- 
etochronology for  demographic  studies.     Copeia  1998: 
23-30. 
Bjorndal,  K.  A.,  A.  B.  Bolten,  T.  Dellinger,  C.  Delgado,  and 
H.  R.  Martins. 

2003.     Compensatory  growth  in  oceanic  loggerhead  sea 
turtles:  response  to  a  stochastic  environment.     Ecology 
84:1237-1249. 
Caetano,  M.  H.,  and  J.  Castanet. 

1993.     Variability  and  microevolutionary  patterns  in  Tritu- 
rus marmoratus  from  Portugal:  age,  size,  longevity  and 
individual  growth.     Amphibia-Reptilia  14:117-129. 
Caetano,  M.  H.,  J.  Castanet,  and  H.  Francillon. 

1985.     Determination  of  the  age  of  Triturus  marmoratus 
marmoratus  from  the  National  Park  of  Peneda  Geres, 
Portugal  by  means  of  skeletochronology.     Amphibia- 
Reptilia  6:117-132. 
Caillouet,  C.  W.,  B.  A.  Robertson,  C.  T.  Fontaine, 
T.  D.  Williams,  B.  M.  Higgins,  and  D.  B.  Revera. 

1997.     Distinguishing  captive-reared  from  wild  Kemp's 
ridleys.     Marine  Turtle  News  Letter  77:1-6. 
Castanet,  J.,  H.  Francillon-Vieillot,  and  R.  C.  Bruce. 

1996.  Age  estimation  in  desmognathine  salaman- 
ders assessed  by  skeletochronology.  Herpetologica 
52:160-171. 

Castanet.  J.,  H.  Francillon-Vieillot,  F  J.  Meunier,  and 
A.  De  Ricqles. 

1993.     Bone  and  individual  aging.     In  Bone,  vol.  7:  Bone 
growth  B  (B.  B.  K.  Hall,  ed.),  p.  245-283.     CRC  press, 
Boca  Raton,  FL. 
Castanet,  J.,  F  J.  Meumier,  and  A.  de  Ricqles. 

1977.     L'enregistrement  de  la  croissance  cyclique  par  le 
tissu  osseux  chez  les  Vertebres  poikilothermes:  donnees 
comparatives  et  essai  de  synthese.     Bull.  Biol.  Fr.  Belg. 
111:183-202. 
Castanet.  J.,  and  E.  Smirina. 

1990.     Introduction  to  the  skeletochronological  method  in 
amphibians  and  reptiles.     Ann.  Sci.  Nat.  Zool.  11:191— 
196. 
Chaloupka,  M.  Y.,  and  J.  A.  Musick. 

1997.  Age,  growth,  and  population  dynamics.  In  The 
biology  of  sea  turtles  (P.  L.  Lutz  and  J.  A.  Musick,  eds.) 
p.  233-276.     CRC  Press,  New  York. 

Chinsamy,  A.,  S.  A.  Hanrahan,  R.  M.  Neta,  and  M.  Seely. 

1995.     Skeletochronological  assessment  of  age  inAngolo- 
saurus  skoogi,  a  cordylid  lizard  living  in  an  aseasonal 
environment.     J.  Herpetol.  29:457-460. 
Coles,  W.  C,  J.  A.  Musick,  and  L.  A.  Williamson. 

2001.  Skeletochronology  validation  from  an  adult  log- 
gerhead (Caretta  caretta).     Copeia  2001:240-242. 


Collins,  E.  P.,  and  G.  H.  Rodda. 

1994.  Bone  layers  associated  with  ecdysis  in  labora- 
tory-reared Boiga  irregularis  (Colubridaei.  J.  Herpetol. 
28:378-381. 

de  Buffrenil,  V,  and  J.  Castanet. 

2000.     Age  estimation  by  skeletochronology  in  the  Nile 
monitor  (Varanus  niloticus),  a  highly  exploited  species. 
J.  Herpetol.  34:414-424. 
Eggert,  C,  and  R.  Guyetant. 

1999.     Age  structure  of  a  spadefoot  toad  Pelobates  fucus 
(Pelobatidae)  population.     Copeia  1999:1127-1130. 
El  Mouden,  P.  E„  H.  Francillon-Vieillot,  J.  Castanet,  and  M.  Znari. 

1997.  Age  individuel  maturite,  croissance  et  longevite 
chez  l'agamide  nord-africain.  Agama  impalearis  Boettger, 
1874,  etudies  a  l'aide  de  la  squelettochronologie. 

Epperly,  S.  P.,  J.  Braun,  and  A.  Veishlow. 

1995.  Sea  turtles  in  North  Carolina  waters.  Conserv. 
Biol.  9:384-394. 

Esteban,  M.,  M.  Garcia-Paris.  and  J.  Castanet. 

1996.  Use  of  bone  histology  in  estimating  the  age  of  frogs 
(Rana  perezi)  from  a  warm  temperate  climate  area. 
Can.  J.  Zool.  74:1914-1921. 

Francillon-Vieillot,  H..  J.  W.  Arntzen,  and  J.  Geraudie. 

1990.  Age,  growth  and  longevity  of  sympatric  Triturus 
cristatus,  T.  marmoratus  and  their  hybrids  I  Amphibia, 
Urodelal.  A  skeletochronological  comparison.  J.  Her- 
petol. 24:13-22. 

Guarino.  F.  M.,  F.  Andreone,  and  F  Angelini. 

1998.  Growth  and  longevity  by  skeletochronological 
analysis  in  Mantidactylus  mierotympanum,  a  rain- 
forest anuran  from  southern  Madagascar.  Copeia 
1998:194-198. 

Guarino,  F.  M.,  F.  Angelini,  and  M.  Cammarota. 

1995.  A  skeletochronological  analysis  of  three  syntopic 
amphibian  species  from  southern  Italy.  Amphibia-Rep- 
tilia 16:297-302. 

Hemelaar,  A.  S.  M.,  and  J.  J.  van  Gelder. 

1980.     Annual  growth  rings  in  phalanges  of  Bufo  bufo 
(Anura,  Amphibia  i  from  the  Netherlands  and  their  use 
for  age  determination.     Neth.  J.  Zool.  30:129-135. 
Hutton,  J.  M. 

1986.     Age  determination  of  living  Nile  crocodiles  from 
the   cortical   stratification   of  bone.     Copeia    1986: 
332-341. 
Klevezal,  G.  A. 

1996.  Recording  structures  of  mammals:  determination 
of  age  and  reconstruction  of  life  history,  p.  49-52.  A. 
A.  Balkema,  Rotterdam. 

Klima,  E.  F„  and  J.  P.  McVey. 

1995.     Headstarting  the  Kemp's  ridley  turtle,  Lepido- 
chelys kempi.     In  The  biology  and  conservation  of  sea 
turtles  (K.  A.  Bjorndal,  eds.),  p.  481-487.     Smithsonian 
Institution,  Washington,  D.C.. 
Klinger,  R.  C,  and  J.  A.  Musick. 

1992.     Annular  growth  layers  in  juvenile  loggerhead 

turtles  (Caretta  caretta).     B.  Mar.  Sci.  51:224-230. 
1995.     Age  and  growth  of  loggerhead  turtles  ( Caretta  caretta ) 
from  Chesapeake  Bay.     Copeia  1995:204-209. 
Kusano,  T.,  K.  Fukuyama,  and  N.  Miyashita. 

1995.  Age  determination  of  the  stream  frog,  Rana  sakuraii, 
by  skeletochronology.     J.  Herpetol.  29:625-628. 

Leclair  Jr.,  R.,  and  G.  Laurin. 

1996.  Growth  and  body  size  in  populations  of  mink  frogs 
Rana  septentrionalis  from  two  latitudes.  Ecography 
19:296-304. 


692 


Fishery  Bulletin  102(4) 


Leclair  Jr.,  M.  H.  Leclair,  J.  Dubois,  and  J-L.  Daoust. 

2000.  Age  and  size  of  wood  frogs.  Rana  sylvatiea.  from 
Kuujjuarapik,  Northern  Quebec.  Can.  Field.  Nat. 
114:381-387. 

Lima,  V.,  J.  W.  Arntzen,  and  N.  M.  Ferrand. 

2000.  Age  structure  and  growth  pattern  in  two  popu- 
lations of  the  golden-striped  salamander  Chioglossa 
lusitanica  (Caudata  Salamandra).  Amphibia-Reptilia 
22:55-68. 

Parham,  J.  F,  and  G.  R.  Zug. 

1997.  Age  and  growth  of  loggerhead  sea  turtles  (Caretta 
caretta)  of  coastal  Georgia:  an  assessment  of  skeletochro- 
nological  age-estimates.     B.  Mar.  Sci.  61:287-304. 

Patnaik,  B.  K„  and  H.  N.  Behera. 

1981.     Age-determination  in  the  tropical  agamid  garden 
lizard,   Calotes  versicolor  iDaudini.  based  on  bone 
histology.     Exp.  Gerontol.  16:295-307. 
Peabody,  F.  E. 

1961.     Annual  growth  zones  in  vertebrates  (living  and 
fossil).     J.  Morphol.  108:11-62. 
Rogers,  K.  L.,  and  L.  Harvey. 

1994.     A  skeletochronology  assessment  of  fossil  and  recent 
Bufo  cognatus  from  south-central  Colorado.     J.  Herpe- 
tol,  28:133-140. 
Sagor,  E.  S.,  M.  Ouellet,  E.  Barten,  and  D.  M.  Green. 

1998.  Skeletochronology  and  geographic  variation  in  age 
structure  in  the  wood  frog,  Rana  sylvatiea.  J.  Herpetol. 
32:469-474. 

Simmons,  D.  J. 

1992.     Circadian  aspects  of  bone  biology.     In  Bone,  vol. 
6:  Bone  growth  A  i  B.  B.  K.  Hall,  ed.  I,  p.  91-128.     CRC 
press,  Boca  Raton,  FL. 
Smirina,  E.  M. 

1994.     Age  determination  and  longevity  in  amphibians. 
Gerontology  40:133-146. 
Swartz,  F.  J. 

1997.  Growth,  maturity,  and  reproduction  of  a  long-term 
captive  male  loggerhead  sea  turtle,  Caretta  caretta  (Che- 
Ionia,  Reptilia)  in  North  Carolina.  Elisha  Mitchell 
Scientific  Society  133:143-148. 


Trenham,  P.  C,  H.  B.  Shaffer,  W.  D.  Koenig.  and 
M.  R.  Stromberg. 

2000.     Life  history  and  demographic  variation  in  the  Cali- 
fornia tiger  salamander  (Ambystoma  californiense).     Co- 
peia  2000:365-377. 
Tucker,  A.  D. 

1997.  Validation  of  skeletochronology  to  determine  age 
of  freshwater  crocodiles  (Crocodylus  johnstoni).  Mar. 
Freshw.  Res.  48:343-351. 

Waye,  H.  L..  and  P.  T.  Gregory. 

1998.  Determining  the  age  of  garter  snakes  (Thamno- 
phis  spp. )  by  means  of  skeletochronology.  Can.  J.  Zool. 
76:288-294. 

Zug,  G.  R,  G.  H.  Balazs,  and  J.  A.  Wetherall. 

1995.  Growth  in  juvenile  loggerhead  sea  turtles  (Caretta 
caretta)  in  the  north  Pacific  pelagic  habitat.  Copeia 
1995:484-487. 

Zug.  G.  R..  G.  H.  Balazs,  J.  A.  Wetherall,  D.  M.  Parker,  and 
S.  K.  K.  Murakawa. 

2002.     Age  and  growth  of  Hawaiian  green  sea  turtles 
(Chelonia   mydas):   and   analysis  based  on  skeleto- 
chronology.    Fish.  Bull.  100:117-127. 
Zug,  G.  R,  and  R.  E.  Glor. 

1998.     Estimates  of  age  and  growth  in  a  population  of 
green  sea  turtles  (Chelonia  mydas)  from  the  Indian  River 
lagoon  system,  Florida:  a  skeletochronological  analysis. 
Can.  J.  Zool.,  76:1497-1506. 
Zug,  G.  R,  H.  J.  Kalb,  and  S.  J.  Luzar. 

1997.     Age  and  growth  in  wild  Kemp's  ridley  sea  tur- 
tles Lepidochelys  kempii  from  skeletochronological 
data.     Biol.  Conserv.  80:261-268. 
Zug,  G.  R,  and  J.  F.  Parham. 

1996.  Age  and  growth  in  leatherback  sea  turtles,  Der- 
mochelys  coriacea  iTestudines.  Dermochelyidae):  a  skel- 
etochronological analysis.  Chelonian  Conserv.  Biol. 
2:244-249. 

Zug,  G.  R.,  A.  H.  Wynn,  and  C.  Ruckdeschel. 

1986.  Age  determination  of  loggerhead  sea  turtles, 
Caretta  caretta,  by  incremental  growth  marks  in  the 
skeleton.  Smithsonian  Institution,  Contrib.  Zool.  427. 
Washington  D.C. 


693 


Abstract— Blue  tCallinectes  sapidus) 
(Portunidae),  lady  (Ovalipes  ocella- 
tus) (Portunidae I,  and  Atlantic  rock 
(Cancer  irroratus)  (Cancridae)  crabs 
inhabit  estuaries  on  the  northeast 
United  States  coast  for  parts  or  all 
of  their  life  cycles.  Their  distribu- 
tions overlap  or  cross  during  cer- 
tain seasons.  During  a  1991-94 
monthly  otter  trawl  survey  in  the 
Hudson-Raritan  Estuary  between 
New  York  and  New  Jersey,  blue  and 
lady  crabs  were  collected  in  warmer 
months  and  Atlantic  rock  crabs  in 
colder  months.  Sex  ratios,  male: 
female,  of  mature  crabs  were  1:2.0 
for  blue  crabs,  1:3.1  for  lady  crabs, 
and  21.4:1  for  Atlantic  rock  crabs. 
Crabs,  1286  in  total,  were  sub- 
sampled  for  dietary  analysis,  and 
the  dominant  prey  taxa  for  all  crabs, 
by  volume  of  foregut  contents,  were 
mollusks  and  crustaceans.  The  pro- 
portion of  amphipods  and  shrimp  in 
diets  decreased  as  crab  size  increased. 
Trophic  niche  breadth  was  widest  for 
blue  crabs,  narrower  for  lady  crabs, 
and  narrowest  for  Atlantic  rock  crabs. 
Trophic  overlap  was  lowest  between 
lady  crabs  and  Atlantic  rock  crabs, 
mainly  because  of  frequent  consump- 
tion of  the  dwarf  surfclam  (Mulinia 
lateralis)  by  the  former  and  the  blue 
mussel  (Mytilus  edulis)  by  the  latter. 
The  result  of  cluster  analysis  showed 
that  size  class  and  location  of  capture 
of  predators  in  the  estuary  were  more 
influential  on  diet  than  the  species 
or  sex  of  the  predators. 


The  Hudson-Raritan  Estuary  as  a  crossroads 
for  distribution  of  blue  (Callinectes  sapidus), 
lady  (Ovalipes  ocellatus), 
and  Atlantic  rock  {Cancer  irroratus)  crabs 


Linda  L.  Stehlik 
Robert  A.  Pikanowski 
Donald  G.  McMillan 

James  J,  Howard  Marine  Sciences  Laboratory 

Northeast  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

74  Magruder  Road 

Highlands,  New  Jersey  07732 

E-mail  address  (for  L  L  Stehlik):  Linda  Stehlika' noaa.gov 


Manuscript  submitted  27  November 
2000  to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
4  May  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:693-710  (20041. 


The  blue  crab  (Callinectes  sapidus) 
(Portunidae),  the  lady  crab  {Ovali- 
pes ocellatus)  (Portunidae),  and  the 
Atlantic  rock  crab  (Cancer  irroratus) 
(Cancridae)  are  the  largest  and  most 
common  brachyuran  crabs  inhabiting 
both  estuaries  and  inner  continen- 
tal shelves  of  the  northeast  coast  of 
North  America.  The  centers  of  abun- 
dance of  these  three  species  over- 
lap in  estuarine  and  coastal  waters 
from  New  York  to  Virginia,  although 
their  ranges  along  the  northwest 
Atlantic  coast  are  broad.  The  blue 
crab  is  nearly  always  an  estuarine 
resident,  except  during  its  larval 
stages,  and  ranges  from  the  waters 
off  Nova  Scotia  to  Argentina  (Wil- 
liams, 1984).  The  northernmost  estu- 
aries where  the  species  is  abundant 
enough  for  commercial  harvest  are 
in  New  Jersey  and  New  York  (Briggs, 
1998;  Stehlik  et  al.,  1998).  The  lady 
crab  is  distributed  from  the  waters 
off  Prince  Edward  Island  to  those 
off  Georgia  but  it  is  most  numerous 
from  Georges  Bank  to  Cape  Hatteras 
(Williams,  1984).  The  Atlantic  rock 
crab  (referred  to  as  "rock  crab"  in 
this  article)  is  distributed  in  waters 
from  off  Labrador  to  Florida  but  is 
most  common  in  estuaries  from  Nova 
Scotia  to  Virginia  (Williams,  1984; 
Stehlik  et  al.,  1991).  Seasonal  migra- 
tions are  common  for  all  three  spe- 
cies. Although  Jonah  crabs  (Cancer 
borealis)  are  present  on  the  continen- 


tal shelf,  they  are  not  included  in  the 
present  study  because  they  are  rare 
within  the  Hudson-Raritan  Estuary 
where  our  study  was  conducted. 

Physiological  tolerances  and  habi- 
tat preferences  of  these  crabs  have 
been  extensively  studied.  In  eastern 
United  States  estuaries  the  blue  crab 
occurs  in  shallow  to  deep,  sandy  to 
muddy  estuaries  and  tributaries 
along  marsh  edges,  and  in  seagrass 
(Van  Engel,  1958;  Milliken  and  Wil- 
liams, 1984;  Hines  et  al.,  1987;  Wil- 
son et  al.,  1990;  van  Montfrans  et 
al.,  1991;  Rountree  and  Able,  1992). 
In  the  colder  portions  of  its  range, 
it  becomes  less  active  at  about  15°C 
(Leffler,  1972),  and  buries  itself,  with- 
out eating,  when  the  temperature  is 
<5°C  (Auster  and  DeGoursey,  1994). 
It  survives  at  34°C  (Leffler,  1972)  and 
at  salinities  from  0  to  50  ppt  (Guerin 
and  Stickle,  1992).  The  lady  crab  is 
most  common  on  sand  substrates 
(Williams,  1984).  It  is  present  on  the 
inner  continental  shelf  from  off  Cape 
Cod  to  off  the  Carolinas  throughout 
the  year  (Stehlik  et  al.,  1991).  Its  tem- 
perature tolerance  is  unknown,  but  it 
does  not  survive  in  <21  ppt  (Birchard 
et  al.,  1982).  The  rock  crab's  optimum 
temperature  range  for  activity  is  14- 
22°C  (Jeffries,  1966);  thus  the  species 
avoids  high  summer  temperatures. 
It  is  found  on  many  substrates,  such 
as  sand,  mud,  bare  rock,  cobble,  and 
algal  beds. 


694 


Fishery  Bulletin  102(4) 


The  diet  of  the  blue  crab  is  generally  mollusks,  crabs, 
and  fish,  depending  on  crab  size  (Virnstein,  1977; 
Laughlin,  1982;  Ryer,  1987;  Hines  et  al„  1990).  The 
diet  of  the  lady  crab  is  mainly  bivalves  such  as  My  a 
arenaria  and  Spisula  solidissima,  and  some  crustaceans 
(McDermott,  1983;  Ropes,  1989;  Stehlik,  1993).  The 
rock  crab  consumes  mollusks,  small  crustaceans,  crabs, 
urchins,  and  fish  (Scarratt  and  Lowe,  1972;  Drummond- 
Davis  et  al.,  1982;  Hudon  and  Lamarche,  1989;  Ojeda 
and  Dearborn,  1991;  Stehlik,  1993).  In  some  of  the 
aforementioned  studies  these  crabs  have  been  consid- 
ered opportunistic  and  as  such  may  be  competitors  for 
the  same  prey  taxa.  However,  differences  in  maximum 
body  size,  chela  structure,  and  the  presence  or  absence 
of  swimming  appendages  among  blue,  lady,  and  rock 
crabs  indicate  that  they  may  have  differences  in  diet 
(Warner  and  Jones,  1976;  Williams.  1984). 

Within  the  Hudson-Raritan  Estuary,  blue,  lady,  and 
rock  crabs  are  all  abundant,  providing  an  opportunity 
to  study  partitioning  of  habitat  and  food  resources  by 
these  species.  The  objectives  of  our  study  were  to  de- 
termine the  temporal  and  spatial  overlap  of  blue,  lady, 
and  rock  crabs  in  this  estuary  and  to  differentiate  the 
composition  of  their  diets  by  the  species,  sex,  and  size 
of  predators,  and  by  location  of  collection. 

This  study  has  potential  practical  applications.  Re- 
source managers  could  use  the  results  to  consider  when 
and  where  crabs  depend  upon  certain  locations  to  com- 
plete their  life  cycles,  if  dredging,  filling,  or  sanctuaries 
were  proposed.  Dietary  analysis  of  these  crabs  could 
indicate  if  they  are  a  cause  of  mortality  for  young  stages 
of  commercially  important  species.  For  instance,  the 
northern  quahog  (Mercenaria  mercenaria)  and  the  soft- 
shell  clam  <M.  arenaria i  recently  have  supported  and 
presently  support  commercial  and  recreational  harvests 
in  the  Hudson-Raritan  Estuary  (MacKenzie.  1990;  1997) 
and  when  young  these  clams  are  consumed  by  crabs. 
The  blue  crab  supports  lucrative  fisheries  in  the  estuary 
(Stehlik  et  al.,  1998)  and  predation  by  various  species  of 
crabs  upon  blue  crab  juveniles  may  affect  recruitment. 


Materials  and  methods 

Study  area 

The  Hudson-Raritan  Estuary,  bordered  by  New  Jersey  on 
the  south  and  Staten  Island  and  Brooklyn,  New  York,  on 
the  north  (Fig.l),  has  a  surface  area  of  about  280  km2. 
The  Hudson,  Raritan,  and  Navesink-Shrewsbury  rivers 
flow  into  the  estuary  from  the  north,  west,  and  south, 
respectively.  The  study  area  is  bounded  on  the  west  by 
the  74°15'  longitude  line;  on  the  east  by  a  line  between 
the  northeast  corner  of  Sandy  Hook,  NJ,  and  the  tip  of 
Rockaway  Point,  NY;  and  on  all  sides  by  the-3  m  con- 
tour. The  area  was  divided  into  nine  strata  according 
to  physiographic  features  (Wilk  et  al.1).  Sandy-bottom 
strata  included  Sandy  Hook  Bay  (stratum  1),  Raritan 
Bay  south  of  Raritan  Channel  (stratum  2),  and  Lower 
Bay  north  of  Raritan  Channel  (stratum  3).  Eastern 


strata  of  more  irregular  depths  were  Romer  Shoals 
(stratum  4),  East  Bank  between  Ambrose  Channel  and 
Rockaway,  NY  (stratum  5),  and  Gravesend  Bay  at  the 
mouth  of  the  Hudson  River  (stratum  6).  Three  strata 
were  channels;  Ambrose  (stratum  7),  Chapel  Hill  (stra- 
tum 8),  and  Raritan  (stratum  9).  Raritan  Channel  is 
maintained  at  a  depth  of  13.7  m,  and  the  average  depth 
of  adjacent  nonchannel  stations  is  7.1  m.  Gravesend  Bay- 
is  more  than  13  m  deep  in  its  center. 

The  bottom  of  the  Hudson-Raritan  Estuary  consists 
mostly  of  soft  sediments  (Jones  et  al.,  1979;  Coch,  1986: 
Wilber2).  The  substrates  in  semi-sheltered  southern 
strata  1  and  2  are  predominantly  fine  sand,  silt,  and 
clay;  those  of  stratum  3  are  mainly  medium  sand,  with 
a  mixture  of  sand,  silt,  and  clay  near  channels;  those 
of  ocean-exposed  strata  4,  5,  6,  and  7  are  gravel,  sand, 
silt,  broken  shell,  and  have  beds  of  blue  mussels  (Myti- 
lus  edulis).  The  bottom  of  Ambrose  Channel  is  silt  and 
clay  near  its  head  and  fine  sand  toward  the  ocean. 
The  sediments  of  the  other  two  channels,  and  their 
immediate  borders,  are  sand,  silt,  and  clay.  Based  on 
physiographic  form,  temperature  ranges,  and  sediments, 
strata  1,  2,  3,  8,  and  9  were  considered  inner  or  river- 
ward  strata,  whereas  4,  5,  6,  and  7  were  considered 
outer  or  oceanic-influenced  strata. 

Collections  and  analyses 

Crabs  were  collected  during  monthly  otter  trawl  sur- 
veys of  the  Hudson-Raritan  Estuary  from  June  1991  to 
December  1994  (Wilk  et  al.1).  Sampling  was  done  from 
the  18-m  research  vessel  Gloria  Michelle  by  towing 
a  9.1-m  otter  trawl  with  a  chain  sewn  to  the  bottom 
opening,  and  a  76-mm  mesh  net  with  a  51-mm  codend 
liner.  Wooden  trawl  doors  were  deployed  to  spread  open 
the  net.  The  net  was  towed  once  per  station  for  10  min 
at  5.6  km/h  to  cover  a  distance  of  approximately  1  km. 
All  tows  were  made  between  8  am  and  2  pm.  During 
1991.  fixed  stations  were  towed  (number  of  stations  in 
1991:  10  in  June,  8  in  July;  11  in  August;  18  in  Sep- 
tember; 22  in  October;  23  in  November;  34  in  Decem- 
ber). Beginning  in  1992,  a  stratified  random  sampling 
design  was  used,  in  which  the  nine  strata  were  divided 
into  190  blocks  of  approximately  0.5  minutes  latitude 
by  0.5  minutes  longitude.  Each  month,  40  blocks  were 
randomly  sampled  without  replication,  and  the  number 
of  blocks  in  each  stratum  was  proportional  to  the  area 
of  the  stratum.  Because  of  conflicting  schedules,  the 
vessel  was  not  available  in  May  or  September  1992  or 
1994.  Temperature  and  salinity  of  water  1  m  above  the 
bottom  were  measured  after  each  tow.  During  1991 


1  Wilk.  S.  J..  E.  M.  MacHaffie.  D.  G.  McMillan,  A.  L.  Pacheco,  R. 
A.  Pikanowski,  and  L.  L.  Stehlik.  1996.  Fish,  megainver- 
tebrates,  and  associated  hydrographic  observations  collected 
in  the  Hudson-Raritan  Estuary,  January  1992-December 
1993,  95  p.  Northeast  Fish.  Sci.  Cent."  Ref.  Doc.  96-14, 
NMFS.  Woods  Hole,  MA. 

-  Wilber,  P.  2000.  Unpubl.  data.  Coastal  Services  Center. 
National  Ocean  Survey.  NOAA,  2234  Hobson  Avenue,  Charles- 
ton, SC,  29405. 


Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Raritan  Estuary 


695 


73°  59' 


74°|15'W 


New  York 


74°|00' 


73°]  59' 


74°|15'W 


Figure  1 

Hudson-Raritan  Estuary,  from  New  York  to  New  Jersey,  with  3-m  contour,  drawn  boundaries, 
and  the  nine  strata  of  the  Hudson-Raritan  Estuary  trawl  survey.  The  inset  shows  the  location 
of  the  estuary  on  the  United  States  coast  off  New  York  and  New  Jersey. 


and  1992  temperature  and  salinity  were  determined  by 
using  a  Niskin  bottle,  a  thermometer,  and  an  induction 
salinometer.  Beginning  in  January  1993  a  Hydrolab" 
Surveyor  III  multiprobe  was  used. 

Fish  and  large  invertebrates  were  counted,  weighed, 
and  measured  (±1.0  cm),  and  sexes  of  the  crabs  were  re- 
corded. Catch  per  unit  of  effort  or  number  per  tow  was 
used  to  estimate  relative  abundance.  Crabs  were  mea- 
sured by  carapace  width  (CW)  between  the  tips  of  the 
anterolateral  teeth.  Specimens  were  saved  for  dietary 
analysis  from  June  1991  through  June  1992.  These 
specimens  were  measured  (±1.0  mm)  and  molt  stages 
were  classified  as  intermolt  (hard-shelled),  premolt  (new 
skin  separates  easily  from  inside  the  carapace),  soft- 
shelled,  or  postmolt  (early  and  late  papershell). 

For  some  analyses,  we  separated  crabs  into  two  size 
classes  based  on  maturity  because  preferred  habitats, 
tolerances,  or  reproductive  needs  may  be  different  for 
different  life  stages.  Most  researchers  use  a  carapace 
width  at  which  250%  or  >80%  of  the  individuals  are 
mature  (produce  viable  eggs  or  sperm)  as  a  separation 
boundary.  Maturity  in  males  is  determined  by  dissec- 
tion or  by  allometric  changes  in  growth  of  appendages 
(Hartnoll,  1978;  Block  and  Rebach,  1998;  de  Lestang 
et  al.,  2003).  Most  female  blue  crabs  in  our  study  area 
and  in  Virginia  had  completed  their  pubertal  molt  and 
thus  could  reproduce  by  12  cm  CW  (Van  Engel,  1958; 


Fisher,  1999;  Stehlik,  unpubl.  data).  In  Virginia,  80% 
of  male  blue  crabs  are  mature  by  11.9  cm  (Van  Engel, 
1990).  In  lady  crabs  from  the  New  York  coast,  nearly  all 
males  are  mature  at  >6  cm,  and  females  at  about  5  cm 
(Briggs  and  Grahn3).  In  the  middle-Atlantic  portion  of 
their  range,  male  rock  crabs  mature  at  5  cm  (Haefner, 
1976)  and  some  females  <5  cm  bear  eggs  (Reilly  and 
Saila,  1978).  We  chose  the  following  CW  boundaries 
for  >80%  maturity:  blue  crabs,  >12  cm  both  sexes;  lady 
crabs,  25  cm  both  sexes;  and  rock  crabs,  males  ^5  cm 
and  females  ^4  cm. 

Data  on  blue,  lady,  and  rock  crabs  from  NEFSC 
(Northeast  Fisheries  Science  Center,  NOAA)  fall  bot- 
tom trawl  surveys  on  the  northeast  United  States  con- 
tinental shelf,  1992-94,  were  used  to  expand  the  geo- 
graphical viewpoint  of  our  study.  The  presence  of  each 
species  in  each  tow  was  plotted  to  show  distributions  in 
a  representative  year.  1992.  The  plots  were  made  with 
Surfer®  (version  6,  Golden  Software  Inc.,  Golden,  CO). 
Methods  on  the  trawl  surveys  are  described  elsewhere 
(Azarovitz,  1981). 


3  Briggs,  P.  T..  and  C.  M.  Grahn.  1996.  Aspects  of  the  fish- 
ery biology  of  the  lady  crab  tOvalipes  ocellatus)  in  New  York 
waters,  8  p.  An  in-house  paper.  New  York  State  Department 
of  Environmental  Conservation,  205  North  Belle  Mead  Road, 
Suite  1,  East  Setauket,  NY,  11733. 


696 


Fishery  Bulletin  102(4) 


Foregut  contents  were  analyzed  as  in  Stehlik  (1993). 
The  foregut  of  each  crab  was  removed  and  preserved  in 
7095  ethanol.  After  opening  the  foregut,  we  estimated 
fullness  of  the  gut  (from  0c/c  to  100%)  visually,  and  prey 
items  were  identified  to  the  lowest  possible  taxon.  The 
proportion  of  the  total  volume  of  the  foregut  contents  con- 
tributed by  each  prey  taxon  was  estimated  visually — a 
less  labor-intensive  modification  of  the  methods  of  Wil- 
liams (1981),  Hyslop  (1980),  and  Steimle  et  al.  (1994). 
The  volume  of  each  prey  taxon  was  multiplied  by  the 
percentage  of  gut  fullness.  Combining  all  foreguts,  the 
volumes  of  prey  taxa  were  listed  in  descending  order. 
The  top  12  prey  categories  on  the  list  (with  the  excep- 
tion of  "unidentified"  and  nonexclusive  categories  such  as 
Mollusca)  were  selected  for  use  in  most  of  the  subsequent 
analyses.  Foreguts  that  did  not  contain  prey  in  any  of  the 
12  categories  were  dropped  from  numerical  analyses. 

The  dietary  data  were  grouped  in  turn  by  predator 
species,  sex,  size  class,  and  collection  stratum,  and  the 
mean  percentage  volumes  of  each  of  the  12  mutually 
exclusive  prey  categories  were  calculated.  For  graphic 
representation  of  ontogenetic  differences  in  diet,  blue 
and  rock  crabs  were  grouped  for  convenience  into  20- 
mm  CW  classes,  and  lady  crabs  were  grouped  in  10-mm 
CW  classes  because  of  their  smaller  size  range.  For 
numerical  analyses,  two  maturity  classes  were  used. 
We  used  Mann-Whitney  tests  to  compare  diets  between 
sexes  within  predator  species  and  between  maturity 
stages  within  predator  species.  The  test  statistic  was  a 
chi  square  approximation. 

Group  average  cluster  analysis  was  used  to  graph  the 
separation  of  diets  by  species,  sexes,  maturity  stages, 
and  strata  by  using  the  12  prey  categories  as  dependent 
variables.  A  Bray-Curtis  similarity  matrix  was  gener- 
ated for  each  of  the  groupings,  cluster  analysis  was 
performed  by  using  Systat®  (version  10.  SPSS  Inc.,  Chi- 
cago, IL),  and  dendrograms  were  generated  by  using  the 
Bray-Curtis  values  as  distance  measures  (Romesburg. 
1984;  Marshall  and  Elliott,  1997).  A  percent  similarity 
level  was  chosen  a  posteriori  that  generated  a  reason- 
able number  of  classes. 

Analysis  of  similarity  (ANOSIM)  was  used  to  test 
for  statistical  significance  of  dietary  differences  among 
predator  species  and  for  sexes  within  species.  Analysis 
of  dissimilarity  (SIMPER)  was  used  to  determine  which 
prey  taxa  contributed  most  to  the  differences  between 
species  pairs  (Clarke  and  Warwick,  1994). 

Spatial,  temporal,  and  trophic  niche  breadth  and  over- 
lap indices  were  calculated  from  the  number  per  tow 
(1992-94)  and  diets  (June  1991-June  1992)  of  each 
crab  species  and  sex.  Temporal  niche  and  overlap  were 
calculated  by  month  for  combined  years.  Female  rock 
crabs  were  dropped  from  consideration  of  trophic  niche 
overlap  due  to  low  sample  size. 

Niche  breadth  (Colwell  and  Futuyama,  1971;  Mar- 
shall and  Elliot,  1997)  is  a  measure  of  exploitation 
within  a  particular  resource  (for  example,  substrates 
or  prey  taxa  within  an  estuary  by  a  species).  Niche 
breadth  values  are  relative  and  can  be  compared  only 
within  one  study.  The  highest  value  corresponds  to  the 


broadest  niche,  or  to  habitat  or  a  diet  generalist  rather 
than  to  a  specialist.  Niche  breadth  (S)  was  calculated 
by  the  formula  of  Colwell  and  Futuyama  (1971),  and 
modified  for  measuring  trophic  niche  breadth  according 
to  Hines  et  al.  (1990): 

B  =  1  /  £(  pk  r  from./  =  1  to  n  . 

where  phl  =  Nkj  I  Yk  ipki  is  the  proportion  of  crabs  of 
species  k  associated  with  resource  state 

j)\ 

j  =  resource  states  (months,  strata,  diet  cat- 
egories); 
n  =  number  of  resource  states; 
Nkj  =  catch  per  tow  of  species  k  at  resource  state 

j;  and 
Yk  =  catch  per  tow  of  species  k  over  all  resource 
states. 

When  trophic  niche  breadth  was  calculated, 

Nh  =  total  volume  of  diet  category  j  consumed  by  preda- 
tor k ; 

Yk  =  total  volume  of  all  diet  categories  consumed  by 
predator  k. 

Niche  overlap  is  a  measure  of  the  joint  use  of  a  resource 
by  two  species  (Colwell  and  Futuyama,  1971).  Niche 
overlap  (CAl)  between  species  h  and  i  was  calculated 
by  the  following  formula  (Colwell  and  Futuyama,  1971; 
Hines  et  al.,  1990): 

C/,,=l-0.5(Xlp,,,-;V)fi-onV  =  lto». 

where  ph/  and  p„  are  calculated  in  the  same  manner  as 
Pk,  above. 

This  index  ranges  from  0  (no  overlap)  to  1  (complete  over- 
lap) and  is  independent  of  sample  size  and  differential 
resource  availability  (Eggleston  et  al..  1998). 


Results 

Temperature  and  salinity 

Bottom  water  temperature  in  the  study  area  followed 
a  temperate  seasonal  cycle.  The  range  during  1992-94 
was  from  0  to  26.6  C.  Using  the  monthly  mean  tempera- 
ture below  or  above  10°C,  and  migration  cycles  of  the 
crabs,  we  grouped  the  months  into  two  seasons:  winter 
(November  through  April)  and  summer  (May  through 
October).  The  mean  temperature  in  the  winter  months 
1992-94  was  5.5°C,  and  that  for  summer  was  18.9°C. 
Temperature  nearest  the  estuary  mouth  was  usually  a 
few  degrees  lower  in  summer  months  and  higher  in  the 
winter  months  each  year,  compared  with  the  average 
throughout  the  estuary. 

Bottom  salinity  in  the  study  area  ranged  from  15.0  to 
33.5  ppt.  The  majority  of  stations  had  salinities  between 


Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Rantan  Estuary 


697 


Table  1 

Species,  sex,  number  collected  (n),  and  sex  ratio  (SR,  ma 

e:female)  of  a 

1  crab 

;  collecte 

d  du 

ring  the 

Hudson-Ra 

■itan  Estuary 

trawl  survey,  June  1991- 

December  1994. 

Mat 

urity  boundaries  are  expla 

ned  in 

the  text 

For  the  subsample  examined  for  stom- 

ach  contents  (June  1991- 

-June  1992),  number 

i/i ),  number 

of  non-empty 

stomachs,  and  the  mean  an 

i  range  of  carapace  width 

(CW,  mm)  are  presented. 

Crab  species, 

SR(m:f) 

SR(m:f) 

stomach 

Subsample  CW 

sex 

collected 

immature 

mature 

opened 

empty 

Mean 

(Range) 

Blue,  male 

2803 

1:1.13 

1:1.97 

167 

120 

112 

(35-185) 

Blue,  female 

4816 

272 

208 

129 

(21-169) 

Lady,  male 

14,903 

1:1.30 

1:2.12 

173 

124 

60 

(34-88) 

Lady,  female 

29.681 

255 

228 

55 

(30-89) 

Atlantic  rock,  male 

15,503 

4.65:1 

21.43:1 

400 

281 

92 

(28-130) 

Atlantic  rock,  female 

822 

19 

14 

51 

(29-80) 

Total 

68,528 

1286 

975 

25  and  30  ppt.  Salinity  decreased  with  distance  from 
the  bay  mouth  and  in  any  one  month,  the  difference 
in  salinity  between  stations  at  the  estuary  mouth  and 
those  at  the  westernmost  part  of  the  study  area  was 
approximately  5-10  ppt. 

Catch  by  species,  size,  and  sex 

From  June  1991  through  December  1994,  more  than 
68,000  blue,  lady,  and  rock  crabs  were  caught  in  1200 
otter  trawl  tows  (Table  1).  Other  mega-invertebrates 
in  the  tows  included  the  northern  moonsnail  (Euspira 
heros),  the  horseshoe  crab  (Limulus  polyphenols),  the 
American  lobster  (Homarus  americanus),  the  portly 
spider  crab  (Libinia  emarginata),  the  flatclaw  hermit 
crab  tPagurus  pollicaris),  mud  crabs  (Xanthidae),  and 
the  sea  star  (Asterias  sp. ) 

Catch  per  tow  of  crabs  by  size  class  increased  as 
they  became  large  enough  to  be  retained  by  the  mesh 
of  the  net  (Fig.  2).  Abundances  of  female  blue  and  lady 
crabs  in  the  study  area  were  greater  than  those  of  the 
males.  In  rock  crabs,  males  predominated  (Table  1).  Im- 
mature blue  and  lady  crabs  had  sex  ratios  fairly  close 
to  1:1  (male:female).  Sex  ratio  in  mature  blue  crabs, 
however,  was  1:1.97,  and  in  mature  lady  crabs,  1:2.12. 
In  all  sizes  of  rock  crabs,  sex  ratio  strongly  favored 
males,  particularly  in  mature  crabs,  in  which  the  ratio 
was  21.43:1. 

Temporal  and  spatial  variation  in  catch 

The  maximum  relative  abundance  of  blue  and  lady 
crabs  occurred  during  the  warm  months  each  year, 
whereas  rock  crabs  were  abundant  only  in  the  cold 
months  (Fig.  3).  Blue  crabs  were  scarce  in  the  otter 
trawls  from  January  through  May  or  June.  We  believe 
that  many  of  them  do  remain  in  the  study  area,  but  are 
relatively  inactive  and  are  not  accessible  to  otter  trawls, 
as  discussed  below.  Lady  crabs  migrated  into  the  estu- 
ary in  April  and  May  and  left  in  October  and  November. 


15 

10 

5 

0 

a" 

^    20 

o 

c 
a> 

cr 
a> 

r    10- 


Blue  crabs 


r*r¥P 


ml 


I 


-^r 


I 


Lady  crabs 


Male 
I  Female 


IL 


15- 

Rock  crabs 

10- 

n  n  I" 

r 

5- 

i 

0- 

■■'Hi- 

II, 

4         8         12       16       20 

Carapace  width  (cmL 

Figure  2 

Carapace  width  (cm)  frequencies  for  male  and 
female  blue,  lady,  and  rock  crabs  collected 
during  the  Hudson-Raritan  Estuary  trawl 
survey,  June  1991-December  1994,  by  percent 
frequency  of  the  total  catch  of  each  species. 


698 


Fishery  Bulletin  102(4) 


Rock  crabs  migrated  into  the  estuary  in  November  and 
gradually  left  during  April,  May,  and  June. 

Hundreds  of  soft  and  postmolt  male  rock  crabs  were 
caught  each  winter  in  the  study  area  (Fig.  4).  The  high- 
est numbers  of  molting  rock  crabs  were  collected  each 
December  and  January,  and  almost  all  of  these  crabs 
had  completed  molting  by  February.  Very  few  molting 
or  postmolt  blue  or  lady  crabs  were  caught. 

The  relative  abundances  of  the  three  species  varied 
by  stratum  (Fig.  5,  A-C;  Fig.  6).  Blue  crabs  of  both 


sexes  were  caught  mainly  in  strata  near  river  mouths 
(strata  1.  2,  and  6),  in  the  Chapel  Hill  and  Raritan 
channels  (strata  8  and  9)  in  summer,  but  mainly  in 
stratum  6  and  in  the  channels  in  winter.  Lady  crabs 
were  widely  distributed  and  were  caught  throughout 
the  study  area,  including  the  outer  strata  close  to  the 
ocean.  Male  rock  crabs  were  most  frequently  collected 
in  and  near  the  channels  and  in  strata  1  and  6,  whereas 
female  rock  crabs  were  sparsely  scattered  throughout 
the  study  area. 


6     8    10  12    2    4     6     8    10  12 
1991  1992 


4    6     8 
1993 


10  1; 


4    6     8    10  12 
1994 


6    8    10  12    2    4     6    8    10  12    2    4    6     8    10  12    2    4     6    8    10  1: 
1991  1992  1993  1994 


10  12    2    4     6     8    10  12 
1991  1992 


6    8    10  1 j 

1993 


4     6     8    10    12 
1994 


O 


Figure  3 

Catch  per  unit  of  effort  (number/towl  of  blue,  lady,  and  rock  crabs  by  month, 
graphed  with  mean  bottom  temperatures.  June  1991-December  1994. 


Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Rantan  Estuary 


699 


Foregut  fullness 

The  total  number  of  blue,  lady,  and  rock  crab  foreguts 
examined  was  1286.  Foregut  fullness  varied  by  month 
in  blue  and  rock  crabs.  The  average  fullness  of  blue 
crabs  was  1%  by  volume  from  January  through  April, 
and  34%  for  the  rest  of  the  year.  Ovigerous  blue  crabs 
(/?  =27)  averaged  40%  full.  Lady  crabs'  average  full- 
ness was  41%  during  the  months  when  they  were  pres- 
ent. The  average  fullness  of  rock  crabs  was  30%  in  all 
months  when  they  were  present;  a  minimum  occurred  in 
January  when  fullness  was  7%.  Of  419  rock  crabs  exam- 
ined, intermolt  crabs  (rc=293)  were  33%  full,  premolt 
crabs  (n  =  9)  were  empty,  soft  crabs  (n=22)  were  empty, 
and  postmolt  crabs  1/2  =  95)  were  20%  full.  Some  rock 
crabs  in  the  late  postmolt  stage  were  full  even  though 
their  chelae  were  not  completely  calcined. 

Diet  composition 

The  number  of  crabs  containing  food  was  975,  and  they 
consumed  44  identifiable  taxa  (Table  2).  Most  of  the 
mollusks  preyed  upon  were  <15  mm  in  shell  length. 
The  crabs  consumed  were  mud  crabs  (Xanthidae)  and 
juvenile  stages  of  other  Anomura  and  Brachyura.  When 
foreguts  were  only  partially  full,  well-digested  remains 
of  prey  frequently  could  be  identified  by  pieces  of  shell 
or  opercula,  mandibles  (for  shrimp),  or  chela  tips  and 
carapace  fragments  (for  crabs)  (Elner  et  al.,  1985).  Rec- 
ognizable prey  taxa  were  grouped  into  12  mutually 
exclusive  categories  (Table  3),  which  contributed  80.1% 
of  the  volume  of  all  prey.  The  prey  category  "CRABS" 
represented  pooled  fragments  of  all  crabs  except  Pagu- 


120 
100 
80 
60- 
40 
20 
0 


V 

Papershell 
■  Soft 
=  Premolt 

Hard 

8     9     10    11    12     1 


3      4      5     6 


1991 


1992 


Figure  4 

Number  of  rock  crabs  at  each  molt  stage  by  month. 
1991-92. 


Figure  5 

Catch  per  unit  of  effort  of  (Al  blue,  (B)  lady,  and  (Cl  rock  crabs,  sexes  combined 
January  1992-December  1994,  mapped  by  the  midpoint  of  each  tow. 


ridae  and  Xanthidae.  Crabs  containing  prey  in  one  or 
more  of  the  12  categories  numbered  713. 

Differences  in  diet  by  predator  species,  sex,  and  size 

Although  the  three  predator  species  shared  most  prey 
taxa,  there  were  differences  in  the  proportions  of  the 
taxa  consumed  (Fig.  7).  Mann-Whitney  tests  comparing 
diets  of  sexes  within  each  species  showed  only  two  sig- 
nificant differences  out  of  36  comparisons.  After  cluster 
analysis  upon  the  12  prey  types  by  species,  sex,  and 
size  class  (immature  and  mature),  the  resulting  den- 
drogram showed  that  diets  were  most  similar  between 
size  classes  within  a  species  (Fig.  8). 
Female  rock  crabs  were  not  included 
because  of  their  small  sample  size. 
When  the  diets  of  the  three  species 
were  compared  by  analysis  of  simi- 
larity (ANOSIM)  they  were  found 
to  be  different  (P=0.067>,  but  the 
data  were  extremely  variable  and 
not  normally  distributed.  No  signifi- 
cant differences  were  found  between 
sexes  within  species  and  we  there- 
fore pooled  sexes  within  species. 

Pairwise  comparisons  of  the  spe- 
cies were  performed  by  analysis  of 
dissimilarity  (SIMPER).  Four  taxa 
contributed  significantly  to  the  dif- 
ference in  diets  of  the  first  pair:  the 
bivalves  M.  edulis  and  M.  lateralis 
were  more  important  in  the  diets 
of  lady  crabs,  and  Xanthidae  and 
CRABS,  were  more  important  in 
the  diets  of  blue  crabs.  The  diets 
of  blue  crabs  and  rock  crabs  were 
significantly  different  in  four  taxa: 
CRABS  and  M.  lateralis  were  more 
important  for  blue  crabs,  and  M. 
edulis  and  Xanthidae  for  rock  crabs. 


700 


Fishery  Bulletin  102(4) 


Rock  crab 

Hudson-Rantan  Estuary 


New  York 


" 


qp  ° 


©  °  ,0  rf> 


None 
o       1-9 
o     10-49 
O    50-99 
O  100-249 
O    2  250 


Figure  5  (continued) 


The  diets  of  lady  crabs  and  rock  crabs  were  significantly 
different  in  two  taxa:  M.  lateralis  for  lady  crabs  and  M. 
edulis  for  rock  crabs. 

Within  the  crab  size  ranges  sampled  adequately  by 
our  gear,  we  found  some  ontogenetic  differences  in  di- 
ets (Fig.  9).  Notably,  amphipods  and  shrimp  were  con- 
sumed by  smaller  sizes  of  all  three  predators.  Certain 
mollusks,  such  as  N.  trivittatus  and  the  Atlantic  jack- 
knife  clam  (Ensis  directus),  increased  in  occurrence  in 
foreguts  with  increasing  crab  size.  Smaller  lady  crabs 
primarily  fed  upon  M.  lateralis,  but  larger  ones  broad- 
ened their  diets  to  include  other  mollusks  such  as  slip- 
persnails  (Crepidula  spp.)  and  M.  edulis.  Blue  and  rock 
crabs  exhibited  two  peaks  in  consumption  of  M.  edulis: 


the  foreguts  of  small  crabs  contained  recently  settled 
mussels,  whereas  those  of  large  crabs  contained  shell 
fragments  and  meat  of  larger  mussels.  Xanthidae  and 
Paguridae,  small  in  body  size,  were  eaten  mostly  by 
intermediate-size  predators. 

Mann-Whitney  tests  showed  that  amphipods  were 
the  only  prey  significantly  different  (P<0.01)  between 
maturity  classes  for  all  three  crab  species. 

Spatial  variability  in  diets 

Cluster  analysis  of  the  diets  by  species  and  stratum 
defined  six  groups  at  50%  similarity  (Fig.  10).  Group  A 
consisted  of  lady  and  rock  crabs  caught  at  oceanward 


Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Raritan  Estuary 


701 


Blue  crab,  winter 


E3  Males 

■  Females 


1 


12 


4      5      6      7      8      9 


Blue  crab,  summer 


fll 

12      3      4      5 


J 


6       7       8       9 


Lady  crab,  summer 


12      3      4       5       6 


l  DO 
80 
60 
40 
20 
0 


Rock  crab,  winter 


n 


q_q_ 


gj 


6- 
4 
2 
0 


Rock  crab,  summer 


^R-n   I 


2mlA 


m 


3      4 


Strata  of  Hudson-Raritan  estuary 

Figure  6 

Catch  per  unit  of  effort  for  male  and  female  crabs  by  seasons  and  strata, 
January  1992-December  1994.  Note  differences  in  scale  among  j  axes. 


100 


Blue 


Lady 


Rock 


m 

FISH 

□ 

XAN 

HI 

PAG 

Hi 

CRAB 

^ 

SHR 

■ 

AMPH 

mi 

MUL 

i 

MYT 

m 

ENS 

□ 

CREP 

g§ 

NASS 

■ 

POLY 

Figure  7 

Percent  volume  of  prey  (of  the  12  chosen  categories)  of 
blue,  lady,  and  rock  crabs  from  the  Hudson-Raritan 
Estuary,  including  all  strata  and  size  classes,  June 
1991-June  1992.  n=713.  Codes  for  prey  taxa  are  from 
Table  3. 


Rock_M_Imm 
Rock_M_Mat 

Blue_M_Imm 
Blue_F_Mat 
Blue_M_Mat 
Blue_F_Imm 
Lady_M_Mat 
Lady_F_Mat 

Lady_M_Imm 
Lady_F_Imm 


i 1 1 1 1 1 1 1 1 

100  90    80    70    60    50    40    30   20 

Percent  similarity  in  diet 

Figure  8 

Cluster  analysis  dendrogram  of  similarities  of  the 
diets  of  species,  sexes  (M=male.  F=female),  and  matu- 
rity stages  (Imm=immature,  Mat  =  mature)  of  blue, 
lady,  and  rock  crabs,  including  all  seasons  and  strata, 
June  1991-June  1992.  Female  rock  crabs  were  not 
included  because  of  the  small  sample  size. 


702 


Fishery  Bulletin  102(4) 


Table  2 

Percent  frequency  of  occurrence  (ci 

FRE)  and  percent 

volume  (%VOL 

)  of  prey  of  blue,  lady. 

and  rock  crabs  collected  du 

"ing  the 

trawl  survey,  June  1991-June  1992.  Dashes  mean  that  the  dietary 

item  was  not  found  in  any  stomachs  of  that  crab 

species. 

"Unid."  means  unidentified:  "other' 

means  uncommon 

identified  taxa 

not  listed  below. 

Blue  crab 

Lady  crat 

Rock  crab 

r£FRE 

<7rVOL 

7rFRE 

^VOL 

%FRE 

rrVOL 

Number  of  nonempty  foreguts 

328 

352 

295 

Plant  material 

1.5 

<0.1 

2.8 

0.3 

4.4 

0.1 

Hydrozoa 

0.6 

<0.1 

1.4 

<0.1 

1.4 

0.1 

Mollusca.  unid. 

6.7 

0.9 

3.4 

0.9 

12.3 

14.2 

Bivalvia,  unid.,  other 

12.2 

1.4 

11.1 

1.6 

7.8 

4.6 

Anadara  transversa 

— 

— 

0.3 

0.1 

— 

— 

Ensis  directus 

2.4 

0.8 

9.9 

4.8 

8.2 

6.2 

Lyonsia  hyalina 

— 

— 

0.3 

<0.1 

— 

— 

Mercenaria  mercenaria 

0.3 

<0.1 

0.3 

<0.1 

— 

— 

Mulinia  lateralis 

9.3 

13.6 

43.5 

33.1 

6.8 

1.4 

Mya  arenaria 

— 

— 

0.3 

0.4 

— 

— 

Mytilus  edulis 

19.6 

14.3 

13.9 

9.8 

28.7 

27.3 

Nucula  proximo 

3.4 

0.4 

5.1 

0.6 

0.7 

0.1 

Petricola  pholadiformis 

1.2 

0.1 

2.3 

0.8 

— 

— 

Pitar  morrhuanus 

— 

— 

0.3 

<0.1 

— 

— 

Spisula  solidissima 

— 

— 

1.7 

0.5 

— 

— 

Tellina  agilis 

4.6 

1.5 

9.4 

2.1 

1.4 

0.2 

Gastropoda,  unid.,  other 

6.4 

0.7 

4.0 

0.2 

1.0 

0.1 

Crepidula  fornieata,  eonvexa 

8.6 

2.5 

5.4 

2.8 

0.3 

<0.1 

Crepidula  plana 

0.6 

<0.1 

0.3 

<0.1 

— 

— 

Nassarius  obsoletus 

1.5 

1.2 

0.3 

0.2 

0.3 

<0.1 

Nassarius  trivittatus 

20.8 

6.8 

15.6 

4.6 

0.3 

<0.1 

Naticidae 

— 

— 

0.6 

<0.1 

— 

— 

Rictaxis  punctostriatus 

— 

— 

0.9 

0.1 

0.3 

<0.1 

Cephalopoda 

0.3 

0.8 

0.6 

0.3 

— 

— 

Polychaeta,  unid.,  other 

2.4 

0.2 

4.3 

0.3 

2.7 

0.9 

Glyceridae 

— 

— 

0.9 

0.3 

— 

— 

Hydroides  dianthus 

— 

— 

0.3 

<0.1 

— 

— 

Nephtyidae 

— 

— 

0.9 

0.3 

— 

— 

Nereidae 

1.5 

0.3 

1.7 

0.1 

0.7 

0.3 

Pherusa  afftnis 

— 

— 

— 

— 

0.3 

0.1 

Pectinaria  gouldii 

2.4 

0.8 

15.9 

2.4 

0.7 

<0.1 

Polynoidae 

— 

— 

0.3 

<0.1 

1.4 

0.1 

Insecta 

0.3 

<0.1 

— 

— 

— 

— 

Crustacea,  unid.,  other 

4.9 

0.6 

3.4 

0.3 

5.8 

1.4 

Amphipoda,  unid.,  other 

2.8 

0.6 

7.7 

1.4 

1.4 

0.2 

Ampelisca  sp. 

1.5 

0.6 

6.8 

1.3 

0.7 

0.4 

Corophium  sp. 

0.6 

0.2 

1.7 

0.7 

— 

— 

Gammarus  sp. 

— 

— 

3.1 

3.3 

— 

— 

Mysidacea 

— 

— 

0.3 

<0.1 

— 

— 

Caridean  shrimp,  unid..  other 

0.6 

<0.1 

2.0 

0.1 

0.7 

<0.1 

Crangon  septemspinosa 

2.8 

0.6 

6.0 

2.3 

3.1 

2.0 

Crabs  unid.,  other' 

17.7 

7.4 

8.8 

2.7 

11.3 

4.2 

Callinectes  sapid  us 

0.3 

<0.1 

0.6 

0.1 

1.4 

1.8 

Cancer  irroratus 

1.5 

1.0 

2.3 

0.5 

2.0 

1.4 

Libinia  sp. 

0.9 

0.6 

0.9 

0.6 

1.0 

0.5 

continued 

Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Rantan  Estuary 


.    703 


Table  2  (continued) 

Blue  ci 

ab 

Lady  crab 

Rock  crab 

^FRE 

%VOL 

<*FRE 

WOL 

<7tFRE 

-2VOL 

Crabs  unid.,  other1  (cont.l 

0.6 

<0.1 

2.0 

0.1 

0.7 

<0.1 

Ovalipes  ocellatus 

3.4 

3.6 

0.6 

0.2 

1.0 

1.0 

Pagurus  longicarpus 

2.1 

1.5 

1.7 

0.6 

— 

— 

Pagurus  sp. 

8.0 

4.3 

5.4 

2.3 

1.7 

0.1 

Xanthidae 

21.1 

20.8 

15.9 

10.6 

21.2 

18.4 

Fish  remains  and  scales 

2.1 

0.9 

3.4 

0.7 

6.4 

3.7 

Inorganic  debris,  sand,  mud 

0.9 

<0.1 

0.9 

0.2 

0.7 

0.1 

Shell  hash 

2.4 

1.5 

— 

— 

— 

— 

Human-made  objects 

4.0 

<0.1 

4.8 

<0.1 

1.7 

<0.1 

Unid.  organic  matter 

— 

9.2 

— 

5.3 

— 

9.0 

Mytilus  byssus 

1.8 

<0.1 

2.0 

<0.1 

3.1 

0.2 

Table  3 

Twelve  mutually  exclusive  prey  categories  that  contributed  80^  of  the  prey  volume  of  all  crabs  examined.  Codes  are  used  in 
Figures  7  and  9.  "Other"  means  uncommon  identified  taxa. 


CODE 


Category 


Identifiable  species 


NASS 

mud  snails 

CREP 

slipper  shells 

ENS 

razor  clam 

MYT 

blue  mussel 

MUL 

dwarf  surfclam 

POLY 

Polychaeta 

AMPH 

Amphipoda 

SHR 

shrimp 

CRAB 

crabs 

PAG 

hermit  crabs 

XAN 

mud  crabs 

FISH 

fish,  fish  scales 

Nassarius  trivittatus,  N.  obsoletus 

Crepidula  fornicata,  C.  convexa,  C.  plana 

Ensis  directus 

Mytilus  edulis 

Mulinia  lateralis 

all 

all 

Crangon  septemspinosa,  unid.,  other 

Libinia  sp..  Cancer  irroratus.  Ovalipes  ocellatus,  Callinectes  sapidus,  crab  unid. 
and  others  excluding  Paguridae  or  Xanthidae 

Pagurus  acadianus,  P.  longicarpus,  unid.,  other 

Xanthidae:  Dyspanopeus  sayi,  unid.,  other 

all 


outer  strata  (4,  5,  and  7)  that  consumed  large  quanti- 
ties of  M.  edulis.  Clumps  of  recently  settled  and  larger 
mussels  were  frequently  collected  in  trawl  nets  in  these 
strata.  Group  B  contained  crabs  from  Gravesend  Bay, 
(stratum  6 )  that  ate  primarily  M.  edulis  and  M.  lateralis. 
Group  C  contained  crabs  caught  in  the  siltier  southern 
strata  and  nearby  channel  (strata  1,  2,  and  9)  that  con- 
sumed mainly  M.  lateralis,  M.  edulis,  and  CRABS.  Group 
D  consisted  of  rock  crabs  collected  at  inner  strata  (2, 
3,  and  8)  that  fed  primarily  upon  E.  directus  and  Xan- 
thidae. Ensis  directus  was  most  common  in  diets  in  the 
northern  sandier  strata  (strata  3,  5,  6,  and  7).  Groups 
E  and  F  consisted  of  lady  and  rock  crabs  that  consumed 


mainly  M.  lateralis.  Four  species-stratum  combinations 
did  not  cluster  with  any  groups. 

Temporal,  spatial,  and  trophic  niche  breadth  and  overlap 

Niche  breadth  and  overlap  were  calculated  for  both  sexes 
of  the  three  crab  species  (Table  4).  Lady  crabs  of  both 
sexes  had  the  narrowest  temporal  niches  (3.896  and 
4.592),  reflecting  their  presence  in  the  estuary  strictly 
in  warm  months.  The  temporal  niche  breadth  of  female 
blue  crabs  (8.187)  was  greatest,  reflecting  their  year-long 
presence  in  the  study  area,  even  in  the  cold  months  when 
many  males  remain  in  rivers.  The  temporal  overlaps  of 


704 


Fishery  Bulletin  102(4) 


Blue    10° 
crab 


Lady 
crab 


Rock 
crab 


Percent  volume 


40-59         60-79         80-99       100-119     120-139     140-159     160-185 
100 


tfiQfiS 
tssss 

FISH 

□ 

XAN 

m 

PAG 

ggg 

CRAB 

^ 

SHR 

■ 

AMPH 

on 

MUL 

□ 

MYT 

ESS 

ENS 

□ 

CREP 

B 

NASS 

■ 

POLY 

IffiS 

FISH 

□ 

XAN 

m 

PAG 

\g% 

CRAB 

m 

SHR 

■ 

AMPH 

m 

MUL 

□ 

MYT 

ss 

ENS 

□ 

CREP 

HI 

NASS 

■ 

POLY 

100 


£3 

FISH 

□ 

XAN 

m 

PAG 

gg 

CRAB 

m 

SHR 

■ 

AMPH 

he 

MUL 

□ 

MYT 

us 

ENS 

□ 

CREP 

n 

NASS 

■ 

POLY 

40-59 


60-79 


80-99 


100-119 


120-139 


Carapace  width  classes  (mm) 


Figure  9 

Percent  volume  of  prey  I  of  the  12  categories  I  by  10-  or  20-mm  size  classes 
of  blue,  lady,  and  rock  crabs,  all  seasons  and  strata.  Codes  for  prey  taxa 
are  from  Table  3. 


male  and  female  lady  crabs  with  male  rock  crabs  were 
the  lowest  in  the  matrix  (0.149  and  0.186). 

The  spatial  niche  breadths  of  lady  crabs  were  larg- 
est (7.320  and  7.324)  (a  result  of  their  nonaggrega- 
tive  distribution  throughout  the  study  area),  whereas 
the  other  two  species  tended  to  aggregate  in  certain 


locations,  particularly  in  or  near  channels.  Female 
rock  crabs  also  had  a  broad  spatial  niche,  although 
they  were  caught  much  less  frequently  than  the  other 
groups.  Spatial  overlap  was  highest  within  species, 
particularly  between  male  and  female  lady  crabs 
(0.908). 


Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Raritan  Estuary 


705 


Predator/ 
stratum 


Lady 
Lady 
Rock 
Rock 
Rock 
Rock 
Blue 
Lady 
Blue 
Lady 
Blue 
Lady 
Blue 
Blue 
Rock 
Blue 
Rock 
Rock 
Rock 
Lady 
Lady 
Rock 
Lady 
Lady 


I 1 1 1 1 1 1 1 1 1 

100  90   80   70    60   50   40   30    20    10 

Percent  similarity  in  diet 

Figure  10 

Cluster  analysis  dendrogram  of  the  diets  of  blue, 
lady,  and  rock  crabs  from  the  nine  strata  of  the 
Hudson-Raritan  Estuary.  The  vertical  line  at  50% 
similarity  defines  groups  A-F. 


42.00 


40  00  - 


38.00 


36.00  - 


Trophic  niche  breadth  was  greatest  in  male  and  fe- 
male blue  crabs  (5.234  and  6.563)  and  male  lady  crabs 
(6.166)  (Table  4).  It  was  narrowest  for  female  rock  crabs, 
but  sample  size  was  low.  Overlap  was  highest  within 
species:  blue  crab  males  and  females  (0.819),  and  lady 
crab  males  and  females  (0.861).  Overlap  was  lowest 
between  lady  and  rock  crabs,  sexes  combined  (0.427  i. 


Discussion 

Temporal  and  spatial  overlap  within  the  estuary 

The  scatter  plots  (Fig.  5)  and  spatial  niche  overlap 
indices  indicate  substantial  likelihood  of  co-occurrence 
and  encounter  among  blue,  lady,  and  rock  crabs  in  the 
Hudson-Raritan  Estuary.  However,  the  species  were  not 
all  active  in  the  study  area  at  the  same  time.  Seasonal 
migration  and  winter  torpor  are  two  mechanisms  that, 
at  times,  prevent  interspecies  encounters.  Rock  crabs  had 
low  temporal  overlaps  with  blue  and  lady  crabs  because 
when  rock  crabs  migrate  in  from  the  coastal  ocean,  lady 
crabs  migrate  out  and  blue  crabs  become  less  active  and 
sometimes  bury  themselves.  Although  otter  trawling  does 
not  adequately  sample  buried  blue  crabs,  commercial  crab 
dredgers  catch  large  numbers  of  overwintering  blue  crabs 
from  December  through  March  in  and  near  the  Raritan 
and  Chapel  Hill  channels  (Stehlik  et  al.,  1998). 

Temporal  overlap  between  blue  crabs  and  lady  crabs 
was  fairly  high  because  of  their  co-occurrence  in  the 


■fr 

it 

Z^r^*-^" 

A 

Qa"" 

'\  O 

-76.00 


-74.00 


-72.00 


-70.00 


-68.00 


-66.00 


Figure  11 

Presence  of  blue,  lady,  and  rock  crabs  at  stations  from  the 
fall  1992  bottom  trawl  survey,  (Northeast  Fisheries  Science 
Center,  Woods  Hole,  MAl.  Each  point  represents  presence  at 
a  station.  Occurrence  inside  the  estuaries! boxed  symbols) 
was  derived  from  the  literature  cited  in  this  article. 


warm  months.  It  was  expected  that  intra-estuarine  spa- 
tial separation  might  minimize  contact  between  these 
species  because  they  are  reported  to  prefer  different 
substrates.  The  blue  crab  is  known  to  occupy  a  variety 
of  substrate  types,  including  sand,  mud,  and  submerged 
vegetation  (Milliken  and  Williams,  1984;  Wilson  et  al. 
1990),  whereas  the  lady  crab  is  primarily  collected  on 
sand  (Williams  and  Wigley,  1977).  The  lady  crab  bur- 
ies itself  in  sand  more  readily  than  in  mud  (Barshaw 
and  Able,  1990)  and  it  is  able  to  forage  more  efficiently 
in  sand  than  in  sand-gravel  or  sand-shell  substrates 
(Sponaugle  and  Lawton,  1990).  However,  as  shown  in 
Figures  5  and  6,  lady  crabs  were  not  confined  to  sandy 
strata  but  were  most  abundant  on  the  fine-grained  sedi- 
ment strata  1,  2,  and  9. 

The  pattern  of  seasonal  estuarine  use  by  blue  and 
lady  crabs  is  not  unique  to  the  Hudson-Raritan  Estuary. 
Other  estuaries  in  which  the  two  Portunidae  are  abun- 
dant in  summer  months  but  uncommon  in  winter  are 
Barnegat  Bay,  NJ  (Milstein  et  al.,  1977;  pers.  observ. ), 
Delaware  Bay  (Winget  et  al.,  1974),  and  Chesapeake 
Bay  (Haefner  and  Van  Engel,  1975). 

Rock  crabs  undergo  seasonal  migrations  from  coastal 
waters  into  and  out  of  estuaries,  but  the  timing  differs 
by  latitude.  In  Canada,  the  Gulf  of  Maine,  and  northern 
Massachusetts,  rock  crabs  are  much  more  abundant  in 
immediate  coastal  waters,  estuaries,  and  in  the  inter- 
tidal  zone  in  warmer  months  (Krouse,  1972;  Scarratt 
and  Lowe,  1972).  Rock  crabs  are  more  numerous  in 
Narragansett  Bay,  Rhode  Island,  in  warmer  months 
(Jeffries,  1966;  Clancy4).  Juveniles  are  present  inside 


Clancy,  M.  2002.  Personal  commun.  Boston  University, 
College  of  General  Studies,  Division  of  Natural  Science, 
Boston,  MA,  02115. 


706 


Fishery  Bulletin  102(4) 


Table  4 

Niche  breadth  and  overlap  for  temporal,  spat 

al,  and  ti 

-ophic  dimensions  among  blue, 

lady,  and  rock  crabs. 

For  temporal  and 

spatial  niches,  all  crabs  (of  all  sizes)  collected 

in  1992- 

94  are  included.  For  trophic  analyses,  only  the  crabs 

containing 

one  or 

more  of  the  12  prey  categories  were  included.  Female  rock  crabs  were 
size. 

not  included  in  trophic  overlap 

because  of  the  small  sample 

Number 

Mean 

Niche 

of  crabs 

CW,  mm 

breadth 

Overlap 

matrices 

Temporal  niche 

{n  =12  months,  1992-94) 

BCF 

LCM 

LCF 

RCM 

RCF 

Blue  crab  male  (BCM) 

2191 

125 

6.104 

BCM 

0.854 

0.604 

0.618 

0.297 

0.463 

Blue  crab  female  (BCF) 

3483 

129 

8.187 

BCF 

0.649 

0.576 

0.376 

0.564 

Lady  crab  male  (LCM) 

11883 

62 

4.592 

LCM 

0.894 

0.186 

0.417 

Lady  crab  female  (LCF) 

25312 

61 

3.896 

LCF 

0.149 

0.339 

Rock  crab  male  (RCM) 

14530 

85 

6.764 

RCM 

0.479 

Rock  crab  female  (RCF) 

778 

51 

5.782 

Spatial  niche 

(ra=9  strata,  1992-94) 

BCF 

LCM 

LCF 

RCM 

RCF 

Blue  crab  male 

2191 

125 

3.927 

BCM 

0.757 

0.720 

0.703 

0.675 

0.474 

Blue  crab  female 

3483 

129 

5.343 

BCF 

0.679 

0.685 

0.746 

0.641 

Lady  crab  male 

11883 

62 

7.324 

LCM 

0.908 

0.677 

0.616 

Lady  crab  female 

25312 

61 

7.320 

LCF 

0.678 

0.638 

Rock  crab  male 

14530 

85 

5.447 

RCM 

0.763 

Rock  crab  female 

778 

51 

7.191 

Trophic  niche 

in=12  prey  categories,  1991-92) 

BCF 

LCM 

LCF 

RCM 

Blue  crab  male 

84 

111.0 

5.234 

BCM 

0.819 

0.570 

0.580 

0.576 

Blue  crab  female 

139 

127.9 

6.563 

BCF 

0.629 

0.651 

0.609 

Lady  crab  male 

98 

59.4 

6.166 

LCM 

0.861 

0.437 

Lady  crab  female 

200 

54.9 

4.655 

LCF 

0.417 

Rock  crab  male 

181 

88.8 

4.139 

Rock  crab  female 

11 

53.7 

2.620 

Trophic  niche,  sexes  combined 

Ui  =  12  prey  categories,  1991-92) 

LC 

RC 

Blue  crab  (BO 

223 

121.6 

6.250 

BC 

0.628 

0.623 

Lady  crab  ( LC ) 

298 

56.4 

5.140 

LC 

0.427 

Rock  crab  ( RC ) 

192 

86.8 

4.008 

that  bay  all  year  (Reilly  and  Saila,  1978).  In  contrast, 
in  Delaware  Bay  and  Chesapeake  Bay  they  occur  in 
coastal  waters  and  estuaries  mainly  in  colder  months 
(Winget  et  al.,  1974;  Haefner  and  Van  Engel,  1975; 
Haefner,  1976).  Our  data  showed  that  rock  crabs  in 
the  Hudson-Raritan  Estuary  conform  to  the  pattern  of 
migration  typical  of  the  latter  southern  bays. 

A  crossroads  or  overlap  in  distribution  of  the  three 
crab  species  is  more  evident  when  a  broader  area  on  the 
continental  shelf  from  Cape  Cod  to  Cape  Hatteras  is  con- 
sidered. Crab  presence  was  plotted  by  using  data  from 
the  fall  1992  continental  shelf  trawl  survey  <Fig.  11). 
Fall  surveys  are  done  in  September  and  October  when 
waters  are  still  warm.  In  the  coastal  waters  off  Rari- 
tan,  Delaware,  and  Chesapeake  Bays,  blue  and  lady 
crabs  were  collected,  whereas  rock  crabs  were  collected 
mainly  on  the  central  shelf.  Estuarine  presence  in  warm 
months,  compiled  from  citations  in  the  present  study,  is 
marked  by  symbols. 


Sex  ratios 

In  the  Hudson-Raritan  Estuary,  sex  ratios  of  blue,  lady, 
and  rock  crabs  were  different  from  1:1.  In  mature  blue 
crabs,  the  sex  ratio  favored  females  because  the  study 
area  is  in  the  deeper  oceanward  portion  of  the  estuarine 
system,  where  females  release  their  eggs  and  overwinter. 
Many  males  spend  their  entire  lives  in  water  of  relatively 
low  salinity  (Van  Engel,  1958),  such  as  is  found  in  the 
nearby  Hudson,  Raritan,  and  Navesink-Shrewsbury 
rivers.  In  the  Navesink  River,  the  sex  ratio  of  male  to 
female  blue  crabs  ;>12  cm  over  a  two-year  period  was 
2.6:1  (Meise  and  Stehlik,  2003). 

In  the  Hudson-Raritan  Estuary,  female  lady  crabs 
a5  cm  outnumbered  males  2:1.  Many  of  these  females 
were  ovigerous  and  therefore  estuarine  use  may  be 
related  to  reproduction.  We  were  unable  to  locate  pub- 
lished reports  of  lady  crabs  or  other  Ovalipes  spp.  mat- 
ing locations,  single-sex  migrations,  or  locations  of  lar- 


Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Raritan  Estuary 


707 


val  release,  any  of  which  might  be  a  reason  for  the  use 
of  the  estuaries  by  female  lady  crabs. 

The  rock  crabs  that  enter  the  estuary  were  predomi- 
nantly males,  and  many  females  may  never  enter  the 
estuary.  Males  use  the  estuary  to  molt,  and  possibly  to 
avoid  predators  offshore.  In  comparison,  on  the  north- 
west Atlantic  continental  shelf,  the  sex  ratio  in  winter 
dredge  collections  was  1:2.2  males:females  (Stehlik  et 
al.,  1991). 

Feeding  periodicity 

Food  consumption  in  crabs  is  affected  by  daily  and 
seasonal  cycles,  temperature  changes,  reproductive 
rhythms,  and  molt  (Warner,  1977;  Stevens  et  al,  1982; 
Ryer,  1987;  Mantelatto,  2001).  In  our  study  area,  blue 
crabs  ate  little  when  inactive  during  the  winter  months, 
as  reported  above.  Choy  (1986)  reported  less  feeding 
during  egg-brooding  in  Portunidae,  but  in  our  study  we 
found  that  fullness  was  about  40%  in  both  egg-bearing 
and  non-egg-bearing  females  in  summer.  A  lack  of  feed- 
ing before  and  during  molt,  until  calcification  has  suf- 
ficiently progressed,  is  typical  of  crabs  (Warner,  1977). 
Empty  stomachs  in  premolt  and  soft  rock  crabs  in  our 
study  supported  this  observation. 

Diet  composition 

We  found  that  in  the  Hudson-Raritan  Estuary,  the  most 
important  prey  items  of  blue  crabs  by  volume  were 
Xanthidae,  then  the  mollusks  M.  edulis  and  M.  late- 
ralis, whereas  only  2%  of  the  prey  volume  was  from 
cannibalism.  In  contrast,  small  blue  crabs  are  of  major 
importance  in  the  diets  of  large  blue  crabs  in  Florida 
(Laughlin,  1982)  and  Maryland  (Hines  et  al.,  1990). 
and  cannibalism  is  the  source  of  more  than  75%  of  the 
mortality  of  juveniles  near  estuarine  shores  (Hines  and 
Ruiz,  1995).  The  major  targets  of  cannibalism,  early 
instars  or  molting  juveniles,  may  be  more  abundant  in 
rivers  adjacent  to  our  study  area  (Meise  and  Stehlik. 
2003). 

The  diets  of  rock  crabs  in  estuarine  and  coastal 
Canada  and  Maine  usually  contained  a  larger  num- 
ber of  prey  categories  than  did  the  diets  in  the  pres- 
ent study  (Scarratt  and  Lowe,  1972;  Drummond-Davis 
et  al..  1982;  Hudon  and  Lamarche,  1989;  Ojeda  and 
Dearborn,  1991).  These  northern  studies  were  done  on 
rock,  boulders,  cobble,  sand,  and  algal  beds,  where  the 
diversity  of  habitats  within  a  study  area  may  offer  a 
larger  assortment  of  potential  prey  than  the  soft-bottom 
habitat  of  our  estuary. 

In  the  Hudson-Raritan  Estuary,  juveniles  of  commer- 
cially or  recreationally  harvested  species  were  rarely 
consumed  by  the  three  species  of  crabs.  Among  mol- 
lusks, M.  arenaria  and  M.  mercenaria  were  scarce  in 
crab  stomachs,  perhaps  because  other  taxa  such  as  M. 
lateralis,  N.  trivittatus,  and  Xanthidae  provided  abun- 
dant prey.  The  other  commercially  important  species 
eaten  by  crabs  was  the  blue  crab  juvenile,  but  infre- 
quently as  mentioned  above. 


Differences  in  diet  among  species,  sexes, 
and  size  classes  of  predators 

Our  data  did  not  support  our  hypotheses,  based  on  exist- 
ing studies,  that  blue,  lady,  and  rock  crabs  would  have 
different  diets  as  a  consequence  of  their  species-specific 
body  and  chela  structures.  Blue  and  lady  crabs  (unlike 
rock  crabs)  swim,  allowing  them  a  greater  foraging 
area  than  rock  crabs.  Chela  structure  affects  the  type 
and  size  of  prey  that  can  be  crushed  (Vermeij,  1978; 
Seed  and  Hughes,  1995;  Behrens  Yamada  and  Bould- 
ing,  1998).  In  Portunidae.  the  long  chelae  (in  relation 
to  their  CW)  have  short  muscle  fibers  better  suited  to 
quick  grabbing  than  to  prolonged  crushing  (Warner 
and  Jones,  1976;  Seed  and  Hughes,  1997).  The  chelae 
of  Cancridae  are  monomorphic  (same  characteristics 
left  and  right  sides),  have  relatively  short,  stout  teeth, 
and  close  relatively  slowly  because  of  their  muscle  fibers 
(Warner  and  Jones,  1976).  Chela  crushing  force  (New- 
tons),  measured  with  a  force  transducer,  is  positively 
correlated  with  chela  height  and  thickness  (Govind 
and  Blundon,  1985;  Block  and  Rebach,  1998).  Although 
the  chela  structures  of  blue  and  rock  crabs  are  quite 
different,  the  chelae  of  mature  rock  crabs  (9-13.5  cm 
CW)  generate  crushing  forces  comparable  to  those  of 
cutter  and  crusher  chelae  of  mature  male  blue  crabs 
(12-16  cm)  (Govind  and  Blundon,  1985). 

Chela  crushing  force  in  mature  blue  and  rock  crabs  is 
likely  to  be  more  than  sufficient  for  successful  foraging 
upon  all  but  the  largest  prey  (Block  and  Rebach,  1998) 
and  may  not  be  a  major  determinant  of  diet.  In  fact, 
crabs  often  prey  upon  small  or  young  bivalves  rather 
than  on  large  sizes,  perhaps  because  the  latter  require 
more  handling  time  and  may  damage  chelae  (Juanes. 
1992;  Seed  and  Hughes,  19951.  Because  Portunidae 
swim  and  have  more  versatile  chelae,  they  may  be  ex- 
pected to  have  broader  trophic  niches  than  Cancridae. 
In  our  study,  blue  crabs  had  the  broadest  trophic  niche, 
lady  crabs  had  an  intermediate  trophic  niche,  and  rock 
crabs  had  the  narrowest  trophic  niche. 

We  found  no  significant  differences  in  diet  by  sex 
within  species.  Sexual  dimorphism  within  a  crab  spe- 
cies accelerates  after  puberty  (Hartnoll.  1978),  but  our 
study  included  many  immature  crabs.  Some  experiment- 
ers using  force  transducers  found  no  significant  differ- 
ence in  crushing  force  between  the  sexes  of  blue  crabs 
of  a  broad  size  range  (Blundon  and  Kennedy,  1982; 
Seed  and  Hughes,  1997),  but  in  blue  crabs  >135  mm, 
males  produced  significantly  more  force  than  females 
(Eggleston,  1990).  Sexual  dimorphism  is  found  in  chela 
length,  but  not  chela  height,  in  lady  crabs  (significantly- 
different  slopes  of  CL/CW  by  regression;  Stehlik,  un- 
publ.  data). 

Carapace  width  and  the  proportion  of  chela  height  to 
carapace  width  are  positively  correlated  with  crushing 
force,  which  makes  it  possible  for  larger  crabs  to  con- 
sume larger,  harder-shelled  mollusks  or  crustaceans 
(Hartnoll,  1978;  Block  and  Rebach,  1998).  The  larg- 
est lady  crabs  do  not  grow  to  the  carapace  widths  or 
chela  lengths  of  mature  blue  crabs;  therefore  the  force 


708 


Fishery  Bulletin  102(4) 


of  their  chelae  cannot  match  those  of  blue  crabs.  As 
they  grow,  Cancridae  and  Portunidae  undergo  shifts  in 
diet,  and  may  be  divided  into  ontogenetically  distinct 
trophic  units  (Laughlin,  1982;  Stevens  et  al.,  1982: 
Stoner  and  Buchanan,  1990;  Rosas  et  al..  1994).  In 
our  study,  larger  crabs  dropped  amphipods  and  shrimp 
from  their  diets,  but  otherwise  only  minor  changes 
occurred  in  prey  identity  and  relative  volumes  of  prey 
taxa  among  size  classes  (Fig.  9).  An  interesting  ontoge- 
netic shift  was  in  the  size  of  prey  eaten:  small  crabs  ate 
small  individuals  of  prey  taxa,  such  as  M.  edulis,  and 
Xanthidae,  and  large  crabs  ate  large  individuals  of  the 
same  taxa.  Thus  in  our  study  the  influence  of  physical 
structure  upon  diet  was  greater  as  body  size  increased 
within  a  species  than  among  species. 

Spatial  variability  and  overlap  in  diets 

The  three  predators  were  scattered  throughout  the 
cluster  diagram  of  diet  among  strata  of  the  estuary 
(Fig.  10),  yet  crabs  from  inner  and  outer  groups  of  strata 
usually  clustered  separately.  We  concluded  that  loca- 
tion influenced  diet  more  than  did  predator  identity. 
The  inner,  outer,  and  channel  strata  differ  in  depth, 
sediment  type,  currents,  and  mean  temperature,  and 
therefore  in  benthic  and  epibenthic  prey  assemblages. 
Our  results  support  the  concept  that  these  species  are 
mainly  opportunistic  in  diet,  as  was  suggested  for  blue 
crabs  (Laughlin,  19821,  and  rock  crabs  (Hudon  and 
Lamarche,  1989).  The  Hudson-Raritan  and  other  nearby 
coastal  and  estuarine  areas  from  Long  Island  Sound  to 
Chesapeake  Bay  are  crossroads  where  blue,  lady,  and 
rock  crabs  share  space  and  resources. 


Acknowledgments 

We  thank  those  who  helped  design  and  carry  out  the 
Hudson-Raritan  Estuary  trawl  surveys,  especially  Stuart 
Wilk,  Anthony  Pacheco,  and  Eileen  MacHaffie.  We  also 
thank  Fred  Farwell,  Sherman  Kingsley,  and  the  NOAA 
Corps  captains  and  crew.  Suellen  Fromm  was  instru- 
mental in  obtaining  data  from  NEFSC  trawl  surveys. 
We  thank  the  scientists  who  shared  their  opinions  and 
unpublished  data.  We  are  indebted  to  colleagues  Mary 
Fabrizio,  Clyde  MacKenzie,  John  Manderson,  Carol 
Meise,  Frank  Steimle,  Allan  Stoner,  and  anonymous 
reviewers  who  helped  improve  the  manuscript.  This 
paper  is  dedicated  to  the  memory  of  Tony  Pacheco. 


Literature  cited 

Auster,  P.  J.,  and  R.  E.  DeGoursey. 

1994.     Winter  predation  on  blue  crabs,  Callinectes  sapi- 
dus, by  starfish  Asterias  forbesi.     J.  Shellfish  Res. 
13:361-366. 
Azarovitz,  T.  A. 

1981.     A  brief  historical  review  of  the  Woods  Hole  Labora- 
tory trawl  survey  time  series.     In  Bottom  trawl  surveys 


(Doubleday.  W.G.,  and  D.  Rivard,  eds.),  p.  62-67.     Can. 
Spec.  Publ.  Fish.  Aquat.  Sci.  58. 
Barshaw,  D.  E„  and  K.  W.  Able. 

1990.     Deep  burial  as  a  refuge  for  lady  crabs.  Ocalipes 
ocellatus:  comparisons  with  blue  crabs,  Callinectes 
sapidus.     Mar.  Ecol.  Prog.  Ser.  66:75-79. 
Behrens  Yamada,  S.,  and  E.  G.  Boulding. 

1998.     Claw  morphology,  prey  size  selection  and  foraging 
efficiency  in  generalist  and  specialist  shell-breaking 
crabs,     j.  Exp.  Mar.  Biol.  Ecol.  220:191-211. 
Birchard,  G.  F.,  L.  Drolet.  and  L.  H.  Mantel. 

1982.     The  effect  of  reduced  salinity  on  osmoregulation 
and  oxygen  consumption  in  the  lady  crab.  Ocalipes  ocella- 
tus (Herbst).     Comp.  Biochem.  Physiol.  71A:321-324. 
Block.  J.  D.,  and  S.  Rebach. 

1998.     Correlates  of  claw  strength  in  the  rock  crab. 
Cancer  irroratus  (Decapoda,  Brachyura).     Crustaceana 
71:468-473. 
Blundon.  J.  A.,  and  V.  S.  Kennedy. 

1982.     Mechanical  and  behavioral  aspects  of  blue  crab, 
Callinectes  sapidus  (Rathbuni,  predation  on  Chesapeake 
Bay  bivalves.     J.  Exp.  Mar.  Biol.  Ecol.  65:47-65. 
Briggs,  P.  T. 

1998.     New  York's  blue  crab  I  Callinectes  sapidus)  fisheries 
through  the  years.     J.  Shellfish  Res.  17:487-491. 
Choy,  S.  C. 

1986.     Natural  diet  and  feeding  habits  of  the  crabs  Lio- 
carcinus  puber  and  L.  holsatus  I  Decapoda.  Brachyura. 
Portunidae).     Mar.  Ecol.  Prog.  Ser.  31:87-99. 
Clarke,  K.  R..  and  R.  M.  Warwick. 

1994.     Change  in  marine  communities:  an  approach  to 
statistical  analysis  and  interpretation,  144  p.     National 
Environmental  Research  Council,  UK. 
Coch.  N.  K. 

1986.  Sediment  characteristics  and  facies  distributions 
in  the  Hudson  system.  In  Sedimentation  in  the  Hudson 
system:  the  Hudson  River  and  contiguous  waterways  iN. 
K.Coch  and  H.  J.  Bokuniewicz,  eds.  I,  p.  109-129.  North- 
eastern Geology  (spec,  issue)  8. 
Colwell,  R.  K.,  and  D.  J.  Futuyama. 

1971.     On  the  measurement  of  niche  breadth  and  overlap. 
Ecology  52:567-576. 
de  Lestang,  S.,  N.  G.  Hall,  and  I.  C.  Potter. 

2003.     Reproductive  biology  of  the  blue  swimmer  crab 
{Portunus  pelagicus,  Decapoda:  Portunidae)  in  five 
bodies  of  water  on  the  west  coast  of  Australia.     Fish. 
Bull.  101:745-757. 
Drummond-Davis,  N.  C,  K.  H.  Mann,  and  R.  A.  Pottle. 

1982.     Some  estimates  of  population  density  and  feeding 
habits  of  the  rock  crab.  Cancer  irroratus,  in  a  kelp  bed  in 
Nova  Scotia.     Can.  J.  Fish.  Aquat.  Sci.  39:636-639. 
Eggleston.  D.  B. 

1990.     Functional  responses  of  blue  crabs  Callinectes 
sapidus  Rathbun  feeding  on  juvenile  oysters  Crassostrea 
virginica  (Gmelin):  effects  of  predator  sex  and  size,  and 
prey  size.     J.  Exp.  Mar.  Biol.  Ecol.  143:73-90. 
Eggleston,  D.  B..  J.  J.  Grover,  and  R.  N.  Lipcius. 

1998.  Ontogenetic  diet  shifts  in  Nassau  grouper:  tro- 
phic linkages  and  predatory  impact.  Bull.  Mar.  Sci. 
63:111-126. 
Elner,  R.  W.,  P.  G.  Beninger.  L.  E.  Linkletter.  and  S.  Lanteigne. 
1985.  Guide  to  indicator  fragments  of  principal  prey  taxa 
in  the  stomachs  of  two  common  Atlantic  crab  species: 
Cancer  borealis  Stimpson,  1859  and  Cancer  irroratus  Say, 
1817.     Can.  Tech.  Rep.  Fish.  Aquat.  Sci.  1403:1-20. 


Stehlik  et  al.:  Distribution  patterns  of  various  crab  species  in  the  Hudson-Raritan  Estuary 


709 


Fisher.  M.  R. 

1999.     Effect  of  temperature  and  salinity  on  size  at 
maturity  of  female  blue  crabs.     Trans.  Am.  Fish.  Soc. 
128:499-506. 
Govind,  C.  K.,  and  J.  A.  Blundon. 

1985.     Form  and  function  of  the  asymmetric  chelae  in 
blue  crabs  with  normal  and  reversed  handedness.     Biol. 
Bull.  168:221-231. 
Guerin.  J.  L.,  and  W.  B.  Stickle. 

1992.     Effect  of  salinity  gradients  on  the  tolerance  and 
bioenergetics  of  juvenile  blue  crabs  {Callinectes  sapidus) 
from  waters  of  different  environmental  salinities.     Mar. 
Biol.  114:391-396. 
Haefner.  P.  A.,  Jr. 

1976.     Distribution,  reproduction  and  moulting  of  the  rock 
crab,  Cancer  irroratus  Say  1917,  in  the  mid-Atlantic 
Bight.     J.  Nat.  Hist.  10:377-397. 
Haefner,  P.  A.,  Jr.,  and  W.  A.  Van  Engel. 

1975.     Aspects  of  molting,  growth  and  survival  of  male 
rock  crabs.  Cancer  irroratus,  in  Chesapeake  Bay.     Ches- 
apeake Sci.  16:253-265. 
Hartnoll,  R.  G. 

1978.  The  determination  of  relative  growth  in  Crus- 
tacea.    Crustaceana  34:281-293. 

Hines,  A.  H.,  R.  N.  Lipcius,  and  A.  M.  Haddon. 

1987.     Population  dynamics  and  habitat  partitioning 
by  sex,  size,  and  molt  stage  of  blue  crabs  Callinectes 
sapidus  in  a  subestuary  of  Chesapeake  Bay.     Mar.  Ecol. 
Prog.  Ser.  36:55-64. 
Hines,  A.  H.,  A.  M.  Haddon,  and  L.  A.  Wiechert. 

1990.     Guild  structure  and  foraging  impact  of  blue  crabs 
and  epibenthic  fish  in  a  subestuary  of  Chesapeake 
Bay.     Mar.  Ecol.  Prog.  Ser.  67:105-126. 
Hines,  A.  H.,  and  G.  M.  Ruiz. 

1995.     Temporal  variation  in  juvenile  blue  crab  mortal- 
ity: nearshore  shallows  and  cannibalism  in  Chesapeake 
Bay.     Bull.  Mar.  Sci.  57:884-901. 
Hudon,  C,  and  G.  Lamarche. 

1989.     Niche  segregation  between  American  lobster  Homa- 
rus  americanus  and  rock  crab  Cancer  irro?'atus.     Mar. 
Ecol.  Prog.  Ser.  52:155-168. 
Hyslop,  E.  J. 

1980.     Stomach  contents  analysis — a  review  of  methods 
and  their  application.     J.  Fish.  Biol.  17:411-429. 
Jeffries,  H.  P. 

1966.     Partitioning  of  the  estuarine  environment  by  two 
species  of  Cancer.     Ecology  47:477-481. 
Jones,  C.  R.,  C.  T.  Fray,  and  J.  R.  Schubel. 

1979.  Textural  properties  of  surficial  sediments  of  Lower 
Bay  of  New  York  Harbor.  Special  Report  21,  113  p.  Ma- 
rine Sciences  Research  Center,  State  University  of  New 
York,  Stony  Brook,  NY. 

Juanes,  F. 

1992.     Why  do  decapod  crustaceans  prefer  small-sized 
molluscan  prey?     Mar.  Ecol.  Prog.  Ser.  87:239-249. 
Krouse,  J.  S. 

1972.     Some  life  history  aspects  of  the  rock  crab,  Cancer 
irroratus,  in  the  Gulf  of  Maine,     J.  Fish.  Res.  Board 
Can.  29:1479-1482. 
Laughlin,  R.  A. 

1982.     Feeding  habits  of  the  blue  crab,  Callinectes  sapidus 
Rathbun,  in  the  Appalachicola  estuary,  Florida.     Bull. 
Mar.  Sci.  32:807-822. 
Leffler,  C.  W. 

1972.     Some  effects  of  temperature  on  the  growth  and 


metabolic  rate  of  juvenile  blue  crabs,  Callinectes  sapidus, 
in  the  laboratory.     Mar.  Biol.  14:104-110. 
MacKenzie,  C.  L.,  Jr. 

1990.  History  of  the  fisheries  of  Raritan  Bay,  New  York 
and  New  Jersey.     Mar.  Fish.  Rev.  52:1-45. 

1997.  The  U.  S.  molluscan  fisheries  from  Massachusetts 
Bay  through  Raritan  Bay,  N.  J.  and  N.  Y.  In  The  his- 
tory, present  condition,  and  future  of  the  molluscan 
fisheries  of  North  and  Central  America  and  Europe. 
Volume  I,  Atlantic  and  Gulf  Coasts  (C.  L.  MacKenzie, 
Jr.,  V  G.  Burrell  Jr.,  A.  Rosenfield,  and  W.  L.  Hobart, 
eds.),  p.  87-118.     NOAA  Tech.  Rep.  127. 

Mantelatto,  F.  L.  M,  and  R.  A.  Christofoletti. 

2001.  Natural  feeding  activity  of  the  crab  Callinectes 
ornatus  (Portunidae)  in  Ubatuba  Bay  (Sao  Paulo,  Brazil): 
influence  of  season,  sex,  size  and  molt  stage.  Mar. 
Biol.  138:585-594. 

Marshall,  S.,  and  M.  Elliott. 

1997.  A  comparison  of  univariate  and  multivariate 
numerical  and  graphical  techniques  for  determining 
inter-  and  intraspecific  feeding  relationships  in  estua- 
rine fish.     J.  Fish  Biol.  51:526-545. 

McDermott,  J.  J. 

1983.  Food  web  in  the  surf  zone  of  an  exposed  sandy  beach 
along  the  mid-Atlantic  coast  of  the  United  States.  In 
Sandy  beaches  as  ecosystems  (A.  McLachlan  and  T. 
Erasmus  eds.l.  p.  529-538.  Dr.  W.  Junk,  The  Hague, 
Netherlands. 

Meise,  C.  J.,  and  L.  L.  Stehlik. 

2003.     Habitat  use,  temporal  abundance  variability, 

and  diet  of  blue  crabs  from  a  New  Jersey  estuarine 

system.     Estuaries  26:731-745. 
Milliken,  M.  R.,  and  A.  B.  Williams. 

1984.  Synopsis  of  biological  data  on  the  blue  crab,  Cal- 
linectes sapidus  Rathbun.  NOAA  Tech.  Rep.  NMFS 
1,  39  p. 

Milstein.  C.  B..  D.  T.  Thomas,  and  associates. 

1977.  Summary  of  ecological  studies  for  1972-1975  in 
the  bays  and  other  waterways  near  Little  Egg  Inlet  and 
in  the  ocean  in  the  vicinity  of  the  proposed  site  for  the 
Atlantic  Generating  Station,  New  Jersey.  Bull.  18,  757  p. 
Ichthyological  Associates,  Inc.,  Ithaca,  NY. 

Ojeda,  F  P.,  and  J.  H.  Dearborn. 

1991.  Feeding  ecology  of  mobile  benthic  predators:  experi- 
mental analyses  of  their  influence  in  Atlantic  rocky 
subtidal  communities  of  the  Gulf  of  Maine.  J.  Exp. 
Mar.  Biol.  Ecol.  149:13-44. 

Reilly,  P.  N.,  and  S.  B.  Saila. 

1978      Biology  and  ecology  of  the  rock  crab,  Cancer  irrora- 
tus Say,  1817,  in  southern  New  England  waters  (Deca- 
poda,  Brachyura).     Crustaceana  34:121-140. 
Romesburg,  H.  C. 

1984.     Cluster  analysis  for  researchers,  334  p.     Lifetime 
Learning  Publications,  Belmont,  CA. 
Ropes,  J.  W. 

1989.     The  food  habits  of  five  crab  species  at  Pettaquams- 
cutt  River,  Rhode  Island.     Fish.  Bull.  87:197-204. 
Rosas,  C,  E.  Lazaro-Chavez,  and  F  Buckle-Ramirez. 

1994.     Feeding  habits  and  food  niche  segregation  of 
Callinectes  sapidus,  C.  rathbunae  and  C.  similis  in  a 
subtropical  coastal  lagoon  of  the  Gulf  of  Mexico.     J. 
Crustacean  Biol.  14:371-382. 
Rountree,  R.  A.,  and  K.  W.  Able. 

1992.  Fauna  of  polyhaline  subtidal  marsh  creeks  in 
southern  New  Jersey:  composition,  abundance  and 
biomass.     Estuaries  15:171-185. 


710 


Fishery  Bulletin  102(4) 


Ryer,  C.  H. 

1987.     Temporal  patterns  of  feeding  by  blue  crabs  (Cal- 
linectes  sapidus)  in  a  tidal-marsh  creek  and  adjacent 
seagrass  meadow  in  the  lower  Chesapeake  Bay.     Es- 
tuaries 10:136-140. 
Scarratt,  D.  J.,  and  R.  Lowe. 

1972.     Biology  of  the  rock  crab  {Cancer  irroratus  i  in 
Northumberland  Strait.     J.  Fish.  Res.  Board  Canada 
29:161-166. 
Seed,  R.,  and  R.  N.  Hughes. 

1995.  Criteria  for  prey  size-selection  in  molliscivorous 
crabs  with  contrasting  claw  morphologies  J.  Exp.  Mar. 
Biol.  Ecol.  193:177-195. 

1997.  Chelal  characteristics  and  foraging  behavior  of 
the  blue  crab  Callinectes  sapidus  Rathbun.  Estuar. 
Coastal  Shelf  Sci.  44:221-229. 

Sponaugle.  S..  and  P.  Lawton. 

1990.     Portunid  crab  predation  on  juvenile  hard  clams: 

effects  of  substrate  type  and  prey  density.     Mar.  Ecol. 

Prog.  Ser.  67:43-53. 
Stehlik,  L.  L. 

1993.  Diets  of  the  brachyuran  crabs  Cancer  irroratus, 
C.  borealis,  and  Ovalipes  ocellatus  in  the  New  York 
Bight.     J.  Crustacean  Biol.  13:723-735. 

Stehlik,  L.  L.,  C.  L.  MacKenzie  Jr.  and  W.  W.  Morse. 

1991       Distribution  and  abundance  of  four  brachyuran 

crabs  on  the  northwest  Atlantic  shelf.     Fish.   Bull. 

89:473-492. 
Stehlik,  L.  L.,  P.  G.  Scarlett,  and  J.  Dobarro. 

1998.  Status  of  the  blue  crab  fisheries  of  New  Jersey.  J. 
Shellfish  Res.  17:475-485. 

Steimle,  F.  W.,  D.  Jeffress,  S.  A.  Fromm,  R.  N.  Reid, 
J.  J.  Vitaliano,  and  A.  B.  Frame. 

1994.  Predator-prey  relationships  of  winter  floun- 
der, Pleuronectes  americanus,  in  the  New  York  Bight 
apex.     Fish.  Bull.  92:608-619. 

Stevens,  B.  G.,  D.  A.  Armstrong,  and  R.  Cusimano. 

1982.     Feeding  habits  of  the  Dungeness  crab  Cancer 
magister   as   determined   by   the   index   of  relative 
importance.     Mar.  Biol.  72:135-145. 
Stoner,  A.  W.,  and  B.  A.  Buchanan. 

1990.     Ontogeny  and  overlap  in  the  diets  of  four  tropical 
Callinectes  species.     Bull.  Mar.  Sci.  46:3-12. 
Van  Engel,  W.  A. 

1958.     The  blue  crab  and  its  fishery  in  Chesapeake 


Bay.     Part  I — Reproduction,  early  development,  growth, 
and  migration.     Commer.  Fish.  Rev.  20:6-17. 

1990.  Development  of  the  reproductively  functional  form 
in  the  male  blue  crab,  Callinectes  sapidus.  Bull.  Mar. 
Sci.  46:13-22. 

van  Montfrans,  J.,  C.  H.  Ryer,  and  R.  J.  Orth. 

1991.  Population  dynamics  of  blue  crabs  Callinectes 
sapidus  Rathbun  in  a  lower  Chesapeake  Bay  tidal 
marsh  creek.     J.  Exp.  Mar.  Biol.  Ecol.  153:1-14. 

Vermeij,  G.  J. 

1978.     Biogeography  and  adaptation,  332  p.     Harvard 
University  Press,  Cambridge,  MA. 
Virnstein,  R.  W. 

1977.     The  importance  of  predation  by  crabs  and  fishes 
on   benthic   infauna   in   Chesapeake   Bay.     Ecology 
58:1199-1217. 
Warner,  G.  F. 

1977.     The  biology  of  crabs,  202  p.     Van  Nostrand  Rein- 
hold  Co.,  New  York,  NY. 
Warner,  G.  F.,  and  A.  R.  Jones. 

1976.  Leverage  and  muscle  type  in  crab  chelae  (Crusta- 
cea: Brachyural.     J.  Zool.,  Lond.  180:57-68. 

Williams,  A.  B. 

1984.     Shrimps,  lobsters,  and  crabs  of  the  Atlantic  coast 

of  the  eastern  United  States,  Maine  to  Florida,  550  p. 

Smithsonian  Institution  Press,  Washington,  DC. 
Williams,  A.  B.,  and  R.  L.  Wigley. 

1977.  Distribution  of  decapod  Crustacea  off  north- 
eastern United  States  based  on  specimens  at  the  North- 
east Fisheries  Center,  Woods  Hole,  MA.  NOAA  Tech. 
Rep.  NMFS  Circ.  407,  44  p. 

Williams,  M.  J. 

1981.     Methods  for  analysis  of  natural  diet  in  portunid 
crabs  (Crustacea,  Decapoda,  Portunidaei.     J.  Exp.  Mar. 
Biol.  Ecol.  52:103-113. 
Wilson.  K.  A.,  K.  W.  Able,  and  K.  L.  Heck  Jr. 

1990.     Habitat  use  by  juvenile  blue  crabs:  a  comparison 
among  habitats  in  southern  New  Jersey.     Bull.  Mar. 
Sci.  46:105-114. 
Winget,  R.  R.,  D.  Maurer,  and  H.  Seymour 

1974.  Occurrence,  size  composition  and  sex  ratio  of  the 
rock  crab,  Cancer  irroratus  Say  and  the  spider  crab. 
Libinia  emarginata  Leach  in  Delaware  Bay.  J.  Nat. 
Hist.  8:199-205. 


711 


Abstract — As  nearshore  fish  popu- 
lations decline,  many  commercial 
fishermen  have  shifted  fishing  effort 
to  deeper  continental  slope  habitats 
to  target  fishes  for  which  biological 
information  is  limited.  One  such  fish- 
ery that  developed  in  the  northeastern 
Pacific  Ocean  in  the  early  1980s  was 
for  the  blackgill  rockfish  (Sebastes 
melanostomits),  a  deep-dwelling 
(300-800  mi  species  that  congre- 
gates over  rocky  pinnacles,  mainly 
from  southern  California  to  southern 
Oregon.  Growth  zone-derived  age  esti- 
mates from  otolith  thin  sections  were 
compared  to  ages  obtained  from  the 
radioactive  disequilibria  of  210Pb,  in 
relation  to  its  parent,  226Ra,  in  otolith 
cores  of  blackgill  rockfish.  Age  esti- 
mates were  validated  up  to  41  years, 
and  a  strong  pattern  of  agreement 
supported  a  longevity  exceeding  90 
years.  Age  and  length  data  fitted 
to  the  von  Bertalanffy  growth  func- 
tion indicated  that  blackgill  rockfish 
are  slow-growing  (A'  =  0.040  females. 
0.068  males  I  and  that  females  grow 
slower  than  males,  but  reach  a  greater 
length.  Age  at  509c  maturity,  derived 
from  previously  published  length-at- 
maturity  estimates,  was  17  years  for 
males  and  21  years  for  females.  The 
results  of  this  study  agree  with  gen- 
eral life  history  traits  already  recog- 
nized for  many  Sebastes  species,  such 
as  long  life,  slow  growth,  and  late 
age  at  maturation.  These  traits  may 
undermine  the  sustainability  of  black- 
gill rockfish  populations  when  heavy 
fishing  pressure,  such  as  that  which 
occurred  in  the  1980s,  is  applied. 


Radiometric  validation  of  age,  growth, 
and  longevity  for  the  blackgill  rockfish 
(Sebastes  melanostomus) 


Melissa  M   Stevens 

Allen  H.  Andrews 

Gregor  M.  Cailliet 

Kenneth  H.  Coale 

Moss  Landing  Marine  Laboratories 

8272  Moss  Landing  Road 

Moss  Landing,  California  95039 

E-mail  address  ((or  A.  H  Andrews,  contact  author):  andrewsiS'mlml  calstate.edu 

Craig  C  Lundstrom 

Department  of  Geology 

University  of  Illinois— Urbana  Champaign 

255  Natural  History  Bldg. 

1301  W.Green  Street 

Urbana,  llmois  61801 


Manuscript  submitted  9  June  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
18  June  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:711-722  (20041. 


The  blackgill  rockfish  (Sebastes  mela- 
nostomus) is  a  deep-water  rockfish 
that  is  found  mainly  along  the  conti- 
nental slope  between  300  and  800  m 
depth  off  central  and  southern  Cali- 
fornia (Moser  and  Ahlstrom.  1978; 
Cross,  1987;  Williams  and  Ralston, 
2002).  Although  not  as  heavily  tar- 
geted in  relation  to  other  commercially 
important  rockfish  species,  a  directed 
commercial  fishery  for  blackgill  rock- 
fish has  existed  since  the  mid-1970s, 
beginning  off  southern  California 
(Point  Conception  area)  and  spreading 
northward  ( Monterey  area)  as  stocks 
of  other  heavily  fished  rockfishes 
declined  (Butler  et  al.,  1999).  Using 
acoustic  sonar  and  set  nets,  the  com- 
mercial fleet  was  able  to  catch  large 
aggregations  of  previously  unexploited 
blackgill  rockfish.  Landings  peaked  in 
1983  with  1346  metric  tons  (t)  caught 
coast-wide,  but  declined  over  the  next 
decade,  presumably  because  of  the  dis- 
appearance of  the  large  concentrations 
that  could  be  located  with  acoustical 
gear  (Butler  et  al.,  1999).  In  2001, 
141  t  were  reportedly  landed  along 
the  entire  west  coast  (PacFIN1) — less 
than  half  of  the  allowable  catch  (343  t; 


NOAA,  2001)  for  blackgill  rockfish 
that  year. 

The  first  stock  assessment  of  black- 
gill rockfish  was  made  by  Butler  et 
al.  (1999).  One  objective  of  this  as- 
sessment was  to  determine  age  and 
growth  characteristics,  which  were 
then  applied  to  estimate  age-at-ma- 
turity,  natural  mortality,  and  stock 
biomass.  Using  conventional  aging 
methods  (i.e.,  otolith  increments), 
we  estimated  that  blackgill  rockfish 
live  at  least  87  years  and  reach  full 
(100%)  maturity  from  13  to  26  years 
for  females,  and  from  13  to  24  years 
for  males.  Although  such  estimates 
are  useful  and  should  be  considered 
whenever  available,  validation  of  the 
age-estimation  procedure  is  needed  to 
be  certain  of  accurate  age  estimates 
(Beamish  and  McFarlane,  1983; 
Campana,  2001).  Inaccurate  age  de- 
terminations in  some  cases  have  led 
to  overharvesting  of  stocks  such  as 
Pacific  ocean  perch  (Sebastes  alutus) 


1  PacFIN  (Pacific  Fisheries  Information 
Network).  2002.  Commercial  fisher- 
ies landing  data.  http://www.PacFIN. 
org.     [Accessed  9  August  2002]. 


712 


Fishery  Bulletin  102(4) 


and  orange  roughy  (Hoplostethus  atlanticus;  Beamish, 
1979;  Archibald  et  al.,  1983;  Mace  et  al.,  1990).  These 
historical  examples  of  fishery  collapses  necessitate  that 
age  validation  be  achieved  before  age  and  growth  infor- 
mation is  applied  to  management. 

In  the  last  decade,  radiometric  age  validation  has 
been  applied  successfully  to  over  20  species  of  rock- 
fishes  and  other  marine  teleosts  (Burton  et  al.,  1999; 
Kastelle  et  al.,  2000;  Andrews  et  al.,  20021.  The  most 
common  technique  uses  the  disequilibria  between  two 
radioisotopes,  radium-226  (226Ral  and  lead-210  (210Pb), 
present  in  the  otolith  (Bennett  et  al.,  1982;  Smith  et  al., 
1991).  Radium-226  is  a  naturally  occurring  radioisotope 
and  calcium  analogue  that  is  incorporated  from  the 
surrounding  seawater  into  the  aragonitic  crystalline 
matrix  of  fish  otoliths.  Radium-226  decays  through  a 
series  of  short-lived  radioisotopes  to  210Pb.  Because 
the  half-lives  of  these  isotopes  are  known,  the  ratio  of 
activity  between  them  (210Pb:226Ra)  gives  a  measure  of 
elapsed  time  since  the  initial  incorporation  of  226Ra  into 
the  otolith  (Campana  et  al.,  1990).  Radium-226  decays 
very  slowly  (a  1600  year  half-life)  in  relation  to  21"Pb 
(a  22  year  half-life),  allowing  the  activity  ratio  of  these 
radioisotopes  to  build  into  secular  equilibrium  (1:1  ra- 
tio; Smith  et  al.,  1991).  Based  on  this  relationship  (also 
referred  to  as  ingrowth),  the  210Pb:  226Ra  activity  ratio 
is  suitable  for  age  determination  in  fishes  up  to  5  half- 
lives  of  210Pb,  or  approximately  120  years  of  age  (An- 
drews et  al.,  1999b;  Campana.  2001).  This  approach  is 
therefore  ideally  suited  to  the  blackgill  rockfish,  whose 
longevity  has  been  estimated  at  almost  90  years  (Butler 
et  al.,  1999). 

The  objectives  of  this  study  were  1)  to  estimate  age 
from  otolith  growth  zone  counts,  2)  to  describe  growth, 
and  3)  to  validate  the  annual  periodicity  of  growth 
zones  used  to  estimate  longevity  for  the  blackgill  rock- 
fish  with  the  radiometric  aging  technique.  An  ancillary 
objective  was  to  create  a  reliable  predictive  relationship 
between  average  otolith  weight  and  estimated  age  for 
use  as  a  timesaving  tool  in  the  management  of  this 
species.  Growth  zones  quantified  in  sectioned  otoliths 
were  used  to  estimate  age,  and  growth  was  described 
by  using  the  von  Bertalanffy  growth  function.  Final  age 
estimates  were  directly  compared  to  radiometric  ages  to 
evaluate  agreement  between  the  two  methods  and  ulti- 
mately were  used  to  validate  age  estimation  procedures, 
age-at-maturity,  and  longevity  for  this  species. 


Materials  and  methods 

Approximately  1210  blackgill  rockfish  sagittal  otoliths 
were  available  for  this  study.  Otoliths  were  collected 
by  National  Marine  Fisheries  Service  (NMFS)  person- 
nel from  commercial  vessels  in  1985  at  ports  along  the 
California  coastline  (Long  Beach  to  Fort  Bragg),  and 
during  NMFS  research  surveys  from  1998  to  2000  from 
central  California  to  the  Oregon-Washington  border. 
Thirty-two  juvenile  blackgill  rockfish,  collected  from 
spot  prawn  traps  along  the  central  California  coast,  were 


provided  by  Robert  Lea  of  the  California  Department  of 
Fish  and  Game  (CDFG).  Fish  total  length  iTL;  cm  or 
mm),  catch  area  (port  or  geographic  location),  and  otolith 
weights  (right  and  left.  1985  samples  only)  were  pro- 
vided. Otoliths  were  first  considered  for  age  estimation 
i  sectioning!,  and  the  remainder  were  reserved  for  radio- 
metric analysis.  Otolith  weights  (left  and  right,  male 
and  female)  were  measured  to  the  nearest  milligram 
and  compared  with  £-tests  to  determine  if  significant 
differences  in  mass  existed  between  sides  or  sexes. 

Estimation  of  age  and  growth 

Based  on  previous  aging  studies  and  the  need  to  con- 
serve samples  for  radiometric  analysis,  approximately 
310  otoliths  (25%  of  the  collection)  were  assumed  to  be 
sufficient  for  age  estimation.  The  left  otolith  from  5  to 
30  fish,  depending  upon  the  number  available  in  each 
50-mm  size  class  (ranging  from  100  mm  to  600  mm), 
was  randomly  chosen  by  using  a  basic  resampling  tool. 
Otoliths  were  thin-sectioned  and  mounted  onto  glass 
slides.  Approximately  50  otoliths  were  damaged  in  the 
sectioning  process,  leaving  260  otoliths  available  for 
age  estimation. 

Sections  were  viewed  by  three  readers  under  magnifi- 
cation (25  and  40x)  with  transmitted  or  reflected  light. 
Each  reader  obtained  age  estimates  by  inspecting  all 
available  growth  axes,  choosing  the  most  discernible 
axis,  and  reading  it  three  times  consecutively.  A  growth 
zone  (here  termed  an  "annulus")  was  defined  as  one  pair 
of  translucent  (winter-forming)  and  opaque  (summer- 
forming)  bands.  A  final  age,  based  on  each  reader's  most 
confident  estimate,  was  chosen.  Precision  between  and 
within  readers  was  compared  by  using  average  percent 
error  (APE;  Beamish  and  Fournier,  1981 1,  index  of  pre- 
cision (D)  and  coefficient  of  variation  (CV;  Chang,  1982). 
Percent  agreement  among  readers  was  also  calculated. 
Reader  1  (author)  determined  the  final  age  estimate  for 
each  section  as  described  in  Mahoney  (2002).  Ages  that 
could  not  be  confidently  resolved  (through  re-examina- 
tion or  discussion)  were  removed  from  analysis. 

Length  and  age  estimates  for  males,  females,  and  sex- 
es combined  were  fitted  to  the  von  Bertalanffy  growth 
function  (VBGF).  A  small  portion  of  juvenile  samples 
(/j  =16)  were  included  in  each  function.  Because  there 
was  strong  agreement  between  facility  aging  techniques 
(MLML  and  NMFS,  La  Jolla.  Butler  et  al.  1999),  ad- 
ditional aged  samples  were  added  to  strengthen  the 
VBGF  and  age  prediction  models  (ra=119).  Estimates  of 
age  at  first,  50%,  and  100^  maturity  were  calculated 
by  inserting  existing  size  at  maturity  data  (Echeverria. 
HIST )  into  the  VBGF  and  solving  for  age  (t). 

Age  prediction,  age  group  determination, 
and  core  extraction 

Campana  et  al.  1 1990 )  was  the  first  to  circumvent  the 
assumption  of  constant  226Ra  uptake  throughout  the 
life  of  the  fish  by  eliminating  younger  growth  layers 
from  adult  otoliths,  leaving  just  the  oldest  layers  of 


Stevens  et  al.:  Radiometric  validation  of  age,  growth,  and  longevity  of  Sebastes  melanostomus 


713 


otolith  growth  (i.e.,  the  core,  representing  the  first  few 
years  of  life).  Radium-226  is  present  at  such  low  activity 
levels,  however,  that  many  otolith  cores  from  fish  of  a 
similar  age  and  same  sex  must  be  pooled  to  acquire  the 
mass  of  material  needed  for  detection  (-0.5  to  1  gram; 
Andrews  et  al.,  1999a,  1999b).  Because  we  possessed 
a  limited  number  of  blackgill  rockfish  otoliths  (-1200), 
an  age  prediction  model  was  created  to  conserve  otolith 
material  for  radiometric  analysis.  It  was  appropriate  to 
assume  from  the  results  of  Francis  (2003)  that  within- 
sample  heterogeneity  with  respect  to  otolith  age  and 
mass  growth  rate  was  negligible  in  the  core  material. 

To  determine  age  groups  for  radiometric  analyses, 
final  ages  for  fish  whose  otoliths  were  sectioned,  along 
with  their  corresponding  average  otolith  weight  (left 
and  right,  «=2),  were  used  to  predict  age  for  the  re- 
maining fish  in  the  collection.  Several  parameters  were 
regressed  to  determine  a  predictive  relationship  be- 
tween average  otolith  weight  (henceforth  termed  "oto- 
lith weight")  and  estimated  age  (i.e.,  section  age).  The 
following  regressions  were  compared  to  estimated  age 
by  using  Kruskal-Wallis  (nonnormal)  ANOVA:  1)  oto- 
lith weight  (to  the  nearest  0.001  g),  2)  otolith  weight 
and  fish  length  (to  the  nearest  1  mm),  and  3)  otolith 
weight  plus  otolith  length  (to  the  nearest  0.001  mm) 
multiplied  by  otolith  weight  (as  an  interaction  term).  A 
power  function  was  also  investigated  but  did  not  result 
in  a  better  fit  than  that  provided  by  a  simple  linear 
regression  (either  log-transformed  or  normal).  A  paired 
sample  (-test  and  student's  (-test  for  slopes  were  used 
to  determine  if  a  significant  difference  existed  between 
male  and  female  otolith  weight,  and  between  male  and 
female  otolith  weight-to-age  regressions,  respectfully. 
The  final  regression  equations  were  applied  to  the  aver- 
age otolith  weight  for  all  individual  remaining  fish  to 
obtain  a  predicted  age.  Age  groups  were  created  if  there 
was  sufficient  otolith  material  from  fish  of  the  same  sex 
and  of  a  similar  predicted  age. 

The  predicted  age  range  for  each  group  was  kept  as 
narrow  as  possible  while  permitting  enough  material  for 
analysis;  approximately  25  to  50  otoliths  were  needed  at 
a  target  core  weight  of  0.02  g.  Fish  that  had  both  oto- 
liths intact  (not  sectioned  or  broken)  were  preferred  to 
reduce  the  number  of  fish  for  each  radiometric  sample. 
To  better  insure  sample  conformity,  90%  confidence 
intervals  with  respect  to  fish  length  and  otolith  weight 
were  used  to  eliminate  from  each  group  dissimilar  fish 
that  may  have  varied  significantly  from  predicted  age. 
In  addition  to  this  discriminating  technique,  groups 
were  further  confined  by  capture  year  and  location. 
Only  samples  caught  in  the  same  year  and  similar  geo- 
graphic location  (based  on  the  majority  of  port  locations 
within  300  miles)  were  included  in  the  same  group. 

Core  size  was  determined  by  viewing  several  whole 
juvenile  blackgill  rockfish  otoliths  with  estimated  ages 
between  1  and  7  years.  The  first  annulus  was  deter- 
mined to  be  approximately  2  mm  wide,  and  a  3-year-old 
otolith  was  measured  at  3  mm  wide,  4  mm  long,  and  1 
mm  thick,  and  having  a  weight  of  0.02  g.  These  dimen- 
sions were  chosen  as  the  target  core  size  because  a  core 


of  this  size  could  be  easily  extracted,  yet  was  young 
enough  to  minimize  the  possible  error  associated  with 
variable  226Ra  uptake  in  the  first  few  years  of  growth. 
Otoliths  from  adult  fish  were  ground  down  to  the  tar- 
get core  size  with  a  lapping  wheel  and  80-  to  120-grit 
silicon-carbide  paper.  Otoliths  from  selected  juveniles, 
if  older  than  age  3  (core  size),  were  also  ground  to  the 
target  core  size. 

Radiometric  analysis 

The  radiometric  analysis  was  conducted  as  described  in 
Andrews  et  al.  (1999a,  1999b).  Because  previous  studies 
have  revealed  extremely  low  levels  of  210Pb  and  226Ra  in 
otolith  samples,  trace  metal  precautions  were  employed 
throughout  sample  cleaning  and  processing  (Bennett  et 
al.,  1982;  Campana  et  al.,  1990;  Andrews  et  al. ,1999a). 
Acids  were  double  distilled  (GFS  Chemicals",  Powell, 
OH)  and  all  dilutions  were  made  using  Millipore-  filtered 
Milli-Q  water  (18  MQ/cm).  Samples  were  thoroughly 
cleaned,  dried,  and  weighed  to  the  nearest  0.0001  g 
prior  to  dissolution.  Whole  juvenile  otoliths  groups  were 
analyzed  first  to  determine  if  exogenous  210Pb  was  a 
significant  factor,  and  to  determine  baseline  levels  of 
226Ra  activity. 

Because  of  the  low-level  detection  problems  associ- 
ated with  (beta)  /3-decay  of  2lnPb,  the  activity  210Pb 
was  quantified  through  the  autodeposition  and  (alpha) 
a-spectrometric  determination  of  its  daughter  proxy, 
polonium-210  (210Po,  half-life=138  days;  Flynn,  1968).  In 
preparation  for  210Po  analysis,  samples  were  dissolved 
in  acid  and  spiked  with  a  calibrated  yield  tracer,  208Po, 
estimated  to  be  5  times  the  activity  of  210Po  in  the  oto- 
lith sample.  Polonium  isotopes  from  the  sample  were 
autodeposited  onto  a  purified  silver  planchet  (A.F  Mur- 
phy Die  and  Machine  Co.,  North  Quincy,  MA)  held  in 
a  rotating  Teflon™  holder  over  a  4-hour  period  (Flynn, 
1968).  The  activity  of  208Po  and  210Po  on  the  planchets 
was  measured  with  ion-implant  detectors  in  a  Tennelec 
(Oak  Ridge,  TN)  TC256  or-spectrometer  interfaced  with 
a  multichannel  analyzer  and  an  eight  channel  digital 
multiplexer.  Counts  were  recorded  with  Nucleus'-'  soft- 
ware (Nucleus  Personal  Computer  Analyzer  II,  The  Nu- 
cleus Inc.,  Oak  Ridge,  TN)  on  an  IBM  computer.  Counts 
measured  over  periods  that  ranged  from  28  to  50  days 
accumulated  from  160  to  919  total  counts.  Lead-210  ac- 
tivity, along  with  uncertainty,  was  calculated  in  a  series 
of  equations  that  corrected  for  background  and  reagent 
counts,  as  well  as  error  associated  with  count  statistics 
and  procedure  (pipetting  error,  yield-tracer  uncertainty, 
etc;  Andrews  et  al.,  1999a).  The  remaining  sample  was 
dried  and  conserved  for  226Ra  analysis. 

Determination  of  226Ra  employed  an  elemental  sepa- 
ration procedure  followed  by  isotope-dilution  thermal 
ionization  mass  spectrometry  (TIMS)  as  described  in 
Andrews  et  al.  (1999a,  1999b).  The  sample  was  spiked 
with  a  known  amount  of  22sRa  yield  tracer  estimated 
to  produce  a  226Ra:228Ra  atom  ratio  close  to  one.  The 
samples  were  dissolved  in  strong  acid  and  dried  re- 
peatedly (~90-100°C)  until  the  sample  color  was  bright 


714 


Fishery  Bulletin  102(4) 


white,  indicating  that  most  organic  material  had 
been  removed.  A  three-step  elemental  separation 
procedure  was  used  to  remove  calcium  and  bari- 
um, elements  that  interfere  with  the  detection  of 
radium  in  the  TIMS  process.  This  involved  pass- 
ing the  samples  through  three  cation  exchange 
columns,  two  containing  a  slurry  of  BioRad  AGE 
50W-X8  resin  (first  and  second  column),  and  one 
containing  EiChroM  Sr®  resin  (third  column). 
The  samples  were  introduced  to  a  highly  acidic 
medium  within  the  columns,  which  separated  the 
elements  according  to  elution  characteristics  (An- 
drews et  al.,  1999b). 

Radiometric  age  for  each  group  was  determined 
by  inserting  the  measured  210Pb  and  22BRa  activi- 
ties into  the  secular  equilibrium  model  (Smith  et 
al.,  1991)  and  correcting  for  the  elapsed  time  be- 
tween capture  and  autodeposition.  Because  these 
activities  were  measured  from  the  same  sample, 
the  calculation  was  independent  of  sample  mass 
(Andrews  et  al.,  1999a,  1999b).  Propagated  uncer- 
tainty associated  with  the  final  210Pb  activity  was 
based  on  count  statistics,  and  procedural  error 
and  uncertainty  for  the  final  226Ra  activity  was 
based  on  procedural  error  and  an  instrumental 
TIMS  analysis  routine  (Wang  et  al.,  1975;  An- 
drews et  al..  1999b).  The  combined  errors  were 
used  to  calculate  high  and  low  radiometric  ages. 

Accuracy  of  age  estimates 

Measured  210Pb:226Ra  activity  ratios  for  each  age 
group,  along  with  their  total  sample  age  (predicted 
age  +  time  since  capture),  were  plotted  with  the 
expected  210Pb:226Ra  growth  curve.  Each  age  group 
range  was  widened  by  multiplying  the  minimum 
and  maximum  age  in  the  range  by  the  age  estimate 
CV,  which  was  determined  from  the  variability  in 
age  estimates  among  three  readers.  Agreement 
between  the  measured  ratio  with  respect  to  esti- 
mated age  and  the  expected  ratio  (ingrowth  curve) 
provided  an  indication  of  the  age  estimate  accuracy. 
Radiometric  age  was  compared  to  the  average  pre- 
dicted age  for  each  group  by  using  two  tests:  1)  a 
paired  sample  £-test  to  determine  if  a  significant 
difference  existed  between  the  two  age  estimates 
for  the  groups  and  2)  predicted  age  was  plotted 
against  radiometric  age  and  the  correlation  was 
compared  to  a  hypothetical  agreement  line  (slope  of 
1)  by  using  r-tests  for  slope  and  elevation. 


Results 


-  __  —  - 

mZ&toSEL 

\ 

1  mm 

jjK, 

yjmiffZ-    iTm 

pp>*^^ 

i 

: 

Figure  1 

Three  images  of  a  blackgill  rockfish  (Sebastes  melariostomus) 
otolith  section  viewed  with  transmitted  light  at  25x  magni- 
fication (top),  40x  (center),  and  80x  magnification  (bottom). 
This  section  was  aged  most  consistently  as  90  years  under  a 
microscope,  but  because  of  finer  digital  resolution  and  contrast, 
the  section  pictured  can  be  aged  as  high  as  102  years. 


Estimation  of  age  and  growth 

Growth  zones  observed  within  otolith  sections  of  most 
blackgill  rockfish  were  difficult  to  interpret.  Distinction 
of  the  first  annulus  was  often  ambiguous,  and  the  band- 
ing pattern  during  the  first  several  years  (1  to  -10)  of 


growth  was,  in  some  sections,  wide  and  inconsistent. 
After  approximately  8  to  12  growth  zones,  the  zone  width 
transitioned  to  a  narrower  zone,  which  became  extremely 
compressed  after  20-40  growth  zones.  In  some  sections, 
these  older  zones  were  beyond  optical  resolution,  whereas 
in  others  they  were  remarkably  clear  (Fig.  1). 


Stevens  et  al.:  Radiometric  validation  of  age,  growth,  and  longevity  of  Sebastes  melanostomus 


715 


Table  1 

Comparison 

of  von  Bertalanffy  growth  function 

pai 

ameters  for  this  studv  and  Butler  et  al. 

(1999 

;  in  parentheses),  for  com- 

bined  sexes, 

females. 

and  males.  All  lengths  are 

total 

lengths  (mm).  Note  that  the  sample 

size  for  females  and  males  does 

not  sum 

to  332  because  the  same  juvenile  samp 

es 

Ul-. 

=  16)  were  used  for  each  sex,  and  onlv 

once 

for  combined  sexes.  N.R.= 

not  reported 

Combined  sexes 

Females 

Males 

L_  (mm) 

509(524') 

548(554') 

448(467') 

95%  CI 

491-528  (N.R.) 

520-576  (N.R.i 

434-462  (N.R.) 

k 

0.045(0.040) 

0.040(0.040) 

0.068(0.060) 

95%  CI 

0.038-0.052 

0.033-0.047 

0.058-0.078 

*o 

-4.86  (-5.02) 

-4.49  (-4.66) 

-2.37l-2.98) 

95%  CI 

-6.60  to -3.12 

-6.30  to  -2.67 

-3.55  to -1.19 

/! 

332(335) 

181 (98) 

167(78) 

r2 

0.81(0.79) 

0.87(0.90) 

0.87(0.92) 

1  Total  ler 

gths 

from  some  samples  in  Butler  et  al.,  1999 

were  estimated  from  fork  length  <  FL  in  mmi  by  using  an 

equation  from  Echeverria  and 

Lenarz 

1984). 

The  most  consistent  axis  in  the  otolith  section  for 
which  confident  interpretations  could  be  made  was 
along  either  the  sulcus  ridge,  or  along  the  dorsoventral 
margin.  Final  age  estimates  were  resolved  for  197  fish, 
or  approximately  76%  of  the  260  successfully  sectioned 
otoliths.  Agreement  among  readers  was  relatively  low: 
approximately  24%  of  age  estimates  were  within  ±1 
year,  61%  were  within  ±5  years,  and  87%  were  within 
±10  years.  The  mean  difference  in  age  estimates  be- 
tween readers  was  2.9  ±4.0  years.  Among  the  three 
readers,  APE  was  10.7%,  D  was  8.4%,  and  CV  was 
14.6%.  Average  percent  error,  D,  and  CV  estimates  were 
comparable  within  readers;  reader  1  APE  was  5.2%,  D 
was  4.1%,  and  CV  was  7.0%.  The  two  oldest  fish  to  be 
aged  were  a  90-year-old  male  (450  mm  TL)  collected  in 
1999  and  an  87-year-old  female  (546  mm  TL)  collected 
in  1985.  Both  individuals  were  caught  south  of  Point 
Conception,  California. 

The  VBGF  fitted  to  age  and  length  data  resulted  in 
distinct  growth  curves  for  male  and  female  blackgill 
rockfish  (Fig.  2).  This  difference  is  also  represented 
by  non-overlapping  confidence  intervals  with  respect 
to  the  primary  VBGF  parameters  (LM,  k;  Table  1).  The 
growth  coefficient,  k,  ranged  from  0.040  (±0.007,  fe- 
male) to  0.068  (±0.010,  male),  and  asymptotic  length 
was  448  ±14  mm  for  males  to  548  ±28  mm  for  females. 
The  asymptotic  length  for  females  was  32  mm  less  than 
the  largest  female  fish  sampled  (580  mm  TL),  and  for 
males,  was  74  mm  less  than  the  largest  male  sampled 
(522  mm  TL).  The  fit  for  all  three  functions  was  satis- 
factory (r2=0.81,  0.87;  Table  1,  Fig.  2).  Estimated  ages 
at  first,  50%,  and  100%  maturity,  derived  from  insert- 
ing published  estimates  of  length-at-maturity  (Echever- 
ria, 1987)  into  the  growth  model  for  each  sex,  were  15, 
21,  and  22  years  for  females  and  13,  17,  and  28  years 
for  males  (Table  2). 


Table  2 

Age  at  maturity  estimates,  in  years,  for  ma 
blackgill  rockfish  (95'i  confidence  intervals 
theses).  Maturity  estimates  were  derived 
published  estimates  of  length  at  maturity 
Bertalanffy  growth  function. 

e  and  female 
are  in  paren- 
by  inserting 
into  the  von 

First 

50% 

100% 

maturity 

maturity 

maturity 

Females 

15(12-22) 

21(16-31) 

22(17-33) 

Length  at 

maturity'  ( mm ) 

300 

350 

360 

Males 

13(11-15) 

17(14-20) 

27(22-35) 

Length  at 
maturity'  (mm) 

290 

330 

390 

'  Echeverria  11987). 

Age  prediction,  age  group  determination, 
and  core  extraction 

A  paired  sample  r-test  indicated  that  there  was  a  sig- 
nificant difference  between  male  and  female  average 
otolith  weight  (f=4.54,  P<0.001),  and  a  student's  r-test 
for  slopes  indicated  a  significant  difference  between  male 
and  female  average  otolith  weight-to-age  regressions 
(rm,  =  1.967,  ?=2.87,  P<0.05).  Therefore,  male  and  female 
age  estimates  and  regressions  were  treated  separately. 
There  was  no  statistical  difference  between  regres- 
sions involving  fish  length  and  average  otolith  weight 
(Kruskal-Wallis  one-way  ANOVA  on  ranks,  #=4.834, 
P=0.089).  A  simple  linear  regression,  with  average  oto- 
lith weight  as  the  independent  variable  and  estimated 


716 


Fishery  Bulletin  102(4) 


age  as  the  dependent  variable,  was  sufficient  to  pre- 
dict age.  Log  normalizing  the  regressions  to  stabilize 
the  variance  in  older  age  estimates  was  unsuccessful 
(Cochran's  test:  «=0.05,  36  df,  C=0.4748,  P=0.486).  The 
final  regressions  are  given  in  Figure  3. 

Fourteen  age  groups  based  on  the  predicted  ages  of 
unsectioned  otoliths  were  chosen.  These  groups  consist- 


700 


600 


500 


400 


■2>     300 


200 


100 


o  males 
X  females 


female 


Female  age  =  548(1  -  e-°<*°('*««»)t  ^=181  (/=0.87) 
Male  age  =  448(1  -  e"0068!'*237!),  n=167  (/=0.87) 


10  20  30  40  50  60 

Otolith  section  age  (yr) 


70 


80 


90 


100 


Figure  2 

Blackgill  rockfish  (Sebastes  melanostomus)  von  Bertalanffy  growth  functions 
plotted  for  males  and  females.  Observed  and  expected  values,  as  well  as  the 
parameters  of  the  equations,  are  given.  Note  that  the  same  juvenile  samples 
(/;  =  16)  were  included  in  both  male  and  female  equations. 


100  - 

males: 

90  - 

y=108.24x-2.65 

r2=0.83,  n=151                                                                     " 

80  - 

,                                                                                                     females 
females:                                                                                                  ♦               J^- 

y=93.803x  +  0.175                                              '  males               .         ^^ 

70  - 

f    60" 
I     50  - 

m 
CO 

a  4°- 
< 

30  - 
20  - 

r2=0.85,  n=165                                                           '    ♦ 

*     ■        l^'  ^ 

■                             0  9              m, _^^°  •» 

"  ■  •    •  ^^^r? 

■  males 

10  - 

•  females 

0     'T      "'                       1 1 1 1 1 1 1 1 1 

0.000      0.100      0.200      0.300      0.400      0.500      0.600      0.700      0.800      0.900 

Average  otolith  weight  (g) 

Figure  3 

Predictive  relationship  between  average  otolith  weight  and  estimated  age  for  blackgill 

rockfish  (Sebastes  melanostomus).  These  regression  equations  were  used  to  predict 

the  age  of  fish  whose  otoliths  were  reserved  for  radiometric  analysis. 

ed  of  four  juvenile  groups,  and  five  male  and  five  female 
adult  groups  I  Table  3).  Fish  lengths  ranged  from  82  mm 
to  580  mm  TL,  and  predicted  age  ranged  from  1  to  69 
years.  The  number  of  otolith  cores  per  age  group  ranged 
from  11  to  59,  representing  7  to  32  fish  per  group.  Total 
sample  weight  for  each  age  group  ranged  from  0.4649  g 
to  1.6424  g.  Whole  otolith  weight  ranged  from  0.041 
to  0.842  g,  and  average  individ- 
ual core  weight  for  the  adult  age 
groups  ranged  from  0.025  g  to 
0.028  g.  The  process  of  extracting 
the  core  inadvertently  destroyed 
some  otoliths  in  the  grinding  pro- 
cess, leading  to  smaller  samples 
for  some  groups. 


Radiometric  analysis 

Radiometric  analysis  of  all  age 
groups  (n  =  14)  resulted  in  the 
successful  determination  of  210Pb 
activity  for  all  samples,  and  lim- 
ited success  for  226Ra  (Table  4). 
Activities  of  210Pb  increased,  as 
expected,  fivefold  from  juvenile  to 
adult  age  groups,  and  ranged  from 
near  0.01  dpm/g  for  the  juvenile 
samples  to  over  0.05  dpm/g  for  the 
oldest  age  groups.  Error  associ- 
ated with  these  measurements 
ranged  from  3.7  to  9.2  %  (Is).  The 
detection  of  226Ra  activity  was 
met  with  some  technical  difficul- 
ties. Because  of  poor  radium 
recovery,  radium  measure- 
ments were  unreliable  in 
three  samples  and  radium 
was  lost  in  four  samples. 
Therefore,  an  average  of  the 
reliable  226Ra  measurements 
was  used  because  of  the  rel- 
ative consistency  of  levels 
measured  in  these  samples 
(0.0643  ±0.0035  dpm/g, 
n=l).  The  use  of  a  single 
estimate  for  226Ra  activ- 
ity was  acceptable  prior  to 
refinement  of  the  technique 
(Andrews  et  al.,  1999b).  Cal- 
culated 210Pb:226Ra  ratios 
increased  as  expected  from 
0.172  to  0.845  and  0.912  for 
the  oldest  groups  (Table  4). 

Age  estimate  accuracy 

Radiometric  ages  were  in 
agreement  with  predicted 
ages,  as  evidenced  by  concor- 
dance of  210Pb:226Ra  activity 


Stevens  et  al.:  Radiometric  validation  of  age,  growth,  and  longevity  of  Sebastes  melanostomus 


717 


Table  3 

Summary  data  for  14  pooled  otolith  age  groups  of  blackgill  rockfish.  The  age  range  and  sample  weight  of  each  age  group  was 
based  on  the  age  prediction  model  and  otolith  availability.  Groups  were  confined  by  year  of  capture,  and  for  the  1985  samples,  by 
port  location.  Mean  total  length  (±1  standard  deviation)  of  individuals  per  group  is  provided,  along  with  the  number  offish  and 
otoliths,  total  sample  weight,  and  average  core  weight. 

Sample 
number 

Age  group 
(yr) 

Sex 

Capture 
year 

Mean 

length 

±cjiTL  mm) 

Number  of 

fish,'  number 

of  otoliths 

Sample 

weight 

(g) 

Avg. 

core  weight 

(g) 

BG1 

1-3 

Juvenile 

1998 

154+26 

7,  ll2 

0.4649 

0.042 

BG2 

4 

Juvenile 

1998 

200  ±8 

10,  82 

1.1687 

0.065 

BG3 

4-5 

Juvenile 

1999 

217  ±9 

15, 192 

1.6630 

0.088 

BG4 

1-7 

Juvenile 

2000 

119  ±37 

25,  362 

0.7854 

0.022 

BG5 

29-31 

Female 

1985 

400  ±20 

25,46 

1.2510 

0.027 

BG6 

26-28 

Male 

1985 

379  ±19 

22.35 

0.8866 

0.025 

BG7 

11-17 

Female 

1998 

276  ±20 

22,33 

0.9018 

0.027 

BG8 

39-41 

Female 

1985 

458  ±22 

31,53 

1.3332 

0.025 

BG9 

48-54 

Male 

1985 

459  ±21 

25,48 

1.2491 

0.026 

BG10 

60-69 

Female 

1985 

525  ±30 

19,30 

0.8254 

0.028 

BG11 

19-23 

Male 

1998 

329  ±16 

21,39 

1.0313 

0.026 

BG12 

56-59 

Female 

1985 

502  ±28 

13.25 

0.6989 

0.028 

BG13 

39-41 

Male 

1985 

428  ±24 

31,59 

1.6424 

0.028 

BG14 

42-47 

Male 

1998 

423  ±26 

32,54 

1.4267 

0.026 

1  Both  otoliths 

were  not  available 

for  every  fish  chosen. 

-  Whole  j 

jveni 

e  otoliths. 

Table  4 

Summary  of  radiometric  results  for  pooled  otolith  age  groups.  Samples  are  listed  in  order  of  increasing  age-group  range.  Activi- 
ties are  expressed  as  disintegrations  per  minute,  per  gram  (dpm/g).  Radium-226  activity  was  averaged  among  samples  with  low- 
analytical  error  (<10%;  n=l)  and  was  determined  to  be  0.0643  (±0.0035)  dpm/g.  This  value  was  then  applied  to  all  samples  to 
gain  an  estimate  of  226Ra  activity  and  radiometric  age.  Agreement  between  radiometric  age  and  predicted  age  was  qualified  by 
the  degree  of  overlap  between  the  two  age  ranges.  Radiometric  age  incorporates  the  time  between  capture  and  analysis. 


Sample 

210Pb  activity 

210pb.226Ea 

Radiometric 

Radiometric 

Predicted  age 

Average 

Age  range 

number 

(dpm/g)  ±%s1 

activity  ratio 

age  (yr) 

age  range  (yr) 

group  range2 

age3  (yr) 

agreement^ 

BG1 

0.0154  ±8.6 

0.234 

7.1 

5.4-8.7 

0-3 

2 

Exceeds 

BG2 

0.0124  ±6.7 

0.193 

5.5 

4.3-6.5 

4-5 

4 

Overlaps 

BG3 

0.0118  ±5.5 

0.184 

5.5 

4.5-6.4 

4-6 

4.5 

Overlaps 

BG4 

0.0111  ±9.2 

0,172 

6.0 

5.3-6.7 

0-8 

3.5 

Encompasses 

BG7 

0.0300  ±5.6 

0.467 

18.0 

15.2-21.4 

9-19 

14 

Overlaps 

BG11 

0.0276  ±5.8 

0.430 

15.7 

13.2-18.7 

16-26 

21 

Overlaps 

BG6 

0.0440  ±4.7 

0.684 

22.3 

16.2-30.3 

22-32 

27 

Overlaps 

BG5 

0.0439  ±4.4 

0.683 

22.1 

16.2-30.3 

25-35 

30 

Overlaps 

BG13 

0.0481  ±3.8 

0.749 

29.3 

21.8-40.4 

33-47 

40 

Overlaps 

BG8 

0.0494  ±4.0 

0.769 

32.1 

23.7-45.1 

33-47 

40 

Overlaps 

BG14 

0.0499  ±3.8 

0.777 

45.8 

37.3-59.1 

36-54 

45 

Overlaps 

BG9 

0.0560  ±3.7 

0.871 

50.7 

35.8-85.1 

41-62 

51 

Encompasses 

BG12 

0.0586  ±4.7 

0.912 

62.9 

40.7-undef. 

48-67 

57 

Encompasses 

BG10 

0.0543  ±4.4 

0.845 

44.8 

31.6-71.6 

51-79 

65 

Overlaps 

'  Error  calculation  based  on  the  standard  deviation  of  210Pb  activity  (Wanget  al.,  1975). 

2  Predicted  age  range  was  extended  by  14.6%  of  coefficient  of  variation  (CV)  associated  with  growth-zone-derived  age  estimates. 

3  The  average  predicted  age  of  each  radiometric  age  group. 

4  Definition  of  terms:  Exceeds  =  radiometric  age  range  is  greater  than  predicted  age  range;  Overlaps  =  radiometric  age  range  partially  agrees  with 
predicted  age  range;  Encompasses  =  radiometric  age  range  was  in  agreement  with  predicted  age  range. 


718 


Fishery  Bulletin  102(4) 


in  otolith  cores  with  expected  ingrowth  curves  through 
time  (Fig.  4).  Of  the  14  pooled  otolith  groups,  three  had 
radiometric  age  ranges  that  fully  encompassed  the  pre- 
dicted age  range,  ten  resulted  in  overlapping  age  ranges, 
and  one  exceeded  predicted  age  (Table  4).  In  addition, 
radiometric  ages  were  in  close  agreement  with  predicted 
ages  in  a  direct  comparison  (r2=0.88;  Fig.  5).  Further 
Ntests  indicated  no  significant  difference  in  slope  (£=1.92, 
P=0.092)  or  elevation  U=0.163,  P=2.201)  between  the 
regression  and  a  hypothetical  agreement  line  (slope  of 
1),  confirming  the  close  agreement  of  radiometric  age 
and  predicted  age. 


Discussion 

Estimation  of  age  and  growth 

The  growth  pattern  present  in  otoliths  of  blackgill  rock- 
fish  was  often  difficult  to  interpret.  Complications  inher- 
ent to  the  growth  pattern  were  the  following:  obscure 
growth  zones  up  to  age  10-15  (the  ages  when  the  otolith 
begins  to  thicken  laterally),  rapid  transition  to  slower 
growth,  conflicting  or  ambiguous  growth  patterns,  and 
poor  resolution  of  extremely  compressed  zones  in  old-age 
fish.  Irregular  patterns  may  have  led  to  enumeration  of 
false  growth  zones  (checks),  and  the  compression  of  the 
outer  layers  may  have  concealed  growth  zones  present  in 
older  fish.  This  finding  has  been  consistent  among  previ- 
ous studies  of  rockfishes  I  Chilton  and  Beamish,  1982). 


1.2 
1  1 
1.0 
0.9 
0.8 
0.7 
06 
0.5 
04 
0.3 
0.2 
0.1 
0.0 


4^ 


^ 


=2 


i 


Expected  Ingrowth  Curve 

a  Juvenile  age  groups 
•  Female  age  groups 
■   Male  age  groups 


20 


40  60 

Sample  age  (yr) 


Figure  4 

Measured  210Pb:226Ra  ratio  plotted  against  total  sample  age  (mean 
predicted  age  plus  time  since  capture!  of  blackgill  rockfish  iSebastes 
melanostomus),  with  respect  to  the  expected  210Pb:226Ra  activity 
ratio.  Horizontal  error  bars  represent  the  predicted  age  range  (based 
on  age  prediction  model  extended  by  14.6%  CV).  Vertical  error 
bars  represent  high  and  low  activity  ratios  based  on  the  analytical 
uncertainty  associated  with  2lnPb  and  226Ra  measurements. 


Because  of  the  difficulty  involved  in  interpreting  growth 
patterns,  aging  of  blackgill  rockfish  otoliths  involved  a 
high  degree  of  individual  subjectivity,  as  evidenced  by 
the  relatively  low  precision  (D=8.4%)  and  high  varia- 
tion (CV=15%)  between  readers.  However,  there  were 
some  remarkably  clear  otoliths  and  for  these  we  were 
highly  confident  of  age  estimates  (Fig.  1).  Overall,  87% 
of  between-reader  age  estimates  were  within  10  years, 
emphasizing  that  although  the  method  of  interpretation 
of  growth  can  be  imprecise,  it  provides  a  reasonable 
indication  of  the  growth  characteristics  and  longevity 
of  this  species. 

The  von  Bertalanffy  growth  parameters  for  male 
and  female  blackgill  rockfish  appear  to  indicate  that 
blackgill  rockfish  possess  distinct  patterns  of  growth 
(Table  1).  Female  blackgill  rockfish  exhibited  a  slower 
growth  rate  than  males  up  to  approximately  25-30 
years  of  age  (Fig.  2).  At  this  point,  the  male  growth 
rate  slows  and  approaches  an  asymptotic  length  of  448 
mm,  but  females  continue  to  grow  in  length,  reaching 
an  asymptotic  length  of  548  mm.  This  trend  of  slower 
growing,  but  ultimately  larger  females  has  been  ob- 
served in  other  slope-dwelling  Sebastes  species,  such 
as  the  darkblotched  (S.  crameri;  Rogers  et  al.,  20001, 
and  splitnose  (S.  diploproa;  Wilson  and  Boehlert.  1990) 
rockfishes.  For  both  sexes,  the  growth  coefficient  is  low 
(k  =  0. 040-0. 068)  when  compared  to  shallower-dwell- 
ing (50-200  m)  rockfishes,  such  as  the  greenstriped 
(S.  elongates,  0.10-0.12;  Love  et  al.,  1990)  and  widow 
(S.  entomelas,  0.20-0.25;  Williams  et  al.,  2000)  rock- 
fishes, but  very  similar  to  other  deep-dwell- 
ing, long-lived  species,  such  as  the  short- 
spine  thornyhead  (Sebastolobus  alaseanus, 
£  =  0.020;  Cailliet  et  al.,  2001),  yelloweye  (S. 
ruberrimus,  £  =  0.046;  Andrews  et  al.,  2002), 
and  bank  (S.  rufus,  £  =  0.041;  Cailliet  et  al., 
2001)  rockfishes. 

Previous  maturity  estimates  for  blackgill 
rockfish  (7-9  yr  males,  6-10  yr  females; 
Echeverria  1987).  based  on  whole  otolith 
counts,  were  much  lower  than  estimates  ob- 
tained from  section  ages  in  the  present  study 
(Table  2).  Maturity  estimates  from  our  study 
support  those  derived  by  Butler  et  al.  (1999). 
largely  because  the  aging  protocol  was  the 
same  between  facilities.  Although  our  growth 
model  included  some  age  estimates  (37%) 
from  Butler  et  al.  (1999),  our  results  further 
confirm  age  at  maturity  (Table  2).  Compared 
to  other  species  of  the  genus,  blackgill  rock- 
fish have  a  late  maturity  that  resides  at  the 
upper  end  of  the  range  for  rockfishes  (Cail- 
liet et  al.,  2001;  Love  et  al.,  2002). 

Extraordinarily  old  ages  in  average-size 
fish  exhibited  by  the  blackgill  rockfish 
should  not  be  dismissed  as  an  anomaly.  In 
this  study  the  oldest  blackgill  rockfish  was 
a  90-year-old  male  (aged  as  high  as  102 
years)  that  was  450  mm  TL.  This  fish  was 
160  mm  less  than  the  maximum  reported 


100 


Stevens  et  al.:  Radiometric  validation  of  age,  growth,  and  longevity  of  Sebastes  melanostomus 


719 


0 


length  (Love  et  al.,  2002).  Accord- 
ing to  an  experienced  rockfish  age 
and  growth  researcher,  "some  of 
the  oldest  specimens  [rockfish]  are 
rarely  the  largest  (lengthwise), 
and  most,  if  not  all,  are  males." 
(Munk2)  The  reasons  for  this  age- 
length  pattern  are  beyond  the 
scope  of  this  study,  but  the  impli- 
cations for  stock  dynamics  and 
management  are  that  it  is  worthy 
of  further  consideration. 

Age  prediction,  age-group 

determination, 

and  core  extraction 


The  use  of  otolith  weight  as  a  proxy 
for  age  has  benefits  over  conven- 
tional otolith  aging  methods  by 
reducing  cost,  increasing  sample 
size,  and  allowing  greater  objectiv- 
ity (Boehlert,  1985;  Pawson,  1990; 
Fletcher,  1991;  Pilling  et  al.,  2003). 
In  this  study,  predicting  ages  from 
otolith  weight  increased  the  number 
of  unsectioned  otoliths  that  could 
be  used  in  the  radiometric  analysis, 
but  the  prediction  model  also  ampli- 
fied the  uncertainty  associated 
with  estimates  of  age.  especially 
in  older  fish.  The  variance  around 
the  regression  line  increased  with 

otolith  weight,  and  log  normalizing  the  data  did  not 
eliminate  this  problem.  Older  predicted  ages,  there- 
fore, were  more  uncertain  than  younger  ages  (Fig.  3). 
Although  limited  to  a  specific  otolith  weight  range,  the 
prediction  model  presented  here  may  provide  managers 
with  a  more  efficient  and  less  costly  way  to  investigate 
the  age  structure  of  blackgill  rockfish  stocks. 

In  an  ideal  study,  otoliths  from  the  entire  estimated 
age  range  for  blackgill  rockfish  would  be  available  in 
the  sample  set.  Otoliths  from  fish  with  predicted  ages 
greater  than  70  years,  however,  were  not  present  in  our 
study  in  sufficient  numbers  to  allow  age  determination 
by  radiometric  methods.  This  was  so,  even  though  more 
than  half  of  the  1200  otolith  pairs  obtained  for  ourstudy 
were  sampled  directly  from  commercial  fishing  vessels 
in  1985  along  the  coast  of  central  and  southern  Cali- 
fornia, where  the  bulk  of  the  fishery  occurred.  Because 
fishermen  often  target  adult  aggregations,  the  absence 
of  these  older  individuals  may  be  an  indication  that  the 
population  had  already  experienced  depletion  of  older 
age  classes  at  the  time  of  sample  collection,  particularly 
if  natural  mortality  is  thought  to  be  low  for  most  rock- 


1:1  line 


Regression 


y=0.823x  + 2.534 


r  =0.89 


20 


40  60 

Predicted  age  (yr) 


80 


100 


Figure  5 

Direct  comparison  of  mean  predicted  age  and  radiometric  age  for  14  pooled 
otolith  age  groups  for  blackgill  rockfish  (Sebastes  melanostomus).  A  regres- 
sion of  the  data  points  and  1:1  line  of  agreement  are  included  for  comparison. 
Horizontal  error  bars  represent  the  error  associated  with  age  estimation 
(average  predicted  age  multiplied  by  14.6%  CV),  plus  the  standard  error  of 
the  regression  (Is)  used  to  predict  age  for  radiometric  samples.  Vertical  error 
bars  represent  high  and  low  radiometric  age  estimates  based  on  the  analytical 
uncertainty  associated  with  210Pb  and  22fiRa  measurements. 


fishes  (Bloeser3).  However,  it  is  possible  that  the  largest, 
oldest  fish  are  naturally  rare,  even  at  the  start  of  an 
intensive  commercial  fishery.  Knowledge  of  blackgill 
rockfish  pre-exploitation  stock  structure  and  population 
dynamics  would  help  to  elucidate  which  (depletion  of 
older  age  classes  or  a  natural  situation  of  low  numbers 
of  older  fish)  is  the  more  likely  scenario. 

Radiometric  analysis 

In  previous  studies  the  analytical  uncertainty  of  226Ra 
was  the  limiting  factor  in  radiometric  age  determina- 
tion (Andrews  et  al.,  1999a).  Typically,  TIMS  determi- 
nation of  226Ra  reduces  error  to  less  than  1-3%  of  the 
determined  value,  but  technical  difficulties  (improperly 
mixed  nitric  acid)  led  to  poor  recovery  and  loss  of  radium 
in  seven  samples.  The  remaining  seven  samples  were 
deemed  reliable  because  of  relatively  high  radium  recov- 
ery, longer  run  times,  and  low  analytical  uncertainty 
as  determined  by  the  TIMS  analysis  routine.  The  226Ra 
activity  determined  for  these  samples  was  consistent 


2  Munk.  K.     2002.     Personal  commun.     Alaska  Department 
of  Fish  and  Game,  P.O.  25526,  Juneau,  AK  99802. 


3  Bloeser,  J.  A.  1999.  Diminishing  returns:  the  status  of 
West  Coast  rockfish,  94  p.  Pacific  Marine  Conservation 
Council,  P.O.  Box  59,  Astoria,  OR  97103. 


720 


Fishery  Bulletin  102(4) 


enough  that  we  could  assume  that  226Ra  activities  were 
similar  among  all  samples  and  that  use  of  an  average 
was  valid  (0.0643  [±0.0035]  dpm/g).  This  approach  is 
acceptable  because  226Ra  activities  measured  in  pre- 
vious radiometric  studies  on  Pacific  rockfishes  were 
relatively  constant.  For  example,  the  activity  of  cored 
yelloweye  rockfish  (S.  ruberrimus)  otoliths  had  a  mean 
226Ra  activity  of  0.0312  (±0.0026)  dpm/g  (n=18;  Andrews 
et  al.,  2002),  and  the  rougheye  rockfish  (S.  aleutianus), 
another  deepwater  species  (to  730  m;  Love  et  al.,  2002), 
had  a  similar  otolith  core  226Ra  activity  averaging  0.065 
(±0.003)  dpm/g  (Kastelle  et  al.,  2000). 

Accuracy  and  uncertainty  of  ages  estimates 

Radiometric  activities  measured  in  blackgill  rockfish 
otoliths  generally  agreed  with  expected  activity  ratios 
for  210Pb  and  226Ra  (Fig.  4),  confirming  the  validity  of 
growth-zone-derived  age  estimates.  In  addition,  a  direct 
comparison  between  radiometric  age  and  predicted  age 
resulted  in  a  strong  agreement  (r2=0.89;  Fig.  5),  which 
was  further  supported  by  slope  and  elevation  tests  that 
revealed  no  significant  difference  from  a  1:1  agreement 
line. 

The  most  critical  sources  of  error  involved  in  age 
estimation,  prediction,  and  radiometric  age  determina- 
tion were  the  following:  1)  age  estimate  uncertainty,  2) 
regression  error  associated  with  predicted  ages,  and  3) 
analytical  uncertainty  associated  with  the  radiometric 
aging  technique  (TIMS  and  a-spectrometry).  Conven- 
tional aging  techniques  are  inherently  subjective  (Boe- 
hlert,  1985;  Campana,  2001)  and  thus  create  uncertain- 
ty associated  with  an  estimated  age.  This  uncertainty 
is  transferred  to  the  prediction  model,  where  the  natu- 
ral variability  associated  with  individual  otolith  weight 
must  also  be  considered.  For  most  samples,  however,  the 
error  bars  either  overlapped  or  were  in  contact  with  the 
agreement  line  (Figs.  4  and  5),  further  confirming  the 
concordance  of  radiometric  age  with  predicted  age. 

Implications  for  management 

When  considering  the  longevity  of  rockfishes  for  which 
a  maximum  age  has  been  reported  (Munk,  2001;  Cail- 
liet  et  al.,  2001),  a  longevity  exceeding  90  years  places 
the  blackgill  rockfish  within  the  top  2Q(7c  of  long-lived 
rockfishes.  There  is  a  trend  for  rockfishes  that  may 
indicate  that  longevity  increases  as  maximum  depth  of 
occurrence  increases,  and  physiological  adaptations  to 
the  environmental  conditions  of  deep-sea  living  could 
provide  an  explanation  (Cailliet  et  al.,  2001).  The  con- 
firmed longevity  and  the  maximum  depth  of  occurrence 
(~800  m)  for  the  blackgill  rockfish  provide  further  sup- 
port for  this  concept. 

Longevity  in  the  rockfishes  has  been  central  to  its 
evolutionary  success  in  relation  to  other  marine  teleosts. 
The  suite  of  life  history  characters  implicit  with  a  long 
lifespan  (slow  adult  growth,  late  age-at-maturity,  low 
adult  natural  mortality)  represent  a  "slow  and  steady" 
adaptive  strategy,  whereby  the  energy  allocated  towards 


individual  growth  is  prolonged,  eventually  contributing 
to  greater  fecundity  (due  to  larger  size  at  maturity) 
over  the  lifespan  of  the  individual.  This  reproductive 
strategy  serves  to  propagate  genetic  material  across 
several  generations,  as  well  as  to  diffuse  the  effect  of 
mortality  associated  with  each  reproductive  event  (Lea- 
man,  1991).  In  this  sense,  longevity  may  act  to  buffer 
the  species  against  short-term  (El  Nino)  and  long-term 
environmental  change  (Pacific  Decadal  Oscillations), 
and  the  stochasticity  inherent  in  the  Pacific  Ocean 
system  (Moser  et  al.,  2000). 

In  the  absence  of  fishing  pressure,  the  genetic  con- 
tribution of  a  slow-growing,  longer-lived  species  may 
be  more  conserved  in  the  collective  species'  gene  pool 
i  Munk2 1.  In  the  presence  of  fishing  pressure,  however, 
this  "slow  and  steady"  adaptation  may  be  detrimental 
(Musick,  1999).  Although  modeling  fish  populations  for 
the  purpose  of  management  typically  involves  some  or 
all  of  these  parameters,  the  focus  is  often  on  deter- 
mining sustainable  biomass  and  this  approach  largely 
ignores  the  unknown  effects  of  changes  in  age  struc- 
ture due  to  removal  of  the  oldest  individuals  from  the 
population  (Craig,  1985),  as  well  as  a  loss  of  genetic 
diversity  that  could  prevent  full  recovery  of  severely 
depleted  populations  (Hauser  et  al,  2002).  Given  the 
current  depressed  condition  of  many  heavily  fished  rock- 
fish stocks,  species-specific  life  history  characteristics, 
such  as  longevity,  growth  rate,  and  age-at-maturity 
estimates,  should  be  given  thorough  consideration  in 
the  development  of  an  effective  management  strategy. 
Management  regulations  that  account  for  these  charac- 
teristics, such  as  a  limited  fishing  season,  or  designa- 
tion of  harvest  refugia  (Yoklavich,  1998),  would  provide 
a  stronger  basis  for  conservation  and  sustainability  of 
the  resource. 


Acknowledgments 

We  wish  to  thank  John  Butler,  Don  Pearson,  and  Cindy 
Taylor  of  the  Southwest  Fisheries  Science  Center,  Mark 
Wilkins,  Jerry  Hoff,  Waldo  Wakefield,  and  Bob  Lauth 
of  the  Alaska  Fisheries  Science  Center,  and  Bob  Lea 
of  CDFG  for  donating  specimens  for  this  study.  Mary 
Yoklavich  (NMFS).  Di  Tracey  and  Larry  Paul  (NIWA, 
New  Zealand),  Kristen  Munk  (ADFG),  Don  Pearson 
(NMFS),  Tom  Laidig  (NMFS),  and  Steve  Campana 
(DFO,  Canada)  provided  valuable  insight  into  black- 
gill rockfish  growth  patterns.  Patrick  McDonald  of  the 
Oregon  Department  of  Fish  and  Wildlife  aged  otolith 
sections.  Pete  Holden  at  the  University  of  California, 
Santa  Cruz,  measured  radium  in  the  refined  samples 
using  TIMS.  The  comments  and  suggestions  of  three 
anonymous  reviewers  were  greatly  appreciated.  This 
work  was  supported  by  the  National  Sea  Grant  College 
Program  of  the  U.S.  Department  of  Commerce's  National 
Oceanic  and  Atmospheric  Administration  under  NOAA 
Grant  number  NA06RG0142,  project  number  R/F-182. 
through  the  California  Sea  Grant  College  Program;  and 
in  part  by  the  California  State  Resources  Agency. 


Stevens  et  al.:  Radiometric  validation  of  age,  growth,  and  longevity  of  Sebastes  melanostomus 


721 


Literature  cited 

Andrews,  A.  H.,  G.  M.  Cailliet,  and  K.  H.  Coale. 

1999a.     Age  and  growth  of  the  Pacific  grenadier  (Cory- 
phaenoides  acrolepis)  with  age  estimate  validation  using 
an  improved  radiometric  ageing  technique.     Can.  J. 
Fish.  Aquat.  Sci.  56:1339-1350. 
Andrews,  A.  H.,  K.  H.  Coale,  J.  Nowicki,  C.  Lundstrom, 
A.  Palacz,  and  G.  M.  Cailliet. 

1999b.  Application  of  an  ion-exchange  separation  tech- 
nique and  thermal  ionization  mass  spectrometry  to  226Ra 
determination  in  otoliths  for  radiometric  age  determi- 
nation of  long-lived  fishes.  Can.  J.  Fish.  Aquat.  Sci. 
56:1329-38. 
Andrews,  A.  H.,  G.  M.  Cailliet,  K.  H.  Coale,  K.  M.  Munk. 
M.  M.  Mahoney  and  V.  M.  O'Connell. 

2002.     Radiometric  age  validation  of  the  yelloweye  rockfish 
{Sebastes  ruberrimus)  from  south-eastern  Alaska.     Mar. 
Freshw.  Res.  53:1-8. 
Archibald,  C.  P.,  D.  Fournier,  and  B.  M.  Leaman. 

1983.     Reconstruction  of  stock  history  and  development 
of  rehabilitation  strategies  for  Pacific  ocean  perch  in 
Queen  Charlotte  Sound,  Canada.     N.  Am.  J.   Fish. 
Manag.  3:283-294. 
Beamish,  R.  J. 

1979.     New  information  on  the  longevity  of  Pacific  ocean 
perch  {Sebastes  alutus).     J.  Fish.  Res.  Board  Can. 
36:1395-1400. 
Beamish.  R.  J.,  and  D.  A.  Fournier. 

1981.  A  method  for  comparing  the  precision  of  a  set 
of  age  determinations.  Can.  J.  Fish.  Aquat.  Sci. 
38:982-983. 

Beamish,  R.  J.,  and  G.  A.  McFarlane. 

1983.     The  forgotten  requirement  for  age  validation  in  fish- 
eries biology.     Trans.  Am.  Fish.  Soc.  112(61:735-743. 
Bennett,  J.  T.,  G.  W.  Boehlert,  and  K.  K.  Turekian. 

1982.  Confirmation  of  longevity  in  Sebastes  diploproa 
(Pisces:  Scorpaenidae)  from  21uPb/226Ra  measurements 
in  otoliths.     Mar.  Biol.  71:209-215. 

Boehlert,  G.  W. 

1985.     Using  objective  criteria  and  multiple  regression 
models  for  age  determination  in  fishes.     Fish.  Bull. 
83:103-117. 
Burton,  E.  J.,  A.  H.  Andrews,  K.  H.  Coale,  and  G.  M.  Cailliet. 
1999.     Application  of  radiometric  age  determination  to 
three  long-lived  fishes  using  210Pb:226Ra  disequilibria 
in  calcified  structures:  a  review.     In  Life  in  the  slow 
lane:  ecology  and  conservation  of  long-lived  marine 
animals  (J.  A.  Musick,  ed.,),  p.  77-87.     Am.  Fish.  Soc. 
Symp.  23. 
Butler.  J.  L..  L.  D.  Jacobson,  and  J.  T.  Barnes. 

1999.  Stock  assessment  for  blackgill  rockfish.  In  Appen- 
dix to  the  status  of  the  Pacific  coast  groundfish  fishery 
through  1998  and  recommended  acceptable  biological 
catches  for  1999:  stock  assessment  and  fishery  evalua- 
tion, 93  p.  Pacific  Fishery  Management  Council,  2130 
SW  Fifth  Avenue,  Suite  224,  Portland,  Oregon,  97201. 
Cailliet,  G.  M„  A.  H.  Andrews,  E.  J.  Burton,  D.  L.  Watters, 
D.  E.  Kline,  and  L.  A.  Ferry-Graham. 

2001.     Age  determination  and  validation  studies  of  marine 
fishes:  do  deep-dwellers  live  longer?     Exp.  Gerontology 
36:739-764. 
Campana,  S.  E. 

2001.  Accuracy,  precision  and  quality  control  in  age  deter- 
mination, including  a  review  of  the  use  and  abuse  of  age 
validation  methods.     J.  Fish  Biology  59:197-242. 


Campana.  S.  E..  K.  C.  Zwanenburg,  and  -J.  N.  Smith. 

1990.  210Pb/226Ra  determination  of  longevity  in 
redfish.     Can.  J.  Fish.  Aquat.  Sci.  47:163-165. 

Chang,  W.  Y.  B. 

1982.     A  statistical  method  of  evaluating  the  reproduc- 
ibility of  age  determination.     Can.  J.  Fish.  Aquat.  Sci. 
48:734-750. 
Chilton,  D.  E.,  and  R.  J.  Beamish. 

1982.     Age  determination  methods  for  fishes  studied  by 
the  Groundfish  Program  at  the  Pacific  Biological  Station, 
102  p.     Can.  Spec. Pub.  Fish.  Aquat.  Sci.  60. 
Craig,  J.  F. 

1985.     Aging  in  fish.     Can.  J.  Zoology  63:1-8. 
Cross,  J.  N. 

1987.     Demersal  fishes  of  the  upper  continental  slope  off 
southern  California.     CalCOFI  Report  28:155-167. 
Echeverria,  T.  W. 

1987.     Thirty-four   species   of  California   rockfishes: 
maturity  and  seasonality  of  reproduction.     Fish.  Bull. 
85:229-249. 
Echeverria.  T.  W.,  and  W.  H.  Lenarz. 

1984.     Conversions  between  total,  fork,  and  standard 
length  in  35  species  of  Sebastes  from  California.     Fish. 
Bull.  82:249-251. 
Fletcher,  W.  J. 

1991.  A  test  of  the  relationship  between  otolith  weight 
and  age  for  the  pilchard  Sardinops  neopilchardus.  Can. 
J.  Fish.  Aquat.  Sci.  48:35-38. 

Flynn,  W.  W. 

1968.     The  determination  of  low  levels  of  polonium-210 
in  environmental  materials.     Anal.  Chimica  Acta.  43: 
221-227. 
Francis,  R.  I.  C.  C. 

2003.     The  precision  of  otolith  radiometric  ageing  offish 
and  the  effect  of  within-sample  heterogeneity.     Can.  J. 
Fish.  Aquat.  Sci.  60:441-447. 
Hauser,  L.,  G.  J.  Adcock,  P.  J.  Smith,  J.  H.  Bernal  Ramirez,  and 
G.  R.  Carvalho. 

2002.     Loss  of  microsatellite  diversity  and  low  effective  pop- 
ulation size  in  an  overexploited  population  of  New  Zea- 
land snapper  {Pagrus  auratus).     Proc.  Nat  Acad.  Sci. 
99(181:11742-11747. 
Kastelle,  C.  R.,  D.  K.  Kimura.  and  S.  R.  Jay. 

2000.     Using  Pb210/Ra226  disequilibrium  to  validate  con- 
ventional ages  in  Scorpaenids  {sic)  (genera  Sebastes  and 
Sebastolobus).     Fish.  Res.  46:299-312. 
Leaman,  B.  L. 

1991.     Reproductive  style  and  life  history  variables 
relative  to  exploitation  and  management  of  Sebastes 
stocks.     Environ.  Biol.  Fishes  30:  253-271. 
Love,  M.  S„  P.  Morris,  M.  McCrae,  and  R.  Collins. 

1990.     Life  history  aspects  of  19  rockfish  species  (Scor- 
paenidae: Sebastes)  from  the  southern  California  Bight, 
38  p.     NOAA  Technical  Report.  NMFS  87. 
Love,  M.  S.,  M.  M.  Yoklavich,  L.  Thorsteinson,  and  J.  Butler. 
2002.     The  rockfishes  of  the  Northeast  Pacific,  405 
p.     Univ.  California  Press,  Berkeley,  CA. 
Mace,  P.  M.,  J.  M.  Fenaughty,  R.  P.  Coburn,  and  I.  J.  Doonan. 
1990.     Growth  and  productivity  of  orange  roughy  {Hoplo- 
stethus  atlanticus)  on  the  north  Chatham  Rise.     New 
Zealand  J.  Mar.  Fresh.  Res.  24:105-109. 
Mahoney,  M.  M. 

2002.  Age,  growth  and  radiometric  age  validation  of 
the  blackgill  rockfish.  Sebastes  melanostomus.  M.Sc. 
thesis,  70  p.  California  State  Univ.,  San  Francisco, 
CA.     [Available  from:  Moss  Landing  Marine  Labora- 


722 


Fishery  Bulletin  102(4) 


tories,  8272  Moss  Landing  Road,  Moss  Landing,  CA 
95039.] 
Moser,  H.  G.,  and  E.  H.  Ahlstrom. 

1978.     Larvae  and  pelagic  juveniles  of  blackgill  rockfish, 
Sebastes  melanostomus,  taken  in  midwater  trawls  off 
southern  California  and  Baja  California.     J.  Fish.  Res. 
Board  Can.  35:981-996. 
Moser,  H.  G.  R.  L.  Charter,  W.  Watson,  D.  A.  Ambrose, 
J.  L.  Butler,  S.  R.  Charter,  and  E.  M.  Sandknop. 

2000.  Abundance  and  distribution  of  rockfish  {Sebastes) 
larvae  in  the  southern  California  Bight  in  relation  to 
environmental  conditions  and  fishery  exploitation.  Cal- 
COFI  Report  41:132-147. 

Musick,  J.  A. 

1999.  Ecology  and  conservation  of  long-lived  marine 
animals.  In  Life  in  the  slow  lane:  ecology  and  con- 
servation of  long-lived  marine  animals  (J.  A.  Musick, 
ed.t,  p.  1-10.     Am.  Fish.  Soc.  Symp.  23. 

Munk,  K.  M. 

2001.  Maximum  ages  of  groundfishes  in  waters  off 
Alaska  and  British  Columbia  and  considerations  of  age 
determination.     Alaska  Fish.  Res.  Bull.  8(1):12-21. 

NOAA  (National  Oceanic  and  Atmospheric  Administration  I. 

2001.     Magnuson-Stevens  Act  Provisions;  Fisheries  off 
West  Coast  states  and  in  the  western  Pacific;  Pacific  coast 
groundfish  fishery;  annual  specifications  and  manage- 
ment measures.     Federal  Register  66:8,  2338-2355. 
Pawson,  M.  G. 

1990.     Using  otolith  weight  to  age  fish.     Fish.  Biol.  36:521- 
531. 
Pilling,  G.  M.,  E.  M.  Grandcourt,  and  G.  P.  Kirkwood. 

2003.     The  utility  of  otolith  weight  as  a  predictor  of  age 
in  the  emperor  Lethrinus  mahsena  and  other  tropical 
fish  species.     Fish.  Res.  60:493-506. 
Rogers,  J.  B„  R.  D.  Methot,  T.  L.  Builder,  K.  Piner,  and 
M.  Wilkins. 

2000.  Status  of  the  darkblotched  rockfish  I  Sebastes  cra- 


meri)  resource  in  2000.  In  Appendix  to  the  status  of 
the  Pacific  Coast  groundfish  fishery  through  2000  and 
recommended  acceptable  biological  catches  for  2001: 
stock  assessment  and  fishery  evaluation,  79  p.  Pacific 
Fishery  Management  Council,  2130  SW  Fifth  Ave.,  Suite 
224,  Portland,  OR  97201. 

Smith,  J.  N.,  R.  Nelson,  and  S.  E.  Campana. 

1991.  The  use  of  Pb-210/Ra-226  and  Th-228/Ra-228  dis- 
equilibria  in  the  ageing  of  otoliths  of  marine  fish  In 
Radionuclides  in  the  study  of  marine  processes  (P.  J. 
Kershaw  and  D.  S.  Woodhead.  eds.l.  p.  350-359.  Else- 
vier Applied  Science,  New  York,  NY. 

Wang,  C.  H.,  D.  L.  Willis,  and  W.  D.  Loveland. 

1975.  Radiotracer  methodology  in  the  biological,  envi- 
ronmental, and  physical  sciences,  480  p.  Prentice  Hali. 
Englewood  Cliffs,  NJ. 

Williams,  E.  H.,  A.  D.  MacCall,  S.  V.  Ralston,  and 
D.  E.  Pearson. 

2000.  Status  of  the  widow  rockfish  resource  in  2000.  In 
Appendix  to  the  status  of  the  Pacific  Coast  ground- 
fish fishery  through  1999  and  recommended  accept- 
able biological  catches  for  2000,  75  p.  Pacific  Fishery 
Management  Council,  2130  SW  Fifth  Ave..  Suite  224, 
Portland,  OR  97201. 

Williams.  E.  K„  and  S.  R.  Ralston. 

2002.  Distribution  and  co-occurrence  of  rockfishes 
(family:  Sebastidael  over  trawlable  shelf  and  slope  habi- 
tats of  California  and  southern  Oregon.  Fish.  Bull. 
100:836-855. 

Wilson.  C.  D..  and  G.  W.  Boehlert. 

1990.  The  effects  of  different  otolith  ageing  techniques 
on  estimates  of  growth  and  mortality  for  the  splitnose 
rockfish,  Sebastes  diploproa,  and  canary  rockfish,  S. 
pmniger.     Calif.  Fish  Game  76:146-160. 

Yoklavich,  M.  (ed). 

1998.  Marine  harvest  refugia  for  West  Coast  rockfish:  a 
workshop,  162  p.     NOAA-Tech.  Memo.  NMFS-SWFSC-255. 


723 


Abstract— Larval  and  juvenile  devel- 
opment of  finescale  menhaden  (Bre- 
voortia  gunteri)  is  described  for  the 
first  time  by  using  wild-caught  indi- 
viduals from  Nueces  Bay,  Texas,  and 
is  compared  with  larval  and  juvenile 
development  of  co-occurring  gulf 
menhaden  (B.  patronus).  Meristics. 
morphometries,  and  pigmentation  pat- 
terns were  examined  as  development 
proceeded.  An  illustrated  series  of 
finescale  menhaden  is  presented  to 
show  changes  that  occurred  during 
development.  For  finescale  menhaden, 
transformation  to  the  juvenile  stage 
was  completed  by  17-19  mm  standard 
length  (SL).  By  contrast,  transfor- 
mation to  the  juvenile  stage  for  gulf 
menhaden  was  not  complete  until  2.3- 
25  mm  SL.  Characteristics  useful  for 
separating  larval  and  juvenile  fines- 
cale menhaden  from  gulf  menhaden 
included  1)  the  presence  or  absence  of 
pigment  at  the  base  of  the  insertion  of 
the  pelvic  fins;  2)  the  standard  length 
at  which  medial  predorsal  pigment 
occurs;  3)  differences  in  the  number 
of  dorsal  fin  ray  elements;  and,  4)  the 
number  of  vertebrae. 


Descriptions  of  larval,  prejuvenile,  and 

juvenile  finescale  menhaden 

(Brevoortia  gunteri)  (family  Clupeidae), 

and  comparisons  to  gulf  menhaden  (B.  patronus) 


James  M.  Tolan 

Texas  Parks  and  Wildlife  Department 

6300  Ocean  Dr.,  NRC  2501 

Corpus  Christ!,  Texas  78412 

E-mail  address:  James  Tolan  @tpwd. state. tx. us 


David  A.  Newstead 

Center  for  Coastal  Studies 
Texas  A&M  University-Corpus  Christi 
6300  Ocean  Dr.,  NRC  3216 
Corpus  Christi,  Texas  78412 


Manuscript  submitted  9  September  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
14  June  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:723-732  (20041. 


Finescale  menhaden  {Brevoortia  gun- 
teri Hildebrand),  one  of  three  recog- 
nized species  of  menhaden  (Reintjes, 
1969;  Hettler.  1984)  found  in  the  Gulf 
of  Mexico,  occurs  in  the  northern  and 
western  Gulf  of  Mexico,  from  Chande- 
leur  Bay,  Louisiana,  to  Campeche  Bay, 
west  of  Punto  Morros  (McEachran 
and  Fechhelm,  1998).  Despite  their 
common  occurrence  in  coastal  and 
estuarine  waters  along  the  Texas  and 
Mexico  coasts  (Simmons,  1957;  Helher, 
1962;  Hoese,  1965;  Whitehead,  1985; 
Castillo-Rivera  and  Kobelkowsky, 
2000),  their  early  development  has 
not  been  described.  Early  develop- 
ment of  gulf  menhaden  (B.  patronus 
Goode),  on  the  other  hand,  has  been 
well  described  (Suttkus,  1956;  Hettler, 
1984;  Ahrenholz,  1991). 

In  coastal  waters  of  the  western 
Gulf  of  Mexico,  finescale  menhaden 
are  spatially  and  temporally  sym- 
patric  with  gulf  menhaden  (Castillo- 
Rivera  et  al.,  1996).  Gulf  menhaden 
are  found  throughout  the  northern 
gulf  from  Florida  Bay  to  Campeche 
Bay.  Yellowfin  menhaden  \B.  smithi 
Hildebrand)  are  found  in  the  eastern 
gulf  from  the  Mississippi  River  Delta 
to  Cape  Lookout,  North  Carolina, 
and  co-occur  with  finescale  menha- 
den only  in  its  extreme  western  range 
(Dahlberg,  1970;  Hoese  and  Moore, 
1977 ).  A  large  amount  of  hybrid  intro- 
gression  occurs  between  gulf  and  yel- 
lowfin menhaden,  although  finescale 


hybrids  (either  finescalexgulf  menha- 
den or  finescalexyellowfin  menhaden) 
have  not  been  reported  (Ahrenholz, 
1991). 

Both  finescale  and  gulf  menhaden 
are  estuarine-dependent  species  in- 
habiting shallow  nursery  areas  for 
their  early  development  (Gunter, 
1945;  Shaw  et  al.,  1985;  Castillo-Ri- 
vera and  Kobelkowsky,  2000).  Gulf 
menhaden  are  intermittent  or  multi- 
ple spawners  (Christmas  and  Waller, 
1975;  Lewis  and  Roithmayr,  1981), 
and  adults  move  offshore  in  late  sum- 
mer and  early  fall.  Spawning  off  the 
coast  of  Texas  is  protracted,  and  the 
spawning  season  begins  at  the  end  of 
August  and  continues  through  April 
(Shaw  et  al..  1985).  Estuarine  immi- 
gration of  gulf  menhaden  ranging  in 
size  from  10  to  32  mm  TL  has  been 
observed  from  late  October  through 
April  (Copeland,  1965:  Gallaway  and 
Strawn,  1974;  Allshouse,  1983).  In 
Nueces  Bay,  the  greatest  densities  of 
gulf  menhaden  larvae  are  seen  from 
late  February  to  early  May,  and  the 
peak  immigration  of  19-26  mm  TL 
individuals  occurs  from  late  April  and 
early  May  (Newstead,  2003).  Fines- 
cale menhaden  spawn  in  estuarine  or 
nearshore  areas  (Gunter,  1945;  Sim- 
mons, 1957)  and  their  spawning  sea- 
son has  been  reported  from  November 
to  March  (Ahrenholz,  1991).  Hellier 
(1962)  reported  25-mm-TL  specimens 
taken  from  the  Upper  Laguna  Madre 


724 


Fishery  Bulletin  102(4) 


on  the  lower  Texas  coast  during  February,  and  Gunter 
(1945),  Simmons  (1957),  and  Hoese  (1965)  have  reported 
postlarval  finescale  menhaden  from  the  middle  and  low- 
er Texas  coasts  from  January  to  May.  Gulf  menhaden 
have  received  considerable  attention  in  fishery  science 
because  of  their  large  population  sizes  and  resulting 
ecological  and  economic  importance  in  the  northern 
Gulf  of  Mexico  (Nelson  and  Ahrenholz,  1981;  Smith, 
19911,  whereas  finescale  menhaden  are  less  numerous 
and  not  directly  sought  by  any  recognized  fishery  (Ahr- 
enholz, 1991).  Our  study  describes  for  the  first  time  the 
development  of  postflexion  (late  larval),  prejuvenile,  and 
juvenile  finescale  menhaden. 


Materials  and  methods 


squares  regression  techniques  (SigmaPlot,  version  5.0, 
SPSS  Inc.,  Chicago,  IL)  in  order  to  graphically  illustrate 
any  development  differences  between  the  two  species. 
Increasing  ratios  (BD/SL,  CP/SL,  and  EYE/SL)  were 
described  with  an  exponential  rise-to-maximum  equation: 


y  =  a(l 


-bx 


), 


whereas,  the  decreasing  ratio  of  PAL/SL  measurements 
were  described  with  a  exponential  decay  equation: 


y  =  ae 


-bx 


In  both  equations,  y  =  body  proportion  ratio;  .r  =  SL: 
a  =  intercept;  and  b  =  SL  specific  exponential  rate  of 
change. 


A  total  of  170  wild-caught  finescale  menhaden  larvae 
and  juveniles  were  used  to  describe  early  development. 
All  specimens  came  from  ichthyoplankton  collections 
in  Nueces  Bay,  Texas  (27.87°N,  97.5TW),  during  May 
and  June  2003.  Individuals  were  collected  in  the  tidal 
channels  of  Nueces  Delta  with  a  side-mounted  push  net 
(60-cm  ring  net,  0.505-mm  mesh).  For  comparison,  357 
wild-caught  gulf  menhaden  larvae  and  juveniles  col- 
lected during  May  and  June  of  1999,  2000,  and  2002 
from  two  nearby  stations  outside  the  delta  (less  than  1.5 
km  away),  in  addition  to  the  tidal  channel  collections  of 
2003.  were  also  studied.  All  individuals  were  initially 
fixed  in  either  10%  formalin  or  95%  ethanol  and  trans- 
ferred to  fresh  95%  ethanol  after  48  hours. 

Pigment  patterns  were  recorded  and  specimens  of  fin- 
escale menhaden  were  illustrated.  Gulf  menhaden  were 
not  illustrated  because  the  figures  in  Hettler  (1984)  are 
adequate. 

Morphometries 

Body  measurements  were  made  to  the  nearest  0.1  mm 
with  an  ocular  micrometer  fitted  to  a  dissecting  micro- 
scope. All  individuals  collected  were  postflexion,  preju- 
venile, or  juvenile  stage  as  defined  in  Leis  and  Rennis 
(1983),  and  standard  length  (SL)  was  measured  as  the 
distance  from  the  tip  of  the  snout  along  the  midline  to 
a  vertical  line  through  the  posterior  edge  of  the  hypural 
plate.  All  lengths  are  SL  unless  otherwise  noted.  Defini- 
tions and  other  terms  are  as  follows: 

BD  =  body  depth;  vertical  depth  at  the  pectoral  sym- 
physis. 

CP  =  caudal  peduncle;  horizontal  distance  from  the 
posterior  edge  of  the  dorsal  fin  base  to  the  pos- 
terior edge  of  the  hypural  plate. 
EYE  =  eye  diameter;  horizontal  distance  between  the 
anterior  and  posterior  edges  of  the  fleshy  orbit. 
PAL  =  preanal  length;  distance  from  the  tip  of  the 
snout  to  the  origin  of  the  anal  fin.  measured 
along  the  midline. 

Ratios  of  these  four  body  proportion  measurements  in 
relation  to  SL  were  fitted  by  means  of  nonlinear  least 


Meristics 

Each  specimen  was  examined  to  determine  whether 
scale  formation  had  been  initiated,  and  a  total  count 
of  ventral  scutes  for  specimens  in  which  they  were  suf- 
ficiently developed  was  obtained.  A  total  of  37  finescale 
and  48  gulf  menhaden  from  the  2003  collections  were 
cleared  and  stained  according  to  Potthoff  ( 1984)  and 
used  for  fin-ray  and  vertebrae  counts.  Because  of  the  dif- 
ficulty in  accurately  counting  myomeres  in  transforming 
clupeids  (Hettler,  1984;  Ditty  et  al.,  1994),  we  chose  to 
count  total  vertebrae  and  use  the  number  of  postdorsal 
and  preanal  vertebrae  instead  of  postdorsal  and  preanal 
myomeres  as  a  potential  diagnostic  character.  Fin-ray 
counts  included  dorsal,  anal,  and  caudal  fins  (both  prin- 
cipal and  procurrent  rays). 


Results 

Morphological  development 

Finescale  menhaden  larvae  were  first  collected  at  9.7  mm 
and  ranged  to  22.5  mm  as  transforming  juveniles  (Fig. 
1).  Transformation  from  the  larval  to  the  juvenile  form 
began  around  14  mm  and  was  completed  by  around  20 
mm  (Fig  2).  Ratios  of  body  depth,  caudal  peduncle,  and 
eye  diameter  all  increased  in  relation  to  standard  length 
as  larvae  grew,  whereas  snout-to-anal  length  decreased 
i  Table  1).  The  decrease  in  snout-to-anal  length  reflected 
the  transformation  from  the  elongate  fusiform  shape  of 
the  larvae  to  the  laterally  compressed  deep-bodied  shape 
of  the  juvenile.  Scales  began  to  form  at  around  15  mm  on 
the  caudal  peduncle  region  and  progressed  forward  along 
the  ventral  and  lateral  surfaces  towards  the  dorsal  sur- 
face. None  of  the  individuals  examined  had  the  enlarged 
and  fringed  median  scales  preceding  the  dorsal  fin, 
which  are  an  adult  characteristic  of  the  genus  Brevoor- 
tia.  Ventral  scutes  also  began  forming  around  15  mm, 
and  the  full  complement  of  27-31  scutes  (McEachran 
and  Fechhelm,  1998)  was  found  by  19  mm. 

Gulf  menhaden  ranged  from  11.7  mm  as  larvae  to 
40.4  mm  juveniles.  For  gulf  menhaden,  body  depth. 


Tolan  and  Newstead:  Larval  and  |uvenile  development  of  Brevoortia  gunten 


725 


Table  1 

Proportional  measurements  in 
development. 

relation  to  standard  length 

(SL)  used  to  describe 

finescale 

menhaden  {Brevoortia 

gunteri)  larval 

Length  class 
(mm,  SL) 

Number  of 
specimens 

Body  depth: 
SL 

Preanal  length: 
SL 

Caudal  peduncle: 
SL 

Ey 

3  diameter: 
SL 

<11.0 

1 

0.100 

0.804 

0.256 

0.054 

11.1-12.0 

2 

0.108 

0.798 

0.259 

0.059 

12.1-13.0 

5 

0.119 

0.778 

0.251 

0.059 

13.1-14.0 

15 

0.139 

0.764 

0.263 

0.064 

14.1-15.0 

28 

0.136 

0.756 

0.269 

0.062 

15.1-16.0 

25 

0.168 

0.735 

0.281 

0.074 

16.1-17.0 

23 

0.208 

0.711 

0.302 

0.082 

17.1-18.0 

17 

0.220 

0.705 

0.311 

0.085 

18.1-19.0 

3 

0.239 

0.700 

0.327 

0.080 

19.1-20.0 

2 

0.219 

0.700 

0.304 

0.078 

>20.1 

1 

0.318 

0.690 

0.304 

0.093 

caudal  peduncle,  and  eye  diameter  ratios  all  similarly 
increased  in  relation  to  standard  length  as  larvae  grew, 
whereas  snout-to-anal  length  decreased  (Table  2).  Scale 
initiation  in  gulf  menhaden  was  not  seen  until  19  mm, 
and  ventral  scutes  did  not  begin  forming  until  around 
18  mm.  The  full  complement  of  scutes  (28-32  scutes; 
McEachran  and  Fechhelm,  1998)  was  seen  by  around 
25  mm.  No  enlarged  median  dorsal  scales  were  noted 
from  the  gulf  menhaden  individuals  examined. 

With  little  overlap  in  the  15-20  mm  size  range  (see 
Fig.  1)  and  a  limited  number  of  juvenile-size  finescale 
menhaden  (SL>20  mm),  it  was  not  possible  to  effectively 
separate  finescale  and  gulf  menhaden  morphometrically 
on  the  basis  of  BD:SL,  PAL:SL,  CP:SL,  and  EYE:SL 
ratios  (Fig  3).  By  25  mm,  proportional  body  measure- 
ments had  become  nearly  constant  for  gulf  menhaden 
whereas  body  measurements  were  still  changing  for 
finescale  menhaden  even  though  they  appeared  to  be 
fully  transformed.  For  a  fish  of  given  size,  finescale 
menhaden  typically  had  a  greater  body  depth,  a  shorter 
preanal  length,  and  a  greater  caudal  peduncle  length 
than  gulf  menhaden. 

Meristic  features 

No  recently  hatched  or  preflexion  finescale  menhaden 
were  examined  and  all  postflexion  individuals  followed 
the  fin  development  sequence  identified  for  other  clupe- 
ids  (Houde  et  al.,  1974:  Hettler,  1984;  Ditty  et  al.,  1994). 
The  caudal  and  dorsal  fins  are  first  to  develop,  followed 
by  the  pelvic  fins,  whereas  the  pectoral  fins  are  the 
last  to  fully  develop  even  though  the  pectorals  are  the 
first  fins  to  form  as  nonrayed  buds.  Only  vertebrae  and 
dorsal-fin  ray  counts  were  useful  in  separating  finescale 
and  gulf  menhaden,  because  most  other  meristics  over- 
lapped (Table  3).  Finescale  menhaden  had  fewer  total 
vertebrae  (  =  43  vs.  46)  and  fewer  dorsal-fin  rays  (median 


125 

100 
75 

B.  gunteri 
n=170 

50 

25 

§     o 

O 

J" 

^^ 

25 

50 

S.  patronus 
n=357 

75 

100 

— 

i 

i                      i                      i 

5                        15                      25                      35                      45 

Body  length  (mm) 

Figure  1 

Length-frequency  histograms  for  Brevoortia  gunteri 

and  B.  patronus  derived  from  ichthyoplankton  col- 

lections in  Nueces  Bay,  Texas,  during  May  and  June 

(1999-2003). 

value  =  18  vs.  21)  than  gulf  menhaden.  Postdorsal  and 
preanal  vertebrae  also  showed  a  high  degree  of  overlap 
between  the  two  species  (Table  3).  The  forward  move- 
ment of  the  anal  fin  in  relation  to  the  dorsal  fin  was 
most  evident  in  fully  transformed  gulf  menhaden,  and 
the  number  of  postdorsal-preanal  vertebrae  decreased 
from  4  to  -3.  The  relative  placement  of  the  anal  fin  also 


726 


Fishery  Bulletin  102(4) 


i'liiiiiiin'illftiiiiumi 


13.7  mm 


5.0  mm 


Figure  2 

Developmental  stages  of  Brevoortia  gunteri.  iAi  Postflexion  larva,  13.7  mm.  (B)  Postflexion 
larva,  15.0  mm.  (C)  Transforming  larva,  17.2  mm.  (D)  Transforming  larva,  19.0  mm.  (Ei 
Transformed  juvenile,  23.9  mm. 


Table  2 

Proportional 

measurements  relative 

to  standard  length 

(SL 

)  used  to  describe 

gulf  menhaden  {Brevoortia 

patronus)  larval 

development. 

Length  class 

Number  of 

Body  depth: 

Preanal  length: 

Caudal  peduncle: 

E 

ve  diameter: 

(mm,  SL) 

specimens 

SL 

SL 

SL 

SL 

<14.0 

1 

0.086 

0.829 

0.229 

0.061 

14.1-15.0 

1 

0.119 

0.786 

0.238 

0.061 

15.1-16.0 

3 

0.136 

0.764 

0.261 

0.063 

16.1-17.0 

4 

0.136 

0.764 

0.261 

0.063 

17.1-18.0 

5 

0.144 

0.749 

0.257 

0.065 

18.1-19.0 

13 

0.169 

0.733 

0.265 

0.067 

19.1-20.0 

26 

0.183 

0.725 

0.274 

0.074 

20.1-21.0 

22 

0.176 

0.735 

0.271 

0.068 

21.1-22.0 

17 

0.236 

0.719 

0.282 

0.084 

22.1-24.0 

9 

0.286 

0.709 

0.298 

0.095 

24.1-26.0 

2 

0.337 

0.730 

0.313 

0.103 

26.1-29.0 

11 

0.352 

0.730 

0.299 

0.098 

29.1-32.0 

4 

0.356 

0.740 

0.308 

0.100 

>32.1 

3 

0.380 

0.736 

0.294 

0.097 

Tolan  and  Newstead:  Larval  and  |uvenile  development  of  Brevoortia  gunten 


727 


gSSgSSgHJfMISHp 


Figure  2  (continued) 


Table  3 

Meristies  in  finescale  menhaden,  Brevoortia  gunteri,  (37  specimens)  and  in  gulf  menhaden.  B.  patronus,  (48  specimens).  Median 
values  are  given  in  parentheses. 


B.  gunteri 

B.  patronus 

Number  in 

full 

complement 

Meristic 

B.  gunteri 

B.  patronus 

Caudal-fin  rays 

Principal      (dorsal) 

10 

10 

10-11 

(ventral  l 

9 

9 

9-10 

Procurrent  (dorsal) 

7-9(8) 

8-9(8) 

8-9 

(ventral  l 

7-8(7) 

5'-8(7) 

7-8 

Dorsal-fin  rays 

17-20  ( 18 ) 

16'-23(21) 

17-20 

21-23 

Anal-fin  rays 

18-24(22) 

20-23(21) 

20-25 

18-22 

Vertebrae 

43-44(43) 

40'-46(46) 

42-44 

45-46 

Postdorsal  and  preanal 

vertebrae 

0-3(2) 

-32-4(l) 

2-43 

1  B.  patronus  larva  SL  =  11.7  mm. 

2  B.  patronus  fully  transformed  individual  SL  =  40.4  mm. 

3  Postdorsal  and  preanal  myomere  counts  in  larvae  10-15  mm  SL  (Ditty  et  al.,  1994) 


728 


Fishery  Bulletin  102(4) 


A 

0.4    - 

°        s^<r^^~°°~ 

(Q      0.3   - 

Q. 
0 

"D 

O      02   - 

m 

/          0(8,         y/ 

A^A   A      ®5Ccb0 

0  1   - 

a*  *aA     / 

a    o         y 

0  0     — i 1 1 ■ 1 1 1 ■ 1 1 1 1 r 

10                     15                     20                     25                     30                     35                     40 

090  i 

B 

0.85   - 

a       B  gunteri 

d                                                                            o      e.  patronus 

_J 

W 

§    080   - 

D) 

c 

QJ 

ra 

ra    0  75  - 

CL 

0  70    - 

a       a\ 

A  \aa  A 

^#V          «o    j^  ^e                                  O         O  _ 

aA^\^  °  °    ^  %                 o    o      o  ° 

a^a|1o°£     0<<?0                o 

A 

0  65   - 

i                                     i                                     1                                     1                                     1                                     1                                     1 

10                     15                     20                     25                     30                     35                     40 

Body  length  (mmj 

Figure  3 

Morphometric  comparisons  (shown  as  a  percentage  of  standard  length 

[SL1 1  for  wild-caught  Brcvoortia  gunteri  and  B.  patronus.  (A)  BD/SL. 

1B1  PAL/SL.  (C)  CP/SL.  (Dl  EYE/SL.  Best  fit  nonlinear  regression 

lines  were  fitted  by  least  squares  estimation. 

changed  in  finescale  menhaden,  although  the  number 
of  postdorsal-preanal  vertebrae  only  decreased  from 
three  to  zero. 

Pigmentation 

Early  pigmentation  patterns  in  finescale  menhaden  (Fig. 
2)  were  similar,  but  not  identical,  to  the  pigmentation 
described  for  gulf  menhaden  (Suttkus,  1956;  Hettler, 
1984).  Both  dorsal  and  ventral  notochord  tip  pigment, 
which  are  diagnostic  for  the  genus  Brevoortia  (Fig.  2), 
were  found  in  all  individuals  examined.  In  specimens 
<14  mm,  pigmentation  was  sparse  and  found  primarily 
along  the  ventral  margin  of  the  caudal  peduncle,  the 
base  of  the  anal  fin,  at  the  end  of  the  gut  near  the  vent. 


and  ventrally  as  two  lines  beginning  at  the  pectoral  fin 
bases  below  the  foregut.  Along  the  dorsal  margin  of  the 
hindgut,  3-6  fine  melanophores  were  usually  present. 
At  the  base  of  the  pelvic  fins,  1-2  small,  paired  stellate 
melanophores  were  present.  Additionally,  all  individuals 
had  a  medial  melanophore  along  the  isthmus  (ventral 
midline  anterior  to  the  cleithrum)  and  most  had  an 
internal  melanophore  at  the  nape.  Other  pigment  pres- 
ent in  the  smallest  finescale  larvae  included  a  series  of 
paired  melanophores  anterior  to  the  dorsal  fin  base  (seen 
in  26^  of  the  larvae  examined).  This  predorsal  mid-line 
pigment  series  increased  both  in  size  and  number  as  the 
larvae  grew.  The  head  was  unpigmented. 

By  16  mm,  pigment  increased  along  the  dorsal  sur- 
face of  the  hindgut,  the  base  of  the  dorsal  fin,  and  the 


Tolan  and  Newstead:  Larval  and  |uvenile  development  of  Brevoortia  gunten 


729 


gunteri 
patronus 


D 


gunten 
patronus 


20  25  30 

Body  length  (mm) 
Figure  3  (continued) 


base  of  the  anal  fin  (Fig.  2).  The  predorsal  midline 
melanophores  series  was  more  prominent  (51%  of  the 
individuals  displayed  this  pigment  pattern).  Additional 
pigmentation  included  paired  lateral  melanophores  near 
the  dorsal  region  of  the  brain  cavity,  internal  pigment 
above  the  notochord  from  the  posterior  margin  of  the 
base  of  the  dorsal  fin  to  the  caudal  fin,  and  internal 
pigment  over  the  gas  bladder.  A  large  melanophore  was 
also  present  above  the  base  of  the  pectoral  fin. 

By  18  mm,  the  dorsal  surface  of  the  head  became 
highly  pigmented  with  up  to  25  small  melanophores. 
The  snout,  lower  jaw,  and  pelvic  fins  were  also  pigment- 
ed by  this  size.  The  dorsal  and  ventral  surfaces  of  the 
caudal  peduncle,  as  well  as  the  medial  predorsal  region 
became  more  densely  pigmented.  Outstanding  features 
at  this  size  included  2-9  ventrolateral  melanophores  in 
a  series  along  the  digestive  tract  level  with  the  pectoral 
fins,  and  isthmus  pigment  was  separated  into  two  spots 


in  many  individuals.  By  20  mm,  the  head  region  was 
heavily  pigmented  and  the  mid-line  predorsal  pigment 
progressed  fully  to  the  head.  Ventrolaterally,  10-30  me- 
lanophores forming  a  triangular  pattern  covered  much 
of  the  digestive  tract.  New  pigmentation  features  at  this 
size  included  1-3  small  melanophores  at  the  base  of  the 
eye  and  an  internal  series  below  the  notochord  from  the 
base  of  the  dorsal  fin  to  the  caudal  fin.  Pigmentation 
at  the  pelvic  fin  insertion  was  lost  in  some  specimens, 
although  it  was  still  visible  internally  in  all  but  one  of 
the  cleared  and  stained  specimens.  Isthmus  pigmenta- 
tion was  also  lost  at  this  size. 

Living  juveniles  (>20  mm)  are  silvery  in  color  over 
most  of  the  body.  The  head,  back,  and  dorsal  and  cau- 
dal fins  are  all  pigmented.  In  preservation,  two  dark, 
slash-like  pigments  spots  were  present  on  the  posterior 
lateral  body  above  and  below  the  urostyle.  In  nearly 
all  other  aspects,  juvenile  finescale  menhaden  closely 


730 


Fishery  Bulletin  102(4) 


resembled  adults  by  this  size.  No  humeral  spots  were 
noted  in  the  two  individuals  examined. 


Discussion 

Finescale  menhaden  larvae  resemble  the  larvae  of  other 
clupeids  (Houde  and  Fore,  1973;  Jones  et  al.,  1978; 
McGowan  and  Berry,  1983;  Hettler,  1984:  Ditty  et  al., 
1994)  in  having  elongate,  slender  bodies,  light  pigmen- 
tation, and  a  small  head  lacking  spines.  They  have  a 
long,  straight  gut,  often  with  striations  along  the  hind- 
gut,  posteriorly  placed  dorsal  and  anal  fins,  and  the 
vent  is  always  posterior  to  the  dorsal  fin  base  (Jones 
et  al.,  1978).  Hettler  (1984)  discussed  the  separation  of 
individual  species  of  Brevoortia,  and  Ditty  et  al.  (1994) 
presented  a  synopsis  of  characters  to  separate  clupeid 
larvae  (<15  mm)  based  on  meristic,  morphometries, 
and  pigmentation.  Finescale  menhaden  have  43-44 
vertebrae,  whereas  gulf  menhaden  have  44-46.  Yel- 
lowfin  menhaden  are  reported  to  have  45-47  vertebrae 
(Houde  and  Swanson,  1975).  The  number  of  vertebrae, 
which  should  approximate  the  number  of  myomeres 
in  larvae  much  smaller  than  those  collected  in  this 
study,  in  conjunction  with  pigment  differences  have  been 
shown  to  be  useful  in  separating  clupeid  species  com- 
plexes (Ditty  et  al.,  1994).  In  the  western  gulf,  counts  of 
43-44  vertebrae  (=myomeres)  would  separate  finescale 
menhaden  from  other  clupeid  larvae  such  as  Sardinella 
aurita  (45-47  vertebrae;  Ditty  et  al.,  1994),  Etrumeus 
teres  (48-50  vertebrae;  Fahay,  1983),  and  Opisthonema 
oglinum  (45-46  vertebrae;  Richards  et  al.,  1974).  Spe- 
cies from  the  western  gulf  with  similar  vertebral  counts 
(Harengula  jaguana,  39-42;  Houde  et  al.,  1974;  and 
Jenkinsia  lamprotaenia,  39-42;  Powles,  1977)  can  be 
distinguished  from  finescale  menhaden  by  their  larger 
PAL:SL  ratio  (>85%  at  15  mm  for  Harengula  vs.  <85l7c 
for  finescale  menhaden.  Table  1)  and  fewer  anal  rays 
(13-14  for  Jenkinsia  vs.  18-24  for  finescale  menhaden, 
Table  3).  Although  vertebral  counts  were  used  suc- 
cessfully in  distinguishing  finescale  menhaden  from 
gulf  menhaden,  the  time  necessary  to  clear  and  stain 
larvae  makes  this  method  impractical  for  distinguishing 
between  large  numbers  of  menhaden. 

In  larval  and  prejuvenile  stages,  finescale  and  gulf 
menhaden  are  morphologically  very  similar.  Propor- 
tional body  measurements  overlapped  too  greatly  to 
reliably  distinguish  the  two  species.  Only  the  presence 
of  medial  predorsal  pigment  prior  to  transformation, 
stellate  melanophores  at  the  pelvic  fin  base,  and  the 
size  at  transformation  were  useful  characters  in  distin- 
guishing the  two  species.  Hettler  (1984)  noted  that  gulf 
menhaden  lack  paired  melanophores  in  the  predorsal 
region  until  initiation  of  transformation.  Pigment  at  the 
pelvic  fin  base  appears  to  be  a  diagnostic  character  for 
the  small-scale  menhadens  because  Houde  and  Swan- 
son  (1975)  also  reported  a  similar  feature  in  yellowfin 
menhaden  as  small  as  12.3  mm.  Although  the  presence 
of  this  pigment  at  the  pelvic  fin  base  is  proposed  to 
be  diagnostic  for  the  small-scale  menhadens  (present 


study),  Hettler's  (1984)  illustration  of  a  16.5-mm  gulf 
menhaden  shows  this  pigment.  Pigmentation  descrip- 
tions for  developing  gulf  menhaden  have  not  specifi- 
cally addressed  melanophores  at  the  pelvic  fin  insertion 
(Hettler,  1984).  Finescale  menhaden  transform  at  a 
smaller  size  (17-19  mm)  than  any  of  the  other  Gulf  of 
Mexico  menhadens.  Gulf  menhaden  did  not  complete 
transformation  until  around  25  mm,  which  is  in  agree- 
ment with  the  reported  lengths  of  25-28  mm  for  both 
laboratory  reared  and  wild-caught  individuals  ( Suttkus. 
1956;  Hettler.  1984).  Yellowfin  menhaden  reach  trans- 
formation at  an  intermediate  size  (20-23  mm;  Houde 
and  Swanson,  1975). 

Even  as  adults,  finescale  menhaden  very  closely  re- 
semble gulf  menhaden  (Hoese  and  Moore.  1977)  and  few 
reliable  characters  effectively  separate  them.  Only  the 
absence  of  striations  on  the  margin  of  the  operculum,  a 
single  humeral  spot  (with  no  hint  of  trailing  spots  along 
the  lateral  margins  I,  and  more  scale  rows  (60-77  in  fi- 
nescale vs.  36-50  in  gulf  menhaden;  Hoese  and  Moore, 
1977;  McEachran  and  Fechhelm,  1998)  distinguish  fin- 
escale from  gulf  menhaden.  All  other  meristics  overlap 
greatly;  i.e.,  counts  of  dorsal-fin  rays,  anal-fin  rays, 
pectoral-fin  rays,  pelvic-fin  rays,  gill-raker  counts,  and 
ventral  scutes.  Although  externally  similar,  significant 
differences  in  internal  structure  between  finescale  and 
gulf  menhaden  have  been  documented.  Finescale  men- 
haden have  fewer  branchiospinules  and  shorter  inter- 
mediate gill  rakers  than  gulf  menhaden  and,  as  such, 
filter  mainly  zooplankton  from  the  water  column;  gulf 
menhaden,  in  contrast,  feed  primarily  on  phvtoplankton 
and  detritus  (Castillo-Rivera  et  al.,  1996). 

Based  on  length-frequency  differences  seen  between 
the  two  species  (Fig.  1),  the  reported  spawning  season 
for  finescale  menhaden  along  the  middle  Texas  coast 
could  be  extended  to  late  May.  We  still  do  not  know  of 
characters  that  would  distinguish  finescale  menhaden 
eggs,  and  yolksac.  preflexion,  and  flexion  larvae  from 
other  species  in  the  genus  Brevoortia.  In  order  to  fully 
describe  the  development  of  finescale  menhaden,  labo- 
ratory spawning  and  rearing  experiments  are  needed 
to  fully  describe  these  early-life  stages.  Houde  (1973) 
presented  relatively  simple  rearing  techniques  that  al- 
low descriptions  of  the  developmental  stages  of  larval 
fish  (from  egg  through  transformation  of  larvae  to  the 
juvenile  stage).  These  methods  have  been  used  success- 
fully for  Atlantic,  gulf,  and  yellowfin  menhaden,  and 
presumably,  finescale  menhaden  could  be  reared  with 
these  same  techniques  if  their  eggs  could  be  obtained. 
The  rearing  of  finescale  menhaden  would  also  allow  the 
effectiveness  of  the  proposed  pigment  characters  used 
to  separate  finescale  menhaden  from  gulf  menhaden  to 
be  tested. 


Acknowledgments 

This  work  was  performed  with  funding  from  the  Coastal 
Bend  Bays  &  Estuaries  Program  under  contract  0203. 
The  Texas  Parks  and  Wildlife  Department,  Coastal 


Tolan  and  Newstead:  Larval  and  |uvenile  development  of  Brevoortia  gunteri 


731 


Studies  Program,  Resource  Protection  Division  provided 
additional  interagency  cooperation  in  the  form  of  equip- 
ment and  field  logistical  support.  This  manuscript  was 
improved  by  comments  from  two  anonymous  reviewers. 


Literature  cited 

Ahrenholz,  D.  W. 

1991.     Population  biology  and  life  history  of  the  North 
American  menhadens,  Brevoortia  spp.     Mar.  Fish.  Rev. 
53(41:3-19. 
Allshouse,  W.  C. 

1983.     The  distribution  of  immigrating  larval  and  post- 
larval  fishes  into  the  Aransas-Corpus  Christi  Bay 
complex.     M.S.  thesis,  118  p.     Corpus  Christi  State 
Univ.,  Corpus  Christi,  Texas. 
Castillo-Rivera,  M.,  and  A.  Kobelkowsky. 

2000.     Distribution  and  segregration  of  two  sympatric 
Brevoortia  species  (Teleostei:Clupeidae).     Estuar.  Coast. 
Shelf  Sci.  50:593-598. 
Castillo-Rivera,  M.,  A.  Kobelkowsky,  and  V.  Zamayoa. 

1996.     Food  resource  partitioning  and  trophic  morphology 
of  Brevoortia  gunteri  and  B.  patronus.     J.  Fish  Biol. 
49:1102-1111. 
Christmas.  J.  Y„  and  R.  S.  Waller. 

1975.     Location  and  time  of  menhaden  spawning  in  the 
Gulf  of  Mexico.  20  p.     Gulf  Coast  Res.  Lab.,  Ocean 
Springs,  MS. 
Copeland,  B.  J. 

1965.     Fauna  of  the  Aransas  Pass  Inlet,  Texas:  I.  Emigra- 
tion as  shown  by  tide  trap  collections.     Publ.  Inst.  Mar. 
Sci.  Univ.  Tex.  10:9-21. 
Dahlberg,  M.  D. 

1970.     Atlantic  and  Gulf  of  Mexico  menhadens,  genus 
Brevoortia  (Pisces:Clupeidae).     Bull.  Fla.  State  Mus., 
Biol.  Sci.  15:91-162. 
Ditty,  J.  G,  E.  D.  Houde,  and  R.  F.  Shaw. 

1994.     Egg  and  larval  development  of  Spanish  sardine, 
Sardinella  aurita  (Family  Clupeidae),  with  a  synopsis  of 
characters  to  identify  clupeid  larvae  from  the  northern 
Gulf  of  Mexico.     Bull.  Mar.  Sci.  54(21:367-380. 
Fahay,  M.  P. 

1983.  Guide  to  the  early  stages  of  marine  fishes  occur- 
ring in  the  western  North  Atlantic  Ocean,  Cape  Hat- 
teras  to  the  southern  Scotian  shelf.  J.  Northwest  Atl. 
Fish.  Sci.  4:1-423. 

Gallaway,  B.  J.,  and  K.  Strawn. 

1974.     Seasonal  abundance  and  distribution  of  marine  fish- 
eries at  a  hot-water  discharge  in  Galveston  Bay,  Texas. 
Contr.  Mar.  Sci.  18:71-137. 
Gunter,  G. 

1945.     Studies  on  marine  fishes  of  Texas.     Publ.  Inst. 
Mar.  Sci.  1(1):1-190. 
Hellier,  T.  R.,  Jr. 

1962.     Fish  production  and  biomass  studies  in  relation  to 
photosynthesis  in  the  Laguna  Madre  of  Texas.     Publ. 
Inst.  Mar.  Sci.  Univ.  Tex.  8:1-22. 
Hettler,  W.  F. 

1984.  Description  of  eggs,  larvae,  and  early  juveniles  of 
gulf  menhaden,  Brevoortia  patronus,  and  comparisons 
with  Atlantic  menhaden,  B.  tyrannus,  and  yellowfin 
menhaden,  B.  smithi.     Fish.  Bull.  82:85-95. 

Hoese,  H.  D. 

1965.     Spawning  of  marine  fishes  in  the  Port  Aransas, 


Texas,  area  as  determined  by  the  distribution  of  young 
and  larvae.     Ph.D.  diss.,  144  p.     Univ.  of  Texas,  Austin, 
TX. 
Hoese,  H.  D.,  and  R.  H.  Moore. 

1977.  Fishes  of  the  Gulf  of  Mexico:  Texas,  Louisiana, 
and  adjacent  waters.  327  p.  Texas  A&M  Univ.  Press, 
College  Station. 

Houde,  E.  D. 

1973.     Some  recent  advances  and  unsolved  problems  in 

the  culture  of  marine  fish  larvae.     Proc.  World  Mari- 

culture  Soc.  3:83-112. 
Houde,  E.  D.,  and  P.  L.  Fore. 

1973.  Guide  to  identify  eggs  and  larvae  of  some  Gulf 
of  Mexico  clupeid  fishes.  Fla.  Dep.  Nat.  Resour.  Mar. 
Res.  Lab.  Leafl.  Ser.  4  (part  1,  no.  23).  14  p. 

Houde,  E.  D.,  W.  J.  Richards,  and  V.  P  .  Sakensa. 

1974.  Descriptions  of  eggs  and  larvae  of  the  scaled  sar- 
dine, Harengula  jaguana.     Fish.  Bull.  72:1106-1122. 

Houde,  E.  D.,  and  L.  J.  Swanson  Jr. 

1975.  Descriptions  of  eggs  and  larvae  of  yellowfin  men- 
haden, Brevoortia  smithi.     Fish.  Bull.  73:660-673. 

Jones,  P.  W..  F.  D.  Martin,  and  J.  D.  Hardy  Jr. 

1978.  Development  of  fishes  of  the  Mid-Atlantic  Bight: 
an  atlas  of  egg,  larvae,  and  juvenile  stages.  Vol.  I: 
Acipenseridae  through  Ictaluridae,  340  p.  U.S.  Dep. 
Inter.,  Biol.  Serv.  Prog. 

Leis,  J.  M.,  and  D.  S.  Rennis. 

1983.     The  larvae  of  Indo-Pacific  coral  reef  fishes,  269 
p.     LTniv.  Hawaii  Press,  Honolulu,  HI. 
Lewis,  R.  M.,  and  C.  M.  Roithmayr. 

1981.     Spawning  and  sexual  maturity  of  gulf  menhaden, 
Brevoortia  patronus.     Fish.  Bull.  78:947-951. 
McEachran,  J.  D..  and  J.  D.  Fechhelm. 

1998.     Clupeidae.     In  Fishes  of  the  Gulf  of  Mexico.     Vol.1: 
Myxiniformes  to  Gasterosteiformes.  p.  328-346.     Univ. 
Texas  Press.  Austin,  TX. 
McGowan,  M.  F.,  and  F.  H.  Berry. 

1983.  Clupeiforms:  development  and  relationships.  In 
Ontogeny  and  systematics  of  fishes,  p.  108-126.  Am. 
Soc.  Ichthyol.  Herpetol.,  Special  Publ.  1 

Nelson,  W.  R.,  and  D  .W.  Ahrenholz. 

1981.     Population  and  fishery  characteristics  of  gulf  men- 
haden, Brevoortia  patronus.     Fish.  Bull.  84:311-325. 
Newstead,  D.  J. 

2003.     Larval  fish  recruitment  to  isolated  nursery  grounds 
in  Nueces  Bay,  Texas.     M.S.  thesis,  61  p.     Texas  A&M 
Univ.-Corpus  Christi,  Corpus  Christi,  TX. 
Potthoff,  T. 

1984.  Clearing  and  staining  techniques.  In  Ontogeny 
and  systematics  of  fishes,  p.  35-37.  Am.  Soc.  Ichthyol. 
Herpetol.,  Special  Publ    1 

Powles,  H. 

1977.     Description  of  larval  Jenkinsia   lamprotaenia 
(Clupeidae.  Dussumieriinaei  and  their  distribution  off 
Barbados,  West  Indies.     Bull.  Mar.  Sci.  27:788-801. 
Reintjes,  J.  W. 

1969.     Synopsis  of  the  biological  data  on  Atlantic  menha- 
den. Brevoortia  tyrannus,  30  p.     U.S.  Dep.  Inter.,  Fish 
Wildl.  Ser.,  Circ."  320. 
Richards,  W.  J.,  R.  V  Miller,  and  E.  D.  Houde. 

1974.     Egg  and  larval  development  of  the  Atlantic  thread 
herring,  Opisthonema  oglinum.     Fish.  Bull.  72:1123- 
1136. 
Shaw,  R.  F.,  J.  H.  Cowan  and  T  L.  Tillman. 

1985.  Distribution  and  density  of  Brevoortia  patronus 


732 


Fishery  Bulletin  102(4) 


(gulf  menhaden!  eggs  and  larvae  in  the  continental 
shelf  waters  of  western  Louisiana.  Bull  Mar.  Sci. 
36(11:96-103. 

Simmons.  E.  G. 

1957.  An  ecological  survey  of  the  upper  Laguna  Madre 
of  Texas.     Publ.  Inst.  Mar.  Sci.  Univ.  Tex.  4:156-200. 

Smith,  J.  W. 

1991.  The  Atlantic  and  gulf  menhaden  purse  seine  fish- 
eries: origins,  harvesting  technologies,  biostatistical 
monitoring,  recent  trends  in  fisheries  statistics,  and 
forecasting.     Mar.  Fish.  Rev.  53(41:28-41. 


Suttkus.  R.  D. 

1956.  Early  life  history  of  the  gulf  menhaden,  Brevoortia 
patronus,  in  Louisiana.  Trans.  North  Am.  Wild.  Conf. 
21:390-407. 

Whitehead,  P.J  . 

1985.  FAO  species  catalogue.  Vol.  7.  Clupeoid  fishes 
of  the  world.  An  annotated  and  illustrated  catalogue 
of  the  herrings,  sardines,  pilchards,  sprats,  anchovies 
and  wolfherrings.  Part  1:  Chirocentridae,  Clupeidae 
and  Pristigasteridae.  FAO  Fish.  Synop.  125,  vol  7,  pt 
1.  303  p.     FAO,  Rome. 


733 


Abstract-The  stomach  contents  of  the 
minimal  armhook  squid  iBerryteuthis 
anonychus)  were  examined  for  338 
specimens  captured  in  the  northeast 
Pacific  during  May  1999.  The  speci- 
mens were  collected  at  seven  stations 
between  145-165°W  and  39-49°N  and 
ranged  in  mantle  length  from  10.3 
to  102.2  mm.  Their  diet  comprised 
seven  major  prey  groups  (copepods, 
chaetognaths,  amphipods.  euphausi- 
ids,  ostracods,  unidentified  fish,  and 
unidentified  gelatinous  prey)  and  was 
dominated  by  copepods  and  chaeto- 
gnaths. Copepod  prey  comprised  four 
genera,  and  86%  by  number  of  the 
copepods  were  from  the  genus  Neo- 
calanus.  Neocalanus  cristatus  was  the 
most  abundant  prey  taxa,  composing 
50%  by  mass  and  35%  by  number  of 
the  total  diet.  Parasagitta  elegans 
(Chaetognatha)  occurred  in  more 
stomachs  (47%)  than  any  other  prey 
taxon.  Amphipods  occurred  in  19%  of 
the  stomachs  but  composed  only  5%  by 
number  and  3%  by  mass  of  the  total 
prey  consumed.  The  four  remaining 
prey  groups  (euphausiids,  ostracods. 
unidentified  fish,  and  unidentified 
gelatinous  prey)  together  composed 
<29c  by  mass  and  <1%  by  number  of 
the  diet.  There  was  no  major  change 
in  the  diet  through  the  size  range 
of  squid  examined  and  no  evidence 
of  cannibalism  or  predation  on  other 
cephalopod  species. 


Diet  of  the  minimal  armhook  squid 
(Berryteuthis  anonychus) 
(Cephalopoda:  Gonatidae) 
in  the  northeast  Pacific  during  spring 

Kazuhisa  Uchikawa 

National  Research  Institute  ot  Far  Seas  Fisheries 
5-7-1  Shimizu-Ondo 
Shizuoka,  424-8633,  Japan 
E-mail  address  stomyctS affrc  go  ip 

John  R.  Bower 

Northern  Biosphere  Field  Science  Center,  Hakodate  Branch 

Hokkaido  University 

3-1-1  Minato-cho,  Hakodate 

Hokkaido  041-8611,  Japan 

Yasuko  Sato 

Department  of  Agriculture,  Forestry  and  Fisheries,  Nngata  Prefecture 

Agriculture  Affairs  Division 

Shlnko-cho 

Nngata  950-8570,  Japan 

Yasunori  Sakurai 

Graduate  School  of  Fisheries  Sciences 
Hokkaido  University 
3-1-1  Minato-cho,  Hakodate 
Hokkaido  041-8611,  Japan 


Manuscript  submitted  2  September  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
29  June  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:733-739  12004). 


The  squid  family  Gonatidae  plays 
an  important  role  in  the  ecosystems 
of  the  North  Pacific.  In  the  Sea  of 
Okhotsk,  the  annual  production  of 
gonatid  squids  is  more  than  half  that 
offish  production  (Lapko,  1996),  and 
in  the  western  and  central  Bering 
Sea,  gonatid  production  is  thought  to 
exceed  that  of  the  dominant  fish  fami- 
lies (Radchenko,  1992).  In  the  sub- 
arctic North  Pacific,  the  gonatids  are 
an  important  link  in  the  pelagic  food 
web  iBrodeur  et  al.,  1999).  To  better 
understand  the  food  web  in  the  North 
Pacific  and  the  processes  influencing 
the  production  of  gonatid  squids  in 
this  region,  information  is  needed  on 
the  feeding  behavior  of  these  squids. 
The  minimal  armhook  squid  iBerry- 
teuthis  anonychus)  (also  known  as  the 
"smallfin  gonate  squid"  [Roper  et  al., 
1984])  is  a  small  gonatid  (maximum 
mantle  length  =  150  mm)  distributed 
mainly  in  the  northeast  Pacific  (Rop- 


er et  al..  1984:  Bower  et  al.,  2002).  It 
is  a  major  prey  for  fishes,  squids,  sea- 
birds,  and  marine  mammals  (Ogi  et 
al.,  1980;  Pearcy  et  al.,  1988;  Pearcy, 
1991;  Kuramochi  et  al.,  1993;  Pearcy 
et  al.,  1993;  Ohizumi  et  al.,  2003)  but 
is  not  targeted  by  any  fishery.  Despite 
the  importance  of  B.  anonychus  in  the 
food  web  of  the  subarctic  North  Pa- 
cific, the  only  published  reports  on 
its  feeding  behavior  are  two  abstracts 
in  the  Russian  literature  (Lapshina, 
1988;  Didenko,  1990).  In  this  arti- 
cle, we  provide  further  information 
on  the  feeding  behavior  of  B.  anony- 
chus by  describing  the  diet  of  a  wide 
size  range  of  squid  collected  from  the 
northeast  Pacific  during  late  spring. 


Methods 

Berryteuthis  anonychus  was  collected 
during  a  United   States  National 


734 


Fishery  Bulletin  102(4) 


65°N 


60°N 


55°N 


50°N 


45  N 


40°  N 


35°N 


180  W 


"I  T 

170°W         160:W         150'W         140  W         130"W         120  W 


Figure  1 

Sampling  stations  in  the  northeast  Pacific  where  Berryteuthis 
anonychus  were  collected  during  May  1999.  Numbers  indicate 
station  numbers. 


Marine  Fisheries  Service  (NMFS)  survey  of  salmon  in 
the  northeast  Pacific  (Carlson  et  al.,  1999;  Bower  et  al., 
2002).  Samples  were  collected  during  6-17  May  1999 
at  seven  stations  between  145-165°W  and  39-49°N 
(Fig.  1).  At  each  station,  a  midwater  trawl  modified  to 
fish  at  the  surface  was  towed  for  1  hour.  The  trawl  was 
198  m  long  and  had  hexagonal  mesh  in  its  wings  and 
body,  and  a  1.2-cm  mesh  liner  was  used  in  the  codend. 
Trawling  speeds  were  7-9  km/h,  and  the  average  net 
dimensions  while  fishing  were  16  m  vertical  spread  and 
45  m  horizontal  spread. 

Squid  samples  were  frozen  on  board  to  -20°C  and 
preserved  in  50%  isopropyl  alcohol  in  the  laboratory. 
The  mantle  length  (ML)  of  each  squid  was  measured 
to  the  nearest  0.1  mm,  and  each  squid  was  weighed  to 
the  nearest  0.01  g.  The  stomach  contents  of  338  squid 
(167  males,  144  females,  27  undetermined)  ranging  in 
ML  from  10.3  to  102.2  mm  (Fig.  2)  were  examined  un- 
der a  stereomicroscope.  A  total  of  359  squid  were  col- 
lected during  the  survey  (Bower  et  al.,  2002),  but  21  of 
these  specimens  were  either  damaged  or  lost,  and  thus 
excluded  from  our  analyses.  Most  prey  items  were  frag- 
mented; therefore  prey  identification  was  usually  based 
on  diagnostic  body  parts  as  described  in  Brodsky  (1950), 
Miller  (1988),  Baker  et  al.  (1990),  and  Vinogradov  et  al. 
(1996).  and  by  comparison  with  zooplankton  specimens 
collected  in  the  same  area.  The  prey  items  were  counted 
and  weighed  to  the  nearest  0.01  mg.  These  wet  mass 
measurements  presumably  underestimated  the  initial  wet 
masses  because  mass  loss  occurs  in  invertebrate  samples 
preserved  in  isopropyl  alcohol  (e.g.,  Howmiller,  1972), 
and  it  was  assumed  all  prey  taxa  were  equally  affected 
by  the  preservation.  The  numbers  of  individuals  of  each 


6CH 


50- 


40- 


Z     30- 


20- 


10-| 
0 


n  =  338 


10     20     30     40     50     60     70     80     90    100 
Mantle  length  (mm) 

Figure  2 

Length-frequency  distribution  for  Berryteuthis 
anonychus. 


prey  taxon  were  estimated  from  the  numbers  of  prey 
parts,  such  as  copepod  mandibles,  amphipod  heads  and 
chaetognath  seizing  hooks.  Because  of  the  difficulty  in 
distinguishing  the  copepods  Neocalanus  plumchrus  and 
N.  flemingeri,  they  were  grouped  as  a  single  taxon,  N. 
plumchrus+flemingeri.  Some  calanoid  copepods  that  could 
not  be  identified  to  genus  level  were  identified  as  either 
a  "specialized  form"  or  a  "generalized  form";  characters 
of  the  specialized  form  included  appendages  that  were 
greatly  enlarged  or  strongly  developed  with  chelae,  spines 
on  the  posterior  corners  of  the  terminal  thoracic  segment, 


Uchikawa  et  al.:  Diet  of  Berryteuthis  anonychus  in  the  northeast  Pacific  during  spring 


735 


Table  1 

Numbers  of  Berryteuthis  anonychus 

stomachs  with  iden- 

tifiable 

prey  remains. 

without  identifiable  remains,  and 

without 

remains  from 

the  northeast  Pacific.  Station  num- 

bers  refer  to  those  shown  in  Figure  1 

With 

Without 

Station 

identifiable 

identifiable 

Without 

no. 

remains 

remains 

remains 

Total 

1 

25 

0 

0 

25 

2 

29 

19 

45 

93 

3 

33 

0 

2 

35 

4 

12 

1 

2 

15 

5 

51 

9 

6 

66 

6 

44 

18 

16 

78 

7 

26 

0 

0 

26 

Total 

220 

47 

71 

338 

and  an  asymmetrically  swollen  genital  segment.  The  gen- 
eralized form  included  calanoid  copepods  of  the  Calanus 
type  that  did  not  share  any  of  these  characters. 

A  stomach-contents  index  (SCI,  %)  was  calculated 
as  SCI=(wet  mass  of  total  stomach  contents/wet  body 
mass)xl00.  For  each  prey  taxon,  the  percentages  by 
number  (N)  and  wet  mass  (WM)  of  the  total  prey,  and 
the  percentage  frequency  of  occurrence  (F)  were  de- 
termined. An  index  of  relative  importance  (IRI)  was 
calculated  for  each  prey  taxon  as  IRf  =  Ft  x  lNt+  WMt) 
(Pinkas  et  al.,  1971),  where  i  denotes  the  taxon.  The  IRI 
for  each  major  group  of  prey  taxa  was  then  standard- 
ized to  '/dRI  (Cortes.  1997): 

n 

%IRI,  =  100  x  IRI,  I  £  IRIt , 

where  /;  is  the  total  number  of  groups  collected. 

Copepod  mandible  size  is  directly  related  to  the  cara- 
pace length  of  several  calanoid  copepods  in  the  North 
Atlantic  (Karlson  and  Bamstedt,  1994);  therefore  man- 
dible width  was  used  as  an  indicator  of  relative  prey 
size  to  compare  copepod  prey  size  with  squid  mantle 
size.  A  total  of  87  mandibles  were  measured  from  the 
stomachs  of  10  squid  measuring  29-102  mm  ML. 


Results 

Of  the  338  stomachs  examined,  267  (79%)  contained 
prey,  and  220  (65%  )  contained  identifiable  prey  (Table  1). 
Individual  SCI  values  ranged  from  0%  to  8.0%  (station 
mean  =  1.0%).  SCI  values  varied  significantly  among 
sampling  times  (Kruskal-Wallis  test,  P<0.001),  and  the 
two  highest  SCI  values  occurred  in  the  afternoon  and 
just  after  sunset  (Fig.  3). 

The  diet  of  B.  anonychus  comprised  seven  major 
prey  groups  and  was  dominated  by  copepods  (A?=70%, 


3.0-1                                                                              T 

t             h> 

2.5- 

{        1 

2.0- 

o      15" 

CO 

1.0- 

263 

-15 

0.5" 

66 
_87                                                           78 

1      1     1     1     1      1     1     1      1 

6         8        10       12       14       16       18      20       22       24 

Time  of  day  (h) 

T 

Sunrise                                                             Sunset 

Figure  3 

Mean  stomach  contents  index  (SCI)  of  Berrvteuthis 

anonychus  collected  in  the  northeast  Pacific  during 

May  1999  at  different  times  of  day.  SE  =  standard 

error  of  the  mean.  Numbers  indicate  squid  sample 

size  for  each  sampling. 

WM=85%,  F=74%,  %IRI=87%)  and  chaetognaths 
(N=24%,  WM=11%,  F=48%,  <7dRI=129c)  (Table  2).  The 
five  other  prey  groups  (amphipods,  euphausiids,  ostra- 
cods,  unidentified  fish,  and  unidentified  gelatinous  prey) 
each  had  a  9cIRI  value  <1%. 

Copepod  prey  comprised  four  genera,  and  86%  by 
number  of  the  copepods  were  from  the  genus  Neocala- 
nus. Neocalanus  cristatus  was  the  most  abundant  prey 
taxa.  composing  50%  by  mass  and  35%  by  number  of 
the  total  diet.  The  three  Neocalanus  taxa  (Neocalanus 
spp.,  N.  plumchrus+flemingeri,  and  N.  cristatus)  com- 
posed 85%  by  mass  and  68%  by  number  of  the  diet. 
Neocalanus  cristatus  was  identified  based  on  the  pres- 
ence of  the  head  crest,  which  develops  at  the  C5  copepo- 
dite  stage  (Brodsky,  1950).  Thus,  this  taxon  comprises 
only  the  C5  and  C6  stages,  and  possible  members  of 
the  Neocalanus  spp.  taxon  include  N.  plumchrus,  N. 
flemingeri,  and  earlier  stages  (C1-C4)  of  N.  cristatus. 
Squid  >60  mm  ML  fed  mainly  on  Neocalanus  crista- 
tus (2V=39%,  WM=53%,  F=50%)  and  Neocalanus  spp. 
(iV=29%,  WM=  31%.  F=40%),  whereas  those  <60  mm 
ML  fed  mainly  on  Neocalanus  spp.  (AT=43%,  WM=53%, 
P=29%)  and  Neocalanus  plumchrus+flemingeri  (N=8%, 
WM=10%,  F=14%),  and  consumed  few  C5-C6  Neocala- 
nus cristatus  (N=4%,  WM=4%,  F=6%).  The  mandible 
size  of  copepod  prey  showed  a  clear  positive  relationship 
with  ML  (Fig.  4),  indicating  that  the  squid  fed  on  larger 
copepods  as  the  squid  grew.  Taxa  from  other  copepod 
genera  (i.e.,  Candacia,  Metridia,  and  Pleuromamma) 
composed  0.5%  of  the  total  prey  number  and  0.1%  of 
the  total  wet  mass  (Table  2). 

Parasagitta  elegans,  the  only  identified  chaetognath, 
occurred  in  more  stomachs  (47%)  than  any  other  prey 
taxon  and  in  58%  of  the  stomachs  from  squid  >60  mm 


736 


Fishery  Bulletin  102(4) 


Table  2 

Prey  items  identified  from  stomach  contents  of  Berry  teu 

this  anonychus  collected  in  the  northeast  Pacific  during  May  1999.  %IRI: 

standardized  index  of  relative  importance.  %IRI  values  in  parentheses  are 

those  for  <60 

mm 

ML  and 

>60  mm  ML  squid.  Fre- 

quency  of  occurrence  was  calculated  from  the  number 

of  stomachs  containing  food.  "- 

means  prey  taxon  was  not  present  in 

stomachs. 

Number 

Wet  mass 

Frequency  of 

%IRI 

Taxon 

(%) 

(%) 

occurrence 

(%) 

(<60  mm  ML,  >60  mm  ML) 

Copepoda 

70.2 

85.3 

74.2 

86.5(80.9,84.81 

Candacia  columbine 

0.2 

0.1 

1.9 

Candaeia  sp. 

<0.1 

<0.1 

0.4 

Metridia  paeifica 

0.2 

<0.1 

2.2 

Neoealanus  cristatus 

35.0 

50.4 

23.2 

Neocalanus  plumchrus+flemingeri 

3.1 

1.8 

12.4 

Neoealanus  spp. 

30.0 

32.3 

33.3 

Pleuromamma  spp. 

0.1 

<0.1 

1.9 

Calanoida  (generalized  form) 

0.5 

0.3 

4.9 

Calanoida  (specialized  form) 

0.1 

0.1 

0.4 

Unidentified  Calanoida 

0.9 

0.3 

14.2 

Unidentified  Copepoda 

0.1 

0.1 

2.6 

Chaetognatha 

23.9 

10.8 

47.6 

12.4(18.1,  13.9) 

Parasagitta  elegens 

23.8 

10.7 

47.2 

Unidentified  Chaetognatha 

0.1 

0.1 

1.1 

Amphipoda 

4.6 

2.5 

19.1 

1.0(1.0.  1.3) 

Hyperia  medusarum 

0.8 

0.9 

2.2 

Themisto  paeifica 

2.5 

0.9 

7.5 

Unidentified  Hyperiidae 

0.4 

0.5 

0.7 

Unidentified  Physocephalata 

<0.1 

<0.1 

0.4 

Unidentified  Hyperiidea 

0.7 

0.2 

7.5 

Unidentified  Amphipoda 

0.1 

<0.1 

1.9 

Euphausiacea 

0.5 

0.9 

4.5 

<0.1  (<0.1,0.1) 

Euphausia  paeifica 

<0.1 

0.4 

0.5 

Thysanoessa  sp. 

<0.1 

<0.1 

0.4 

Unidentified  Euphausiacea 

0.5 

0.5 

3.7 

Ostracoda 

<0.1 

<0.1 

1.1 

<0.1  (<0.1,— ) 

Unidentified  fish 

<0.1 

0.8 

0.4 

<0.1  (— ,  <0.1) 

Unidentified  gelatinous  prey 

<0.1 

<0.1 

0.4 

<0.1  (<0.1,— ) 

Unidentified  Crustacea 

0.1 

<0.1 

1.1 

Unidentified  material 

0.6 

0.1 

18.7 

Neocalanus  plumchrus  and  N.  flemingeri 

were  grouped  as  a  si 

ogle  taxon  tN.  plumchrus+flemingeri)  because  of  d i f ri c l 

Ity  in  distinguishing  these 

species  in  partly  digested  materials. 

Neocalanus  cristatus  comprises  stages  C5 

and  C6  only. 

Neocalanus  spp.  =  N.  cristatus  (stages  Cl- 

C4 ).  N.  plumchrus,  and  N.  flemingeri. 

Calanoida  (specialized  form!  =  unidentified  individuals  with  markedly  enlarged  appendages.  strongU 

deve 

loped  chel 

ae,  a  spine  on  the  posterior 

corner  of  the  terminal  thoracic  segment,  or  asymmetrically  swollen  genital  segments. 

Calanoida  (generalized  form)  =  unidentifi 

?d  Calanus-type  individuals  that  share  none 

of  the  characters  of  the  special 

zed  form. 

ML.  P.  elegans  was  the  third  most  abundant  prey  taxon, 
composing  24%  by  number  and  11%  by  mass  of  the  total 
diet  (Table  2). 

Amphipods  (mainly  Themisto  paeifica  and  Hyperia 
medusarum)  were  consumed  by  199r  of  the  squid  but 
composed  only  5%  by  number  and  3%  by  wet  mass  of 


the  total  prey  consumed.  The  four  other  prey  groups 
combined  composed  <2re  by  mass  and  <1%  by  number 
of  the  diet.  There  were  no  major  changes  in  %IR1  val- 
ues through  the  size  range  of  squid  examined  (Table  1) 
and  no  evidence  of  cannibalism  or  predation  on  other 
cephalopod  species. 


Uchikawa  et  al.:  Diet  of  Berryteuthis  anonychus  in  the  northeast  Pacific  during  spring 


737 


0,  KM  >  i — i~i — ■  i  ■ — r-1 — i  ■  i — ■  i  ■ — r-1 — i 

20    30   4(1   50   60    70    80   90  100  110 

Mantle  length  (mm) 

Figure  4 

Relationship  between  mandible  width  of 
copepod  prey  and  mantle  length  of  Ber- 
ryteuthis anonychus. 


Discussion 

The  diet  of  Berryteuthis  anonychus  collected  in  the  north- 
east Pacific  during  May  was  dominated  by  calanoid 
copepods  and  chaetognaths.  During  early  July  in  this 
area,  B.  anonychus  larger  than  those  examined  in  the 
present  study  (ML:  75-127  mm  vs.  10-102  mm)  fed  on 
a  wider  variety  of  prey,  including  primarily  calanoid 
copepods,  hyperiid  amphipods,  pteropods,  and  euphau- 
siids  (Lapshina,  1988).  Possible  causes  for  this  change 
in  diet  include  seasonal  change  in  prey  availability  and 
an  ontogenetic  change  in  the  squid's  ability  to  capture 
prey. 

The  zooplankton  composition  in  the  upper  150  m  of 
the  subarctic  North  Pacific  is  highly  seasonal.  Neocala- 
nus  copepods,  the  major  prey  of  B.  anonychus,  dominate 
the  epipelagic  zooplankton  community  during  spring 
and  early  summer  (Mackas  and  Tsuda.  1999).  They 
then  descend  from  the  upper  layer  to  spend  the  late 
summer,  autumn,  and  early  winter  at  400-2000  m,  well 
below  the  depth  range  of  B.  anonychus  (0-200  m;  Nesis, 
1997).  As  a  result  of  this  ontogenetic  descent,  the  upper 
ocean  zooplankton  biomass  decreases  greatly,  and  the 
community  is  then  dominated  by  a  different  group  of 
species.  This  group  includes  euphausiids  (Mackas  and 
Tsuda,  1999),  which  are  consumed  by  more  B.  anony- 
chus in  July  (28%;  Lapshina,  1988)  than  in  May  (5%; 
present  study).  Other  prey  that  show  a  large  increase 
in  frequency  of  occurrence  between  May  and  July  are 
amphipods  (19%  in  May,  52%  in  July)  and  pteropods 
(0%  in  May,  40%  in  July). 

Oceanic  squids  such  as  B.  anonychus  generally  feed 
on  small  crustaceans  as  juveniles  and  then  shift  their 
diet  to  larger  fish  and  other  cephalopods  as  they  grow 
(Rodhouse  and  Nigmatullin,  1996).  We  observed  no  such 
ontogenetic  shift  within  the  size  range  examined,  but 
copepod  prey  size  was  found  to  increase  with  growth. 


These  data  are  consistent  with  those  for  other  squids 
in  that  prey  size  increases  during  development  (Nixon, 
1987;  Hanlon  and  Messenger,  1996).  Most  gonatids 
undergo  ontogenetic  vertical  descent  (Roper  and  Young, 
1975;  Nesis,  1997),  and  a  clear  shift  in  the  diet  can  ac- 
company this  habitat  shift  (e.g.,  as  seen  in  Berryteuthis 
magister;  Nesis,  1997).  Nesis  (1997),  however,  suggested 
that  B.  anonychus  does  not  undergo  ontogenetic  descent; 
therefore  no  such  habitat-change-related  shift  in  diet 
would  be  expected  to  occur  in  this  species. 

Highest  feeding  intensities  were  recorded  in  the  after- 
noon and  just  after  sunset,  which  would  indicate  that 
B.  anonychus  feeds  both  day  and  night.  Such  a  feed- 
ing scenario  is  supported  by  the  high  overlap  in  depth 
distributions  of  B.  anonychus  (day:  50-200  m,  night: 
0-150  m;  Nesis,  1997)  and  its  main  prey,  Neocalanus 
cristatus;  during  spring,  N.  cristatus  occurs  mainly  at 
50-150  m,  and  like  the  other  Neocalanus  species,  shows 
no  evidence  of  diel  vertical  migration  (Mackas  et  al., 
1993).  Therefore  B.  anonychus  and  N.  cristatus  occupy 
nearly  the  same  depth  range  both  day  and  night. 

The  chaetognath  Parasagitta  elegans  was  the  third 
most  abundant  prey  taxon  and  was  consumed  by  more 
squid  than  any  other  taxon.  Parasagitta  elegans  forms 
an  important  fraction  of  the  springtime  macrozooplank- 
ton  community  in  the  North  Pacific  (Brodeur  and  Ter- 
azaki,  1999)  and  inhabits  mainly  the  epipelagic  layer 
(0-200  m)  (Kotori,  1976;  Terazaki  and  Miller,  1986); 
therefore  predation  on  P.  elegans  could  also  occur  both 
day  and  night.  Another  gonatid  squid.  Gonatus  mado- 
kai,  has  also  been  found  to  prey  on  Parasagitta  sp. 
(Kubodera  and  Okutani,  1977). 

There  was  no  evidence  of  cannibalism,  which  com- 
monly occurs  in  many  gonatids,  particularly  Berryteu- 
this magister  and  Gonatopsis  borealis  (Lapko,  1996; 
Nesis,  1997).  Cannibalism  in  squids  appears  to  occur 
less  frequently  when  prey  are  abundant  (Shchetinnikov, 
1992;  Santos  and  Haimovici,  1997),  as  is  the  case  in 
the  North  Pacific  during  spring.  In  addition,  at  nearly 
every  station  sampled,  squid  of  a  small  size  range  were 
collected  (Bower  et  al.,  2002);  therefore  it  seems  that  op- 
portunities for  intercohort  cannibalism  were  limited. 

The  large  stock  size  of  B.  anonychus  in  the  North 
Pacific  (Nesis,  1997)  and  its  importance  in  the  diet  of 
higher  predators  may  indicate  that  the  food  chain  from 
copepods  through  squids  and  these  higher  predators  is 
an  important  trophic  pathway  in  the  pelagic  food  web  of 
the  Subarctic  Pacific  during  spring.  The  large  seasonal- 
ity in  zooplankton  composition  in  the  upper  150  m  may 
indicate  that  these  trophic  pathways  will  show  similar 
seasonal  variations. 


Acknowledgments 

We  thank  the  late  H.  Richard  Carlson  for  providing 
us  with  squids  collected  during  the  May  1999  NMFS 
salmon  survey  aboard  the  FV  Great  Pacific.  We  also 
thank  Chingis  Nigmatullin  and  the  late  Kir  Nesis  for 
translating  two  Russian  abstracts  into  English,  H.  Sugi- 


738 


Fishery  Bulletin  102(4) 


saki  and  M.  Terazaki  for  helping  identify  prey,  K.  Ichige 
for  helping  in  the  laboratory,  and  the  three  anonymous 
reviewers  of  the  manuscript. 


Literature  cited 

Baker  A.  de  C,  B.  P.  Boden,  and  E.  Brinton. 

1990.     A  practical  guide  to  the  euphausiids  of  the  world, 
96  p.     Natural  History  Museum,  London. 
Bower,  J.  R.,  J.  M.  Murphy,  and  Y.  Sato. 

2002.     Latitudinal  gradients  in  body  size  and  maturation 
of  Berry  tetuhis  anonychus  (Cephalopoda:  Gonatidae)  in 
the  northeast  Pacific.     Veliger  45:309-315. 
Brodeur,  R.  D.,  and  M.  Terazaki. 

1999.     Springtime  abundance  of  chaetognaths  in  the  shelf 
region  of  the  northern  Gulf  of  Alaska,  with  observa- 
tions on  the  vertical  distribution  and  feeding  of  Sagitta 
elegans.     Fish.  Oceanogr.  8:93-103. 
Brodeur,  R.,  S.  McKinnell,  K.  Nagasawa,  W.  Pearcy, 
V.  Radchenko,  and  S.  Takagi. 

1999.     Epipelagic  nekton  of  the  North  Pacific  subarctic 
and  transition  zones.     Prog.  Oceanogr.  43:365-397. 
Brodsky,  K.A. 

1950.     Calanoida  of  the  far  eastern  seas  and  Polar 
basin   of  the   U.S.S.R.     Dokl.   Akad.    Nauk.   SSSR. 
35:1-442.     USSR  Israel  Program  Scient.  Transl.  (1967), 
Jerusalem.  Transl.  No.  TT-65-51200,  440  p. 
Carlson,  H.  R.,  J.  M.  Murphy,  C.  M.  Kondzela,  K.  W.  Myers, 
and  T  Nomura. 

1999.     Survey  of  salmon  in  the  northeastern  Pacific  Ocean, 
May  1999.     North  Pacific  Anadromous  Fish  Commission, 
Document  450,  37  p.     Natl.  Mar.  Fish.  Serv.,  Juneau,  AK. 
Cortes,  E. 

1997.     A  critical  review  of  methods  of  studying  fish  feeding 
based  on  analysis  of  stomach  contents:  application  to 
elasmobranch  fishes.     Can.  J.  Fish.  Aquat.  Sci.  54:726- 
738. 
Didenko,  V.  D. 

1990.  Prospects  of  fishery  for  the  squid  Berryteuthis 
anonychus  in  the  north-eastern  Pacific.  In  "Vth  all- 
USSR  conference  on  commercial  invertebrates,  Minsk- 
Naroch.  Abstracts  of  Communications,"  p.  82-83.  [In 
Russian. I  VNIRO  Publishing,  Moscow.  Russia. 
Hanlon,  R.  T.  and  J.  B.  Messenger. 

1996.     Cephalopod  behaviour,  232  p.     Cambridge  Univ. 
Press,  Cambridge,  UK. 
Howmiller,  R.  P. 

1972.     Effects  of  preservatives  on  weights  of  some  common 
macrobenthic  invertebrates.     Trans.  Am.  Fish.  Soc. 
101:743-746. 
Karlson,  K.,  and  U.  Bamstedt. 

1994.     Planktivorous  predation  on  copepods.  Evaluation 
of  mandible  remains  in  the  predator  guts  as  a  quan- 
titative estimate  of  predation.     Mar.  Ecol.  Prog.  Ser. 
108:79-89. 
Kotori,  M. 

1976.  The  biology  of  Chaetognatha  in  the  Bering  Sea 
and  the  northern  North  Pacific  Ocean,  with  emphasis 
on  Sagitta  elegans.  Mem.  Fac.  Fish.  Hokkaido  Univ. 
23:95-183. 

Kubodera,  T,  and  T  Okutani. 

1977.  Description  of'a  new  species  of  gonatid  squid.  Gona- 
tus  madokai  n.sp.  from  the  north-west  Pacific,  with  notes 
on  morphological  changes  with  growth  and  distribution 


in  immature  stages  (Cephalopoda:  Oegopsidal.     Venus 
Jap.  J.  Malacol.  36i  3 ':  123-151. 

Kuramochi,  T.  T  Kubodera,  and  N.  Miyazaki. 

1993.  Squids  eaten  by  Dall's  porpoises,  Phoeoenoides  dalli 
in  the  northwestern  North  Pacific  and  in  the  Bering 
Sea.  //)  Recent  advances  in  cephalopod  fisheries  biol- 
ogy (T  Okutani,  R.  K.  O'Dor,  and  T  Kubodera,  eds.l, 
p.  229-240.     Tokai  Univ.  Press,  Tokyo. 

Lapko,  V.  V. 

1996.  The  role  of  squids  in  the  Sea  of  Okhotsk  com- 
munities.    Oceanology  35:672-677. 

Lapshina,  V.  I. 

1988.  On  the  feeding  of  the  squid  Berryteuthis  anony- 
chus Pearcy  and  Voss  from  the  area  of  north-eastern 
Pacific.  Deposited  MS,  Dep.  VINITI  25.10.88,  No.  7673- 
B  88,  1-5.     [In  Russian. 1 

Mackas,  D.  L.,  H.  Sefton,  C.  B.  Miller,  and  A.  Raich. 

1993.  Vertical  habitat  partitioning  by  large  calanoid 
copepods  in  the  oceanic  subarctic  Pacific  during 
spring.     Prog.  Oceanogr.  32:259-294. 

Mackas,  D.  L.,  and  A.  Tsuda. 

1999.  Mesozooplankton  in  the  eastern  and  western 
subarctic  Pacific:  community  structure,  seasonal  life 
histories,  and  interannual  variability.  Prog.  Oceanogr. 
43:335-363. 

Miller,  C.  B. 

1988.  Neocalanus  flemingeri,  a  new  species  of  Calani- 
dae  (Copepoda:  Calanoida)  from  the  subarctic  Pacific 
Ocean,  with  a  comparative  redescription  of  Neocala- 
nus plumchrus  (Marukawal  1921.  Prog.  Oceanogr. 
20:223-273 

Nesis,  K.  N. 

1997.  Gonatid  squids  in  the  subarctic  North  Pacific: 
ecology,  biogeography.  niche  diversity  and  role  in  the 
ecosystem.     Adv.  Mar.  Biol.  32:243-324. 

Nixon,  M. 

1987.  Cephalopod  diets.  In  Cephalopod  life  cycles,  vol.  II: 
comparative  reviews  (P.  R.  Boyle,  ed.).  p.  201-219.  Aca- 
demic Press,  London. 

Ogi,  H.,  T  Kubodera,  and  K.  Nakamura. 

1980.     The  pelagic  feeding  ecology  of  the  short-tailed 
shearwater  Puffinus  tenuirostris  in  the  subarctic  Pacific 
region.     J.  Yamashina  Inst.  Ornithol.  12:157-182. 
Ohizumi,  H.,  T  Kuramochi,  T  Kubodera.  M.  Yoshioka.  and 
N.  Miyazaki. 

2003.  Feeding  habits  of  Dall's  porpoises  (Phoeoenoides 
dalli)  in  the  subarctic  North  Pacific  and  the  Bering 
Sea  basin  and  the  impact  of  predation  on  mesopelagic 
micronekton.  Deep-Sea  Res.  (I  Oceanogr.  Res.  Pap.). 
50:593-610. 
Pearcy,  W.  G. 

1991.     Biology  of  the  transition  region.     In  Biology,  ocean- 
ography, and  fisheries  of  the  North  Pacific  Transition 
Zone  and  Subarctic  Frontal  Zone  (J.  A.  Wetherall,  ed.), 
p.  39-55.     NOAA  Tech.  Rep.  NMFS  105. 
Pearcy,  W.  G.,  R.  D.  Brodeur,  J.  M.  Shenkcr,  W.  W.  Smoker, 
and  Y.  Endo. 

1988.  Food  habits  of  Pacific  salmon  and  steelhead  trout, 
midwater  trawl  catches  and  oceanographic  conditions 
in  the  Gulf  of  Alaska.  1980-1985.  Bull.  Ocean  Res. 
Inst.  Univ.  Tokyo.  26:29-78. 

Pearcy,  W  (I..  .1,  I',  Fisher,  and  M.  M.  Yoklavich. 

1993.  Biology  of  the  Pacific  pomfret  ( lirmna  japonica  I 
in  the  North  Pacific  Ocean.  Can.  J.  Fish.  Aquat.  Sci. 
50:2608-2625. 


Uchikawa  et  al.:  Diet  of  Berryteuthis  anonychus  in  the  northeast  Pacific  during  spring 


739 


Pinkas  L.,  M.  S.  Oliphant,  and  I.  L.  K.  Iverson. 

1971.     Food  habits  of  albacore,  bluefin  tuna,  and  bonito 
in  California  waters.     Calif.  Dep.  Fish  Game  Fish.  Bull. 
152:1-105. 
Radchenko,  V.  I. 

1992.     The  role  of  squids  in  the  pelagic  ecosystem  of  the 
Bering  Sea.     Oceanology  32:762-767. 
Rodhouse,  P.  G.,  and  Ch.  M.  Nigmatullin. 

1996.     Role  as  consumers.     Phil.  Trans.  R.  Soc.  Lond. 
B.  351:1003-1022. 
Roper,  C.  F.  E.,  and  R.  E.  Young. 

1975.     Vertical  distribution  of  pelagic  cephalopods.     Smith- 
son.  Contrib.  Zool.  209:1-51. 
Roper,  C.  F.  E.,  M.  J.  Sweeney,  and  C.  E.  Nauen. 

1984.  FAO  species  catalogue.  Vol.  3.  Cephalopods  of  the 
world.  An  annotated  and  illustrated  catalogue  of  species 
of  interest  to  fisheries.  FAO  Fish.  Synop.  125,  vol.  3, 
277  p.     FAO,  Rome. 


Santos,  R.  A.,  and  M.  Haimovici. 

1997.     Food  and  feeding  of  the  short-finned  squid  Illex 
argentinus  (Cephalopoda:  Ommastrephidae)  off  southern 
Brazil.     Fish.  Res.  33:139-147. 
Shchetinnikov,  A.  S. 

1992.     Feeding  spectrum  of  squid  Sthenoteuthis  oualani- 
ensis  (Oegopsidai  in  the  eastern  Pacific.     J.  Mar.  Biol. 
Assoc.  U.K.  72:849-860. 
Terazaki,  M.,  and  C.  B.  Miller. 

1986.     Life  history  and  vertical  distribution  of  pelagic 
chaetognaths  at  Ocean  Station  P  in  the  subarctic 
Pacific.     Deep-Sea  Res.  33:323-337. 
Vinogradov,  M.  E.,  A.  F.  Volkov,  and  T.  N.  Semenova. 

1996.  Hyperiid  amphipods  (Amphipoda,  Hyperiidea)  of 
the  world  oceans  (D.  Siegel-Causey,  ed.),  632  p.  Science 
Publishers,  Lebanon,  NH. 


740 


Abstract— The  relative  abundance 
of  Bristol  Bay  red  king  crab  (Para- 
lithodes  camtschaticus)  is  estimated 
each  year  for  stock  assessment  by 
using  catch-per-swept-area  data  col- 
lected on  the  Alaska  Fisheries  Sci- 
ence Center's  annual  eastern  Bering 
Sea  bottom  trawl  survey.  To  estimate 
survey  trawl  capture  efficiency  for  red 
king  crab,  an  experiment  was  con- 
ducted with  an  auxiliary  net  (fitted 
with  its  own  heavy  chain-link  foot- 
rope  I  that  was  attached  beneath  the 
trawl  to  capture  crabs  escaping  under 
the  survey  trawl  footrope.  Capture 
probability  was  then  estimated  by 
fitting  a  model  to  the  proportion  of 
crabs  captured  and  crab  size  data. 
For  males,  mean  capture  probability 
was  727  at  95  mm  (carapace  length), 
the  size  at  which  full  vulnerability  to 
the  survey  trawl  is  assigned  in  the 
current  management  model:  84. 1^  at 
135  mm,  the  legal  size  for  the  fish- 
ery; and  939c  at  184  mm,  the  maxi- 
mum size  observed  in  this  study.  For 
females,  mean  capture  probability  was 
707c  at  90  mm,  the  size  at  which  full 
vulnerability  to  the  survey  trawl  is 
assigned  in  the  current  manage- 
ment model,  and  777  at  162  mm, 
the  maximum  size  observed  in  this 
study.  The  precision  of  our  estimates 
for  each  sex  decreased  for  juveniles 
under  60  mm  and  for  the  largest 
crab  because  of  small  sample  sizes. 
In  situ  data  collected  from  trawl- 
mounted  video  cameras  were  used  to 
determine  the  importance  of  various 
factors  associated  with  the  capture  of 
individual  crabs.  Capture  probabil- 
ity was  significantly  higher  when  a 
crab  was  standing  when  struck  by  the 
footrope,  rather  than  crouching,  and 
higher  when  a  crab  was  hit  along  its 
body  axis,  rather  than  from  the  side. 
Capture  probability  also  increased  as 
a  function  of  increasing  crab  size  but 
decreased  with  increasing  footrope 
distance  from  the  bottom  and  when 
artificial  light  was  provided  for  the 
video  camera. 


Capture  probability  of  a  survey  trawl  for 
red  king  crab  (Paralithodes  camtschaticus) 


Kenneth  L.  Weinberg 

Alaska  Fisheries  Science  Center 

National  Marine  Fisheries  Service,  NOAA 

7600  Sand  Point  Way  N.E 

Seattle,  Washington  98115 

E-mail  address  ken  Weinberg  gnoaa  gov 

Robert  S.  Otto 

Kodiak  Fisheries  Research  Center 
National  Marine  Fisheries  Service,  NOAA 
301  Research  Court 
Kodiak,  Alaska  99615 

David  A.  Somerton 

Alaska  Fisheries  Science  Center 
National  Marine  Fisheries  Service,  NOAA 
7600  Sand  Point  Way  N.E. 
Seattle,  Washington  98115 


Manuscript  submitted  5  September  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
28  April  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:740-749(2004) 


Regulations  limit  the  annual  har- 
vest of  Bristol  Bay  red  king  crab 
(RKC;  Paralithodes  camtschaticus)  to 
males  >135  mm  in  carapace  length1 
(6.5  inches  carapace  width),  and  the 
size  of  the  harvest  is  dependent  upon 
the  estimated  biomasses  of  mature 
males  and  females.  For  stock  assess- 
ments of  RKC,  area-swept  abundance 
estimates  are  determined  from  the 
data  from  annual  eastern  Bering  Sea 
(EBS)  bottom  trawl  surveys  conducted 
by  the  National  Marine  Fisheries  Ser- 
vice, Alaska  Fisheries  Science  Center 
(AFSC),  and  these  estimates  are  used 
as  input  into  a  length-based  assess- 
ment model  (Zheng  et  al.,  1995)  to 
compute  the  total  allowable  catch  for 
each  annual  fishing  season. 

It  is  assumed  with  the  current  as- 
sessment model  that  all  male  RKC 
>95  mm  and  all  female  RKC  >90  mm 
within  the  path  of  the  survey  trawl 
(wingtip  to  wingtip)  are  captured. 
This  assumption  seems  reasonable 
because  the  survey  trawl  uses  a  small 
diameter  footrope  designed  to  stay 
close  to  the  bottom  and  red  king  crab 
are  quite  large.  However,  video  pho- 
tography taken  following  the  2000 
EBS  survey  revealed  that  a  consider- 
able number  of  large  (>90  rami  RKC 


pass  under  the  footrope  of  the  survey 
trawl. 

To  assess  the  potential  impact  of 
escaping  crab  on  the  calculation  of 
crab  biomass,  we  conducted  an  ex- 
periment to  estimate  the  size-related 
capture  efficiency  of  the  standard 
survey  bottom  trawl  for  Bristol  Bay 
RKC.  In  this  experiment,  crab  pass- 
ing beneath  the  survey  trawl  were 
subsequently  captured  with  an  aux- 
iliary net  that  was  attached  under- 
neath and  behind  the  footrope  of 
the  survey  trawl  (Engas  and  Godo, 
1989;  Walsh,  1992).  Experimental 
nets  like  the  one  used  for  this  study 
have  been  used  previously  in  trawl 
efficiency  studies  for  flatfish  (Munro 
and  Somerton,  2002),  as  well  as  for 
snow  iChionoecetes  opilio)  and  Tan- 
ner (C.  bairdi)  crabs  (Somerton  and 
Otto,  1999).  Trawl  catch  data  alone, 
however,  tell  little  about  the  details 
involved  with  escapement.  Therefore, 
we  deployed  a  video  camera  on  the 
trawl  to  observe  crab  behavior  and 
analyzed  a  combination  of  trawl-per- 


1  All  references  to  measured  crab  lengths 
are  carapace  length. 


Weinberg  et  al.:  Capture  probablity  of  a  survey  trawl  for  Parahthodes  camtschaticus 


741 


trawl  codend 


auxiliary  codend 


trawl  footrope 


wrapped  wire 

chain  hangings 

twine  hangings 

fishing  line 


auxiliary  footrope 


Figure  1 

The  83/112  Eastern  bottom  trawl  with  auxiliary  net  used  in  the  2002 
red  king  crab  capture  efficiency  experiment  (figure  adapted  from  Munro 
and  Somerton,  2002,  with  permission  from  Elsevierl. 


formance  and  crab-behavioral  variables  to  help  us  un- 
derstand the  escapement  process. 


Materials  and  methods 

Description  of  trawl  gear 

The  83/112  Eastern  bottom  trawl  has  been  used  by  the 
AFSC  in  annual  surveys  to  assess  EBS  crab  and  shelf 
groundfish  stocks  since  1982  (Armistead  and  Nichol, 
1993).  For  the  present  experiment,  an  auxiliary  net  with 
an  independent  footrope  constructed  of  heavy  chain-link 
and  a  separate  codend  were  attached  to  the  bottom  of 
the  survey  trawl  to  capture  epibenthic  animals  passing 
beneath  the  trawl  footrope  (Fig.  1).  Briefly,  the  83/112 
Eastern  is  a  low-rise  trawl  that  has  a  25.3-m  long  hea- 


drope  strung  with  48  floats  giving  it  approximately  102 
kg  of  lift  and  a  34.1  m  long,  5.2-cm  diameter  footrope 
constructed  of  1.6-cm  stranded  wire  rope  protected  with 
a  single  wrap  of  polypropylene  line  and  split  rubber  hose. 
The  net  is  constructed  with  nylon  twine:  10.1-cm  stretch 
mesh  throughout  the  wing  and  throat  sections;  8.9-cm 
stretch  mesh  in  the  intermediate  section;  a  double  layer 
of  8.9-cm  stretch  mesh  in  the  codend;  and  a  3.1-cm 
stretch  mesh  liner  in  the  codend.  It  is  fished  with  a  pair 
of  1.8x2.7-m  steel  V-doors  weighing  approximately  816 
kg  apiece. 

The  auxiliary  net  attaches  to  the  wingtips  and  to  the 
bottom  of  the  survey  trawl  so  that  the  bottom  panel  of 
the  trawl  serves  as  the  top  panel  of  the  auxiliary  net 
up  to  the  beginning  of  the  intermediate  section.  At  this 
point,  the  two  nets  part  and  the  auxiliary  net  then  has 
a  top  panel  of  8.9-cm  stretch  mesh  and  a  double  layer 


742 


Fishery  Bulletin  102(4) 


61  "N- 


60  N- 


59  N- 


58  N- 


57°N- 


56N- 


55N- 


54N- 


166  W 


164  =  W 


162  W 


160  W 


158  W 


Figure  2 

The  annual  eastern  Bering  Sea  survey  station  grid  showing  the 
number  of  successful  tows  per  station  block  made  during  the  2002 
red  king  crab  capture  efficiency  study.  Each  block  represents  a 
400-nmi2  area. 


codend  with  a  3.1-cm  stretch  mesh  liner.  The  38.2  m  long 
auxiliary  footrope  constructed  of  heavy  16-mm-long  link 
trawl  chain  was  designed  to  drag  through  soft  bottom 
and  presumably  captures  all  escaping  crabs.  Munro  and 
Somerton  (2002)  provided  detailed  construction  plans  of 
this  experimental  gear  in  their  appendices. 

Experimental  design 

Operations  were  conducted  from  21  to  29  July  2002, 
aboard  the  FV  Arcturus,  one  of  two  commercial  stern 
trawlers  chartered  by  the  AFSC  since  1993  to  carry 
out  annual  Bering  Sea  groundfish  surveys.  Trawling 
took  place  in  Bristol  Bay  (Fig.  2)  at  depths  from  41  to 
77  m  and  followed  standardized  survey  protocols  that 
included  towing  during  daylight  hours  at  a  1.5  m/sec  (3 
knots)  vessel  speed  and  using  locked  winches  and  stan- 
dardized lengths  of  trawl  warp  (scope)  at  each  towing 
depth.  Acoustic  net  mensuration  equipment  was  used 


to  measure  wing  spread  for  each  tow.  Bottom  contact 
sensors  were  used  on  the  centers  of  both  the  trawl  and 
auxiliary  footropes  to  measure  the  distance  (in  centime- 
ters) between  the  footropes  and  the  bottom  (Somerton 
and  Weinberg.  2001).  A  silicon-intensified  tube  (SIT) 
camera,  which  uses  ambient  light,  was  attached  to  the 
center  of  the  trawl  to  view  RKC  interaction  with  the 
footrope.  On  some  of  our  trial  tows,  however,  a  30-W 
quartz  halogen  light  was  also  used  to  increase  contrast 
between  ambient  light  and  the  sea  floor. 

Two  departures  from  standardized  survey  protocol 
were  necessary  for  this  experiment.  First.  27.5-m  long 
bridles  were  used  instead  of  the  survey  standard  55- 
m  long  bridles  to  help  offset  the  loss  of  wing  spread 
caused  by  the  added  drag  of  the  auxiliary  net  (Munro 
and  Somerton,  2002).  Second,  tow  length  was  shortened 
from  the  survey  standard  of  30  min  to  20  min  to  mini- 
mize the  decrease  of  path  width  over  time  due  to  in- 
creased drag  from  large  catches  in  the  auxiliary  net. 


Weinberg  et  al.:  Capture  probablity  of  a  survey  trawl  for  Paralithodes  camtschaticus 


743 


Towing  sites  were  selected  according  to  catch  rates 
and  carapace  lengths  obtained  from  the  recently  com- 
pleted 2002  EBS  survey  (Stevens2).  Tows  were  made  in 
pairs,  one  in  a  northerly  direction,  the  other  in  a  south- 
erly direction  and  were  offset  to  the  east  or  west  by  a 
minimum  of  0.1  nmi;  the  initial  direction  was  chosen 
randomly  in  order  to  mitigate  any  bias  that  the  current 
flow  might  have  on  footrope  contact  with  the  bottom 
(Weinberg,  2003).  Increased  effort  was  given  to  sites 
producing  favorable  numbers  and  crab  lengths  by  add- 
ing additional  towing  pairs.  For  each  tow  the  total  catch 
of  all  species  from  each  net  was  first  weighed  before  all 
RKC  were  removed  from  the  catch,  weighed,  coded  by 
sex,  and  measured  to  the  nearest  millimeter. 

Data  analysis 

Trawl  geometry  Trawl  geometry  for  standard  survey 
nets  and  experimental  nets  was  measured  to  confirm 
that  the  two  gear  types  fished  similarly.  Average  wing 
spreads  and  footrope  heights  off-bottom  for  experimental 
tows  were  compared  to  those  from  33  standard  survey 
gear  tows  taken  at  the  same  or  nearby  sampling  loca- 
tions. Because  the  depth  of  sampling  varied,  wing  spread 
and  footrope  height  were  linearly  regressed  on  scope, 
a  factor  variable  indicating  gear  type  (i.e.,  survey  or 
experimental),  and  their  interaction.  Two-tailed  r-tests 
were  used  to  test  for  the  difference  in  the  slopes  and  the 
intercepts  between  gear  types.  Significance  of  the  inter- 
action term  indicated  that  slopes  differed  between  gear 
types.  For  nonsignificant  interaction,  significance  of  the 
intercepts  indicated  that  wing  spread  or  footrope  height 
differed  between  gear  types  by  a  constant  amount. 

Capture  probability  Capture  probability  for  the  experi- 
mental gear  was  estimated  from  catch  data  of  the  trawl 
and  the  auxiliary  net  as  a  function  of  carapace  length 
(L)  for  both  male  and  female  crab.  Based  on  the  assump- 
tion that  the  auxiliary  net  allows  no  escapement,  the 
probability  of  capture  at  the  footrope  was  modeled  as  a 
logistic  function  (Munro  and  Somerton,  2002)  by  using 
SPLUS  software  (version  6.1,  Insightful  Corporation, 
Seattle,  WA).  Two  models  were  considered:  the  first, 
a  two-parameter  model  which  reaches  an  asymptotic 
maximum  of  1  (unity): 


PlLh 


l+e-<a+pLr 


(1) 


and  the  second,  a  three-parameter  model  which  reaches 
an  asymptotic  maximum  less  than  1: 


PiD- 


1  +  e 


-la+/3L) 


(2) 


2  Stevens,  B.  G.,  R.  A.  Macintosh,  J.  A.  Haaga,  C.  E.  Armistead, 
and  R.  S.  Otto.  2002.  Report  to  industry  on  the  2002 
eastern  Bering  Sea  crab  survey.  AFSC  Proc.  Rep.  2002-5, 
59  p.  Alaska  Fish.  Sci.  Cent.,  Natl.  Mar.  Fish.  Serv.,  NOAA, 
Kodiak  Fishery  Research  Center,  301  Research  Court,  Kodiak 
AK  99615. 


Because  crab  capture  at  the  footrope  is  a  binomial  pro- 
cess (i.e.,  crabs  are  either  captured  or  they  escape),  the 
models  were  fitted  to  the  capture  and  length  data  by 
using  maximum  likelihood  (Millar,  1992;  Munro  and 
Somerton,  2001)  and  the  data  were  pooled  across  tows. 
For  each  sex,  both  models  were  fitted  to  the  data,  and  the 
best  of  the  competing  models  was  selected  according  to 
the  lowest  obtained  value  of  the  Akaike  information  cri- 
terion (AIC;  Burnham  and  Anderson,  1998),  defined  as 

AIC  =  -2(log  likelihood )  +  2(  number  of  parameters). 

After  choosing  a  model  for  each  sex,  we  examined  whether 
the  capture  probability  curves  differed  between  sexes  by 
fitting  a  model  to  the  data  for  both  sexes  combined,  and 
then  comparing  the  value  of  AIC  for  this  model  to  the 
sum  of  the  AIC  values  for  the  models  fitted  to  each  sex, 
again  using  the  minimum  AIC  value  to  objectively  select 
the  better  of  the  two  models. 

Bootstrapped  confidence  intervals  were  constructed  for 
the  mean  capture  probability  for  each  1-mm  length  cat- 
egory, between  the  smallest  and  the  largest  individuals 
(Efron  and  Tibshirani,  1993)  by  resampling  the  catch- 
at-size  data  from  individual  hauls  1000  times.  Empirical 
95%  confidence  intervals  were  then  determined  as  the 
range  between  the  25th  highest  and  the  25th  lowest  of 
the  bootstrap  capture  probability  estimates. 

Video  data  analyses 

To  understand  the  factors  associated  with  crab  escape- 
ment under  the  footrope,  a  video  camera  was  mounted 
on  the  trawl  to  observe  RKC  interaction  with  the  center 
of  the  trawl  footrope.  These  in  situ  video  observations 
included  tows  made  in  2000  on  the  standard  trawl  and  in 
2002  on  the  experimental  trawl.  Artificial  light  was  pro- 
vided for  all  of  the  2000  tows,  and  for  some  of  the  2002 
trial  tows  made  before  the  capture  efficiency  experiment 
began.  All  the  2002  experimental  tows  were  made  under 
natural  light  conditions.  RKC  encounters  observed  on 
the  videotapes  were  counted  from  the  time  the  footrope 
settled  to  the  bottom  until  the  time  the  footrope  was 
lifted  off-bottom  at  the  end  of  the  tow.  The  probability  of 
capture  was  predicted  as  a  function  of  several  explana- 
tory variables.  Variables  observed  and  codes  (in  paren- 
theses) for  each  individual  included  the  following: 

1  the  capture  event — the  crab  escaped  beneath  the 
footrope  (0)  or  was  captured  (1); 

2  use  of  artificial  light — the  tow  was  made  with  (0) 
or  without  artificial  light  (1); 

3  estimated  mean  footrope  height  above  the  sea  floor 
over  the  course  of  the  entire  tow  was  based  on  the 
bottom  contact  sensor  and  was  expressed  in  centi- 
meters (0-5); 

4  body  height — the  crab  was  observed  to  be  crouching 
with  its  legs  either  tucked  beneath  the  carapace  or 
stretched  out  so  that  the  carapace  was  very  close 
to  the  bottom  (0)  or  standing  upright  on  its  dactyls 
(1); 


744 


Fishery  Bulletin  102(4) 


Table  1 

Regression  coefficients  of  wing  spread  (widthl  and  footrope  distance  from  the  bottom  (footropel  as  a  function  of  scope  and  gear 
type  based  on  tows  from  the  capture  efficiency  experiment  and  tows  using  standard  survey  gear.  Also  provided  are  the  results  of 
two-tailed  Mests  testing  for  the  difference  in  intercept  between  gear  types. 


Slope 


Intercept 


Experimental  gear  (n  =  43) 


Standard  gear  (n  =  33l 


P intercept 


Width  (ml 
Footrope  (cm) 


0.0038 
0.0087 


15.3206 
-1.2582 


16.0713 
-0.4241 


<0.0000 
0.0023 


5  body  orientation  at  the  point  of  contact  with  the 
footrope — footrope  contact  occurred  along  the  body 
axis  (1)  or  from  the  side  (2); 

6  crab  size — small  (0),  medium  (1),  or  large  (2),  where 
size  is  expressed  as  an  approximation  based  on 
visual  comparison  of  carapace  length  to  the  dimen- 
sions of  trawl  parts,  such  as  mesh  or  chain  links. 
Corresponding  length  intervals  were  approximately 
<90  mm,  90-135  mm,  and  >135  mm. 

As  a  general  rule,  crabs  could  be  seen  in  the  videos  1-2 
seconds  prior  to  contact  with  the  footrope.  Assignment 
of  codes  was  typically  straightforward.  However,  in  some 
instances,  several  reviews  of  the  encounter  were  neces- 
sary in  order  to  determine  a  crab's  position  or  orientation 
in  relation  to  the  footrope. 

The  probability  of  capture  was  estimated  by  using 
stepwise  generalized  linear  modeling  (GLM;  Venables 
and  Ripley,  1994)  to  fit  a  logistic  model  describing  the 
probability  of  capture  as  a  function  of  crab  size,  where 
body  height,  body  orientation,  average  footrope  distance 
from  the  bottom  during  the  tow,  the  use  of  artificial 
light,  and  all  possible  first  order  interactions  were  con- 
sidered as  additional  potential  terms.  The  model  fitting 
procedure  (with  data  from  crabs  for  which  all  variables 
were  observed)  entailed  a  stepwise  backward  model  se- 
lection process.  The  process  began  with  fitting  the  model 
to  all  interaction  terms  less  one,  and  then  calculating 
and  comparing  the  resulting  AIC  values.  The  interaction 
producing  the  largest  decrease  in  AIC  was  subsequently 
eliminated.  Next,  the  procedure  was  repeated  with  the 
remaining  terms  until  no  interaction  term  could  be 
eliminated  without  increasing  the  AIC.  Then,  the  above 
process  was  repeated  for  the  main  effects.  For  the  main 
effects  having  an  interaction  term,  both  the  main  effects 
and  the  interaction  term  were  eliminated  together  as  a 
unit.  The  final  model  chosen  contained  those  terms  that 
produced  the  minimum  AIC  value. 


Results 

Effect  of  the  auxiliary  net  on  trawl  geometry 

Regressions  of  wing  spread  and  footrope  height  on 
scope,  gear  type,  and  their  interaction  were  compared 


to  determine  how  closely  the  two  gear  types  fished.  The 
interaction  term  was  not  significant  for  wing  spread 
(P=0.08)  nor  for  footrope  height  (P=0.82),  indicating  that 
the  slopes  did  not  differ  between  gear  types.  However. 
tests  of  the  intercepts  were  significant  for  both  wing 
spread  and  footrope  height  and  indicated  that  trawl 
geometry  differed  between  survey  and  experimental 
trawls  (Table  1).  Predicted  standard  survey  wing  spreads 
for  the  minimum  (137  m),  median  (229  m),  and  maxi- 
mum (320  m)  scopes  used  were  16.6,  16.9.  and  17.3  m 
—  approximately  0.8  m  more  than  the  experimental 
gear  at  the  same  scopes.  Predicted  footrope  distances 
off  the  bottom  were  0.8,  1.6,  and  2.4  cm,  at  the  above 
three  scope  values — approximately  0.8  cm  greater  than 
the  experimental  gear.  Although  we  detected  statistical 
differences  in  the  trawl  geometry  between  the  two  gear 
types,  the  actual  difference  in  physical  measurements 
was  small  and  presumably  had  only  a  nominal  effect  on 
the  results  of  the  capture  efficiency  experiment. 

Our  assumption  that  the  auxiliary  net  caught  all 
escaping  crabs  was  reinforced  by  two  observations:  1) 
the  data  from  the  bottom  contact  sensor  on  the  chain 
footrope  indicated  consistent  contact  with  the  sea  floor; 
and  2)  the  auxiliary  net  consistently  had  large  catches 
of  benthic  organisms  other  than  crab,  such  as  starfish 
and  shells,  and  produced  enough  drag  on  the  system  to 
reduce  wing  spread.  The  effectiveness  of  the  auxiliary 
footrope  at  capturing  escaping  crab  is  in  part  due  to 
its  weight  and  small  diameter  that  enable  it  to  sweep 
beneath  the  crabs  and  in  part  due  to  the  suspension  of 
benthic  organisms  initiated  by  the  turbulence  created 
by  the  passing  of  the  first  footrope. 

Length-based  capture  probability 

Capture  probability  was  estimated  from  length  mea- 
surements (n=3233)  collected  from  43  successful 
experimental  tows  (21  north,  22  south)  made  within 
11  standard  EBS  survey  station  blocks  (Fig.  2).  Male 
samples  (n  =  1667)  ranged  in  size  from  23  to  184  mm 
(Fig.  3).  Female  samples  (/?  =  1566)  ranged  in  size  from 
51  to  162  mm. 

The  two-parameter  model  (model  1)  of  capture  prob- 
ability was  selected  over  the  three-parameter  model 
(model  2)  because  it  had  a  lower  AIC  value  for  both 
male  and  female  RKC  (Table  2).  For  the  comparison  of 


Weinberg  et  al.:  Capture  probabhty  of  a  survey  trawl  for  Paralithodes  camtschaticus 


745 


100  i 

Males 

80  ■ 

P  Captures 

1 

n 

n 

60  ■ 

■  Escapes 

' 

n 

40- 

20  - 

1     1 

,     1 

1 

1 

_nJl 

1 

ll 

1 

1 

1 

l 

1 

1 

1 

■   L  n    n. 

20 

30       40       50       60       70       80       90       100      110     120      130      140      150      160      170      180 

0) 

0) 
LL 

150 

n 

125 

Females 

□  Captures 

100  ■ 
75- 

■  Escapes 

n 

50- 

1 

25  - 

—  ill 

1 

1, 

1 

1 

J 

1    1          .        ~. 

20 

30       40       50       60       70       80       90       100     110     120     130     140      150     160     170      180 

Carapace  length  (5  mm  bins) 

Figure  3 

Size  frequency  of  red  king  crabs  taken  in  the  survey  trawl  (captures!  and  the 

auxiliary 

let  (escapes)  during  the  2002  capture  efficiency  study.  Crab  have 

been  binned  into  5-mm  carapace  length  intervals. 

Table  2 

Estimated  parameter  and  Akaike's 
and  the  3-parameter  logistic  models 
strapping  1000  replicates  are  provi 
Eastern  survey  bottom  trawl. 

information  criterion  (AIC)  values  for  the  maximum 
for  males,  females,  and  for  sexes  combined.  Parameter 
ied  for  the  final  model  used  to  estimate  red  king  crab 

likelihood  fit  of  the  2-parameter 
variance  estimates  based  on  boot- 
:apture  probability  for  the  83/112 

Sex 

Model 

a 

P 

7 

AIC 

Var  a 

Var/3 

Covariance  a/3 

Male 

1/(1  +e-(a  +  W) 

-0.7366 

0.0178 

— 

1725.3 

0.113 

1.28xl0-5 

-1.14xl0-3 

//(l  +e-(a  +  ft') 

-0.8826 

0.0209 

0.9622 

1727.1 

— 

— 

— 

Female 

l/(l  +  e-(a  +  ») 

0.3540 

0.0054 

— 

1875.7 

0.279 

2.58xl0-5 

-2.63xl0-3 

y/(l  +e-lo  +  b]) 

0.3540 

0.0054 

0.9999 

1877.7 

— 

— 

— 

Combined  sexes 

1/(1  +e-(«  +  W) 

-0.4569 

0.0143 

— 

3612.7 

— 

— 

— 

selectivity  curves  for  males  and  females  with  model  1, 
we  found  the  summed  AIC  for  separate  curves  to  be 
lower  than  the  AIC  for  sexes  combined.  Consequently, 
separate  selectivity  curves  were  estimated  for  males 
and  females  (Table  2). 


The  fitted  model  predicted  male  capture  probabil- 
ity to  be  41.9%  at  23  mm  (size  of  the  smallest  male 
observed);  72.2%  at  95  mm  (size  at  which  full  vulner- 
ability to  the  survey  trawl  is  assigned  in  the  current 
management  model);  80.2%  at  120  mm  (size  assigned 


746 


Fishery  Bulletin  102(4) 


in  the  current  management  model  for  male  maturity); 
84.1%  at  135  mm  (legal  size  for  males);  and  92.7%  at 
184  mm  (size  of  the  largest  male  crab  encountered  in 


1.0- 

o 

Males 

o 

0.8- 

°    ■■c^o 

0.6- 

o 

0.4- 

0.2- 

O 


50  100  150 

Carapace  length  (mm) 

Figure  4 

A  2-parameter  logistic  model  (solid  line)  and  959S  confi- 
dence bounds  idashed  lines)  estimating  male  red  king  crab 
capture  probability  for  the  83/112  Eastern  survey  bottom 
trawl.  Symbols  are  scaled  to  the  sample  length  frequency 
summed  over  all  tows  and  binned  into  5-mm  carapace  length 
intervals.  Symbols  range  in  size  from  the  smallest  circle 
representing  a  single  individual  to  the  largest  circle  repre- 
senting 138  males. 


1  0n 


0.8- 


o 

o.     0.6 

0 


0.4 


0.2 


o 

-~9 

Females 

0^ 

o 

■  06 

o 

o 

0 

0 

50  100  150 

Carapace  length  (mm) 

Figure  5 

A  2-parameter  logistic  model  isolid  line)  and  959!  confi- 
dence bounds  idashed  linesl  estimating  female  red  king  crab 
capture  probability  for  the  83/112  Eastern  survey  bottom 
trawl.  Symbols  are  scaled  to  the  sample  length  frequency 
summed  over  all  tows  and  binned  into  5-mm  carapace  length 
intervals.  Symbols  range  in  size  from  the  smallest  circle 
representing  two  individuals  to  the  largest  circle  represent- 
ing 206  females. 


our  experiment  [Fig.  4]).  The  fitted  model  predicted 
female  capture  probability  to  be  65.27c  at  51  mm  (size 
of  the  smallest  female  observed);  69.8%  at  90  mm  (size 
at  which  both  full  vulnerability  to  the  survey  trawl 
and  50%  female  maturity  are  assigned  in  the  cur- 
rent management  model);  74.7%  at  135  mm  (same 
size  at  which  males  enter  the  fishery);  and  77.4% 
at  162  mm  (size  of  the  largest  female  crab  encoun- 
tered in  our  experiment  [Fig.  5]).  Estimated  capture 
probability  for  both  male  and  female  crab  was  equal 
at  88  mm  (69.9%).  Model  variability,  as  indicated 
by  the  95%  confidence  bounds,  was  greatest  at  the 
extremes  of  our  size  ranges  because  of  low  sample 
frequency.  This  was  especially  true  for  small  crabs, 
and  the  uncertainty  was  so  large  that  extrapolation 
of  the  capture  probability  functions  to  either  males 
or  females  below  <60  mm  is  not  recommended. 

Factors  influencing  escapement 

Modeling  the  effect  of  various  factors  on  capture 
probability  was  based  on  observations  of  RKC 
(ra=248)  from  videotapes  collected  during  28  EBS 
tows.  Approximately  two-thirds  of  the  counted  crabs 
were  captured.  The  influence  of  artificial  lighting, 
body  height,  footrope  distance  from  the  bottom, 
crab  size,  body  orientation,  and  the  interaction  of 
body  height  and  body  orientation  were  significant 
(Table  3).  Capture  probability  decreased  when  lights 
were  used  and  when  the  distance  between  the  foot- 
rope  and  the  bottom  increased.  Capture  probability 
increased  when  crabs  were  standing  up  on  their 
legs,  with  increased  body  size,  and  when  the  footrope 
contact  was  made  along  the  body  axis  rather  than 
from  the  side  of  the  crab. 

Capture  probability,  based  on  direct  observation, 
was  predicted  by  the  fitted  logistic  models  to  illus- 
trate how  the  various  explanatory  variables  affect 
the  capture  outcome.  We  present  two  examples.  In 
the  first  case,  capture  probability  in  natural  light 


Table  3 

Model  coefficients  for  predicting  red  king  crab  cap- 
ture probability  from  counts  obtained  with  a  trawl- 
mounted  video  camera. 

Standard 

Value 

error 

Intercept 

-2.014 

0.676 

Light  variable 

-0.959 

0.526 

Body  height  variable 

3.789 

0.624 

Body  orientation  variable 

0.207 

0.613 

Crab  size  variable 

1.506 

0.315 

Footrope  height  variable 

-0.286 

0.197 

Body  height  and  orientation 

-1.436 

0.785 

interaction 

Weinberg  et  al.:  Capture  probablity  of  a  survey  trawl  for  Paralnhodes  camtschat/cus 


747 


conditions  and  when  crab  are  oriented  sideways 
to  the  oncoming  footrope,  was  predicted  as  a  func- 
tion of  footrope  distance  off  the  bottom  for  each 
size  class,  and  for  both  standing  and  crouching 
crab  (Fig.  6).  For  all  size  groups,  capture  prob- 
ability decreased  with  increasing  footrope  height 
from  the  bottom.  The  importance  of  whether  a 
crab  was  standing  or  crouching  diminishes  with 
decreasing  crab  size  because  the  footrope  is  more 
likely  to  pass  completely  over  smaller  crab.  In 
contrast,  the  importance  of  standing  was  higher 
for  large  crab  because  they  were  more  likely  to  be 
undercut  by  the  footrope  and  captured,  whereas 
crouching  crab  were  more  susceptible  to  hav- 
ing their  legs  first  pinned  down  by  the  footrope, 
which  exerted  a  downward  pressure  on  their 
carapace  and  allowed  the  footrope  to  pass  over 
the  crab.  Capture  probability  of  medium-size  in- 
dividuals, which  included  a  large  proportion  of 
egg-bearing  females,  was  more  dependent  upon 
the  body  height  of  the  crab.  Footrope  contact  be- 
low the  carapace  typically  resulted  in  capture; 
however  contact  above  the  legs  often  forced  the 
crab's  carapace  down,  causing  the  crab  to  roll 
forward  and  pass  beneath  the  footrope. 

In  the  second  case,  capture  probability  was 
predicted  for  natural  light  conditions  when  the 
footrope  is  1  cm  off-bottom,  as  a  function  of  crab 
size,  by  body  orientation  to  the  footrope,  and 
for  standing  and  crouching  individuals  (Fig.  7). 
Under  these  conditions,  capture  probability  was 
greater  for  crab  contacted  along  their  body  ax- 
is than  for  crab  hit  from  the  side.  In  addition, 
capture  probability  increased  with  crab  size  for 
both  standing  and  crouching  crab,  regardless 
of  whether  the  footrope  first  contacted  the  crab 
along  their  body  axis  or  from  the  side.  When  the 
footrope  was  1  cm  off  bottom,  the  difference  in 
the  body-orientation  effect  on  capture  probability 
for  standing  crab  was  greater  for  smaller  crab 
than  for  medium  and  larger  individuals,  but  rela- 
tively equal  for  crouching  crab  of  all  sizes. 


1  0- 

Standing 

08- 

^~~~ . _^^ 

06- 

Medium       ~~~    — -- ________^^ 

04- 

*■"-— _ 

0.2- 

^r— _____^^ 

5      00- 

O 

Q- 

i                      i                      i                      i                      i 

12                            3                           4                            5 

CD 
3 

CO 

O 

Crouching 

0.8- 

Large"""        — ______^ 

0.6- 

^^— -^__^ 

0.4- 

___^                                                                                ^ 

Medium""^" — . 

02- 

~~~~- — _ 

0.0- 

Small 

■                             i                             i                             i                             i 
12                            3                            4                            5 

Footrope  height  (cm) 

Figure  6 

Estimated  red  king  crab  capture  probability  (based  on  direct 

observations!  by  size  class  as  a  function  of  footrope  height  above 

the  bottom  for  standing  and  crouching  individuals.  For  this 

model,  it  was  assumed  that  no  artificial  light  was  used  for  the 

camera  and  that  crab  were  oriented  sideways  to  the  footrope 

upon  initial  contact. 

Discussion 

Our  observations  confirmed  that  adult  Bristol  Bay  red 
king  crab  can  escape  beneath  the  footrope  of  the  AFSC's 
83/112  Eastern  survey  bottom  trawl  under  normal 
towing  conditions.  Capture  probability  increased  with 
size  but  did  not  reach  100%  for  the  largest  crab  caught. 
For  the  current  management  model  used  for  RKC  stock 
assessments,  100%  capture  probability  is  assumed  for 
adult  crabs  and  should  be  revised.  A  recruitment  ogive 
is  used  in  the  calculation  of  the  total  spawning  bio- 
mass  for  defining  overfishing  under  the  Magnuson- 
Stevens  Fishery  Conservation  and  Management  Act 
(Stevens2).  Revised  computations  of  vulnerability  will  be 
required  for  this  purpose  as  well.  Survey  trawl  selectiv- 
ity, although  similar  between  the  two  sexes  at  prerecruit 


sizes,  was  generally  15%  higher  for  legal-size  males 
than  for  equal-size  females.  This  between-sex  difference 
in  capture  probability  may  be  explained  by  behavioral 
differences  (for  instance,  egg-bearing  females  stand  dif- 
ferently from  large  males).  Unfortunately,  crab,  when 
viewed  from  above,  mask  their  gender;  thus  sex  was 
excluded  from  our  modeling  exercise  of  video  data. 

Survey  catch  statistics  for  RKC  are  routinely  included 
in  the  management  modeling  procedure  to  estimate  the 
abundance  of  legal-size  males  (>135  mm),  male  prere- 
cruits  (95-134  mm),  the  effective  spawning  biomass  of 
males  (>120  mm),  and  the  spawning  biomass  of  females 
(>90  mm,  as  determined  from  size  at  50%  maturity).  We 
estimated  capture  probability  for  legal-size  males  (up 
to  184  mm)  to  range  from  84%  to  93%,  for  prerecruit 
males  from  73%  to  84%,  and  for  the  mature  portion 
of  the  male  spawning  population  (up  to  184  mm)  from 


748 


Fishery  Bulletin  102(4) 


80%  to  93%  Our  estimated  capture  probability  for  the 
survey  trawl  on  the  female  portion  of  the  spawning 
RKC  population  ranged  from  70%  to  77%  for  crab  up 
to  162  mm.  A  review  of  the  AFSC  database  for  EBS 
crab  surveys  showed  that  the  largest  male  and  female 
crabs  taken  were  200  mm  and  172  mm.  Corresponding 
capture  probabilities  estimated  by  the  model  for  these 
size  crabs  were  94%  and  78%,  respectively. 

Two  main  factors  affect  the  overall  capture  efficiency 
of  epibenthic  species  by  a  bottom  trawl:  1)  horizontal 
herding,  defined  as  movement  into  the  path  of  the  trawl 
between  the  wingtips  in  response  to  stimuli  produced  by 
the  doors  or  bridles;  and  2)  escapement,  defined  as  the 
avoidance  of  capture  once  the  crab  is  within  the  path  of 
the  trawl.  We  believe  herding  is  negligible  because  our 
observations  of  crab  movement,  which  were  consistent 
with  those  reported  by  Rose  (1999),  indicated  that  RKCs 
are  slow-moving  animals  that  can  travel  only  slight  dis- 
tances before  being  overtaken  by  a  trawl  approaching 


1.0- 

Standing                       ______ 

0.8- 

^       ^^^^ 

0.6- 

j^^^^^ 

04- 

0.2- 

Capture  probability 
o                      o 

■                                                    i                                                    i 
small                                             medium                                             large 

Crouching 

0.8- 

0.6- 

/^^ 

0.4- 

^^^^ 

0.2- 

^  ad"** 

0.0- 

I"                                                                                                        1                                                                                                            1 

small                                             medium                                              large 

Crab  size 

Figure  7 

Estimated  red  king  crab  capture  probability  (based  on  direct 

observations)  by  body  orientation  at  the  time  of  footrope  contact 

as  a  function  of  crab  size  for  standing  and  crouching  individuals. 

For  this  model,  it  was  assumed  that,  no  artificial  light  was  used 

for  the  camera  and  that  the  footrope  was  1  cm  off  bottom. 

at  1.5  m/sec.  Our  video  observations  of  the  trawl  bridle 
revealed  that  RKCs  consistently  passed  over  the  top  of 
the  bare  cable,  with  one  exception — where  a  few  crabs 
were  seen  sliding  along  the  bridle,  legs  entangled,  to  the 
wingtip  before  being  cast  outside  the  path  of  the  trawl. 
Escapement  is  likely  restricted  to  footrope  escapement 
because  mesh  escapement  is  impeded  by  the  spiny  sur- 
face and  long  legs  of  the  crab  and  could  only  occur  for 
the  smallest  individuals,  which  we  encountered  in  low 
numbers  and  which  could  not  be  predicted  reliably  by 
our  model. 

We  recognize  from  the  analysis  of  our  in  situ  data 
that  capture  probability  is  influenced  not  only  by  trawl 
performance  but  also  by  crab  behavior.  For  instance, 
crabs  standing  upright,  such  as  moving  or  migrating 
individuals,  are  more  susceptible  to  capture  than  those 
with  their  bodies  resting  on  the  substrate.  Crab  density 
could  also  affect  capture  probability  as  seen  for  some 
species  offish  (Godo  et  al.,  1999).  The  crabs  we  observed 
with  our  video  cameras  were  fairly  dispersed  and 
the  maximum  number  of  crabs  seen  in  any  single 
video  frame  was  two  (twice  observed).  Crabs  in 
relatively  low  abundance  are  likely  to  react  di- 
rectly to  the  gear,  but  in  areas  of  high  abun- 
dance, crabs  may  react  to  each  other  in  response 
to  the  stimuli  from  the  approaching  gear,  causing 
them  to  crouch  or  conversely  move  away  from 
perceived  danger.  Both  of  these  responses  would 
result  in  a  different  capture  probability. 

Our  estimates  of  capture  probability  apply 
to  the  conditions  in  which  the  EBS  survey  is 
conducted;  that  is,  relatively  disperse  offshore 
populations  encountered  during  daylight  hours  on 
sandy  bottom  during  the  summer  months.  There 
are  other  behavioral  factors  or  environmental 
conditions  that  we  did  not  consider  in  the  present 
study  but  which  could  affect  the  efficiency  of  the 
survey  trawl.  These  include,  but  are  not  limited 
to,  the  following:  trawling  where  the  substrate  is 
substantially  different;  crabs  that  are  either  ag- 
gregated into  pods  or  are  buried  (Dew3);  and  tem- 
peratures or  tidal  currents  that  would  affect  the 
migratory  or  feeding  behavior,  and  therefore  the 
body  height  of  crab  (Dew,  1990).  Our  estimates 
of  capture  probability  are  also  based  on  the  as- 
sumption that  the  auxiliary  net  is  100%  efficient 
at  capturing  crab  escaping  beneath  the  footrope 
of  the  survey  trawl.  We  have  no  direct  evidence  to 
believe  otherwise.  However,  if  crabs  also  escaped 
the  auxiliary  net,  then  our  estimates  of  capture 
probability  would  be  too  large. 

In  conclusion,  we  wish  to  clarify  to  users  of 
our  findings  that,  although  these  experimen- 
tally determined  selectivity  models  indicate  an 
upward  correction  in  spawning  biomass  of  red 
king  crab  may  be  in  order,  we  find  no  reason 


3  Dew.  C.  B.  2003.  Personal  commun.  Alaska 
Fisheries  Science  Center,  7600  Sand  Point  Wav  NE, 
Seattle,  WA  98115. 


Weinberg  et  al.:  Capture  probablity  of  a  survey  trawl  for  Para/ithodes  camtschaticus 


749 


to  claim  that  the  stock  is  in  any  better  condition  than 
the  condition  that  was  determined  by  the  most  recent 
assessment.  The  foremost  utility  of  the  AFSC  annual 
EBS  surveys  is  to  monitor  distribution  and  abundance 
trends  through  time.  The  survey  accomplishes  this  by 
maintaining  strict  protocols  and  consistency  in  trawling 
methods,  in  computation  of  area-swept  abundance,  and 
in  nonenvironmentally  affected  trawl  efficiency.  The 
survey  times  series  is  designed  to  detect  changes  in 
abundance,  signaling  advances  in  the  population's  re- 
building processes,  regardless  of  whether  crab  are  100% 
or  80%  vulnerable  to  the  survey  trawl.  We  advocate  that 
careful  consideration  be  given  to  the  other  factors  that 
drive  the  management  model,  along  with  the  results  of 
our  capture  efficiency  experiment,  to  ensure  that  the 
stock  rebuilding  process  remains  uninterrupted. 


Acknowledgments 

We  are  thankful  to  Captain  Glenn  Sullivan  and  the  crew 
of  the  FV  Arcturus  for  their  professional  attitudes  and 
relentless  attention  to  detail;  to  scientists  Chris  John- 
ston, Frank  Shaw,  and  Kerim  Aydin  for  their  assistance 
at  sea  following  the  2002  survey;  to  Craig  Rose  and  Scott 
McEntire  for  technical  support;  to  Dave  King  and  Jim 
Smart  for  preparation  of  the  experimental  trawl  gear; 
and  to  Gary  Walters,  Braxton  Dew,  Doug  Pengilly,  Jie 
Zheng,  and  our  anonymous  reviewers  for  their  helpful 
comments  during  the  manuscript  review  process. 


Literature  cited 

Armistead,  C.  E.,  and  D.  G.  Nichol. 

1993.     1990  bottom  trawl  survey  of  the  eastern  Bering 
Sea  continental  shelf.     NOAA  Tech.  Memo.  NMFS- 
AFSC-7,  190  p. 
Burnham,  K.  P..  and  D.  R.  Anderson. 

1998.     Model  selection  and  inference:  a  practical  infor- 
mation-theoretic approach,  .353  p.     Springer-Verlag, 
New  York,  NY. 
Dew,  C,  B. 

1990.  Behavioral  ecology  of  podding  red  king  crab, 
Paralithodes  eamtsehatica.  Can.  J.  Fish.  Aquat.  Sci. 
47:1944-1958. 


Efron,  B.,  and  R.  Tibshirani. 

1993.  An  introduction  to  the  bootstrap,  436  p.  Chapman 
and  Hall,  New  York,  NY. 

Engas,  A.,  and  O.  R.  Godo. 

1989.     Escape  offish  under  the  fishing  line  of  a  Norwegian 
sampling  trawl  and  its  influence  on  survey  results.     J. 
Cons.  Int.  Explor.  Mer  45:269-276. 
God0,  O.  R..  S.  J.  Walsh,  and  A.  Engas. 

1999.     Investigating  density-dependent  catchability  in 
bottom  trawl  surveys.     ICES  J.  Mar.  Sci.  56:292- 
298. 
Millar,  R.  B. 

1992.     Estimating  the  size-selectivity  of  fishing  gear  by 
conditioning  on  the  total  catch.     J.  Am.  Stat.  Assoc. 
87:962-968. 
Munro,  R  T.,  and  D.  A.  Somerton. 

2001.  Maximum  likelihood  and  non-parametric  methods 
for  estimating  trawl  footrope  selectivity.  ICES  J.  Mar. 
Sci.  58:220-229. 

2002.  Estimating  net  efficiency  of  a  survey  trawl  for  flat- 
fishes.    Fish.  Res.  55:267-279. 

Rose.  C.  R. 

1999.     Injury  rates  of  red  king  crab,  Paralithodes  camts- 
chaticus, passing  under  bottom-trawl  footropes.     Mar. 
Fish.  Rev.  61(21:72-76. 
Somerton,  D.  A.,  and  R.  S.  Otto. 

1999.     Net  efficiency  of  a  survey  trawl  for  snow  crab, 
Chionoecetes  opilio,  and  Tanner  crab,  C.  bairdi.     Fish. 
Bull.  97:617-625. 
Somerton,  D.  A.,  and  K.  L.  Weinberg. 

2001.     The  affect  of  speed  through  the  water  on  footrope 
contact  of  a  survey  trawl.     Fish.  Res.  53:17-24. 
Venables,  W.  N,  and  B.  D.  Ripley. 

1994.  Modern  applied  statistics  with  S-plus,  452  p. 
Springer-Verlag,  New  York,  NY. 

Walsh,  S.  J. 

1992.  Size-dependent  selection  at  the  footgear  of  a  ground- 
fish  survey  trawl.  N.  Am.  J.  Fish.  Manag.  12:625- 
633. 

Weinberg,  K.  L. 

2003.  Change  in  the  performance  of  a  Bering  Sea  survey 
trawl  due  to  varied  trawl  speed.  Alaska  Fish.  Res. 
Bull.  10(11:42-49. 

Zheng,  J.,  M.  C.  Murphy,  and  G.  H.  Kruse. 

1995.  A  length-based  population  model  and  stock-recruit- 
ment relationships  for  red  king  crab,  Paralithodes  camts- 
chaticus, in  Bristol  Bay.  Alaska.  Can.  J.  Fish.  Aquat. 
Sci.  52  (61:1229-1246. 


750 


Evidence  of  shark  predation  and  scavenging  on 
fishes  equipped  with  pop-up  satellite  archival  tags 


David  W.  Kerstetter 

School  of  Marine  Science 
Virginia  Institute  of  Marine  Science 
College  of  William  and  Mary 
Gloucester  Point,  Virginia  23062 
E-mail  address:  bailey@vims.edu 

Jeffery  J.  Polovina 

Pacific  Islands  Fisheries  Science  Center 
National  Marine  Fisheries  Service 
Honolulu,  Hawaii  96822 

John  E.  Graves 

School  of  Marine  Science 
Virginia  Institute  of  Marine  Science 
College  of  William  and  Mary 
Gloucester  Point,  Virginia  23062 


Over  the  past  few  years,  pop-up  sat- 
ellite archival  tags  (PSATs)  have 
been  used  to  investigate  the  behav- 
ior, movements,  thermal  biology, 
and  postrelease  mortality  of  a  wide 
range  of  large,  highly  migratory  spe- 
cies including  bluefin  tuna  (Block  et 
al.,  2001),  swordfish  (Sedberry  and 
Loefer,  2001),  blue  marlin  (Graves  et 
al.,  2002),  striped  marlin  (Domeier 
and  Dewar,  2003),  and  white  sharks 
(Boustany  et  al.,  2002).  PSAT  tag 
technology  has  improved  rapidly, 
and  current  tag  models  are  capable 
of  collecting,  processing,  and  stor- 
ing large  amounts  of  information  on 
light  level,  temperature,  and  pressure 
(depth)  for  a  predetermined  length  of 
time  before  the  release  of  these  tags 
from  animals.  After  release,  the  tags 
float  to  the  surface,  and  transmit  the 
stored  data  to  passing  satellites  of  the 
Argos  system. 

A  problem  noted  by  several  au- 
thors using  early  PSAT  models  was 
the  occasional  occurrence  of  tags 
that  did  not  transmit  data.  Clearly, 
a  tag  attached  to  a  moribund  fish 
that  would  sink  to  a  depth  exceeding 
the  pressure  limit  of  the  tag  casing 
would  be  destroyed.  To  prevent  the 
loss  of  tags  due  to  mortality  events, 
tag  manufacturers  and  researchers 


have  developed  mechanisms  that  re- 
lease tags  from  dead  or  dying  fish 
before  the  structural  integrity  of  the 
tag  is  compromised  at  depth.  These 
mechanisms  include  both  mechani- 
cal devices  that  sever  the  monofila- 
ment tether  that  attaches  the  tag  to 
the  fish  upon  reaching  a  given  depth 
and  internal  software  subroutines 
that  activate  the  normal  electronic 
release  mechanism  if  the  tag  either 
reaches  a  certain  depth  or  maintains 
a  constant  depth  for  a  predetermined 
length  of  time. 

Despite  the  addition  of  these  re- 
lease mechanisms  to  PSATs,  some 
tags  still  fail  to  transmit  data.  Such 
failure  could  result  from  any  of  the 
following  events  or  conditions:  me- 
chanical failure  of  a  critical  tag  com- 
ponent; destruction  by  fishing  crews 
unaware  of  or  not  participating  in 
the  present  research;  excessive  epi- 
faunal  growth  that  makes  the  tag 
negatively  buoyant  or  prevents  the 
tag  from  floating  with  the  antenna 
in  a  vertical  position;  or  fouling  of 
the  tag  on  the  fish,  fishing  gear,  or 
flotsam.  Another  cause  of  failure  is 
that  the  tags  could  be  lost  as  a  re- 
sult of  ingestion.  For  example,  a  free- 
swimming  white  marlin  (Tetrapturus 
albidus)  was  observed  mouthing  and 


almost  swallowing  a  free-floating 
PSAT  off  the  Dominican  Republic  in 
May  2002  (Graves,  personal  observ.). 
Alternately,  the  tag  could  be  ingested 
incidentally  with  part  of  the  tagged 
fish,  as  described  by  Jolley  and  Irby 
(1979)  who  reported  that  an  acoustic 
tag  on  a  sailfish  (Istiophorus  platyp- 
terus)  was  eaten  along  with  the  fish 
by  an  undetermined  species  of  shark. 
In  this  note,  we  present  data  from 
PSATs  deployed  on  two  white  marlin 
in  the  western  North  Atlantic  Ocean 
and  on  an  opah  (Lampris  guttatus) 
in  the  central  Pacific;  the  data  from 
these  tags  indicate  that  the  tags  were 
consumed  by  sharks. 


Materials  and  methods 

White  marlin  1  (WM1) 

At  approximately  10:00  am  local  time 
on  1  September  2002,  a  white  marlin 
was  observed  on  pelagic  longline 
gear  set  during  the  night  near  the 
southeastern  edge  of  Georges  Bank. 
The  fish,  which  had  been  caught 
on  a  slightly  offset,  straight-shank 
J-style  hook  (size  9/0),  was  manu- 
ally guided  with  the  leader  along- 
side the  vessel.  A  PTT-100  HR  model 
PSAT  (Microwave  Telemetry,  Inc., 
Columbia,  MD)  was  attached  to  the 
dorsal  musculature  approximately  5 
cm  below  the  base  of  the  dorsal  fin 
with  a  large  nylon  anchor  according 
to  the  procedure  and  tether  design 
described  in  Kerstetter  et  al.  (20031. 
The  tag  was  activated  shortly  after 
the  white  marlin  was  first  identified, 
although  approximately  one  hour  is 
required  following  activation  for  this 
tag  model  to  begin  collecting  data. 
The  tag  was  programmed  to  record 
point  measurements  of  temperature, 
light,  and  pressure  (depth  I  in  four- 
minute  time  intervals  and  to  detach 
from  the  animal  after  10  days.  After 
release  from  the  fish,  the  positively 
buoyant  tag  was  expected  to  float  to 
the  surface  and  transmit  stored  and 
real-time  data.  For  both  white  marlin 


Manuscript  submitted  27  April  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
7  June  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:750-756120041. 


NOTE     Kerstetter  et  al.:  Shark  predation  and  scavenging  on  fishes  equipped  with  satellite  archival  tags 


751 


Table  1 

Comparison  of  depths 

and  temperatures  recorded  by  three 

pop-up  sa 

tellite  archiva 

tags  (PSATs)  before  and 

after  the  tags 

were 

ingested  by 

an  organism. 

Befo 

•e  ingestion 

Afte 

r  ingestion 

Depth 

Depth 

Temp. 

Temp 

Depth 

Depth 

Temp 

Temp 

range 

mean 

range 

mean 

range 

mean 

range 

mean 

Animal 

(m) 

(SD) 

CO 

(SD) 

;; 

(m) 

(SD) 

CO 

(SD) 

n 

WM1 

145.2 

145.2 

(±0.00) 

11.6-11 

11.7 
(±0.07l 

179 

0-564.9 

130.0 

(±237.50) 

12.1-26.5 

24.1 
(±0.84) 

2755 

WM2 

0-26.9 

5.9 
(±4.44) 

19.8-27.8 

24.7 
(±0.91) 

207 

0-699.3 

131.0 

(±162.61) 

18.9-29.5 

27.3 
(±1.20) 

1683 

Opah 

32-456 

221.81 
(±92.20) 

8-25.6 

16.68 
(±4.21) 

360 

0-524 

170.56 
(±133.83) 

26.2-30.6 

28.64 
(±0.67) 

168 

tags,  minimum  straight-line  distances  were  calculated 
between  the  point  of  release  and  the  first  clearly  trans- 
mitted location  of  the  tag  following  its  release  (pop-off) 
(Argos  location  codes  0-3). 

At  the  time  of  tagging,  the  longline  hook  used  to  cap- 
ture the  fish  was  not  visible  in  the  mouth  of  the  white 
marlin.  The  leader  was  therefore  cut  as  close  as  possible 
to  the  fish  before  the  fish  was  released,  following  the 
standard  operating  procedure  for  the  domestic  pelagic 
longline  fleet.  The  fish  was  maintained  alongside  the 
vessel  for  less  than  three  minutes  for  the  application  of 
the  PSAT  and  a  conventional  streamer  tag.  Although  the 
white  marlin  was  initially  active  at  the  side  of  the  vessel, 
some  light  bleeding  from  the  gills  was  noted.  After  re- 
lease, the  fish  swam  away  slowly  under  its  own  power. 


the  dorsal  musculature  with  a  Wildlife  Computers  tita- 
nium anchor.  The  tag  was  programmed  to  record  the 
temperature  and  depth  occupied  by  the  fish  in  binned 
histograms,  and  the  minimum  and  maximum  tempera- 
tures and  depths  for  12-hour  time  periods.  However, 
these  12-hour  bins  encompassed  both  day  and  night 
periods.  The  tag  was  programmed  to  be  released  six 
months  after  deployment.  In  the  event  of  a  premature 
release,  the  tag  was  programmed  to  begin  transmitting 
stored  data  if  it  remained  at  the  surface  for  longer  than 
three  days.  The  opah  was  lively  and  quickly  dived  after 
it  was  released. 


Results 


White  marlin  2  (WM2) 

At  9:05  am  on  2  August  2003,  a  white  marlin  was 
observed  on  pelagic  longline  gear  with  the  same  configu- 
ration in  the  same  approximate  area  of  Georges  Bank  as 
WM1.  The  fish  was  caught  by  a  circle  hook  (size  16/0)  in 
the  right  corner  of  the  mouth,  and  although  the  stomach 
was  everted,  the  fish  appeared  to  be  in  excellent  physi- 
cal condition.  A  PTT-100  HR  tag  had  been  activated  at 
6:30  am  that  morning,  and  was  therefore  collecting  data 
at  the  time  of  tagging.  After  the  fish  was  brought  to  the 
side  of  the  vessel,  both  the  PSAT  and  a  conventional 
streamer  tag  were  attached  to  this  fish  in  less  than  three 
minutes  by  using  the  same  protocol  as  that  described  for 
WM1,  and  the  fish  swam  strongly  away  from  the  vessel 
after  release  without  any  evident  bleeding. 

Opah 

At  5:52  pm  local  time  on  21  November  2002,  a  female 
opah  was  observed  on  pelagic  longline  gear  set  during 
the  day  east  of  the  Island  of  Hawaii.  The  fish  was  brought 
to  the  side  of  the  fishing  vessel  and  a  Wildlife  Computers 
(Redmond.  WA)  PAT2  model  tag  was  attached  through 


WM1 

Release  of  the  PSAT  was  expected  to  occur  on  10  Sep- 
tember 2002  and  the  tag  was  expected  to  begin  transmit- 
ting data  on  that  date,  but  the  first  transmission  was 
not  received  until  almost  two  days  later.  At  the  time  of 
first  transmission,  the  PSAT  was  81.3  km  (43.9  nmi) 
west-southwest  of  the  tagging  location.  A  total  of  81.5% 
of  the  archived  light  level,  temperature,  and  pressure 
(depth)  data  was  recovered. 

The  light  level,  temperature,  and  pressure  (depth) 
readings  over  time  are  presented  in  Fig.  1  (A-C)  and 
summarized  in  Table  1.  The  first  light  level  measure- 
ments indicated  that  the  fish  was  already  in  relatively 
dark  waters  within  one  hour  following  its  release.  Light 
levels  continued  to  drop  to  almost  zero  during  the  next 
ten  hours  and  remained  at  that  level  for  the  next  nine 
days  (Fig.  1A).  During  the  next  seven-day  surface  trans- 
mission period,  the  tag  recorded  real-time  day  and  night 
differences  in  light  levels,  which  indicated  that  the  light 
sensor  was  functioning  properly. 

Sea  surface  temperatures  in  the  area  where  the  gear 
was  set  and  hauled  back,  varied  from  25.2°  to  26.7°C 
(D.  Kerstetter.  unpubl.  data)  and  the  first  temperature 


752 


Fishery  Bulletin  102(4) 


Q. 
CD 
D 


A 

WM1 

flu 

1 

200  • 

i 

i"f 

f 

400  • 

'A 

i 

1  !l 

i: 

600  • 

i 

9/1/02        9/4/02        9/7/02       9/10/02      9/13/02      9/16/02      9/19/02 


B 


9/1/02        9/4/02        9/7/02       9/10/02      9/13/02      9/16/02      9/19/02 


c 

1.0' 

™ 

"" 

"T  r 

np 

08. 

06- 

04. 

02. 

1 

~ 

9/1/02         9/4/02         9/7/02        9/10/02       9/13/02       9/16/02       9/19/02 


WM2 


8/3/03       8/6'03       8/9/03     8/12/03    8/15/03    8/18/03     8/21/03     8/24/03 


E 


32 

28- 

24- 
20- 


8/3/03       8/6/03       8/9/03      8/12/03    8/15/03     8/18/03    8/21/03    8/24/03 


8/3/03       8/6/03       8/9/03      8  12/03    8/15/03     8/18/03    8/21/03    8  24,03 


Figure  1 

Graphs  of  data  on  and  depth  (A  and  Di.  temperature  (B  and  E),  and  light  index  iC  and  Fl  for  tags  WM1 
and  WM2.  Lighter  lines  and  points  are  prior  to  programmed  release  date,  whereas  darker  lines  and  points 
are  "real-time"  surface  condition  measurements  transmitted  by  the  tag  in  addition  to  the  archived  data. 


recording  by  the  PSAT  (one  hour  after  activation)  was 
11°C  (Fig.  IB).  The  temperature  remained  fairly  con- 
stant at  11°C  for  a  period  of  approximately  ten  hours 
after  which  there  was  a  rapid  rise  to  25°C.  The  temper- 
ature of  the  PSAT  remained  between  22.5°  and  26. 5C 
for  the  next  nine  days  (until  the  programmed  release 
date),  with  the  exception  of  one  brief  decrease  to  20  C 
on  8  September.  When  the  tag  began  transmitting  on 
12  September,  the  real-time  surface  temperature  was 
23.6°C. 

The  pressure  data  (Fig.  IC)  indicated  that  the  tag  was 
at  a  depth  of  approximately  145  m  at  one  hour  following 
release.  The  PSAT  remained  at  this  depth  for  a  little 
more  than  ten  hours  after  which  the  data  suggested 
that  there  was  a  rapid  rise  to  the  surface.  For  the  next 
nine  days,  the  tag  reported  considerable  vertical  move- 


ment between  the  surface  and  depths  to  565  m.  The 
tag  was  at  the  surface  when  it  began  transmitting  both 
archived  and  real-time  data  on  12  September. 

WM2 

The  tag  reported  data  as  expected  on  13  August  2003 
and  transmitted  57.3rr  of  the  archived  data.  At  the  time 
of  first  transmission,  the  PSAT  was  600.1  km  (324.0 
nmi)  east-southeast  of  the  tagging  location.  Summary 
depth  and  temperature  data  recorded  by  the  PSAT  are 
included  in  Table  1. 

From  the  depth  and  temperature  data,  it  appears 
that  the  fish  survived  for  approximately  24  hours  af- 
ter release,  at  which  point  the  light  readings  dropped 
to  zero  (see  Fig,  ID)  and  remained  at  that  level  for 


NOTE     Kerstetter  et  al.:  Shark  predation  and  scavenging  on  fishes  equipped  with  satellite  archival  tags 


753 


the  next  eight  days.  The  depth  record  following 
this  change  in  light  level  was  marked  by  several 
discrete  diving  events,  and  depths  (see  Fig.  IF) 
ranged  between  the  surface  and  over  699  m. 
Recorded  temperatures  for  this  period  varied 
between  18.9°  and  29.5°C,  although  sea  surface 
temperatures  in  the  area  where  gear  was  set  and 
hauled  back  varied  from  20.9°  to  26.0°C  (Ker- 
stetter, unpubl.  data).  On  12  August,  the  light 
level  returned  to  its  maximum  value  and  the  tag 
remained  at  the  surface  for  approximately  one 
day  until  its  scheduled  release  date  (13  August) 
when  it  began  transmitting  data. 

Opah 

The  PAT2  satellite  tag  was  expected  to  pop-up 
6  months  after  deployment,  but  the  first  trans- 
mission was  received  after  only  34  days  from  a 
location  about  330  km  (178  nmi)  northwest  of  the 
deployment  site.  All  the  archived  binned  light 
level,  temperature,  and  pressure  (depth)  data 
from  this  period  were  recovered  (see  Table  1). 
This  tag  model  collected  eight  temperature  and 
depth  samples  during  each  12-hour  period,  result- 
ing in  16  values  per  day  or  528  total  values  for 
the  deployment  period.  The  two  12-hour  blocks 
were  removed  from  all  analyses  to  more  accu- 
rately represent  the  differences  in  data  between 
specimens:  1)  the  12-hour  block  after  tagging  in 
order  to  allow  for  the  recovery  of  the  animal,  and 
2)  the  12-hour  block  during  which  the  predation 
event  putatively  occurred  in  order  to  clarify  the 
potentially  distant  depth  and  temperature  char- 
acteristics of  the  ingesting  animal. 

The  measured  sea  surface  temperature  during 
the  tagging  of  the  opah  was  25.9°C.  The  ranges 
of  dive  depths,  temperature,  and  light  based  on 
minimum  and  maximum  values  over  the  12-hour 
day  and  night  periods  showed  two  distinct  pat- 
terns (Fig.  2).  During  the  first  period  (23  days), 
the  dive  depths  ranged  from  about  32  to  456 
m  (Fig.  2A).  Water  temperatures  encountered 
by  the  tag  during  this  period  ranged  from  8.0° 
to  25.6°C  (Fig.  2B)  and  the  light  index  values 
ranged  from  about  50  to  150  (Fig.  2C).  During 
the  second  period  (11  days),  the  dive  depths  ranged  from 
0  to  524  m,  temperature  ranged  from  26.2°  to  30.6°C 
(higher  than  the  24.2-24.8°C  SST  recorded  by  the  tag 
after  it  was  released  from  the  fish),  and  the  light  index 
recorded  persistently  low  values. 


Discussion 

WM1 

Our  interpretation  of  these  data  is  that  the  PSAT 
on  WM1  was  ingested  by  an  animal  scavenging  the 
marlin  carcass.  The  first  PSAT  readings  for  WM1, 


Dive  minimums  and  maximums 

0 

100 

E 

200 

Q 

300 
400 
500 

' 

11/24/02        11/30/02         12/6/02         12/12/02        12/18/02        12/24/02 

Dive  minimums  and  maximums 

28 

'.•■'   Ii''   I'-'-'l'l 

O 

24- 

o 

Q. 

E 
|2 

20  - 
16  - 
12  - 

R 

O                              i                  '                  i                  '                  1                  '                  l                  '                  I                  '                  i 

11/24/02        11/30/02         12/6/02         12/12/02        12/18/02        12/24/02 

Light  minimums  and  maximums  from  location  data 

250 

200- 

en 
0 
=J 
CO 

> 

150 

100 

_i 

50-            '           I    I 
0 

,  |                        ..,11... 

I                      I                                            I           '           1           '           I 
11/24/02        11/30/02         12/6/02         12/12/02        12/18/02       12/24/02 

Date 

Figure  2 

G 

raphs  of  depth  (A),  temperature  (B),  and  light  index  (C)  for 

the  opah  PAT  tag  from  deployment  until  transmission. 

recorded  about  one  hour  after  its  release,  indicated 
that  the  marlin  was  already  dead  or  moribund  by 
that  time  and  was  descending  to  the  ocean  floor.  For 
the  next  ten  hours,  the  tag  and  carcass  remained  at 
a  constant  depth  of  145  m  (the  depth  of  the  nearest 
sounding  at  the  site  of  release,  according  to  NOAA 
depth  chart  13003  [1998],  was  approximately  160  m) 
and  at  a  temperature  of  11°C.  The  light  level  steadily 
decreased  at  approximately  4:30  pm,  corresponding  to 
changes  in  ambient  light  from  the  setting  of  the  sun. 
At  approximately  9:00  pm  local  time,  there  was  a  dra- 
matic change  in  conditions  when  temperature  rapidly 
rose  to  near  26°C  and  depths  began  to  vary  between 
the  surface  and  600  m. 


754 


Fishery  Bulletin  102(4) 


We  cannot  attribute  these  chang- 
es to  a  resuscitation  of  the  fish  for 
three  reasons.  1)  The  measured  light 
levels  indicated  that  the  tag  was  in 
complete  darkness  for  a  period  often 
days,  even  though  it  was  at  the  sur- 
face during  daylight  hours.  A  mal- 
functioning light  sensor  cannot  ex- 
plain this  observation  because  the  tag 
recorded  day  and  night  differences 
in  light  levels  at  the  surface  during 
the  seven-day  transmission  period  af- 
ter it  was  released  from  the  fish.  2) 
After  a  rapid  increase,  the  tempera- 
ture remained  relatively  constant, 
between  23°  and  26°C,  even  when 
the  tag  was  at  depths  in  excess  of 
300  m.  Although  dive  behavior  may 
be  affected  by  location-specific  con- 
ditions, previous  PSAT  observations 
of  more  than  20  other  white  marlin 
indicated  that  temperature  ranges  of 
individual  dive  events  rarely  exceed 
8°C  when,  it  is  assumed,  animals 
make  foraging  dives  to  depth  (Horo- 
dysky  et  al.,  in  press).  3)  The  PSAT 
recorded  several  dives  in  excess  of  400  m,  and  previous 
observations  of  white  marlin  have  revealed  no  dives  in 
excess  of  220  m  (Horodysky  et  al.,  in  press).  Finally, 
the  PSAT  was  scheduled  to  be  released  from  WM1  after 
ten  days  on  10  September.  Although  archiving  of  light, 
temperature,  and  pressure  data  ceased  on  that  date,  the 
tag  did  not  begin  transmitting  until  12  September. 

WM2 

The  shallow  dive  patterns  reported  by  this  fish  may 
indicate  that  it  survived  for  approximately  24  hours 
following  its  release.  Between  12:45  and  3:07  pm  (local 
time),  the  light  level  fell  abruptly  from  the  maximum 
light  level  value  to  zero.  At  3:08  pm,  the  temperature 
was  19.8°C  at  166  m  depth;  by  4:37  pm,  the  tempera- 
ture was  above  24°C  and  remained  above  this  value  for 
the  remainder  of  the  deployment  period.  At  5:58  pm  on 
12  September,  the  light  levels  returned  to  maximum 
strength  from  zero — an  indication  that  the  tag  had 
likely  been  egested.  For  the  19  hours  remaining  of  the 
programmed  deployment  period  prior  to  pop-off,  the 
depth,  light,  and  temperature  data  all  indicated  that 
the  tag  was  floating  at  the  surface. 

Opah 

Based  on  recovered  data,  our  conjecture  is  that  the  tag 
was  attached  to  the  live  opah  for  the  first  23  days.  Then, 
sometime  during  the  12-hour  period  from  2:00  pm  13 
December  to  2:00  am  14  December  the  tag  was  ingested. 
From  our  data,  we  cannot  discern  whether  1)  the  tag 
was  detached  prematurely  from  the  opah  and  was  float- 
ing on  the  surface  when  it  was  ingested,  2)  an  animal 


D 
CD 

■o 


Temperature 
Depth 


Figure  3 

Delayed  temperature  changes  recorded  by  tag  WM1  following  deep  dive 
events  on  the  morning  of  2  September  2002.  Arrows  indicate  the  lowest 
temperatures  recorded  in  association  with  a  movement  of  the  animal  to 
depth;  note  that  these  temperatures  were  often  recorded  while  the  animal 
was  at  or  near  the  surface  and  therefore  represent  a  delay  between  depth 
and  temperature. 


attacked  the  opah  and  ingested  the  tag  incidentally, 
or  3)  an  animal  ingested  the  tag  alone.  However,  it  is 
unlikely  that  the  opah  died,  sank  to  the  ocean  floor,  and 
was  scavenged  because  the  ocean  floor  in  the  area  where 
the  opah  was  tagged  is  below  2000  m.  We  have  observed 
from  other  tags  on  opahs  what  we  believe  are  mortalities; 
these  occur  shortly  after  tagging  and  show  that  the  tag 
reaches  depths  in  excess  of  1000  m  before  detaching  when 
the  emergency  pressure  release  in  the  tag  is  triggered.  We 
did  not  observe  depths  below  600  m  at  any  time  during 
this  record,  and  therefore  the  pressure-induced  detach- 
ment mechanism  on  the  tag  was  not  triggered. 

The  ingestion  hypothesis  for  the  failure  of  these  three 
tags  to  transmit  data  is  supported  by  several  lines  of 
evidence.  First,  the  light  level  readings  were  consistent 
with  a  tag  residing  in  the  complete  darkness  of  an 
alimentary  canal.  Second,  although  temperature  varia- 
tions occurred  during  the  deployment  period,  the  delay 
in  temperature  changes  during  dives  to  depths  indicates 
that  the  tags  were  not  directly  exposed  to  ambient  wa- 
ter (see  Fig.  3  for  an  example  from  WM1,  as  well  as  the 
comparisons  in  Table  1)  and  further  may  indicate  that 
the  scavenger  was  either  endothermic  or  of  large  enough 
size  to  mitigate  heat  loss  at  depth. 

There  are  several  organisms  that  could  have  eaten 
these  PSATs,  whether  by  scavenging  a  carcass  or  at- 
tacking a  moving  fish.  Clearly,  each  of  these  organisms 
was  sufficiently  large  to  ingest  the  tag  without  seri- 
ously damaging  it.  It  is  unlikely  that  a  cetacean  was 
responsible  for  any  of  these  events  because  internal 
temperatures  for  odontocete  whales  (including  killer 
whales,  Orcinus  orca)  range  between  approximately  36° 
and  38°C  (Whittow  et  al.,  1974)— well  above  the  range 
of  temperatures  recorded  by  the  PSATs. 


NOTE     Kerstetter  et  al.:  Shark  predation  and  scavenging  on  fishes  equipped  with  satellite  archival  tags 


755 


The  only  other  natural  predators  of  large  pelagic 
fishes  are  various  species  of  sharks.  Several  species  of 
lamnid  sharks  maintain  elevated  body  temperatures, 
including  the  shortfin  mako  {Isurus  oxyrinchus)  and  the 
white  shark  (Carcharodon  carchariasl,  both  of  which  are 
found  in  the  area  of  Georges  Bank  (Cramer,  2000)  and 
the  Central  Pacific  (Compagno,  1984).  Several  shortfin 
makos  were  caught  by  the  same  longline  vessel  during 
the  week  following  each  white  marlin  PSAT  deployment 
(WM1:  n=4,  95-189  cm  FL;  WM2:  n=3,  94-199  cm  FL) 
(Kerstetter,  unpubl.  data).  The  opah  tag  record  closely 
resembles  the  relatively  constant  temperature  noted  for 
lamnid  sharks,  despite  the  independence  of  stomach 
temperature  with  ambient  water  for  these  endothermic 
sharks  as  reported  by  Carey  et  al.  (1981).  It  is  also 
interesting  to  note  that  although  precipitous  tempera- 
ture fluctuations  were  generally  absent,  a  rapid  drop  in 
temperature  from  24°  to  20°C  was  observed  with  tag 
WM1  on  8  September  at  32.3  m  depth— a  fluctuation 
that  could  have  resulted  from  another  feeding  event 
that  brought  cool  food  matter  into  the  stomach.  Simi- 
lar reductions  in  stomach  temperatures  due  to  feeding 
have  been  noted  for  white  sharks  (McCosker,  1987).  The 
range  of  temperatures  recorded  by  each  of  the  two  white 
marlin  tags  appears  rather  broad  for  an  endothermic 
shark,  however,  and  although  the  temperature  at  depth 
was  not  measured,  the  delay  in  stomach  temperature 
closely  resembles  the  pattern  of  blue  shark  internal 
temperatures  {Prionace  glauca)  measured  in  the  Mid- 
Atlantic  (Carey  and  Scharold,  1990). 

The  diving  behavior  recorded  by  the  three  tags  also 
corroborates  ingestion  of  the  tags  by  sharks.  Carey  et  al. 
(1982)  reported  that  a  tagged  white  shark  off  Long  Is- 
land, New  York,  made  frequent  dives  to  the  bottom  dur- 
ing a  3.5-day  acoustic  tracking  period.  White  sharks  are 
known  to  dive  to  depth  while  scavenging  whale  carcasses 
(Dudley  et  al.,  2000;  Carey  et  al.,  1982).  A  juvenile  white 
shark  also  tracked  by  Klimley  et  al.  (2002)  spent  far 
more  extended  times  at  depth  than  either  white  marlin 
tag.  Although  the  programming  of  the  tag  on  the  opah 
precludes  such  fine-scale  analyses  of  diving  behavior,  the 
available  data  are  not  inconsistent  with  the  mako  tracks 
in  the  study  of  Klimley  et  al.  (2002).  However,  the  short 
duration  dives  with  frequent  returns  to  the  surface  seen 
with  the  two  white  marlin  tags  most  closely  resemble 
those  of  blue  sharks  (Carey  and  Scharold,  1990)  and 
were  notably  missing  from  the  tracks  of  three  shortfin 
makos  observed  by  Klimley  et  al.  (2002). 

If  sharks  were  indeed  the  scavenging  animals,  it 
is  likely  that  the  tags  were  regurgitated,  rather  than 
egested  through  the  alimentary  canal,  whereupon  the 
PSAT  floated  to  the  surface  and  was  able  to  transmit 
the  archived  data.  The  narrow  diameter  of  the  spiral 
valve  in  the  elasmobranch  gastrointestinal  tract  would 
likely  be  too  narrow  to  allow  the  undamaged  passage 
of  an  object  the  size  of  a  PSAT,  even  for  a  large  shark. 
Although  the  available  literature  describing  regurgita- 
tion abilities  of  pelagic  sharks  is  rather  limited,  Hazin 
et  al.  (1994)  reported  that  35%  of  blue  sharks  brought 
aboard  for  scientific  study  had  everted  and  protruding 


stomachs.  Economakis  and  Lobel  (1998)  also  stated 
their  belief  that  regurgitation  of  ingested  ultrasonic 
tags  was  the  primary  cause  of  lost  tracks  for  grey  reef 
sharks  iCarcharhinus  amblyrhynchos)  on  Johnston  Atoll 
in  the  central  Pacific  Ocean. 


Conclusions 

The  temperatures  and  dive  depths  recorded  by  the  opah 
tag  and  both  white  marlin  tags  after  apparent  ingestion 
share  similarities,  yet  also  contain  sufficient  information 
to  indicate  the  different  identities  of  the  ingesting  organ- 
isms. The  dive  depths  in  all  cases  ranged  from  the  surface 
to  over  500  m,  whereas  the  temperatures  remained  rela- 
tively constant  at  several  degrees  above  the  background 
SST,  even  during  deep  dive  events.  Temperature  ranges 
alone  strongly  indicate  sharks  rather  than  odontocete 
whales  were  the  ingesting  organisms.  However,  limited 
literature  on  the  internal  stomach  temperatures  of  the 
various  pelagic  sharks  forces  us  to  rely  on  telemetered 
diving  behavior  data  for  further  species  identification, 
which  we  used  in  the  present  study  to  suggest  that  blue 
sharks  ingested  the  two  white  marlin  tags  (on  account 
of  the  broad  range  of  recorded  temperatures)  and  that 
an  endothermic  shark  ingested  the  opah  tag. 

It  is  not  possible  to  account  for  all  of  the  factors  that 
may  result  in  the  failure  of  satellite  tags  to  transmit 
data,  but  the  results  from  these  three  PSATs  indicated 
that  biological  activities  such  as  predation  and  scaveng- 
ing may  play  an  important  role.  We  believe  that  the 
most  consistent  explanation  for  the  data  transmitted 
by  these  three  tags  is  that  they  were  ingested  by  large 
sharks.  One  cannot  calculate  the  probability  that  a 
tag  could  be  engulfed  whole  without  physical  damage 
to  the  tag,  survive  for  several  days  in  the  caustic  en- 
vironment of  a  digestive  system,  and  be  regurgitated 
with  sufficient  battery  power  to  transmit  data  to  the 
Argos  satellites,  but  we  suspect  that  the  probability  is 
not  very  great.  We  expect  that  a  far  greater  number  of 
tags  may  have  had  similar  fates,  that  is  to  say,  they 
were  damaged  by  predation  or  scavenging  and  digestion 
processes  or  were  regurgitated  later  in  the  transmis- 
sion cycle,  when  the  PSAT  batteries  had  insufficient 
remaining  power  for  successful  data  transmission.  The 
failure  of  satellite  tag  to  transmit  data  is  frequently 
considered  to  be  the  result  of  internal  tag  malfunction 
or  user  error.  However,  these  three  data  sets  clearly 
indicate  that  the  failure  of  PSATs  to  function  may  also 
be  due  to  predation  or  scavenging  events. 


Acknowledgments 

The  authors  would  like  to  thank  the  Captain  of  the  FV 
Sea  Pearl  and  Captain  Greg  O'Neill  of  the  FV  Carol  Ann, 
Don  Hawn  (University  of  Hawaii),  who  deployed  the  tag 
on  the  opah,  Evan  Howell  (PIFSC)  for  analyses  of  the 
opah  data,  Andrij  Horodysky  (VIMS),  who  provided  a 
critical  review  of  the  manuscript.  Melinda  Braun  (Wild- 


756 


Fishery  Bulletin  102(4) 


life  Computers),  who  suggested  the  predation  hypothesis 
to  explain  the  opah  data,  and  Lissa  Werbos  (Microwave 
Telemetry.  Inc.),  who  independently  suggested  the  scav- 
enging hypothesis  for  the  WM1  data.  This  research  was 
supported  in  part  by  the  National  Marine  Fisheries 
Service,  the  NOAA  Ocean  Exploration  Program,  and 
the  University  of  Hawaii  Pelagic  Fisheries  Research 
Program  (PFRP). 


Literature  cited 

Block,  B.  A.,  H.  Dewar,  S.  B.  Blackwell.  T.  D.  Williams, 
E.  D.  Prince,  C.  J.  Farwell,  A.  Boustany,  S.  L.  H.  Teo, 

A.  Seitz,  A.  Walli,  and  D.  Fudge. 

2001.  Migratory  movements,  depth  preferences  and 
thermal  biology  of  Atlantic  bluefin  tuna.  Science 
293:1310-1314. 

Boustany,  A.  M.,  S.  F.  Davis,  P.  Pyle,  S.  D.  Anderson, 

B.  J.  LeBoeuf,  and  B.  A.  Block. 

2002.  Expanded  niche  for  white  sharks.  Nature  415: 
35-36. 

Carey,  F.  G.,  J.  W.  Kanwisher,  O.  Brazier,  G.  Gabrielson, 
J.  G.  Casey,  and  H.  L.  Pratt. 

1982.     Temperature  and  activities  of  a  white  shark,  Car- 
charodon  carcharias.     Copeia  1982(21:254-260. 
Carey,  F.  G.,  and  J.  V.  Scharold. 

1990.     Movements  of  blue  sharks  iPrionace  glauca)  in 
depth  and  course.     Mar.  Biol.  106:329-342. 
Carey,  F.  G.,  J.  M.  Teal,  and  J.  W.  Kanwisher. 

1981.     The  visceral  temperatures  of  mackerel  sharks 
(Lamnidae).     Physiol.  Zool.  54(3):334-344. 
Compagno,  L.  J.  V. 

1984.     Sharks  of  the  world:  an  annotated  and  illustrated 
catalogue  of  shark  species  known  to  date.     Part  1. 
Hexanchiformes  to  Lamniformes.     U.N.  FAO  species 
synopsis  125,  vol.  4,  pt.  I,  249  p.     FAO,  Rome. 
Cramer,  J. 

2000.     Species  reported  caught  in  the  U.S.  commercial 
pelagic  longline  and  gillnet  fisheries  from  1996  to 
1998.     NMFS  Sustainable  Fisheries  Division  SFD- 
99/00-78:1-33. 
Domeier.  M.  L,  and  H.  Dewar. 

2003.     Post-release  mortality  rate  of  striped  marlin  (  Tet- 
rapterus  audax)  caught  with  recreational  tackle.     Mar. 
Freshw.  Res.,  54(41:435-445. 
Dudley,  S.  F.  J.,  M.  D.  Anderson-Reade,  G.  S.  Thompson,  and 
P.  B.  McMullen. 

2000.     Concurrent  scavenging  off  a  whale  carcass  by  great 


white  sharks,  Carcharodon  carcharias,  and  tiger  sharks. 
Galeocerdo  cuvier.     Fish.  Bull.  98:646-649. 
Economakis.  A.  E.,  and  P.  S.  Lobel. 

1998.     Aggregation  behavior  of  the  grey  reef  shark,  Car- 
charhinus  amblyrhynchos,  at  Johnston  Atoll,  central 
Pacific  Ocean.     Environ.  Biol.  Fishes  51(21:129-139. 
Graves,  J.  E..  B.  E.  Luckhurst,  and  E.  D.  Prince. 

2002.     An  evaluation  of  pop-up  satellite  tags  for  estimating 
post-release  survival  of  blue  marlin  {Makaira  nigricans) 
from  a  recreational  fishery.     Fish.  Bull.  100:134-142. 
Hazin,  F.,  R.  Lessa,  and  M.  Chammas. 

1994.     First  observations  on  stomach  contents  of  the 
blue    shark,   Prionace   glauca,    from    southwestern 
equatorial  Atlantic.     Revista  Brasileira  de  Biologia 
54(21:195-198. 
Horodysky,  A.  Z„  D.  W.  Kerstetter,  and  J.  E.  Graves. 

In  press.  Habitat  preferences  and  diving  behavior  of 
white  marlin  (Tetrapturus  albidus)  released  from  the 
recreational  rod-and-reel  and  commercial  pelagic  longline 
fisheries  in  the  western  North  Atlantic:  implications  for 
habitat-based  stock  assessment  models.  International 
Commission  for  the  Conservation  of  Atlantic  Tunas, 
Coll.  Vol.  Sci.  Pap.,  SCRS  2003/033. 
Jolley,  J.  W.,  Jr.,  and  E.  W.  Irby  Jr. 

1979.     Survival  of  tagged  and  released  Atlantic  sailfish 
ilstwphorus  platypterus:  Istiophoridae)  determined  with 
acoustical  telemetry.     Bull.  Mar.  Sci.  29(21:155-169. 
Kerstetter,  D.  W.,  B.  E.  Luckhurst,  E.  D.  Prince,  and 
J.  E.  Graves. 

2003.     Use  of  pop-up  satellite  archival  tags  to  demonstrate 
survival  of  blue  marlin  (Makaira  nigricans)  released  from 
pelagic  longline  gear.     Fish.  Bull.  101(41:939-948. 
Klimley,  A.  P.,  S.  C.  Beavers,  T.  H.  Curtis,  and  S  .J.  Jorgensen. 
2002.     Movements  and  swimming  behavior  of  three  spe- 
cies of  sharks  in  La  Jolla  Canyon,  California.     Environ. 
Biol.  Fishes  63:117-135. 
McCosker.  J.  E. 

1987.     The  white  shark,  Carcharodon  carcharias,  has  a 
warm  stomach.     Copeia  1987(11:195-197. 
NOAA  (National  Oceanographic  and  Atmospheric 
Administration). 

1998.     Atlantic  coast  chart  13003:  Cape  Sable  to  Cape 
Hattaras.     National  Ocean  Service,  Washington.  D.C. 
Sedberry,  G  .R.,  and  J.  K.  Loefer. 

2001.     Satellite  telemetry  of  swordfish,  Xiphias  gla- 
dius,  off  of  the  eastern  United  States.     Mar.  Biol. 
139:355-360. 
Whittow.  G.  C,  I.  F.  G.  Hampton.  D.  T.  Matsuura,  C.  A.  Ohata. 
R.  M.  Smith,  and  J.  F.  Allen. 

1974.     Body  temperature  of  three  species  of  whales.     J. 
Mammol.  55(31:653-656 


757 


Survival  rates  for  rays  discarded  by  the 

bottom  trawl  squid  fishery  off  the  Falkland  Islands 


Vladimir  V.  Laptikhovsky 

Falkland  Islands  Government  Fisheries  Department 

P.O.  Box  598 

Stanley,  FIQQ  1ZZ 

Falkland  Islands 

E-mail  address:  vlaptikhovskyigfisheries.gov  fk 


Waters  off  the  Falkland  Islands  are 
subject  to  a  specialized  multispecies 
ray  fishery  and  were  first  fished  by 
a  Korean  fleet  in  1989.  More  than 
twenty  different  rajid  species  have 
been  recorded  from  catches  around 
the  islands,  and  five  species  accounted 
for  87.04%  of  the  total  catch  during 
1993-2002.  Catches  peaked  in  1993 
at  8523  metric  tons,  and  specific  fish- 
ing licenses — R  (second  season)  and  F 
(first  season) — were  first  introduced 
in  1994  and  in  1995,  respectively 
(Agnew  et  al.  2000;  Falkland  Islands 
Government,  2002;  Wakeford  et  al., 
in  press). 

In  addition  to  the  licensed  ray  fish- 
ery, rays  are  taken  as  bycatch  in  the 
bottom  trawl  fishery  that  targets  the 
squid  Loligo  gahi  and,  to  a  lesser  ex- 
tent, by  the  trawl  fishery  that  targets 
finfish.  A  10%  bycatch  of  nontarget 
species  is  allowed  in  both  these  fish- 
eries. In  2000-2002,  the  reported 
ray  bycatch  of  trawlers  not  licensed 
to  catch  rays  represented  between 
20.2%  and  31.9%  of  the  total  ray 
catch.  However,  under-reporting  of 
elasmobranch  bycatch  is  a  common 
practice  for  trawl  fisheries  where 
sharks  and  rays  are  discarded  (Ste- 
vens et  al.,  2000),  and  the  reported 
chondrichthyan  catch  is  only  about 
half  of  the  estimated  actual  global 
catch  (Bonfil,  1994).  The  actual  ray 
bycatch  in  Falkland  waters  may  be 
much  higher  than  reported  because 
only  large  rays  are  processed  (and 
therefore,  reported)  onboard  trawl- 
ers. This  situation  makes  ray  fishery 
management  in  the  Falkland  Islands, 
which  is  already  difficult  because  of 
the  nature  of  the  multispecies  tar- 
get, even  more  complicated.  However, 


good  management  is  of  primary  im- 
portance because  sharks  and  rays 
appear  to  be  particularly  vulnerable 
to  over-exploitation  because  of  their 
late  attainment  of  sexual  maturity, 
long  life  span,  both  low  fecundity  and 
natural  mortality,  and  close  relation- 
ship between  recruitment  and  paren- 
tal stock  (Stevens  et  al.,  20001.  In 
the  Falkland  trawl  fisheries  (which 
includes  most  trawlers  licensed  to 
catch  rays),  rays  smaller  than  ap- 
proximately 30  cm  disk  width  are 
discarded  after  spending  between  5 
min  and  4  hours  in  the  fish  bin  and 
passing  through  the  factory  sorting 
line  together  with  other  catch.  Some 
rays  that  have  been  caught,  stored, 
and  then  discarded  still  show  signs 
of  life.  In  contrast  to  other  marine 
organisms  whose  survival  after  be- 
ing discarded  has  been  investigated, 
ray  survival  has  been  studied  only 
in  Australian  waters  (Stobutzki  et 
al..  2002).  The  aim  of  this  study  was 
to  investigate  the  survival  rates  of 
discarded  rays  onboard  trawlers  in 
the  Falkland  waters. 


Materials  and  methods 

The  research  was  conducted  onboard 
the  Falkland  Islands  registered 
trawler  Sil  (length  of  78.5  m,  gross 
tons  (GRT)  of  2156  t,  net  tons  (NT) 
of  647  t).  The  vessel  used  a  bottom 
trawl  with  a  vertical  opening  of  5  m, 
horizontal  opening  of  30  m,  and  a 
codend  mesh  size  of  110  mm.  Trawl- 
ing speed  varied  between  3.8  and  and 
4.2  kn.  Fishing  occurred  at  a  depth 
of  80-190  m  during  the  day  and  the 
early  part  of  the  night.  The  surface 


temperature  was  8.7-9.2°C;  the  near 
bottom  temperature  was  6.8-7.6°C. 
Up  to  four  hauls  occurred  daily.  Each 
catch  was  released  from  the  codend 
into  the  fish  bin,  which  had  a  continu- 
ous supply  of  sea  water,  and  the  catch 
immediately  began  to  be  sorted  on  a 
conveyor  belt.  Squids  and  commercial 
fish  were  separated  from  the  noncom- 
mercial discarded  bycatch  and  were 
frozen.  Of  a  total  of  4306.2  kg  of  rays 
caught  during  the  observed  period, 
67.0%  were  discarded  and  only  the 
large  rays  were  processed.  The  time 
taken  to  sort  the  catch  was  between 
1  and  3  hours. 

A  total  of  66  rays  that  had  been 
discarded  by  fishermen  were  sampled 
randomly  from  the  conveyor  belt  and 
put  into  a  40-liter  (44x35x26  cm) 
or  a  60-liter  (31x76x26  cm)  fish  box 
that  contained  running  seawater. 
For  each  animal,  the  species  and  sex 
was  identified  and  total  length  (TL) 
and  disk  width  (DW)  were  measured 
within  1  cm.  Their  "stamina  index" 
was  assigned  according  to  four  major 
categories: 

A     alive,  flapping  wings. 

I  immobile,  but  alive,  reacting  to 
irritation,  spiracles  beginning 
to  work  actively  after  being 
placed  in  seawater. 

D  dead;  immobile,  but  spiracles  begin 
to  move  slowly  and  irregularly 
after  being  placed  in  seawater. 

DD  dead;  paralyzed,  body  stiffened 
and  wings  curved  but  may 
resume  breathing  after  being 
placed  in  seawater. 

Each  ray  (including  those  evident- 
ly dead)  was  kept  in  these  boxes  ei- 
ther until  its  death  was  evident  (no 
breathing)  or  it  fully  recovered  and 
began  to  try  to  swim  actively.  In 
some  rays  the  rate  of  spiracle  con- 
tractions was  episodically  recorded. 


Manuscript  submitted  5  June  2003 
to  the  Scientific  Editor's  Office. 

Manuscript  approved  for  publication 
30  June  2004  by  the  Scientific  Editor. 

Fish.  Bull.  102:757-759  (2004). 


758 


Fishery  Bulletin  102(4) 


Table  1 

Species  composition  and 

survival  of  sampled 

rays.  DW=  disk  width. 

Species 

n 

TL,  cm 

DW.  cm 

T 

ime  spent  in  fish  bin  (min.) 

Survival  rate  (%) 

Bathyraja  albomaculata 

14 

36-61 

26-44 

20-110  (mean  45) 

71.4 

B.  brachiurops 

11 

15-67 

9-49 

31-145  (mean  72) 

54.6 

B.  griseocauda 

3 

62-83 

47-60 

30-75  (mean  60) 

0.0 

B.  macloviana 

2 

36-42 

24-29 

70-135 

0.0 

B.  magellanica 

5 

30-44 

20-30 

50-125  (mean  90) 

60.0 

Bathyraja  sp. 

16 

24-104 

21-74 

5-120  (mean  52 ) 

75.0 

Psammobatis  sp. 

15 

29-47 

18-33 

30-200  (mean  98) 

60.0 

Table  2 

Ray  survival  (S%  ),  mean  recovery  time  (RT,  min.),  and  occurrence  of  the  four 
of  time  (T,  min.)  spent  in  the  fish  bin.  T=time  (minutes).  A=alive;  I=immobile; 

'stamina  index"  categories  after  different 
D=presumed  dead;  DD  =  dead. 

periods 

T                                     n                          S 

RT 

Occurrence  of  categories  c'<  l 

A 

I                           D 

DD 

5-30                             16                      87.5 

31-60                             20                      75.0 

65-120                            24                       41.7 

125-200                            6                      16.7 

38.2 

55.5 

102.2 

20' 

18.75 
10.0 

0 

0 

25                         18.75 
30.0                     40.0 
20.8                     50.0 
16.7                     83.3 

37.5 
20.0 
29.2 

0 

'  Only  one  individual  (Psammobathis  sp.  1. 

Results 

The  sampled  rays  belonged  to  eight  species  (Table  1).  Of 
the  66  sampled  rays,  a  total  of  21  were  dead  at  sampling, 
four  recovered  breathing  but  then  died,  and  39  survived. 
Two  rays  recorded  as  category  DD  in  the  "stamina 
index"  were  released  before  full  recovery  after  being  held 
for  4  to  9  hours  in  running  water.  Even  though  these 
individuals  were  still  breathing,  both  were  considered 
dead  because  they  still  had  stiffened  bodies  and  curved 
wings.  If  they  had  been  in  such  a  state  for  a  long  time 
in  their  natural  habitat,  they  almost  certainly  would 
have  been  consumed  by  scavengers  or  caught  again  by 
another  trawler.  The  overall  survival  rate  was  59.1%, 
female  survival  rate  was  66.7%,  and  male  survival  rate 
was  56.4%. 

All  five  rays  assigned  to  the  "stamina  index"  category 
A  were  sampled  between  5  and  30  min  (mean  20  min) 
after  the  catch  was  poured  into  the  fish  bin.  All  five 
individuals  began  immediately  to  breathe  normally  and 
recovered  within  5  to  20  minutes. 

Of  a  total  of  18  rays  assigned  to  the  "stamina  index" 
category  I,  which  were  sampled  between  15  and  145  min 
(mean  55.7  min)  after  haul,  88.9%  (n=16)  survived.  The 
breathing  of  these  specimens  at  the  time  of  sampling 
was  usually  slow,  although  occasionally  normal.  Spira- 
cle contraction  rates  gradually  increased  from  an  initial 


rate  of  5-15  bit/min  to  25-28  bit/min  for  B.brachiurops 
specimens  and  to  35-38  bit/min  for  individuals  of  B. 
albomaculata  and  Bathyraja  sp.  Upon  attaining  normal 
breathing,  they  remained  immobile,  but  fully  recovered 
between  15  minutes  and  3  hours. 

The  survival  rate  of  28  rays  that  were  assigned  to 
the  "stamina  index"  category  D  was  39.3%  («=11).  Of 
the  remaining  individuals,  two  rays  died  after  15  and 
45  minutes  after  being  placed  in  running  seawater  and 
15  rays  were  dead  at  the  time  of  sampling.  The  skates 
were  sampled  between  30  and  200  min  (mean  84.2  min) 
after  the  haul.  Those  that  survived  took  5-80  minutes 
to  recover  normal  breathing  and  between  15  and  315 
minutes  to  attain  full  recovery. 

A  total  of  15  rays  were  assigned  the  "stamina  index" 
category  DD.  However  seven  of  them  (46.7%)  survived. 
These  individuals  were  sampled  between  20  and  115 
minutes  (mean  63.9  min.)  after  the  haul  and  fully  re- 
covered within  40  to  150  minutes. 

Survival  rate  varied  substantially  among  the  eight 
species  sampled  (Table  1).  In  general,  ray  survival  dras- 
tically decreased  and  recovery  time  increased  with  the 
time  spent  in  the  fish  bin  (Table  2).  The  critical  dura- 
tion in  the  fish  bin  appeared  to  be  between  one  and  two 
hours;  only  one  Psammobathis  sp.  survived  more  than 
two  hours  in  the  fish  bin  and  exhibited  a  surprisingly 
fast  recovery. 


NOTE     Laptikhovsky:  Survival  rates  for  rays  by  the  bottom  trawl  squid  fishery 


759 


Discussion 


Acknowledgments 


The  survival  of  discarded  rays  during  trawling  opera- 
tions in  the  Falkland  waters  is  quite  important.  Although 
65.2%  of  the  individuals  were  initially  assigned  as  dead, 
the  actual  mortality  was  40. 99c,  although  it  took  some 
rays  up  to  six  hours  to  recover.  Survival  of  shallow-water 
shelf  species  such  as  Psammobatis  sp.,  in  particular,  but 
also  B.  brachiurops  and  B.  magellaniea,  was  somewhat 
higher  than  relatively  deep-water  species  such  as  B. 
albomaculata,  B.  griseocauda,  and  Bathyraja  sp.,  which 
inhabit  the  shelf  edge  and  upper  part  of  the  slope.  This 
survival  rate  was  most  likely  related  to  the  greater 
resilience  to  environmental  changes  for  shallow-water 
species,  whose  habitat  is  more  changeable  both  season- 
ally and  spatially.  Male  survival  was  lower,  which  is  in 
accordance  with  data  for  rays  and  skates  obtained  in 
northern  Australian  waters  (Stobutzky  et  al.,  2002). 

Recent  data  from  a  tropical  prawn  fishery  off  northern 
Australia  showed  that  on  average  449c  of  individuals  of 
a  number  of  ray  and  shark  species  survived  a  trawl- 
ing event  (Stobutzky  et  al.,  2002).  The  Falkland  ray 
survival  rate  was  higher.  This  difference  may  be  due 
either  to  the  higher  metabolic  rates  of  tropical  ray  spe- 
cies (and  therefore  a  higher  vulnerability  to  asphyxia), 
or  to  an  overestimation  of  their  mortality,  which  was 
assessed  immediately  after  individuals  where  landed  on 
deck  (unlike  the  recovery  time  allowed  in  the  present 
study).  The  latter  factor  is  more  probable  because  in  the 
present  study  41.9%  of  rays  initially  recorded  as  dead 
(D  and  DD)  eventually  recovered. 

Despite  the  demonstrated  ability  of  skates  to  survive 
after  being  caught  and  stored  in  fish  bins,  their  contin- 
ued survival  is  not  guaranteed  once  they  are  discarded. 
They  may  fall  prey  to  the  hundreds  of  albatrosses  and 
other  scavenging  birds  that  are  associated  with  trawl- 
ers (author's  pers.  obs.).  The  consumption  of  differ- 
ent discarded  fish  species  and  squids  from  trawlers  in 
Falkland  waters  by  seabirds,  primarily  by  black-browed 
albatrosses,  has  been  studied  (Thompson,  1992),  but  it 
is  not  known  whether  rays  are  also  taken  by  sea  birds 
and  to  what  extent.  Despite  the  great  abundance  of 
seabirds  around  vessels  in  the  Southwest  Atlantic,  it 
is  likely  that  they  consume  a  minor  part  of  discards  as 
found  in  Australia  (Hill  and  Wassenberg,  2000).  Most 
of  the  discarded  fish  probably  fall  to  the  sea  floor  and 
attract  and  are  consumed  by  bottom  scavengers  and 
bottom  dwellers  (Laptikhovsky  and  Fetisov,  1999;  Lap- 
tikhovsky and  Arkhipkin,  2003).  Consequently,  even 
after  recovering  and  successfully  avoiding  the  seabirds, 
the  discarded  skates  may  be  consumed  or  mortally  in- 
jured by  these  bottom  scavengers  during  the  recovery 
time,  which  appears  to  be  about  0.5-1.5  hours. 


I  would  like  to  thank  the  crew  of  FV  Sil  for  their  valu- 
able help  during  sampling  procedures  and  their  hospital- 
ity onboard;  the  Director  of  Fisheries,  John  Barton,  for 
supporting  this  work;  A.  I.  Arkhipkin  and  an  anonimous 
reviewer  for  valuable  comments;  and  Helen  Otley  (FIFD) 
for  language  editing. 


Literature  cited 

Agnew,  D.  J.,  C.  P.  Nolan,  J.  R.  Beddington.  and 
R.  Baranovski. 

2000.     Approaches  to  the  assessment  and  management  of 
multispecies  skate  and  ray  fisheries  using  the  Falkland 
Islands  fishery  as  an  example.     Can.  J.  Fish.  Aquat. 
Sci.  57:429-440. 
Bonn],  R. 

1994.     Overview  of  world  elasmobranch  fisheries.     FAO 
Fish.  Tech.  Pap.  341.  119  p. 
Falkland  Islands  Government. 

2002.  Fisheries  Department  fisheries  statistics,  7, 
1993-2002.  Falkland  Islands  Government  Fisheries 
Department,  Stanley,  Falkland  Islands. 

Hill,  B.  J.,  and  T.  J.  Wassenberg. 

2000.  The  probably  fate  of  discards  from  prawn  trawl- 
ers fishing  near  coral  reefs.  A  study  in  northern  Great 
Barrier  Reef,  Australia.     Fish.  Res.  48:277-286. 

Laptikhovsky,  V..  and  A.  Arkhipkin  . 

2003.  An  impact  of  seasonal  squid  migrations  and  fishery 
on  the  feeding  spectra  of  subantarctic  notothenioids 
Patagonotothen  ramsayi  and  Cottoperca  gobio  around 
the  Falkland  Islands.     J.  Appl.  Ichthyol.  19:35-39. 

Laptikhovsky,  V.,  and  A.  Fetisov 

1999.  Scavenging  by  fish  of  discards  from  the  Patagonian 
squid  fishery.     Fish.  Res.  41:93-97. 

Stevens,  J.  D.,  R.  Bonfil,  N.  K.  Dulvy,  and  P.  A.  Walker. 

2000.  The  effects  of  fishing  on  sharks,  rays  and  chimae- 
ras  (chondrichthyans).  and  the  implication  for  marine 
ecosystems.     ICES  J.  Mar.  Sci.  57:476-494. 

Stobutzki,  I.  C,  M.  J.  Miller,  D.  S.  Heales,  and  D.  T.  Brewer. 

2002.     Sustainability  of  elasmobranchs  caught  as  bycatch 
in  a  tropical  prawn  (shrimp)  trawl  fishery.     Fish.  Bull. 
100:800-821. 
Thompson  K.  R. 

1992.     Quantitative  analysis  of  the  use  of  discards  from 
squid  trawlers  by  black-browed  albatrosses  Diomedea 
melanophris  in  the  vicinity  of  the  Falkland  Islands.     Ibis 
134:11-21. 
Wakeford,  R  .C,  D.  A.  J.  Middleton,  D.  J.  Agnew, 
J.  H.  W.  Pompert,  and  V.  V.  Laptikhovsky. 

In  press.  Management  of  the  Falkland  Islands  mul- 
tispecies ray  fishery:  addressing  sustainability  and 
diversity.     Can.  J.  Fish.  Aquat.  Sci. 


760 


Acknowledgment  of  reviewers 

The  editorial  staff  of  Fishery  Bulletin  would  like  to  acknowledge  the  scientists 
who  reviewed  articles  published  in  2003-2004.  Their  contributions  have  helped 
ensure  the  publication  of  quality  science. 


Dr.  G.R.  Abbe 

Dr.  Pere  Abello 

Mr.  Douglas  H.  Adams 

Dr.  Vera  N.  Agostini 

Dr.  Juergen  Alheit 

Dr.  Robert  J.  Allman 

Mr.  Michael  D.  Arendt 

Dr.  Alexander  I.  Arkhipkin 

Dr.  David  A.  Armstrong 

Dr.  Colin  Attwood 

Ms.  Larisa  Avens 

Mr.  M.  Scott  Baker  Jr. 

Dr.  Donald  M.  Baltz 

Dr.  A.  Banner 

Dr.  Jay  Barlow 

Dr.  Steve  Berkeley 

Dr.  Eric  R  Bjorkstedt 

Dr.  James  A.  Bohnsack 

Ms.  Genevieve  Briand 

Dr.  Richard  W.  Brill 

Dr.  Alejandro  M.  Brockmann 

Dr.  Fiona  M.  Brook 

Dr.  Elizabeth  Brooks 

Dr.  Nancy  Brown-Peterson 

Dr.  Jay  Burnett 

Mr.  Michael  Burton 

Dr.  Morgan  S.  Busby 

Dr.  Michael  Canino 
Dr.  John  K.  Carlson 
Dr.  Milani  Y.  Chaloupka 
Dr.  David  M.  Checkley  Jr. 
Dr.  Susan  J.  Chivers 
Dr.  Phillip  J.  Clapham 
Dr.  William  Coles 
Mr.  L.  Alan  Collins 

Dr.  Craig  Dahlgren 

Dr.  Marilyn  E.  Dahlheim 

Dr.  Louis  B.  Daniel  III 

Dr.  Jana  L.D.  Davis 

Dr.  Earl  G.  Dawe 

Dr.  Edward  E.  DeMartini 

Dr.  Heidi  Dewar 

Dr.  Robert  W.  Elner 
Dr.  Tomo  Eguchi 
Dr.  Charles  E.  Epifanio 
Dr.  Sheryan  P.  Epperly 


Dr.  S.  Frantini 

Dr.  Gregory  L.  Fulling 

Ms.  Moira  Galbraith 

Dr.  Francisco  J.  Garcia-Rodriguez 

Mr.  Bert  Geary 

Dr.  Harry  J.  Grier 

Dr.  Churchill  B.  Grimes 

Dr.  Donald  R.  Gunderson 

Dr.  Chris  Habicht 
Dr.  Lewis  J.  Haldorson 
Dr.  J.  Mark  Hanson 
Dr.  E.  Brian  Hartwick 
Dr.  James  T.  Harvey 
Dr.  Jonathan  Heifetz 
Dr.  Kevin  T.  Hill 
Dr.  Simeon  Hill 
Dr.  David  B.  Holts 

Dr.  J.  Jeffrey  Isely 

Dr.  George  D.  Jackson 
Ms.  Nadine  Johnston 
Dr.  Lindsay  Joll 

Dr.  Michel  J.  Kaiser 
Mr.  Craig  R.  Kastelle 
Dr.  Izhar  A.  Khan 
Dr.  J.  King 
Dr.  A.  Peter  Klimley 
Dr.  Suzanne  Kohin 

Dr.  Thomas  E.  Laidig 
Dr.  Richard  W.  Langton 
Ms.  Amy  Lapolla 
Dr.  Robert  N.  Lea 
Dr.  Christopher  M.  Legault 
Dr.  Steven  T.  Lindley 
Dr.  Romuald  Lipcius 
Dr.  Kwang  Ming  Liu 
Dr.  Kai  Lorenzen 
Dr.  Milton  S.  Love 
Mr.  Mark  S.  Lowry 
Dr.  Mark  Luckenbach 

Dr.  R.  Bruce  MacFarlane 
Dr.  William  K.  Macy 
Dr.  Richard  McBride 
Dr.  Susanne  McDermott 
Ms.  Kitty  Mecklenburg 
Dr.  David  A.  Milton 


Dr.  T.J.  Minello 
Mr.  Karl  W.  Mueller 
Dr.  Ashley  Mullen 
Dr.  Keith  D.  Mullin 
Dr.  Michael  D.  Murphy 
Dr.  Kate  Myers 

Dr.  Wallace  J.  Nichols 

Dr.  Peter  F.  Olesiuk 

Dr.  Ernst  Peebles 
Dr.  Karl  M.  Polivka 
Dr.  Kenneth  H.  Pollock 
Dr.  Allyn  B.  Powell 

Dr.  Hans-Joachim  Raetz 
Dr.  Stephen  Ralston 
Mrs.  Tone  Rasmussen 
Dr.  Sherrylynn  Rowe 
Dr.  Peter  Rubec 
Mr.  D.E.  Ruzzante 

Dr.  Yvonne  Sadovy 

Dr.  Bernard  Sainte-Marie 

Dr.  Kurt  M.  Schaefer 

Dr.  George  R.  Sedberry 

Dr.  Jeffrey  Seminoff 

Mr.  Lawrence  Settle 

Dr.  James  B.  Shaklee 

Dr.  Alan  Sinclair 

Dr.  Oscar  Sosa-Nishizaki 

Dr.  Gretchen  Steiger 

Dr.  David  L.  Stein 

Dr.  Allan  W.  Stoner 

Dr.  D.P  Swain 

Dr.  Yonat  Swimmer 

Dr.  Yuji  Tanaka 
Dr.  Sven  Thatje 
Dr.  A.M.  Tokranov 
Dr.  M.J.  Tremblay 
Dr.  Marc  Trudel 

Dr.  Fred  M.  Utter 

Dr.  Peter  Van  Tamelen 

Dr.  Michael  Vecchione 

Dr.  Claire  M.  Waluda 

Mr.  William  Watson 

Dr.  George  Watters 

Dr.  Elizabeth  L.  Wenner 

Mr.  A.J.  Winship 

Dr.  Sabine  Petra  Wintner 

Dr.  Bernd  Wursig 

Dr.  Orio  Yamamura 
Dr.  Richard  E.  Young 


761 


Fishery  Bulletin  Index 

Volume  102(1-4),  2004 
List  ot  titles 


102(1) 


142  Growth,  mortality,  and  hatchdate  distributions  of 
larval  and  juvenile  spotted  seatrout  (Cynoscion 
jiebulosus)  in  Florida  Bay,  Everglades  National 
Park,  by  Allyn  B.  Powell,  Robin  T.  Chesire,  Elisabeth 
H.  Laban,  James  Colvocoresses,  Patrick  O'Donnell, 
and  Marie  Davidian 


1  The  effects  of  size-selective  fisheries  on  the  stock 
dynamics  of  and  sperm  limitation  in  sex-changing 
fish,  by  Suzanne  H.  Alonzo  and  Marc  Mangel 

14  An  environmentally  based  growth  model  that  uses 
finite  difference  calculus  with  maximum  likeli- 
hood method:  its  application  to  the  brackish  water 
bivalve  Corbicula  japonica  in  Lake  Abashiri,  Japan, 
by  Katsuhisa  Baba,  Toshifumi  Kawajiri,  Yasuhuro 
Kuwahara,  and  Shigeru  Nakao 

25  Juvenile  salmonid  distribution,  growth,  condition, 
origin,  and  environmental  and  species  associations 
in  the  Northern  California  Current,  by  Rick  D. 
Brodeur,  Joseph  P.  Fisher,  David  J.  Teel,  Robert  L. 
Emmett,  Edmundo  Casillas,  and  Todd  W.  Miller 

47  Spatial  and  temporal  variation  in  the  diet  of  the 
California  sea  lion  (Zalophus  californianus)  in  the 
Gulf  of  California,  Mexico,  by  Francisco  J.  Garcia- 
Rodriguez  and  David  Auarioles-Gamboa. 

63  Recruitment  and  spawning-stock  biomass  distribu- 
tion of  bay  anchovy  {Anchoa  mitchilli )  in  Chesapeake 
Bay.  by  Sukgeun  Jung  and  Edward  D.  Houde 

78  Coupling  ecology  and  economy:  modeling  optimal 
release  scenarios  for  summer  flounder  (Paralichthys 
dentatus)  stock  enhancement,  by  Todd  G.  Kellison 
and  David  B.  Eggleston 

94  Sex-specific  growth  and  mortality,  spawning  season, 
and  female  maturation  of  the  stripey  bass  (Lutjanus 
carponotatus)  on  the  Great  Barrrier  Reef,  by  Jacob 
T.  Krtizer 


156  Age  determination  and  growth  of  the  night  shark 
(Carcharhinus  signatus)  off  the  northeastern  Brazil- 
ian coast,  by  Francisco  M.  Santana  and  Rosangela 
Lessa 

168  Distribution  and  biology  of  prowfish  (Zaprora  sile- 
nus)  in  the  northeast  Pacific,  by  Keith  R.  Smith, 
David  A.  Somerton,  Mei-Sun  Yang,  and  Daniel  G. 
Nichol 

179  Fish  lost  at  sea:  the  effect  of  soak  time  on  pelagic 
longline  catches,  by  Peter  Ward,  Ransom  A.  Myers, 
and  Wade  Blanchard 

196  Effects  of  density-dependence  and  sea  surface  tem- 
perature on  interannual  variation  in  length-at-age  of 
chub  mackerel  (Scomber  japonicus )  in  the  Kuroshio- 
Oyashio  area  during  1970-1997,  by  Chikako  Wata- 
nabe  and  Akihiko  Yatsu 

207  Latitudinal  and  seasonal  egg-size  variation  of 
the  anchoveta  (Engraulis  ringens)  off  the  Chilean 
coast,  by  Llanos-Rivera,  Alejandra,  and  Leonard  R. 
Castro 

213  Molecular  methods  for  the  genetic  identification 
of  salmonid  prey  from  Pacific  harbor  seal  iPhoca 
vitulina  richardsi)  scat,  by  Maureen  Purcell,  Greg 
Mackey.  Eric  LaHood,  Harriet  Huber,  and  Linda 
Park 

221  Diel  vertical  migration  of  the  bigeye  thresher  shark 
(Alopias  superciliosus),  a  species  possessing  orbital 
retia  mirabilia,  by  Kevin  C.  Weng  and  Barbara  A. 
Block 


108  Examination  of  the  foraging  habits  of  Pacific  harbor 
seal  (Phoca  vitulina  richardsi)  to  describe  their  use 
of  the  Umpqua  River,  Oregon,  and  their  predation  on 
salmonids,  by  Anthony  J.  Orr,  Adria  S.  Banks,  Steve 
Mellman,  Harriet  R.  Huber,  Robert  L.  DeLong,  and 
Robin  F.  Brown 

118  Larval  development  of  the  sidestriped  shrimp  (Pan- 
dalopsis  dispar  Rathbun)  (Crustacea,  Decapoda, 
Pandalidae)  reared  in  the  laboratory,  by  Wongyu 
Park,  R.  Ian  Perry,  and  Sung  Yun  Hong 

127  Sources  of  age  determination  errors  for  sablefish 
(Anoplopoma  fimbria)  by  Donald  E.  Pearson  and 
Franklin  R.  Shaw 


102(2) 

233  Annual  estimates  of  the  unobserved  incidental  kill 
of  pantropical  spotted  dolphin  (Stenella  attenuata 
attenuata)  calves  in  the  tuna  purse-seine  fishery  of 
the  eastern  tropical  Pacific,  by  Frederick  Archer,  Tim 
Gerrodette,  Susan  Chivers,  and  Alan  Jackson 

245  A  remarkable  new  species  of  Psednos  (Teleostei: 
Liparidae )  from  the  western  North  Atlantic  Ocean, 
by  Natalia  V.  Chernova  and  David  L.  Stein 

251  Age  and  growth  of  sailfish  (Istiophorus  platypterus) 
in  waters  off  eastern  Taiwan,  by  Wei-Chuan  Chiang, 
Chi-Lu  Sun,  Su-Zan  Yeh,  and  Wei-Cheng  Su 


762 


Fishery  Bulletin  102(4) 


264  A  habitat-use  model  to  determine  essential  fish 
habitat  for  juvenile  brown  shrimp  {Farfantepenaeus 
aztecus)  in  Galveston  Bay,  Texas,  by  Randall  D. 
Clark,  John  D.  Christensen,  Mark  E.  Monaco,  Philip 
A.  Caldwell,  Geoffrey  A.  Matthews,  and  Thomas  J. 
Minello 


102(3) 

407  Testicular  development  in  migrant  and  spawning 
bluefin  tuna  (Thunnus  thynnus  (L.>)  from  the  east- 
ern Atlantic  and  Mediterranean,  by  Francisco  J. 
Abascal.  Cesar  Megina,  and  Antonio  Medina 


278  Translocation  as  a  strategy  to  rehabilitate  the  queen 
conch  {Strombus  gigas)  population  in  the  Florida 
Keys,  by  Gabriel  A.  Delgado,  Claudine  T  Bartels, 
Robert  A.  Glazer,  Nancy  J.  Brown-Peterson,  and 
Kevin  J.  McCarthy 

289  Genetic  differentiation  among  Atlantic  cod  (Gadus 
morhua)  from  Browns  Bank,  Georges  Bank,  and 
Nantucket  Shoals,  by  Christopher  Lage,  Kristen 
Kuhn,  and  Irv  Kornfield 

298  Conserving  oyster  reef  habitat  by  switching  from 
dredging  and  tonging  to  diver  harvesting,  by  Hunter 
S.  Lenihan  and  Charles  H.  Peterson 

306  Fecundity,  egg  deposition,  and  mortality  of  market 
squid  iLolilgo  opalescens),  by  Beverly  J.  Macewicz, 
John  R.  Hunter,  Nancy  C.  H.  Lo,  and  Erin  L.  LaCasella 


418  Maturity,  ovarian  cycle,  fecundity,  and  age-specific 
parturition  of  black  rockfish  (Sebastes  melanops),  by 
Stephen  J.  Bobko  and  Steven  A.  Berkeley 

430  Maori  octopus  ( Octopus  maorum  )  bycatch  and  south- 
ern rock  lobster  (Jasus  edwardsii)  mortality  in  the 
South  Australian  lobster  fishery,  by  Daniel  J.  Brock 
and  Timothy  M.  Ward 

441  Small-boat  surveys  for  coastal  dolphins:  line-tran- 
sect surveys  of  Hector's  dolphins  (Cephalorhymhus 
hectori),  by  Stephen  Dawson,  Elisabeth  Slooten, 
Sam  DuFresne,  Paul  Wade,  and  Deanna  Clement 

452  Description  and  growth  of  larval  and  pelagic  juvenile 
pygmy  rockfish  (Sebastes  wilsoni)  (family  Sebasti- 
dae),  by  Thomas  E.  Laidig,  Keith  M.  Sakuma,  and 
Jason  A.  Stannard 


328  The  dusky  rockfishes  (Teleostei:  Scorpaeniformes I 
of  the  North  Pacific  Ocean:  resurrection  of  Sebastes 
variabilis  (Pallas,  1814)  and  a  redescription  of 
Sebastes  ciliatus  iTilesius,  1813),  by  James  Wilder 
Orr  and  James  E.  Blackburn 

349  Recruitment  as  a  evolving  random  process  of  aggre- 
gation and  mortality,  by  Joseph  E.  Powers 

366  Diet  shifts  of  juvenile  red  snapper  (Lutjanus 
campechanus)  with  changes  in  habitat  and  fish  size, 
by  Stephen  T  Szedlmayer  and  Jason  D.  Lee 


464  Estimating  the  emigration  rate  of  fish  stocks  from 
marine  sanctuaries  using  tag-recovery  data,  by 
Richard  McGarvey 

473  Reproductive  dynamics  of  female  spotted  seatrout 
(Cynoscion  nebulosus)  in  South  Carolina,  by  William 
A.  Roumillat  and  Myra  C.  Brouwer 

488  Estimating  Dungeness  crab  (Cancer  magistvr^ 
abundance:  crab  pots  and  dive  transects  compared, 
by  S.  James  Taggart,  Charles  E.  O'Clair,  Thomas  C. 
Shirley,  and  Jennifer  Mondragon 


376  Individual  growth  rates  and  movement  of  juvenile 
white  shrimp  (Litopenaeus  setiferus)  in  a  tidal  marsh 
nursery,  by  Stacey  Webb  and  Ronald  T.  Kneib 

389  Does  the  California  market  squid  (Loligo  opalescens) 
spawn  naturally  during  the  day  or  at  night?  A  note 
on  the  successful  use  of  ROVs  to  obtain  basic  fisher- 
ies biology  data,  by  John  Forsythe,  Nuutti  Kangas, 
and  Roger  T.  Hanlon 

393  Incidental  capture  of  loggerhead  (Caretta  caretta) 
and  leatherback  (Dermochelys  coriacea)  sea  turtles 
by  the  pelagic  longline  fishery  off  southern  Brazil, 
by  Jorge  E.  Kotas,  Silvio  dos  Santos,  Venancio  G. 
de  Azevedo,  Berenice  M.  G.  Gallo.  and  Paulo  C.  R. 
Barata 

400  Diet  changes  of  Pacific  cod  {Gadus  macrocephalus) 
in  Pavlof  Bay  associated  with  climate  changes  in  the 
Gulf  of  Alaska  between  1980  and  1995,  by  Mei-Sun 
Yang 


498  A  method  to  improve  size  estimates  of  walleye  pol- 
lock (Theragra  chalcogramma)  and  Atka  mackerel 
(Pleurogrammus  monopterygius)  consumed  by  pin- 
nipeds: digestion  correction  factors  applied  to  bones 
and  otoliths  recovered  in  scats,  by  Dominic  J.  Tollit. 
Susan  G.  Heaslip,  Tonya  K.  Zeppelin,  Ruth  Joy, 
Katherine  A.  Call,  and  Andrew  W.  Trites 

509  Sizes  of  walleye  pollock  (Theragra  chalcogramma) 
and  Atka  mackerel  (Pleurogrammus  monopter- 
ygius) consumed  by  the  western  stock  of  Steller  sea 
lions  (Eumetopias  jubatus)  in  Alaska  from  1999 
to  2000,  by  Tonya  K  Zeppelin,  Dominic  J.  Tollit, 
Katherine  A  Call,  Trevor  J.  Orchard,  and  Carolyn 
J.  Gudmundson 

522  Sizes  of  walleye  pollock  (Theragra  chalcogramma) 
consumed  by  the  eastern  stock  of  Steller  sea  lions 
{Eumetopias  jubatus)  in  Southeast  Alaska  from  1994 
to  1999,  by  Dominic  J.  Tollit,  Susan  G.  Heaslip,  and 
Andrew  Trites 


List  of  titles 


763 


533  Multidirectional  movements  of  sportfish  species 
between  an  estuarine  no-take  zone  and  surrounding 
waters  of  the  Indian  River  Lagoon,  Florida,  by  Derek 
M.  Tremain,  Christopher  W.  Harnden,  and  Douglas 
H.  Adams 

545  Distribution,  age,  and  growth  of  young-of-the  year 
greater  amberj ack  (Seriola  dumerili )  associated  with 
pelagic  Sargassum,  by  R.  J.  David  Wells  and  Jay  R. 
Rooker 

555  Identification  of  formalin-preserved  eggs  of  red 
sea  bream  iPagrus  major)  (Pisces:  Sparidae)  using 
monoclonal  antibodies,  by  Shingo  Hiroishi,  Yas- 
utaka  Yuki,  Eriko  Yuruzume,  Yosuke  Onishi,  Tomoji 
Ikeda,  Hironobu  Komaki,  and  Muneo  Okiyama 


102(4) 

563  Distribution  and  abundance  of  humpback  whales 
(Megaptera  novaeangliae)  and  other  marine  mam- 
mals off  the  northern  Washington  coast,  by  John 
Calambokidis,  Gretchen  H.  Steiger,  David  K.  Ellifrit, 
Barry  L.  Troutman,  and  C.  Edward  Bowlby 

581  Reproductive  biology  of  male  franciscanas  (Ponto- 
poria  blainvillei)  (Mammalia:  Cetacea)  from  Rio 
Grande  do  Sul,  southern  Brazil,  by  Daniel  Danile- 
wicz,  Juan  A.  Claver,  Alejo  L.  Perez  Carrera,  Edu- 
ardo  R.  Secchi,  and  Nelson  F.  Fontoura 

593  Red  snapper  (Lutjanus  campechanus)  demographic 
structure  in  the  northern  Gulf  of  Mexico  based  on 
spatial  patterns  in  growth  rates  and  morphomet- 
ries, by  Andrew  J.  Fischer,  M.  Scott  Baker  Jr.,  and 
Charles  A.  Wilson 

604  Elemental  signatures  in  otoliths  of  larval  walleye 
pollock  (Theragra  chalcogramma)  from  the  north- 
east Pacific  Ocean,  by  Jennifer  L.  FitzGerald.  Simon 
R.  Thorrold.  Kevin  M.  Bailey,  Annette  L.  Brown,  and 
Kenneth  P.  Severin 

617  A  sudden  collapse  in  distribution  of  Pacific  sar- 
dine (Sardinops  saga.x}  off  southwestern  Australia 
enables  an  objective  re-assessment  of  biomass 
estimates,  by  Daniel  J.  Gaughan,  Timothy  I.  Leary, 
Ronald  W.  Mitchell,  and  Ian  W.  Wright 

634  Fish  recolonization  in  temperate  Australian  rock- 
pools:  a  quantitative  experimental  approach,  by 
Shane  P.  Griffiths,  Ron  J.  West,  Andy  R.  Davis,  and 
Ken  G.  Russell 


661  Fishery  dynamics  of  the  California  market  squid 
(Loligo  opalescens),  as  measured  by  satellite  remote 
sensing,  by  Michael  R.  Maxwell,  Annette  Henry, 
Christopher  D.  Elvidge.  Jeffrey  Safran,  Vinita  R. 
Hobson.  Ingrid  Nelson,  Benjamin  T.  Tuttle,  John  B. 
Dietz.  and  John  R.  Hunter 

671  Magnitude  and  distribution  of  sea  turtle  bycatch  in 
the  sea  scallop  (Placopecten  magellanicus)  dredge 
fishery  in  two  areas  of  the  northwestern  Atlantic 
Ocean,  2001-2002,  by  Kimberly  T.  Murray 

682  Validation  and  interpretation  of  annual  skeletal 
marks  in  loggerhead  (Caretta  caretta)  and  Kemps 
ridley  (Lepidochelys  kempii)  sea  turtles,  by  Melissa 
L.  Snover  and  Aleta  A.  Hohn. 

693  The  Hudson-Raritan  Estuary  as  a  crossroads  for  dis- 
tribution of  blue  iCallinectes  sapidus),  lady  (Ovali- 
pes  ocellatus),  and  Atlantic  rock  (Cancer  irroratus) 
crabs,  by  Linda  L.  Stehlik.  Robert  A.  Pikanowski, 
and  Donald  G.  McMillan 

711  Radiometric  validation  of  age,  growth,  and  longevity 
for  the  blackgill  rockfish  (Sebastes  melanostomus), 
by  Melissa  M.  Stevens,  Allen  H.  Andrews,  Gregor 
M.  Cailliet.  Kenneth  H.  Coale.  and  Craig  C. 
Lundstrom 

723  Descriptions  of  larval,  prejuvenile,  and  juvenile 
finescale  menhaden  (Brevoortia  gunteri)  (family 
Clupeidae),  and  comparisons  to  gulf  menhaden 
(B.  patronus),  by  James  M.  Tolan  and  David  A. 
Newstead 

740  Capture  probability  of  a  survey  trawl  for  red  king 
crab  (Paralithodes  camtschaticus),  by  Kenneth  L. 
Weinberg,  Robert  S.  Otto,  and  David  A.  Somerton 

733     Diet  of  the  minimal  armhook  squid  {Berryteuthis 

anonychus)  (Cephalopoda:  Gonatidae)  in  the  north- 
east Pacific  during  spring,  by  Kazuhisa  Uchikawa, 
John  R.  Bower,  Yasuko  Sato,  and  Yasunori  Sakurai 

750  Evidence  of  shark  predation  and  scavenging  of  fishes 
equipped  with  pop-up  satellite  archival  tags,  by 
David  W.  Kerstetter,  Jeffery  J.  Polovina,  and  John 
E.  Graves 

757  Survival  rates  of  rays  discarded  by  the  bottom  trawl 
squid  fishery  off  the  Falkland  slands,  by  Vladimir  V. 
Laptikhovsky 


648  Factors  influencing  the  timing  and  frequency  of 
spawning  and  fecundity  of  the  goldlined  seabream 
iRhabdosargus  sarba)  (Sparidae)  in  the  lower 
reaches  of  an  estuary,  by  S.  Alexander  Hesp,  Ian  C. 
Potter,  and  Sonja  R.  M.  Schubert 


764 


Fishery  Bulletin  102(4) 


Fishery  Bulletin  Index 

Volume  102(1-4),  2004 
List  ot  authors 


Abascal,  Francisco  J.    407 
Adams,  Douglas  H.    533 
Alonzo,  Suzanne  H.     1 
Andrews,  Allen  H.    711 
Archer,  Frederick    233 
Auarioles-Gamboa,  David    47 
Azevedo,  Venancio  G.  de    393 

Baba,  Katsuhisa    14 
Bailey,  Kevin  M.    604 
Baker  Jr.,  M.  Scott    593 
Banks,  Adria  S.    108 
Barata,  Paulo  C.  R.    393 
Bartels,  Claudine  T.    278 
Berkeley,  Steven  A.    418 
Blackburn,  James  E.    328 
Blanchard,  Wade    179 
Block,  Barbara  A.    221 
Bobko,  Stephen  J.    418 
Bower,  John  R.    733 
Bowlby,  C.  Edward    563 
Brock,  Daniel  J.    430 
Brodeur,  Rick  D.    25 
Brouwer,  Myra  C.    473 
Brown,  Annette  L.    604 
Brown,  Robin  F.    108 
Brown-Peterson,  Nancy  J.    278 

Cailliet,  Gregor  M.    711 
Caldwell,  Philip  A.    264 
Call,  Katherine  A.    498,  509 
Calambokidis,  John    563 
Casillas,  Edmundo    25 
Castro,  Leonard  R.    207 
Chernova,  Natalia  V.    245 
Cheshire,  Robin  T.     142 
Chiang,  Wei-Chuan    251 
Chivers,  Susan    233 
Christensen,  John  D.    264 
Clark,  Randall  D.    264 
Claver,  Juan  A.    581 
Clement,  Deanna    441 
Coale,  Kenneth  H.    711 
Colvocoresses,  James    142 

Danilewicz,  Daniel    581 
Davidian,  Marie    142 
Davis,  Andy  R.    634 
Dawson,  Stephen    441 
Delgado,  Gabriel  A.    278 
DeLong,  Robert  L.    108 


Dietz,  JohnB.    661 
DuFresne,  Sam    441 

Eggleston,  David  B.    78 
Ellifnt,  David  K    563 
Elvidge,  Christopher  D.    661 
Emmett,  Robert  L.    25 

Fischer,  Andrew  J.    593 
Fisher,  Joseph  P.    25 
FitzGerald,  Jennifer  L.    604 
Fontoura,  Nelson  F    581 
Forsythe,  John    389 

Gallo,  Berenice  M.  G.    393 
Garcia-Rodriguez,  Francisco  J.    47 
Gaughan,  Daniel  J.    617 
Gerrodette,  Tim    233 
Glazer,  Robert  A.    278 
Graves,  John  E.    750 
Griffiths,  Shane  P.    634 
Gudmundson,  Carolyn  J.    509 

Hanlon,  Roger  T,    389 
Harnden,  Christopher  W.    533 
Heaslip,  Susan  G.    498,  522 
Henry,  Annette    661 
Hesp,  S.  Alexander    648 
Hiroishi,  Shingo    555 
Hobson,  Vinita  R.    661 
Hohn,AletaA.    682 
Hong,  Sung  Yun    118 
Houde,  Edward  D.    63 
Huber,  Harriet  R.    108,213 
Hunter,  John  R.    306,  661 

Ikeda,  Tomoji    555 

Jackson,  Alan    233 
Joy,  Ruth    498 
Jung,  Sukgeun    63 

Kangas,  Nuutti    389 
Kawajiri,  Toshifumi    14 
Kellison,  Todd  G.    78 
Kerstetter,  David  W.    750 
Kneib,  Ronald  T    376 
Komaki,  Hironobu    555 
Kornfield,  Irv    289 
Kotas,  Jorge  E.    393 
Kritzer,  Jacob  P.    94 


Kuhn,  Kristen    289 
Kuwahara,  Yasuhiro    14 

Laban,  Elisabeth  H.    142 
LaCasella,  Erin  L.    306 
Lage,  Christopher    289 
LaHood,  Eric    213 
Laidig,  Thomas  E.    452 
Laptikhovsky,  Vladimir  V.    757 
Leary,  Timothy  I.    617 
Lee,  Jason  D.    366 
Lenihan,  Hunter  S.    298 
Lessa,  Rosangela    156 
Llanos-Rivera,  Alejandra    207 
Lo,  Nancy  C.  H.    306 
Lundstrom,  Craig  C.    711 

Macewicz,  Beverly  J.    306 
Mackey,  Greg    213 
Mangel,  Marc    1 
Matthews,  Geoffrey  A.    264 
Maxwell,  Michael  R.    661 
McCarthy,  Kevin  J.    278 
McGarvey,  Richard    464 
McMillan,  Donald  G.    693 
Medina,  Antonio    407 
Megina,  Cesar    407 
Mellman,  Steve    108 
Miller,  Todd  W.    25 
Minello,  Thomas  J.    264 
Mitchell,  Ronald  W.    617 
Monaco,  Mark  E.    264 
Mondragon,  Jennifer    488 
Murray,  Kimberly  T    671 
Myers,  Ransom  A.    179 

Nakao,  Shigeru    14 
Nelson,  Ingrid    661 
Newstead,  David  A.    723 
Nichol,  Daniel  G.    168 

O'Clair,  Charles  E.    488 
O'Donnell,  Patrick    142 
Okiyama,  Muneo    555 
Onishi,  Yosuke    555 
Orchard,  Trevor  J.    509 
Orr,  Anthony  J.     108 
Orr,  James  W.    328 
Otto,  Robert  S.    740 

Park,  Linda    213 
Park,  Wongyu  R.     118 
Pearson,  Donald  E.    127 
Perez  Carrera,  Alejo  L.    581 
Perry,  R.  Ian    118 
Peterson,  Charles  H.    298 
Pikanowski,  Robert  A.    693 
Polovina,  Jeffery  J.     750 


List  of  authors 


765 


Potter,  Ian  C.    648 
Powell,  Allyn  B.    142 
Powers,  Joseph  E.    349 
Purcell,  Maureen    213 

Rooker,  Jay  R.    545 
Roumillat,  William  A.    473 
Russell,  Ken  G.    634 

Safran,  Jeffrey    661 
Sakuma,  Keith  M.    452 
Sakurai,  Yasunori    733 
Santana,  Francisco  M.    156 
Santos,  Silvio  dos    393 
Sato,Yasuko    733 
Secchi,  Eduardo  R.    581 
Severin,  Kenneth  P.    604 
Shaw,  Franklin  R.     127 
Shirley,  Thomas  C.    488 
Schubert,  Sonja  R.  M.    648 
Slooten,  Elisabeth    441 


Smith,  Keith  R.    168 
Snover,  Melissa  L.    682 
Somerton,  David  A.    168,  740 
Stannard,  Jason  A.    452 
Stehlik,  Linda  L.    693 
Steiger,  Gretchen  H.    563 
Stein,  David  L.    245 
Stevens,  Melissa  M.    711 
Su,  Wei-Cheng    251 
Sun,  Chi-Lu    251 
Szedlmayer,  Stephen  T.    366 

Taggart,  S.  James    488 
Teel,  David  J.    25 
Tolan,  James  M.    723 
Tollit,  Dominic  J.    498,  509,  522 
Thorrold,  Simon  R.    604 
Tremain,  Derek  M.    533 
Trites,  Andrew  W    498,  522 
Troutman,  Barry  L.    563 
Tuttle,  Benjamin  T.    661 


Uchikawa,  Kazuhisa    733 

Wade,  Paul    441 
Ward,  Peter    179 
Ward.  Timothy  M.    430 
Watanabe,  Chikako    196 
Webb,  Stacey    376 
Wells,  R.  J.  David    545 
Weng,  Kevin  C.    221 
Weinberg,  Kenneth  L.    740 
West,  Ron  J.    634 
Wilson,  Charles  A.    593 
Wright,  Ian  W.    617 

Yang,  Mei-Sun    168,  400 
Yatsu,  Akihiko    196 
Yeh,  Su-Zan    251 
Yuki,  Yasutaka    555 
Yuruzume,  Eriko    555 

Zeppelin,  Tonya  K.    498,  509 


766 


Fishery  Bulletin  102(4) 


Fishery  Bulletin  Index 

Volume  102(1-4),  2004 
List  ot  subjects 


Abundance 

Crab,  Dungeness  488 

dolphin.  Hector's  441 

estimates  (pelagic  longline)  179 

harbor  seal.  Pacific   108 

whales  563 
Age 

and  growth 

amberjack,  greater  533 
rockfish,  blackgill  711 
sailfish  251 
shark,  night   156 

determination 

problems  with   127 
sablefish  127 

validation  711 
Aggregation  349 
Alabama  366 
Alaska  488,  498,  509,  522 
Aleutian  Islands   168 
Alopias  superciliosus  -  see  shark, 

bigeye  thresher 
Alternative  fishing  practices  298 
Amberjack,  greater  545 
Anchoa  mitchilli  -  see  anchovy,  bay 
Anchoveta  207 
Anchovy,  bay  63 

Anoplopoma  fimbria  -  see  sablefish 
Antibodies,  monoclonal  555 
Assemblages  634 
Atlantic  245,407,671 
Australia  430,  464,  634,  581,  617. 

634 

Bass,  stripey  94 

Batch  fecundity  473 

Bering  Sea  509 

Berryteuthis  anonychus  -  see  squid, 

minimal  armhook 
Biomass 

distribution  63 

estimation  (sardine)  617 

spawning  stock  63 
Bones  498 

Bottom  trawl   168,  757,  740 
Brazil   156,393,581 
Brevoortia 

gunteri  -  see  menhaden,  finescale 

patronus  -see  menhaden,  gulf 
Browns  Bank  289 
Bycatch 

in  longline  fishery  393 

in  lobster  fishery  430 


in  sea  scallop  fishery  671 

in  squid  fishery  757 

in  tuna  purse-seine  fishery  233 

of  dolphins  233 

of  octopus  430 

of  sea  turtles  393,671 

California  25,  306,  389,  453 
Callineetes  sapidus  -  see  crab,  blue 
Cancer 

irroratus  -  see  crab,  Atlantic  rock 

magister  -  see  crab,  Dungeness 
Capture 

incidental  (sea  turtles)  393 

probability  740 

-recapture  563 
Carcharhinus  signatus  -  see  shark, 

night 
Caretta  caretta  -  see  sea  turtles, 

loggerhead 
Catch  rate  430 
Catchability  740 
Cephalorhynchus  hectori  -  see 

dolphin.  Hector's 
Cetaceans  661 
Chesapeake  Bay  63 
Chile  207 

Climate  change  400 
Clupeidae  723 
Cod 

Atlantic  289 

Pacific  400 
Conch,  queen  278 
Controlled  access  areas  563 
Corbicula  japonica    14 
CPUE  489 
Crab 

Atlantic  rock  693 

blue  693 

Dungeness  489 

lady  693 

red  king  740 
Crassostrea  virgin ica  -see  oyster, 

American 
Crustacea   118,489 
Cynoscion  nebulosus  -  see  seatrout, 

spotted 

Daytime  spawning  389 
Demographic  structure  593 
Density  (population)   179 
Dermochelys  coriacea  -  see  sea  turtle, 
leatherback 


Descriptions  (taxonomic) 

Brevoortia 
gunteri  723 
patronus  723 

Psednos  245 

Sebastes 
ciliatus  328 
variabilis  328 
wilsoni  452 
Diet 

cod.  Pacific  400 

prey  volume  366 

prowfish   168 

sea  lion 

California  47 
Steller  498,  509,  522 

seal,  harbor  108 

snapper,  red  366 

squid  733 
Digestion  correction  factor  498 
Dissolved  oxygen  63 
Distribution 

amberjack,  greater  533 

crab  693 

hatchdate  ( seatrout )  142 

prowfish   168 

salmon,  juvenile  25 

whale  563 
Diurnal  cycle  389 
Dive  transects  489 
Dissolved  oxygen  (DO)  63 
Dolphin 

Hector's  441 

Spotted,  pantropical  233 
Dorsal-fin  spine   251 

Egg 

deposition  306 

identification  (seabream)  555 

-size  variation  207 
Elasmobranch  156,  221,  757,  750 
Emigration  rate  464 
Engraulis  ringens  -  see  anchoveta 
Escapement  740 
Estuaries  376,533,648 
Eumetopias jubatus  -see  sea  lion. 

Steller 
Everglades   142 

Falkland  Islands  757 
Farfantepenaeus  aztecus  -  see  shrimp, 

brown 
Fecundity  473,  648 

batch   648 

relative  473 

rockfish,  black  418 

seabream  648 

squid,  market   306 


List  of  sublets 


767 


Finite  difference  calculus   14 
Fish  size  (red  snapper)  366 
Fisheries 

interaction  509 

management  581 

size-selective   1 

tuna  233 
Fishery  dynamics  (squid)  661 
Fishery  reserve  533 
Florida 

Bay  142 

eastern  coast  533 

Keys  278 
Flounder,  summer  78 
Footrope  740 
Formalin  555 
Foraging  (harbor  seal)  108 
Franciscana  581 

Gadus 

macroeephalus  -  see  cod,  Pacific 

morhua  -  see  cod,  Atlantic 
Galveston  Bay  264 
Generalized  linear  model  563 
Generalized  additive  model  563 
Genetic 

differentiation  (among  cod)  289 

identification   108,  213 
Georges  Bank  289 
GIS  264 
Glacier  Bay  489 
GLOBEC  25 
Goldlined  seabream  648 
Gonadosomatic  index  94 
Great  Barrier  Reef  94 
Growth 

amberjack,  greater  545 

curve  156 

prowfish   168 

rockfish452,  711 

seatrout,  spotted   142 

sex-specific  94 

shark,  night   156 

shrimp,  white  376 

snapper,  red  593 
Gulf  of  Alaska  168,400,509 
Gulf  of  California  47 
Gulf  of  Mexico  593 

Habitat 

conservation  298 

use  of  264 
Hand  harvesting  298 
Harvesting  techniques  298 
Hatching  date 

seatrout,  spotted  545 

snapper,  red  366 

histology  407 
Hudson-Raritan  Estuary  693 
Humerus  682 


Hypergeometric  likelihood  464 

Identification  (fish  eggs)  555 

Incidental  kill  233 

Indeterminate  648 

Indian  River  Lagoon  533 

Intertidal  634 

Istiophorus  platypterus  -  see  sailfish 

Japan  14,  196 

Jasus  edwardsii  -  see  lobster, 

southern  rock 
Juvenile  studies  25,  142,  264 

menhaden  723 

mortality  233 

pollock  604 

rockfish  453 

shrimp  264,  376 

snapper  366 

Kuroshio-Oyashio   179 

Lake   14 
Larval 

description  723 

development  118,  723 

rockfish  453 
Lead210  711 
Length 

at  age,  chub  mackerel   179 

frequency,  shark  156 
Life  history 

model   1 

seatrout.  spotted  142 
Line-transect  survey  441,  563 
Liparidae  245 
Litopenaeus  setiferus  -see  shrimp, 

white 
Lobster,  southern  rock  430 
Loligo  opalescens  -  see  squid,  market 
Longevity  ( rockfish )  711 
Longline  fishery   179,393 
Lutjanus 

campechanus  -  see  snapper,  red 

carponotatus  -  see  bass,  stripey 

Mackerel 

Atka  498,  509 

chub   179,  196 
Mammals  47,108,  213,  581,  563 
Marginal  increment  analysis   156 
Mark-recapture  78,  376 
Marlin,  white  750 
Marine  sanctuaries  464 
Maturity 

bass,  stripey  94 

rockfish,  black  418 
Maximum  likelihood  method   14 
Mediterranean  407 


Megaptera  novaeangliae  -see  whale, 

humpback 
Menhaden 

finescale  723 

gulf  723 
Mexico  47,  593 
Microwire  tags  376 
Mid-water  trawl  survey  63 
Migration 

diel  221 

from  marine  reserves  533 

shark  221 

vertical  221 
Models 

growth   14 

habitat  use  264 

life  history   1 

optimal  release  78 

recruitment  349 
Mortality  349 

amberjack,  greater  545 

dolphin,  pantropical  spotted  233 

lobster  430 

longline  fishery  179 

seatrout,  spotted  142 

squid,  market  306 
Movement 

shrimp  376 

sportfish  533 
MtDNA 

rockfish  453 

salmonids  213 
Multidimensional  scaling  36 

Nantucket  Shoals  289 
National  marine  sanctuary  563 
New  South  Wales  634 
New  species  245 
New  Zealand  441 
North  Carolina  (flounder)  78 
Northern  California  Current  25 
Nursery  habitat  366,  376 

Octopus,  maori  430 

Octopus  maorum  -  see  octopus,  maori 

Oncorhynchus 

kisutch  -  see  salmon,  coho  108 

Opah  750 

Orbital  retia  mirabilia  221 

Orcinus  orca  -  see  whale,  killer 

Oregon   108,418 

Otoliths 

amberjack,  greater  545 
elemental  signature  in  604 
in  fecal  samples  47,  108,  498,  522 
rockfish  453 

Ovalipes  ocellatus  -  see  crab,  lady 

Ovarian  cycle  418 

Oxytetracycline   127 

Oyster,  American  298 


768 


Fishery  Bulletin  102(4) 


Pacific  Ocean  168,  233,  328,  733,  604, 

733 
Pagrus  major  -  see  sea  bream,  red 
Pandalopsis  clispar  -  see  shrimp, 

sidestriped 
Paralichthys  dentatus  -see  flounder, 

summer 
Paralithodes  eamtschaticus  -see  crab, 

red  king 
Parturition  (rockfish)  418 
PavlofBay  400 
Penaeidae  376 
Phoca  vitulina  richardsi  -  see  seal, 

harbor 
Photo-identification  (marine 

mammals)  563 
Pinnipeds  498,  509,  522 
Phocoenoides  dalli  -  see  porpoise, 

Dall's 
Placopecten  magellanicus  -  see  sea 

scallop 
Pleurogrammus  monopterygius  -  see 

mackerel,  Atka 
Poikilotherm  682 
Pollock,  walleye  498,  509,  522,  604 
Pontoporia  blainvillei  -see 

franciscana 
Pop-up  satellite  archival  tags  221,  750 
Population 

dynamics  661 

structure  604 
Porgy,  red   1 
Porpoise,  Dall's  563 
Postovulatory  follicle  473 
Prey-size  selectivity  509,  522 
Protogynous  1 
Prowfish    168 

Psednos  rossi  -  see  snailfish 
Purse-seine  fishery  233 

Radiometric  age  711 

Radium226  711 

Rays  757 

Recolonization  (fish  in  rockpools)  634 

Recovery  278 

Recruitment 

anchovy,  bay  63 

processes  of  349 
Reef  habitat  298 
Remote  sensing  661 
Remotely  operated  vehicle   389 
Reproduction 

bass,  stripey  94 

behavior  389 

franciscana  581 

prowfish   168 

rockfish,  black  418 

seatrout,  spotted  473 


squid  389 

tuna,  bluefin  407 
Residence  time  376 
Restoration 

conch  278, 

oyster  298 
Rhabdosargus  sarba — see  seabream, 

goldlined 
Rio  Grande  do  Sul  581 
Rockfish 

black  418 

blackgill  711 

dusky  328 

pygmy  452 

rockpools  634 
ROV  (remotely  operated  vehicle)  389 

Sablefish   127 
Sailfish  251 
Salmon 

coho  25 

decline  in  Onchorynchus  spp.   108 

juvenile  25 

predation  on   108 

prey  of  213 
Salmonids  25,  213 

juvenile  25 
Sampling  bias  488 
Sardine,  Pacific  581,  617 
Sardinops  sagax  -  see  sardine. 

Pacific 
Sargassum   545 
Satellite 

archival  tags  750 

remote  sensing  661 
Scale-free  networks  349 
Scat  47,  108,  213,  498,  522 
Scomber  japonicus  -  see  mackerel, 

chub 
Scorpaenidae  711 
SCUBA  366 
Sea  bream 

goldined  648 

red  555 
Sea  lion 

California  47 

Steller  498,  509,  522 
Sea  scallop  671 
Sea  surface  temperature  179 
Seal,  harbor  108,  213 
Seatrout,  spotted   142,  473 
Sebastes 

ciliatus  -  see  rockfish,  dusky 

melanops  -see  rockfish,  black 

melanostomus  -  see  rockfish, 
blackgill 

variabilis  -  see  rockfish,  dusky 

uilsoni  -  see  rockfish,  pygmy 
Sea  turtle 

Kemp's  ridley  682 


leatherback  393 

loggerhead  393,682 
Seriola  dumerili  -  see  amberjack, 

greater 
Sex 

change  1 

ratio  693 
Sexual  maturity  418 
Sharks 

thresher,  bigeye  221 

night  156 

predation  750 
Shrimp 

brown  264 

sidestriped  118 

white  376 
Size-selective  fisheries   1 
Skeletochronology  (sea  turtles)  682 
Snailfish  245 
Snapper 

red  366,  593 

Spanish  flag  93 
Soak  time  (gear)   179 
South  Carolina  473 
Sparidae  555,648 
Spawning 

anchovy,  bay  63 

bass,  stripey  94 

diel  389 

frequency  473 

seabream  648 

season  418,  648 

squid  306 

stock  biomass  63 

tuna,  bluefin  407 
Species 

protogynous   1 
Squid  73 

fishery  757 

market  306,389,661 

minimal  armhook  733 
SteneUa  attenuata  attenuata  -see 

dolphin,  spotted,  pantropical 
Stock  enhancement  78 
Strombus  gigas  -  see  conch,  queen 
Surveys 

dolphin  441,489 

with  fish  traps   127 
Survival  rates   179,  757 

Tagging  464,  750,  604 

marine  reserves  464,  533 

natural  604 

shark  221,  750 
Taiwan  251 

Temperate  estuaries  693 
Temperature 

Effect  on  fish  size  I  mackerel)   196 


List  of  subjects 


769 


Testicular  development  407 
Tetrapturus  albidus  -  see  marlin, 

white 
Theragra  chalcogramma  -  see  pollock, 

walleye 
Thunnus  thynnus  -  see  tuna,  bluefin 
Translocation  (for  rehabilitation)  278 
Traps  430 
Trawl  efficiency  740 
Trophic  niche  breadth  693 


Tuna,  bluefin  407 
Turtle  -  see  sea  turtle 

Umpqua  River  108,  213 
Underwater  video  740 

Vertebral  sections   156 

Washington  563 

Western  North  Atlantic  245 


Whale 

humpback  563 
killer  563 

Young-of-the-year  (amberjack)  545 

Zalophus  californianus  -  see  sea  lion, 

California 
Zaprora  silenus  -see  prowfish 


770 


Fishery  Bulletin  102(4) 


Superintendent  of  Documents  Publications  Order  Form 
*5178 

I I  YJl/O,  please  send  me  the  following  publications: 


Subscriptions  to  Fishery  Bulletin 
for  S55.00  per  year  ($68.75  foreign) 


The  total  cost  of  mv  order  is  $  . 


Prices  include  regular  domestic 


postage  and  handling  and  are  subject  to  change. 


I  Company  or  Personal  Name  i 


(Please  type  or  print! 


i Additional  address/attention  line) 


I  Street  address) 


(City,  State,  ZIP  Code  I 


I  Daytime  phone  including  area  codei 


i  Purchase  Order  No. ) 


Charge 
your 
order. 

IT'S 
EASY! 


Please  Choose  Method  of  Payment: 

]  Check  Payable  to  the  Superintendent  of  Documents 
□  GPO  Deposit  Account   I    I    I    I    I    I    I    I 
]  VISA  or  MasterCard  Account 


To  fax 

vour  orders 

(2021  512-2250 


(Credit  card  expiration  date) 


(Authorizing  Signature) 

Mail  To:    Superintendent  of  Documents 

P.O.  Box  371954,  Pittsburgh,  PA  15250-7954 


Thank  you  for 
your  order! 


This  statement  is  required  by  the  Act  of  August  12, 
1970,  Section  3685,  Title  39,  U.S.  Code,  showing 
ownership,  management,  and  circulation  of  the 
Fishery  Bulletin,  publication  number  366-370, 
and  was  filed  on  02  September,  2004.  The  Bulletin 
is  published  quarterly  (four  issues  annually) 
with  an  annual  subscription  price  of  $55.00 
(sold  by  the  Superintendent  of  Documents,  U.S.  Gov- 
ernment Printing  Office,  Washington.  DC  20402). 
The  complete  mailing  address  of  the  office  of  publica- 
tion is  N'MFS  Scientific  Publications  Office,  NOAA, 
7600  Sand  Point  Way  NE,  BIN  C15700,  Seattle,  WA 
98115.  The  complete  mailing  address  of  the  head- 
quarters of  the  publishing  agency  is  National  Marine 
Fisheries  Service.  N'l  >.VV  Department  nil  'nmmerce. 
1335  East-West  Highway,  Silver  Spring,  MD  20910. 
Tin  name  of  the  publisher  is  Willis  Hobart  and  the 
managing  editor  is  Sharvn  Matnotli;  their  mailing 
address  i-  NMFS  Scientific  Publications  Office,  7600 
Sand  Point  Way  NE.  BIN  15700,  Seattle.  WA98 1 15 


The  owner  is  the  U.S.  Department  of  Commerce, 
14th  St.  N.W,  Washington,  DC  20230;  there  are 
no  bondholders,  mortgages,  or  other  security  hold- 
ers. The  purpose,  function,  and  nonprofit  status  of 
the  organization  (agency)  and  the  exempt  status 
for  Federal  income  tax  purposes  has  not  changed 
during  the  preceding  12  months.  The  extent  and 
nature  of  circulation  is  as  follows:  total  number 
of  copies  (A)  (average  number  of  copies  of  each 
issue  during  the  preceding  12  months)  was  1371 
and  the  actual  number  of  copies  of  the  single 
issue  pu  I>1]  sheil  iiriirest  to  ill,-  lib  updates  was  l.'iSL! 
Paid  circulation  (B)  is  handled  by  the  U.S.  Govern- 
in.  -ill  I 'nut  me,  <  I  If  ice.  Washington.  IK1  L'D  Id'J.  anil 
the  total  number  printed  lor  sales  i  mail  subscrip- 
tions and  individual  sales  '  was  17")  tot  the  average 
number  of  copies  each  issue  during  the  preceding 
12  months  and  475  the  actual  number  of  copies  of 
( In-  Bingle  issue  published  nearest  to  the  filing  date 
(C)   Vix-v  distribution  (D)  by  mail;  samples,  compli- 


mentary and  other  free  copies  (average  number  of 
copies  each  issue  during  the  preceding  12  months) 
was  792  and  the  actual  number  of  copies  of  the 
single  issue  published  nearest  to  the  filing  date 
was  761.  Free  distribution  outside  the  mail  (El 
by  carriers  or  other  means  was  0  for  both  average 
number  of  copies  and  actual  number  of  copies. 
Total  free  distribution  iF)  was  0  for  both  average 
number  of  copies  and  actual  number  of  copies  of 
the  single  issue  published  nearest  the  filing  date. 
The  total  distribution  iG:  sum  of  D  and  B)  (average 
number  of  copies  each  issue  during  the  preceding 
12  months)  was  1267  and  the  actual  number  of 
copies  of  the  single  issue  published  nearest  to  the 
filing  date  was  1236.  There  were  65  copies  (avg. 
annual)  not  distributed  iHi.  The  total  (I:  sum  of  G 
and  H  i  is  equal  to  the  net  press  run  figures  shown  in 
Item  A:  1332  and  1301  copies,  respectively.  I  certify 
that  the  statements  made  by  me  above  are  correct 
and  complete:  (Signed1  Willis  Hobart.  Publisher. 


*C  "k 


Fishery  Bulletin 

Guidelines  for  contributors 


Content  of  papers 

Articles 

Articles  are  reports  of  10  to  30  pages  (double 
spaced)  that  describe  original  research  in  one  or 
a  combination  of  the  following  fields  of  marine 
science:  taxonomy,  biology,  genetics,  mathematics 
(including  modeling),  statistics,  engineering,  eco- 
nomics, and  ecology. 

Notes 

Notes  are  reports  of  5  to  10  pages  without  an 
abstract  that  describe  methods  and  results  not 
supported  by  a  large  body  of  data.  Although  all 
contributions  are  subject  to  peer  review,  responsi- 
bility for  the  contents  of  articles  and  notes  rests 
upon  the  authors  and  not  upon  the  editor  or  the 
publisher.  It  is  therefore  important  that  authors 
consider  the  contents  of  their  manuscripts  care- 
fully. Submission  of  an  article  is  un-derstood  to 
imply  that  the  article  is  original  and  is  not  being 
considered  for  publication  elsewhere.  Manuscripts 
must  be  written  in  English.  Authors  whose  native 
language  is  not  English  are  strongly  advised  to 
have  their  manuscripts  checked  for  fluency  by 
English-speaking  colleagues  prior  to  submission. 

Preparation  of  papers 
Text 

Title  page  should  include  authors'  full  names  and 
mailing  addresses  (street  address  required)  and 
the  senior  author's  telephone,  fax  number,  e-mail 
address,  as  well  as  a  list  of  key  words  to  describe  the 
contents  of  the  manuscript.  Abstract  must  be  less 
than  one  typed  page  (double  spaced)  and  must  not 
contain  any  citations.  It  should  state  the  main  scope 
of  the  research  but  emphasize  the  author's  con- 
clusions and  relevant  findings.  Because  abstracts 
are  circulated  by  abstracting  agencies,  it  is  impor- 
tant that  they  represent  the  research  clearly  and 
concisely.  General  text  must  be  typed  in  double- 
spaced  format.  A  brief  introduction  should  state  the 
broad  significance  of  the  paper;  the  remainder  of 
the  paper  should  be  divided  into  the  following  sec- 
tions: Materials  and  methods,  Results,  Discussion 
(or  Conclusions),  and  Acknowledgments.  Headings 
within  each  section  must  be  short,  reflect  a  logical 
sequence,  and  follow  the  rules  of  multiple  subdi- 
vision (i.e.  there  can  be  no  subdivision  without  at 
least  two  subheadings).  The  entire  text  should  be 
intelligible  to  interdisciplinary  readers;  therefore, 
all  acronyms  and  abbreviations  should  be  written 
out  and  all  lesser-known  technical  terms  should  be 
defined  the  first  time  they  are  mentioned.  The 
scientific  names  of  species  must  be  written  out  the 
first  time  they  are  mentioned;  subsequent  mention 
of  scientific  names  may  be  abbreviated.  Follow  Sci- 
entific style  and  format:  CBE  manual  for  authors, 
editors,  and  publishers  (6th  ed.)  for  editorial  style 
and  the  most  current  issue  of  the  American  Fish- 
eries Society's  common  and  scientific  names  of 
fishes  from  the  United  States  and  Canada  for 
fish  nomenclature.  Dates  should  be  written  as  fol- 
lows: 11  November  1991.  Measurements  should  be 
expressed  in  metric  units,  e.g.  metric  tons  (t).  The 
numeral  one  ( 1 )  should  be  typed  as  a  one,  not  as  a 
lower-case  el  (1). 

Footnotes 

Use  footnotes  to  add  editorial  comments  regarding 
claims  made  in  the  text  and  to  document  unpub- 


lished works  or  works  with  local  circulation.  Foot- 
notes should  be  numbered  with  Arabic  numerals 
and  inserted  in  10-point  font  at  the  bottom  of  the 
first  page  on  which  they  are  cited.  Footnotes  should 
be  formatted  in  the  same  manner  as  citations. 
If  a  manuscript  is  unpublished,  in  the  process 
of  review,  or  if  the  information  provided  in  the 
footnote  has  been  conveyed  verbally,  please  state 
this  information  as  ''unpubl.  data,"  "manuscript 
in  review,"  and  "personal  commun.,"  respectively. 
Authors  are  advised  wherever  possible  to  avoid  ref- 
erences to  nonstandard  literature  (unpublished  lit- 
erature that  is  difficult  to  obtain,  such  as  internal 
reports,  processed  reports,  administrative  reports, 
ICES  council  minutes,  IWC  minutes  or  working 
papers,  any  "research"  or  "working"  documents, 
laboratory  reports,  contract  reports,  and  manu- 
scripts in  review).  If  these  references  are  used, 
please  indicate  whether  they  are  available  from 
NTIS  (National  Technical  Information  Sendee)  or 
from  some  other  public  depository.  Footnote  format: 
author  (last  name,  followed  by  first-name  initials); 
year;  title  of  report  or  manuscript;  type  of  report 
and  its  administrative  or  serial  number;  name  and 
address  of  agency  or  institution  where  the  report  is 
filed. 

Literature  cited 

The  literature  cited  section  comprises  works  that 
have  been  published  and  those  accepted  for  pub- 
lication (works  in  press)  in  peer-reviewed  jour- 
nals and  books.  Follow  the  name  and  year  system 
for  citation  format.  In  the  text,  write  "Smith  and 
Jones  ( 1977 )  reported"  but  if  the  citation  takes 
the  form  of  parenthetical  matter,  write  "(Smith 
and  Jones,  1977)."  In  the  literature  cited  section, 
list  citations  alphabetically  by  last  name  of  senior 
author:  For  example,  Alston,  1952;  Mannly,  1988; 
Smith,  1932;  Smith,  1947;  Stalinsky  and  Jones, 
1985.  Abbreviations  of  journals  should  conform 
to  the  abbreviations  given  in  the  Serial  sources 
for  the  BIOSIS  previews  database.  Authors  are 
responsible  for  the  accuracy  and  completeness  of 
all  illations.  Literature  citation  format:  author 
(last  name,  followed  by  first-name  initials);  year; 
title  of  report  or  article;  abbreviated  title  of  the 
journal  in  which  the  article  was  published,  volume 
number,  page  numbers.  For  books,  please  provide 
publisher,  city,  and  state. 

Tables 

Tables  should  not  be  excessive  in  size  and  must  be 
cited  in  numerical  order  in  the  text.  Headings  in 
tables  should  be  short  but  ample  enough  to  allow 
the  table  to  be  intelligible  on  its  own.  All  unusual 
symbols  must  be  explained  in  the  table  legend. 
Other  incidental  comments  may  be  footnoted  (use 
italic  arabic  numerals  for  footnote  markers).  Use 
asterisks  only  to  indicate  probability  in  statistical 
data.  Place  table  legends  on  the  same  page  as  the 
table  data.  We  accept  tables  saved  in  most  spread- 
sheet software  programs  (e.g.  Microsoft  Excel). 
Please  note  the  following: 

•  Use  a  comma  in  numbers  of  five  digits  or  more 
(e.g.  13,000  but  3000). 

•  Use  zeros  before  all  decimal  points  for  values 
less  than  one  (e.g.  0.31). 

Figures 

Figures  include  line  illustrations,  computer-gener- 
ated line  graphs,  and  photographs  (or  slides).  They 


must  be  cited  in  numerical  order  in  the  text.  Line 
illustrations  are  best  submitted  as  original  draw- 
ings. Computer-generated  line  graphs  should  be 
printed  on  laser-quality  paper.  Photographs  should 
be  submitted  on  glossy  paper  with  good  contrast. 
All  figures  are  to  be  labeled  with  senior  author's 
name  and  the  number  of  the  figure  (e.g.  Smith, 
Fig.  4).  Use  Helvetica  or  Arial  font  to  label  ana- 
tomical parts  (line  drawings)  or  variables  (graphs) 
within  figures;  use  Times  Roman  bold  font  to  label 
the  different,  sections  of  a  figure  (e.g.  A,  B,  C). 
Figure  legends  should  explain  all  symbols  and 
abbreviations  seen  within  the  figure  and  should  be 
typed  in  double-spaced  format  on  a  separate  page 
at  the  end  of  the  manuscript.  We  advise  authors  to 
peruse  a  recent  issue  of  Fishery  Bulletin  for  stan- 
dard formats.  Please  note  the  following: 

•Capitalize  the  first  letter  of  the  first  word  of 
axis  labels. 

•  Do  not.  use  overly  large  font  sizes  to  label  axes 
or  parts  within  figures. 

•  Do  not  use  boldface  fonts  within  figures. 

•  Do  not  create  outline  rules  around  graphs. 

•  Do  not  use  horizontal  lines  through  graphs. 

•  Do  not  use  large  font  sizes  to  label  degrees  of 
longitude  and  latitude  on  maps. 

•  Indicate  direction  of  degrees  longitude  and 
latitude  on  maps  (e.g.  170  El 

•  Avoid  placing  labels  on  a  vertical  plane 
(except  on y  axis). 

•Avoid  odd  (nonstandard)  patterns  to  mark 
sections  of  bar  graphs  and  pie  charts. 

Copyright  law 

Fishery  Bulletin .  a  U.S.  government  publication,  is 
not  subject  to  copyright  law.  If  an  author  wishes  to 
reproduce  any  part  of  Fishery  Bulletin  in  his  or  her 
work,  he  or  she  is  obliged,  however,  to  acknowledge 
the  source  of  the  extracted  literature. 

Submission  of  papers 

Send  four  printed  copies  (one  original  plus  three 
copies ) — clipped,  not  stapled — to  the  Scientific  Edi- 
tor, at  the  address  shown  below.  Send  photocopies 
of  figures  with  initial  submission  of  manuscript. 
Original  figures  will  be  requested  later  when  the 
manuscript  has  been  accepted  for  publication. 
Do  not  send  your  manuscript  on  diskette  until 
requested  to  do  so. 

Dr.  Norman  Bartoo 

National  Marine  Fisheries  Service,  NOAA 

8604  La  Jolla  Shores  Drive 

La  Jolla,  CA  92037 

Once  the  manuscript  has  been  accepted  for  publi- 
cation, you  will  be  asked  to  submit  a  software  copy 
of  your  manuscript.  The  software  copy  should  be 
submitted  in  WordPerfect  or  Word  format  (in 
Word,  save  as  Rich  Text  Format).  Please  note  that 
we  do  not  accept  ASCII  text  files.  Color  figures 
must  be  CMYK  files. 

Reprints 

Copies  of  published  articles  and  notes  are  avail- 
able free  of  charge  to  the  senior  author  (50  copies) 
and  to  his  or  her  laboratory  (50  copies).  Additional 
copies  may  be  purchased  in  lots  of  100  when  the 
author  receives  page  proofs.