Skip to main content

Full text of "Fishery bulletin"

See other formats


U.S.  Department 
of  Commerce 

Volume  108 
Number  2 
April  2010 


U.S.  Department 
of  Commerce 

Gary  Locke 

Secretary  of  Commerce 

National!  Oceanic 
and  Atmospheric 
Administration 

Jane  Lubchenco,  Ph.D. 

Administrator  of  NOAA 


National  Marine 
Fisheries  Service 

James  W.  Balsiger,  Ph.D. 

Acting  Assistant  Administrator 
for  Fisheries 


The  Fishery  Bulletin  (ISSN  0090-0656) 
is  published  quarterly  by  the  Scientific 
Publications  Office,  National  Marine 
Fisheries  Service,  NOAA,  7600  Sand 
Point  Way  NE,  BIN  C15700,  Seattle,  WA 
98115-0070.  Periodicals  postage  is  paid 
at  Seattle,  WA.  POSTMASTER:  Send 
address  changes  for  subscriptions  to  Fish- 
ery Bulletin,  Superintendent  of  Docu- 
ments, Attn.:  Chief,  Mail  List  Branch, 
Mail  Stop  SSOM,  Washington,  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  Secretary  of  Commerce  has  deter- 
mined that  the  publication  of  this  peri- 
odical is  necessary  according  to  law  for 
the  transaction  of  public  business  of  this 
Department.  Use  of  funds  for  printing  of 
this  periodical  has  been  approved  by  the 
Director  of  the  Office  of  Management  and 
Budget. 

For  sale  by  the  Superintendent  of 
Documents,  U.S.  Government  Printing 
Office,  Washington,  DC  20402.  Subscrip- 
tion price  per  year:  $36.00  domestic  and 
$50.40  foreign.  Cost  per  single  issue: 
$21.00  domestic  and  $29.40  foreign.  See 
back  for  order  form. 


Scientific  Editor 

Richard  D.  Brodeur,  Ph.D. 

Associate  Editor 
Julie  Scheurer 

National  Marine  Fisheries  Service 
Northwest  Fisheries  Science  Center 
2030  S.  Marine  Science  Dr. 
Newport,  Oregon  97365-5296 


Managing  Editor 

Sharyn  Matriotti 

National  Marine  Fisheries  Service 
Scientific  Publications  Office 
7600  Sand  Point  Way  NE 
Seattle,  Washington  98115-0070 


Editorial  Committee 


John  Carlson 
Kevin  Craig 
Jeff  Leis 
Rich  McBride 
Rick  Methot 
Adam  Moles 
Frank  Parrish 
Dave  Somerton 
Ed  Trippel 
Mary  Yoklavich 


National  Marine  Fisheries  Service,  Panama  City,  Florida 

Florida  State  University,  Tallahassee,  Florida 

Australian  Museum,  Sydney,  New  South  Wales,  Australia 

National  Marine  Fisheries  Service,  Woods  Hole,  Massachusetts 

National  Marine  Fisheries  Service,  Seattle,  Washington 

National  Marine  Fisheries  Service,  Auke  Bay,  Alaska 

National  Marine  Fisheries  Service,  Honolulu,  Hawaii 

National  Marine  Fisheries  Service,  Seattle,  Washington 

Department  of  Fisheries  and  Oceans,  St.  Andrews,  New  Brunswick,  Canada 

National  Marine  Fisheries  Service,  Santa  Cruz,  California 


Fishery  Bulletin  web  site:  www.fisherybulletin.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  1941.  Separates  were  issued  as  documents  through 
volume  46;  the  last  document  was  No.  1103.  Beginning  with  volume  47  in  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  in  exchange 
for  other  scientific  publications. 


U.S.  Department 
of  Commerce 

Seattle,  Washington 

Volume  108 
Number  2 
April  2010 


Fishery 

Bulletin 


Contents 


Articles 


119-135  Tribuzio,  Cindy  A.,  Gordon  H.  Kruse,  and  Jeffrey  T.  Fujioka 

Age  and  growth  of  spiny  dogfish  (Squalus  acanthias) 
in  the  Gulf  of  Alaska:  analysis  of  alternative  growth  models 


Companion  articles 


136-144  Rose  Craig  S.,  Carwyn  F.  Hammond,  and  John  R.  Gauvin 

Effective  herding  of  flatfish  by  cables  with  minimal  seafloor  contact 

145-154  Ryer,  Clifford  H.,  Craig  S.  Rose,  and  Paul  J.  Iseri 

Flatfish  herding  behavior  in  response  to  trawl  sweeps: 
a comparison  of  diel  responses  to  conventional  sweeps 
and  elevated  sweeps 


The  National  Marine  Fisheries 
Service  (NMFS)  does  not  approve, 
recommend,  or  endorse  any  proprie- 
tary product  or  proprietary  material 
mentioned  in  this  publication.  No 
reference  shall  be  made  to  NMFS, 
or  to  this  publication  furnished  by 
NMFS,  in  any  advertising  or  sales 
promotion  which  would  indicate  or 
imply  that  NMFS  approves,  recom- 
mends, or  endorses  any  proprietary 
product  or  proprietary  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. 

The  NMFS  Scientific  Publications 
Office  is  not  responsible  for  the  con- 
tents of  the  articles  or  for  the  stan- 
dard of  English  used  in  them. 


155-161  Mateo,  Ivan,  Edward  G.  Durbin,  David  A.  Bengtson, 

Richard  Kingsley,  Peter  K.  Swart,  and  Daisy  Durant 

Spatial  and  temporal  variation  in  otolith  chemistry  for  tautog 
( Toutoga  onitis)  in  Narragansett  Bay  and  Rhode  Island  coastal  ponds 

162-173  Masuda,  Reiji,  Masami  Shiba,  Yoh  Yamashita,  Masahiro  Ueno, 
Yoshiaki  Kai,  Asami  Nakanishi,  Masaru  Torikoshi, 
and  Masaru  Tanaka 

Fish  assemblages  associated  with  three  types  of  artificial  reefs:  density 
of  assemblages  and  possible  impacts  on  adiacent  fish  abundance 


174-192  Lo,  Nancy  C.  H.,  Beverly  J.  Macewicz,  and  David  A.  Griffiths 

Biomass  and  reproduction  of  Pacific  sardine  (Sardinops  sagax) 
off  the  Pacific  northwestern  United  States,  2003-2005 


193-207  Hernandez  Jr.,  Frank  J.,  Sean  P.  Powers,  and  William  M .Graham 
Seasonal  variability  in  ichthyoplankton  abundance  and  assemblage 
composition  in  the  northern  Gulf  of  Mexico  off  Alabama 


II 


Fishery  Bulletin  108(2) 


208-217 

Stevenson,  Duane  E.,  and  Kristy  A.  Lewis 

Observer-reported  skate  bycatch  in  the  commercial  groundfish  fisheries  of  Alaska 

218-225 

Fergusson,  Emily  A.,  Molly  V.  Sturdevant,  and  Joseph  A.  Orsi 

Effects  of  starvation  on  energy  density  of  juvenile  chum  salmon  (Oncorhynchus  keta) 
captured  in  marine  waters  of  Southeastern  Alaska 

226-232 

Fruh,  Erica  L.,  Aimee  Keller,  Jessica  Trantham,  and  Victor  Simon 
Accuracy  of  sex  determination  for  northeastern  Pacific  Ocean  thornyheads 
( Sebastolobus  altivelis  and  5.  alascanus) 

233-247 

Jacobson,  Larry  D.,  Kevin  D.  E.  Stokesbury,  Melissa  A.  Allard,  Antonie  Chute,  Bradley  P.  Harris, 
Deborah  Hart,  Tom  Jaffarian,  Michael  C.  Marino  II,  Jacob  1.  Nogueira,  and  Paul  Rago 

Measurement  errors  in  body  size  of  sea  scallops  ( Placopecten  magellanicus) 
and  their  effect  on  stock  assessment  models 

248 

Guidelines  for  authors 
Subscription  form  (inside  back  cover) 

119 


Age  and  growth  of  spiny  dogfish 
{Squalus  accmthias ) in  the  GuSff  of  Alaska: 
analysis  of  alternative  growth  models 

Cindy  A.  Tribuzio  (contact  author)1 
Gordon  H.  Kruse1 
Jeffrey  T.  Fujioka2 

Email  address  for  contact  author:  cindy.tribuzio@noaa.gov 

1 School  of  Fisheries  and  Ocean  Sciences,  Juneau  Center 
University  of  Alaska  Fairbanks 

17101  Pt.  Lena  Loop  Road 
Juneau,  Alaska  99801 

Present  address  for  contact  author:  National  Oceanic  and  Atmospheric  Administration 

National  Marine  Fisheries  Service 
Alaska  Fisheries  Science  Center 
Auke  Bay  Laboratories 
17109  Pt.  Lena  Loop  Road 
Juneau,  Alaska  99801 

2 National  Oceanic  and  Atmospheric  Administration 
National  Marine  Fisheries  Service 

Alaska  Fisheries  Science  Center 
Auke  Bay  Laboratories 
17109  Pt.  Lena  Loop  Road 
Juneau,  Alaska  99801 


Abstract — Ten  growth  models  were 
fitted  to  age  and  growth  data  for  spiny 
dogfish  ( Squalus  acanthias ) in  the 
Gulf  of  Alaska.  Previous  studies  of 
spiny  dogfish  growth  have  all  fitted 
the  t0  formulation  of  the  von  Berta- 
lanffy  model  without  examination 
of  alternative  models.  Among  the 
alternatives,  we  present  a new  two- 
phase  von  Bertalanffy  growth  model 
formulation  with  a logistically  scaled 
k parameter  and  which  estimates  L0. 
A total  of  1602  dogfish  were  aged 
from  opportunistic  collections  with 
longline,  rod  and  reel,  set  net,  and 
trawling  gear  in  the  eastern  and  cen- 
tral Gulf  of  Alaska  between  2004  and 
2007.  Ages  were  estimated  from  the 
median  band  count  of  three  indepen- 
dent readings  of  the  second  dorsal 
spine  plus  the  estimated  number  of 
worn  bands  for  worn  spines.  Owing  to 
a lack  of  small  dogfish  in  the  samples, 
lengths  at  age  of  small  individuals 
were  back-calculated  from  a subsam- 
ple of  153  dogfish  with  unworn  spines. 
The  von  Bertalanffy,  two-parameter 
von  Bertalanffy,  two-phase  von  Ber- 
talanffy, Gompertz,  two-parameter 
Gompertz,  and  logistic  models  were 
fitted  to  length-at-age  data  for  each 
sex  separately,  both  with  and  without 
back-calculated  lengths  at  age.  The 
two-phase  von  Bertalanffy  growth 
model  produced  the  statistically  best 
fit  for  both  sexes  of  Gulf  of  Alaska 
spiny  dogfish,  resulting  in  L00  = 87.2 
and  102.5  cm  and  £ = 0.106  and  0.058 
for  males  and  females,  respectively. 


Manuscript  submitted  17  February  2009. 
Manuscript  accepted  3 November  2009. 
Fish.  Bull.  108:119-135  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


The  spiny  dogfish  (Squalus  acanthias ) 
is  a small  long-lived  shark  common 
among  temperate  coastal  areas  in 
the  Atlantic  and  Pacific  oceans  (Com- 
pagno,  1984).  This  species  has  been 
the  target  of  commercial  fisheries  over 
much  of  its  range,  in  some  cases  for 
over  a century  (Ketchen,  1986).  In 
some  areas,  severe  declines  in  popu- 
lation abundance  and  stock  structure 
have  occurred  (e.g.,  Rago  et  al.,  1998). 
Many  elasmobranchs,  including  spiny 
dogfish,  are  “equilibrium  strategists” 
that  are  highly  susceptible  to  over- 
fishing because  of  their  slow  growth 
rates,  low  fecundity,  and  late  matu- 
ration (King  and  McFarlane,  2003), 
all  of  which  are  directly  related  to 
recruitment  and  parental  stock  sizes 
(Holden,  1974;  1977).  Off  the  west 
coast  of  North  America,  spiny  dogfish 
were  depleted  by  intense  fisheries  in 
the  1940s,  owing  to  the  quantity  and 
quality  of  vitamin  A in  their  livers 
(Ketchen,  1986);  the  fishery  demand 
decreased  by  1950  with  the  develop- 
ment of  synthetic  vitamin  A (Ketchen 
et  al.,  1983).  Since  the  1970s,  spiny 
dogfish  have  continued  to  be  targeted 
by  commercial  fisheries  in  British 


Columbia  and  the  state  of  Washing- 
ton for  human  consumption. 

Although  not  targeted,  spiny  dog- 
fish is  a common  bycatch  species  in 
many  fisheries  in  both  state  and  fed- 
eral waters  off  the  coast  of  Alaska. 
In  the  Gulf  of  Alaska  (GOA)  spiny 
dogfish  are  taken  in  Pacific  salmon 
(Oncorhynchus  spp.)  gillnet  fisher- 
ies, sablefish  ( Anoplopoma  fimbria) 
fisheries,  Pacific  halibut  (Hippoglos- 
sus  stenolepis)  longline  fisheries,  and 
groundfish  trawl  fisheries  (Boldt, 
2003).  Although  an  estimated  aver- 
age of  482.1  metric  tons  (t)  of  spiny 
dogfish  was  taken  annually  from 
1997  to  2007  in  observed  fisheries 
(Tribuzio  et  al.,  2008),  the  bycatch 
in  state  waters  is  unknown  and  the 
bycatch  rates  in  federally  managed 
fisheries  are  likely  underestimated 
because  of  unobserved  fisheries  (e.g., 
the  halibut  individual  fishing  quota, 
IFQ).  Nearly  all  of  this  unintended 
bycatch  was  and  still  is  discarded  at 
sea.  Even  though  estimated  catch  is 
<1%  of  estimated  spiny  dogfish  bio- 
mass (Courtney  et  al.,  2006),  the  po- 
tential development  of  a commercial 
fishery  demands  further  investigation 


120 


Fishery  Bulletin  108(2) 


of  the  effect  of  total  fishing  mortality  on  biomass  and 
an  investigation  of  spiny  dogfish  life  history  character- 
istics in  Alaska. 

Biological  reference  points  (e.g.,  BMS Y,  F35%)  are 
benchmarks  against  which  stock  abundance  or  fishing 
mortality  rates  can  be  compared  to  determine  stock 
status.  Most  commonly  used  reference  points  are  func- 
tions of  stock  productivity,  such  as  growth,  recruitment, 
and  natural  mortality  (Bonfil,  2005);  thus  accurate  es- 
timates of  age  and  growth  are  important.  For  instance, 
estimates  of  age  and  the  growth  coefficient  (k)  are  criti- 
cal for  estimating  natural  mortality  (M),  where  a lack 
of  data  prevent  direct  estimation  of  M,  abundance,  and 
appropriate  harvest  rates.  In  the  GOA,  biological  refer- 
ence points,  such  as  those  from  age  and  growth  models, 
have  yet  to  be  determined  for  spiny  dogfish. 

Extension  of  life  history  parameters  from  other  re- 
gions to  Alaska  may  be  inappropriate  because  age  and 
growth  characteristics  of  spiny  dogfish  vary  widely  over 
its  geographic  range.  For  example,  maximum  age  in  the 
northwest  Atlantic  Ocean  is  35-40  years  (Nammack 
et  al.,  1985),  but  in  the  eastern  North  Pacific,  spiny 
dogfish  have  been  aged  to  over  80  years  (Saunders  and 
McFarlane,  1993).  Growth  characteristics  also  vary 
widely  throughout  the  North  Pacific  and  North  Atlantic 
oceans  (Ketchen,  1975;  Nammack  et  al.,  1985).  Even 
within  the  North  Pacific  basin,  biological  parameters, 
such  as  k,  can  vary  with  latitude  (Vega,  2006). 

The  selection  of  an  appropriate  growth  model  is  im- 
portant when  estimating  regionally  specific  parameters. 
Elasmobranch  age  and  growth  studies  have  generally 
focused  on  fitting  length-at-age  data  to  the  von  Berta- 
lanffy  (vB)  growth  equation,  irrespective  of  goodness- 
of-fit  or  alternative  growth  models  (Carlson  and  Bare- 
more,  2005).  Despite  its  common  use,  the  vB  growth 
equation  may  not  be  the  best-fit  growth  model  for  all 
elasmobranch  species.  For  example,  the  logistic  model 
fitted  best  among  four  models  tested  for  the  spinner 
shark  ( Carcharhinus  brevipinna,  Carlson  and  Baremore, 
2005),  and  a two-phase  vB  model  fitted  best  among  five 
models  for  the  piked  spurdog  ( Squalus  megalops,  Brac- 
cini  et  al.,  2007).  A model  that  is  not  the  best  descriptor 
of  a species’  growth  could  have  compounding  effects  on 
demographic  analyses,  stock  assessment,  and  fishery 
management. 

Typical  growth  models  involve  parameters  of  asymp- 
totic length  (L^),  k,  and  t0  (Cailliet  et  ah,  2006).  The  t0 
parameter  is  biologically  difficult  to  interpret  because 
it  is  not  measurable  and  testable  in  wild  animals  (Be- 
verton  and  Holt,  1957).  This  parameter  is  the  age  at 
which  the  animal  is  of  zero  length  and  is  based  on  an 
assumption  of  a fixed  growth  curve  from  fertilization 
through  life  (Beverton  and  Holt,  1957).  It  is  generally 
interpreted  to  represent  the  period  of  gestation  in  tele- 
ost  fish  species,  but  this  assumption  is  violated  for  elas- 
mobranchs  (Driggers  et  al.,  2004).  For  instance,  when 
considering  males  and  females  separately,  models  will 
estimate  different  t0  values.  If  f0  is  truly  representative 
of  gestation  time,  then  it  leads  to  the  incorrect  infer- 
ence that  male  and  female  pups  have  different  gestation 


periods.  For  these  reasons,  growth  models  that  use  size 
at  birth  (L0)  instead  of  t0  may  be  more  appropriate  for 
elasmobranchs  (Cailliet  and  Goldman,  2004). 

The  purpose  of  this  study  was  to  estimate  best-fit 
growth  models  for  male  and  female  spiny  dogfish  in 
the  GOA.  Resultant  growth  equations  provide  critical 
parameters  for  a better  understanding  of  spiny  dogfish 
biology,  estimation  of  biological  reference  points  includ- 
ing indirect  estimates  of  M,  improved  stock  assess- 
ments, and  development  of  sound  fishery  management 
plans  for  this  species  in  waters  off  Alaska. 

Materials  and  methods 
Sample  collection 

Spiny  dogfish  were  collected  by  targeted  sampling 
cruises,  state  and  federal  assessment  surveys,  and  oppor- 
tunistic fishery  bycatch  samples  between  July  2004  and 
April  2007  across  the  GOA  (Fig.  1,  Table  1 (delete  bold 
font  after  placing  tables).  All  spiny  dogfish  were  sexed 
and  length  was  measured  to  the  nearest  centimeter 
(total  length  extended=T,Lej.,;  total  length  natural^TL^; 
precaudal  length=PCL;  and  fork  length=FL;  Tribuzio  et 
al.,  2009).  Here,  length  measurements  are  reported  as 
total  length  extended  (TLext).  The  posterior  dorsal  spine 
was  removed  and  stored  frozen  for  laboratory  analyses. 
In  the  laboratory,  spines  were  cleaned  by  thawing,  by 
boiling  briefly,  and  the  loose  tissue  was  scraped  free. 
Spines  were  allowed  to  dry  overnight  and  then  stored  in 
individual  paper  envelopes  for  subsequent  age  reading. 

Sampling  bias  was  examined  because  we  sampled 
with  multiple  gear  types  in  different  locations.  To  test 
for  potential  bias,  a chi-squared  (x2)  test  was  conducted 
to  test  for  statistically  significant  (P<0.05)  differences 
in  the  mean  length  at  age  by  sex  for  each  gear  (trawl, 
setnet,  longline,  rod  and  reel)  and  region  (Cook  Inlet, 
Prince  William  Sound,  Yakutat  Bay,  and  Gulf  of  Alas- 
ka). Statistically  significant  differences  among  different 
gears  would  provide  evidence  of  sampling  bias.  However, 
statistically  significant  differences  among  different  geo- 
graphic areas  would  provide  equivocal  evidence  of  bias 
because  the  possibility  of  true  underlying  differences  in 
size  distributions  by  area  could  not  be  dismissed. 

Age  determinations 

The  posterior  dorsal  spines  were  read  in  the  laboratory 
according  to  the  methods  of  Ketchen  (1975)  and  Beamish 
and  McFarlane  (1985).  Each  band  pair  (hereafter  termed 
“band”),  consisting  of  one  dark  and  one  light  band,  was 
counted  as  one  year  or  annulus  (Cailliet  et  al.,  2006). 
Aging  was  conducted  by  two  scientists  at  the  Washington 
Department  of  Fish  and  Wildlife’s  age  laboratory  and  by 
the  lead  author  at  the  University  of  Alaska  Fairbankans. 
Ease  of  age  reading  was  categorized  from  1 (easiest)  to  3 
(most  difficult).  Spines  were  photographed  on  a lxl  mm 
grid  to  standardize  measurements.  All  measurements 
were  rounded  to  the  nearest  0.01  mm  by  using  Bersoft 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


121 


Locations  where  spiny  dogfish  ( Squalus  acanthias ) were  sampled  in  the  Gulf 
of  Alaska  in  2004-07.  The  size  of  the  circle  is  proportional  to  the  number  of 
spiny  dogfish  sampled  at  each  location. 


Table  1 

Locations,  gear  types,  and  sample  sizes  for  male  and  female  spiny  dogfish  (Squalus  acanthias)  collected  during  2004-07.  “Sport” 
gear  refers  to  hook-and-line  fishing  with  rod  and  reel,  “longline”  refers  to  multiple  hooks  on  a groundline,  “trawl”  denotes  either 
bottom  or  pelagic  trawls,  and  “set  net”  refers  to  a stationary  floating  gill  net,  generally  anchored  at  one  end  to  the  shore. 


Year 

Area 

Gear 

Males  (n) 

Females  ( n ) 

2004 

Yakutat  Bay 

Sport 

21 

91 

2004 

Gulf  of  A laska  (GOA) 

Longline 

52 

85 

2005 

Southeast  Alaska  (SEAK) 

Longline 

1 

13 

2005 

Yakutat  Bay 

Longline 

11 

23 

2005 

Yakutat  Bay 

Sport 

0 

15 

2005 

Cook  Inlet 

Sport 

6 

25 

2005 

Yakutat  Bay 

Longline 

41 

95 

2005 

GOA 

Longline 

112 

204 

2005 

Cook  Inlet 

Sport 

8 

12 

2005 

Yakutat  Bay 

Sport 

1 

72 

2005 

Prince  William  Sound 

Longline 

27 

69 

2005 

GOA 

Trawl 

83 

125 

2006 

Kamishak  Bay 

Trawl 

24 

26 

2006 

Cook  Inlet 

Set  net 

50 

90 

2006 

Copper  River 

Set  net 

9 

5 

2006 

Yakutat  Bay 

Set  net 

4 

57 

2006 

Icy  Point  (SEAK) 

Trawl 

0 

1 

2006 

Prince  William  Sound 

Longline 

87 

91 

2006 

Cherikoff  Island  (SW  GOA) 

Trawl 

28 

13 

2007 

Cherikoff  Island  (SW  GOA) 

Trawl 

20 

16 

122 


Fishery  Bulletin  108(2) 


EBD 
— 


Figure  2 

Measurements  taken  on  spiny  dogfish  ( Squalus  acanthias)  spines. 
Last  readable  point  (LRP)  is  the  point  where  the  bands  are  no  longer 
visible  on  the  leading  edge  of  the  spine  (upper  edge  in  this  picture). 
EBD  = enamel  base  diameter,  SBD  = spine  base  diameter,  BL  = base 
length,  and  TL  = spine  total  length,  which  only  applies  to  spines  that 
are  unworn.  All  measurements  were  taken  in  millimeters. 


Image  Measurement  vers  5.0  software 
(Bersoft,  Inc.,  http://bersoft.com).  Mea- 
surements included  spine  base  diameter 
(SBD),  enamel  base  diameter  (EBD),  last 
readable  point  (LRP,  also  called  the  no- 
wear point);  and,  for  nonworn  spines,  base 
length  (BL),  and  spine  total  length  (TL, 

Fig.  2)  were  also  measured  to  the  near- 
est 0.01  mm.  Nonworn  spines  were  those 
spines  with  a LRP<2Ab  mm  (McFarlane 
and  King,  2009),  which  is  the  EBD  at 
birth. 

Aging  bias  and  precision  were  evalu- 
ated for  all  three  readers.  Pair-wise  age- 
bias  plots  were  used  to  compare  each 
reader  against  the  other  two  (Campana 
et  al.,  1995)  and  a %2  test  for  symmetry 
was  used  to  test  for  statistically  significant  systematic 
bias  among  the  three  readers  (Hoenig  et  al.,  1995). 
Readers  were  considered  to  be  in  agreement  when  ages 
were  within  10%  of  each  other  rather  than  within  some 
fixed  1-  or  2-year  age  interval.  For  instance,  if  reader 
X counted  10  bands,  then  reader  Y’s  count  would  have 
to  have  been  between  9-11  bands  to  be  in  agreement, 
but  if  reader  X counted  40  bands,  then  reader  Y’s  count 
would  have  to  be  between  36-44  to  be  in  agreement. 
We  contend  that  the  use  of  a percentage  to  define  the 
interval  size  is  more  appropriate  for  this  long-lived  spe- 
cies. Finally,  the  coefficient  of  variation  (CV)  between 
readers  was  calculated  according  to  Campana’s  methods 
(2001). 

Spiny  dogfish  ages  are  not  always  equal  to  the  num- 
ber of  counted  bands  for  two  reasons:  1)  bands  are  de- 
posited during  embryonic  development,  and  2)  because 
the  external  spines  can  become  worn  or  can  break  off. 
This  problem  was  addressed  by  a correction  method 
for  estimating  the  number  of  missing  bands  that  was 
based  on  a regression  of  band  counts  on  the  SBD  of 
unworn  spines  (Ketchen,  1975).  This  method  was  sub- 
sequently re-examined  and  accepted  as  the  best  avail- 
able method  for  the  original  samples  plus  additional 
samples  from  the  same  geographic  region  (McFarlane 
and  King,  2009). 

Various  regression  approaches  were  compared  to  de- 
termine which  method  resulted  in  the  best  model  for 
estimating  the  number  of  worn  bands  in  spiny  dog- 
fish collected  from  the  GOA,  including:  nonlinear  least 
squares  regression  (NLS,  Eq.  1),  and  ordinary  least 
squares  (OLS,  Eq.  2): 

Band  count  = b0EBDt>1  (1) 

In  (Band  count)  = ln(60)  + ln(EBD)61,  (2) 

where  b0  and  bx  are  estimated  parameters  (based  on 
Ketchen  1975,  McFarlane  and  King  2009).  Also,  we 
fitted  parameters  for  Equations.  1 and  2 with  weighted 
nonlinear  least  squares  (WNLS)  and  weighted  ordi- 
nary least  squares  (WOLS),  where  weights  were  applied 
to  the  residuals  as  follows:  spines  in  readability  cat- 


egory 1 were  given  a weight  of  1,  those  in  category  2 
were  weighted  by  0.5,  and  those  in  category  3 by  0.3. 
These  values  were  chosen  to  discount  the  contribution 
of  individual  length  at-age  data  points  to  the  estimation 
process  based  on  the  degree  of  uncertainty  in  the  age 
estimates  for  difficult-to-read  spines.  As  an  alternative 
to  this  weighting  scheme,  we  explored  the  weighting 
process  by  using  the  inverse  of  the  variance  in  assigned 
ages  for  each  readability  category.  Ages  of  worn  spines 
were  then  estimated  by  equating  the  LRP  to  the  EBD 
in  the  best-fit  model  from  Equations  1-4  and  by  adding 
the  resultant  number  of  bands  to  the  median  band  count 
from  the  three  readings  and  by  subtracting  two  years 
(for  bands  deposited  during  gestation)  to  obtain  the  final 
estimated  age  of  the  animal  (Ketchen,  1975).  In  the  case 
of  nonworn  spines,  age  was  estimated  by  the  median 
band  count  minus  two  years.  Data  for  males  and  females 
were  combined  for  these  worn  band  models. 

Fitting  of  growth  models 

A total  of  10  growth  model  variations  were  fitted  sepa- 
rately to  the  length-at-age  data  for  males  and  females 
(Table  2).  The  growth  models  included  1)  the  vB  growth 
model  for  estimating  /0;  2)  the  two-parameter  vB  with 
fixed  L0;  3)  the  two-phase  vB  with  L0  (used  in  the  present 
study);  4)  the  Gompertz;  5)  the  two-parameter  Gompertz; 
and  6)  the  logistic.  For  comparison  with  previous  studies 
L0  is  estimated  for  model  1 by  setting  /=0.  An  estimate 
of  L0  (i.e.,  the  size  at  birth)  for  GOA  spiny  dogfish  was 
not  available;  therefore  model  2 was  run  with  L0  fixed 
at  26.2  cm  (size  at  birth  for  spiny  dogfish  from  British 
Columbia;  Ketchen,  1972).  Models  3 and  5 were  run  in 
three  different  ways:  1)  L0  was  estimated  by  the  model; 

2)  with  L0  set  at  the  value  estimated  from  model  1;  and 

3)  with  L0  set  at  26.2  cm.  Model  3 is  an  adaptation  of 
the  two-phase  vB  model  (Soriano  et  al.,  1992).  Standard 
fitting  procedures  with  the  two-phase  model  resulted  in 
the  At  parameter  from  Soriano  et  al.  (1992)  changing  for 
a brief  time  period  and  then  returning  to  its  original 
value.  To  correct  this  we  reformulated  the  At  parameter 
from  Soriano  et  al.  (1992);  this  treatment  changes  k, 
depending  on  the  age  of  the  dogfish,  so  that  At  would 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


123 


Table  2 

Growth  models  fitted  to  spiny  dogfish  ( Squalus  acanthias)  length-at-age  (Lt)  data.  Parameters  are:  asymptotic  length  (La),  the 
growth  coefficient  (k),  length  at  birth  (L0),  age  at  size  zero  (f0),  a phase  change  parameter  (At)  for  the  two-phase  model,  age  at 
transition  (th),  magnitude  of  the  maximum  difference  between  model  1 and  the  two  phase  model  (h),  time  increment  from  previ- 
ous t value  (5),  and  the  inflection  point  of  the  logistic  curve  (a). 


Model  number 

Model  name 

Model  equation 

Reference 

1 

vB  1 

L, 

= Z,oo(l-e"t('"'o)) 

von  Bertalanffy  (1938) 

2 

vB  2 

L 

t 

ll 

i 

1 

TO 

a- 

Fabens (1965) 

3a-3c 

Two-phase  vB  with  L0 

L 

= A-s+b.-Aj*(  i-v4-'*""-'’). 

This  study 

1 + /ope«h-<) 


4 

Gompertz 

L 

t 

= Le  1 1 

Ricker  (1975) 

L 

5a-5c 

Two-parameter  Gompertz 

Lt 

= V ]>G  = 

In  — 
L 

Mollet  et  al.  (2002) 

L 

6 

Logistic 

L 

t 

, -k(t-a) 

l + e [ 1 

Ricker  (1979) 

follow  a logistic  pattern  and  remain  in  the  second  phase. 
Another  problem  we  encountered  fitting  the  two-phase 
model  was  that  the  typical  differential  form  of  the  vB 
equation  can  result  in  a decrease  in  length  at  the  tran- 
sition between  phases.  To  prevent  this  unlikely  result 
the  difference  equation  form  of  the  vB  equation  (Gulland 
1969)  was  used  in  this  analysis. 

Model  parameters  for  equations  describing  the  num- 
ber of  worn  bands  or  growth  were  fitted  by  nonlinear 
least-squares  regression  or  ordinary  least-squares  re- 
gression, and  confidence  intervals  were  estimated  by 
a bootstrap  procedure  with  5000  replicates  by  using  R 
statistical  software  (R,  vers.  2.10.0,  www.r-project.org). 
Confidence  intervals  (95%)  for  parameter  estimates 
were  based  on  the  lower  and  upper  2.5th  percentile  of 
the  bootstrap  replications.  Parameters  were  considered 
significantly  different  if  the  95%  confidence  intervals 
did  not  overlap.  To  evaluate  best  model  fit  for  the  male 
and  female  datasets,  Akaike  information  criteria  (AIC) 
and  model  summary  statistics  were  calculated  (Burn- 
ham and  Anderson,  2004). 


Fraser-Lee  back-calculation  methods  (Francis,  1990; 
Campana,  1990;  Goldman  et  al.,  2006).  The  Fraser-Lee 
method  produced  results  that  on  an  individual  level 
could  be  quite  unreasonable  (large  negative  ages),  but 
on  average  were  more  biologically  reasonable  than  either 
of  the  Dahl-Lea  methods.  Further,  growth  model  results 
with  either  of  the  Dahl-Lea  methods  were  unreasonable 
(Lx  of  >150  cm  TLext),  therefore,  we  used  the  Fraser-Lee 
method  for  our  data.  Thus,  the  following  equation  was 
used  to  estimate  back-calculated  length-at-age  data: 


TL,  = TL,  + 


[EBD,-EBDc)(TLc-TLbh 


birth  , 


EBDc  - EBDbirth 


(3) 


where  TLi  = the  back  calculated  length; 

TLC  = the  length  at  capture; 

TL  birth  = the  length  at  birth; 

EBDi  = the  enamel  base  diameter  at  band  /; 

EBDc  - the  enamel  base  diameter  at  capture;  and 
EBDhirth  = the  enamel  base  diameter  at  birth. 


Back-calculation  methods 

Owing  to  a paucity  of  specimens  with  EBD< 3.5  mm, 
back-calculation  methods  were  used  to  fill  in  the  size 
range  missing  from  samples.  The  spine  diameter  at  each 
band  along  the  spine  (hereafter  called  “band  diameters”) 
was  measured  from  a random  subsample  of  153  unworn 
spines  for  use  in  the  estimation  of  worn  bands  (Eqs.  1-4); 
spiny  dogfish  with  unworn  spines  tend  to  be  smaller  and 
younger  than  those  with  worn  spines.  We  examined  the 
Dahl-Lea,  linear  Dahl-Lea,  and  size  at  birth  modified 


Results 

Sample  collection 

A total  of  1713  spiny  dogfish  were  sampled  over  the  four 
years  of  the  study  (585  males,  1128  females,  Table  1)  of 
which  537  male  and  1062  female  spines  were  usable. 
Lengths  ranged  from  56  to  99  cm  TLext  for  males,  and 
56  to  123  cm  TLext  for  females.  The  x2  test  revealed  no 
significant  differences  between  the  mean  length  at  age 


124 


Fishery  Bulletin  108(2) 


45 
40  - 
35  - 
30  - 
25 
20  - 
15  - 
10  - 
5 
0 


d 


k 


0 


10  20  30 

Band  count  reader  1 


40 


Band  count  reader  1 


45 
40 
35 
30  - 
25  - 
20 
15 
10  - 
5 - 
0 


3 


0 


10  20  30 

Band  count  reader  3 


40 


Band  count  reader  3 


Figure  3 

A comparison  of  age  counts  among  readers.  (A)  Reader  2’s  mean  band  counts  (y-axis)  in 
relation  to  the  band  counts  of  reader  1;  (B)  Reader  3’s  mean  band  counts  in  relation  to 
the  band  counts  of  reader  1;  and  (C)  Reader  2’s  mean  band  counts  in  relation  to  the  band 
counts  of  reader  3.  Vertical  lines  are  95%  confidence  intervals  and  the  diagonal  line  is 
the  1:1  relationship  line.  (D)  Percent  agreement  and  coefficient  of  variation  for  reader  2 
(Rd  2)  compared  to  reader  1.  The  percent  agreement  (±10%)  is  represented  by  the  solid 
line  and  circles  and  the  coefficient  of  variation  (CV)  by  the  dashed  line  and  open  circles. 
(E)  Percent  agreement  and  coefficient  of  variation  of  reader  3 (Rd  3)  compared  to  reader 
1;  and  (F)  Percent  agreement  and  coefficient  of  variation  of  reader  2 (Rd  2)  compared  to 
those  of  reader  3. 


of  any  of  the  data  groupings  (P>0.99,  0.019<x2<4.525). 
Thus,  we  failed  to  find  evidence  of  sampling  bias  or 
geographic  differences  in  average  size  at  age. 

Age  determinations 

Sampled  dogfish  ranged  in  age  from  8 to  50  years  old. 
The  x2  test  and  the  age-bias  plots  indicated  no  signifi- 


cant systematic  bias  between  the  three  readers  (x2=241, 
206,  and  259  between  readers  2 and  1,  readers  2 and 
3,  and  readers  3 and  1,  respectively;  all  P>0.05;  Fig. 
3,  A-C).  The  percent  agreement  between  readers  2 
and  1 (Fig.  3D)  and  readers  3 and  1 (Fig.  3E)  was  high 
for  band  counts  less  than  30  but  was  more  variable 
or  decreased  for  band  counts  greater  than  30  (Fig.  3, 
D-F).  For  readers  2 and  3,  the  percent  agreement  was 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


125 


Table  3 

Summary  of  the  parameters  used  in  the  worn-band  estimation  models  and  model  fits  for  spiny  dogfish  (Squalus  acanthias).  The 
observed  data  are  sample  data,  the  band-diameter  data  were  determined  from  a subsample  of  unworn  spines  where  the  diam- 
eter of  each  band  was  measured  to  simulate  bound  count  at  spine  size  for  younger  animals  that  were  not  sampled  in  this  study. 
Regression  models  are  ordinary  least  squares  (OLS),  weighted  ordinary  least  squares  (WOLS),  nonlinear  least  squares  (NLS) 
and  weighted  nonlinear  least  squares  (WNLS).  Estimated  model  parameters  (95%  confidence  intervals  in  parentheses)  and 
goodness-of-fit  indicator  AIC,  the  Akaike  information  criteria. 


Model 

Parameter 

Observed  sample  data 
ti  = 685 

Observed  band-diameter  data 
71  = 3877 

Estimate 

AIC 

Estimate 

AIC 

OLS 

bo 

2.690(1.952-3.708) 

6.205 

0.211  (0.199-0.223) 

3.738 

K 

1.135  (0.949-1.322) 

2.867(2.825-2.910) 

WOLS 

\ 

2.471  (1.788-3.415) 

6.219 

0.212  (-0.201-0.224) 

3.721 

K 

1.179(0.991-1.367) 

2.856(2.814-2.898) 

NLS 

bo 

4.325  (3.400-5.444) 

4.016 

0.539(0.487-0.594) 

3.781 

b. 

0.955  (0.807-1.111) 

2.241  (2.178-2.309) 

WNLS 

bo 

4.009  (3.106-5.231) 

4.018 

0.528  (0.475-0.586) 

3.763 

K 

0.998(0.826-1.164) 

2.247  (2.180-2.318) 

more  variable  for  band  counts  less  than  20  (Fig.  3F). 
The  CV  between  all  three  readers  was  generally  low 
(<30%)  for  band  counts  less  than  30,  and  there  was 
a notable  increase  in  the  variability  and  CV  for  band 
counts  greater  than  30. 

Spiny  dogfish  spines  grow  in  a predictable  pattern 
with  age  (Fig.  4).  The  brownish-black  banded,  enameled 
portion  of  the  spine  grows  in  length  at  a faster  rate 
than  the  white  base  portion. 

Inclusion  of  the  back-calculated  band  diameter 
data  dramatically  changed  the  worn  band  estima- 
tion models  (Fig.  5),  and  therefore  further  worn  band 
estimations  were  made  with  both  the  observed  and 
back-calculated  band  diameter  data.  There  were  no 
significant  differences  between  the  estimated  worn- 
band  model  parameters,  but  the  WOLS  model  had 
the  lowest  AIC  value  and  therefore  was  chosen  as  the 
best-fit  model  (Table  3).  Alternative  fits  to  the  WOLS 
and  WLNS  models,  based  on  weightings  by  using  the 
inverse  variance  in  assigned  ages  for  each  readability 
category,  yielded  very  similar  parameter  values  and 
nominally  poorer  fits  indicated  by  slightly  larger  AIC 
values  (not  shown).  A high  degree  of  natural  varia- 
tion resulted  in  wide  95%  confidence  intervals  for  all 
parameters.  Moreover,  parameter  confidence  inter- 
vals for  the  WOLS  GOA  model  widely  overlapped  the 
parameter  confidence  intervals  for  the  Hecate  Strait 
and  Strait  of  Georgia  models  (McFarlane  and  King, 
2009).  Although  the  parameters  were  not  statistically 
significantly  different,  the  GOA,  Hecate  Strait,  and 
Strait  of  Georgia  models  appear  to  represent  biologi- 
cally meaningful  differences  in  growth  (Fig.  5).  The 
Hecate  Strait  and  Strait  of  Georgia  models  tend  to 
overestimate  the  band  count  for  larger  spines  and 
underestimate  for  smaller  spines  of  spiny  dogfish  col- 
lected from  the  GOA. 


0 10  20  30  40  50  60  70  80  90  100 

Size  class  (cm) 

Figure  4 

Relationship  between  mean  second  dorsal  spine 
length  and  fish  size  determined  from  unworn  spines 
from  spiny  dogfish  ( Squalus  acanthias)  collected 
in  the  Gulf  of  Alaska.  The  top  line  is  spine  total 
length  ( TL ) and  bottom  line  is  base  length  ( BL ) in 
millimeters.  Numbers  above  upper  line  represent  the 
sample  size  for  each  10-cm  size  class.  Solid  vertical 
lines  represent  95%  confidence  intervals.  The  dashed 
vertical  line  represents  the  approximate  size  at  birth 
(Ketchen,  1972). 


Fitting  of  growth  models 

The  two-phase  vB  models  fitted  the  observed  data  best 
for  males  and  females  based  on  AIC  values  (Fig.  6,  A 
and  D,  Tables  4 and  5).  For  males,  the  two-phase  model, 
where  L0  was  used  from  model  1 (model  3b),  was  the 
best  fit  and  for  females,  it  was  the  model  where  L0  was 
estimated  from  model  1 (model  3b).  Estimated  (and  95% 


126 


Fishery  Bulletin  108(2) 


~o 


Enamel-base  diameter  (EBD)  (mm) 

Figure  5 

Relationship  of  band  count  to  enamel-base  diameter  for  spiny 
dogfish  ( Squalus  acanthias ) collected  in  the  Gulf  of  Alaska  (GOA) 
between  2004  and  2007.  The  best-fit  model  (weighted  ordinary 
least  squares  [WOLS])  for  (A)  the  observed  data  only  and  (B) 
the  observed  data  with  the  band-diameter  data;  both  sections  A 
and  B show  the  published  best-fit  relationships  for  spiny  dogfish 
collected  from  Hecate  Strait  and  the  Strait  of  Georgia,  British 
Columbia  (McFarlane  and  King,  2009)  for  comparison. 


confidence  limits)  asymptotic  lengths  (LJ  were 
87.2  cm  (range  85.3-90.0  cm)  and  102.5  cm 
(range  99.9-106.3  cm)  and  growth  coefficients 
( k ) were  0.106  (range  0.097-0.117)  and  0.058 
(range  0.052-0.063)  for  males  and  females, 
respectively.  After  including  the  back-calculated 
data  and  the  mean  back-calculated  data,  the 
two  phase  models  were  no  longer  the  best  fit 
for  males.  The  best-fit  model  with  inclusion  of 
back-calculated  data  was  model  2,  and  model 
1 fitted  best  for  the  data  including  the  mean 
back-calculated  data.  Similarly,  for  females  the 
two-phase  models  were  not  the  best-fit  based 
on  AIC  values  after  the  inclusion  of  back-cal- 
culated and  mean  back-calculated  data:  model 
6 was  the  best  fit  with  inclusion  of  back-calcu- 
lated data,  and  model  5c  (with  L0  from  model 
1)  was  the  best  fit  for  the  data  including  the 
mean  back-calculated  data  (Tables  4 and  5,  Fig. 

6,  B,  C,  E,  F). 

Predicted  length  at  age  was  similar  for  males 
and  females  for  the  observed  data,  up  to  about 
age  15,  when  a transition  between  growth 
phases  occurred  (Fig.  6).  After  the  transition, 
females  continued  to  grow  at  a faster  rate  and 
to  larger  sizes  than  males  (Fig.  6,  A and  D).  At 
the  point  of  transition  in  the  two-phase  models 
growth  increased  for  about  five  years  before 
slowing,  for  both  sexes. 

Discussion 

The  model  fits  for  all  10  examined  growth 
models  were  similar  and  had  very  small  dif- 
ferences in  AIC,  but  the  estimated  parame- 
ters differed  substantially;  for  example,  the 
growth  coefficient  ( k ) was  significantly  different 
between  some  models  and  thus  could  impact 
estimates  of  natural  mortality  and  subsequent 
demographic  analyses.  The  values  of  k tended 
to  fall  into  two  groupings  (in  both  data  sets), 
and  those  models  that  estimated  the  higher  k 
were  also  those  that  estimated  lower  estimates 
for  Lx.  Interestingly,  even  with  the  significantly 
different  estimates  of  k,  these  estimates  were  still  at  the 
lower  range  of  reported  growth  rates  for  different  types 
of  shark  species  (Cailliet  and  Goldman,  2004). 

Cailliet  et  al.  (2006)  recommended  considering  more 
than  one  form  of  evaluation  of  model  performance  and 
considering  biological  interpretations  along  with  statis- 
tical fit  when  choosing  the  best  model.  Mean  squared 
error  and  the  correlation  coefficient  (r2)  were  also  cal- 
culated for  each  model,  but  determinations  of  best  fit 
by  the  above  criteria  did  not  differ  from  those  where 
AIC  was  used  and  therefore  are  not  reported.  For  the 
observed  data  models  3a  and  3b  were  the  statistical 
best  fit  for  males  and  females,  respectively.  However, 
the  two-phase  models  tended  to  be  unstable  and  would 
converge  at  different  localized  minima,  depending  on 


the  starting  value.  A further  consideration  for  the  two- 
phase  models  is  that  the  growth  curve  indicates  a pe- 
riod of  rapid  growth  immediately  following  the  age  at 
transition. 

The  purpose  of  a two-phase  model  is  to  incorporate 
changes  in  energy  allocation  as  animals  grow:  imma- 
ture fish  use  surplus  energy  for  growth,  whereas  ma- 
ture fish  use  surplus  energy  for  reproduction  (Soriano 
et  al.,  1992).  Thus,  the  rate  of  growth  changes  after 
maturation.  In  our  case,  the  transition  between  the  two 
growth  phases  occurred  before  the  age  at  50%  maturity 
for  both  males  and  females  The  early  age  at  transition 
and  the  period  of  rapid  growth  after  transition  indi- 
cate that  for  female  spiny  dogfish  there  is  a “growth 
spurt”  about  15  years  before  age  at  50%  maturity.  For 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


127 


Female 


Age  (years) 


Figure  6 

Model  fits  for  male  (A-C)  and  female  (D-F)  spiny  dogfish  ( Squalus  acanthias)  length-at-age  data. 
(A  and  D)  Best-fit  growth  models  based  on  the  observed  sample  data;  (B  and  E)  best-fit  growth 
models  based  on  the  observed  sample  data  and  the  back-calculated  data;  and  (D  and  F)  best-fit 
growth  models  based  on  the  observed  sample  data  and  the  mean  back-calculated  data.  nobs  is  the 
number  of  samples,  nback  is  the  number  of  data  points  created  through  back  calculation  of  the  ages 
from  band-diameter  data,  and  is  the  number  of  mean  back-calculated  data  points. 

’ mean  1 


males,  the  pattern  was  similar,  but  occurred  just  before 
age  at  50%  maturity.  This  finding  does  not  follow  the 
theory  behind  the  two-phase  model  and  indicates  that 
a two-phase  model  may  not  be  most  appropriate  in  this 
situation. 

The  two-phase  vB  model  by  Soriano  et  al.  (1992)  has 
been  examined  with  data  sets  from  many  species  of 
sharks  to  determine  whether  it  is  an  adequate  descrip- 
tor of  shark  growth  (Araya  and  Cubillos,  2006).  Where- 
as the  two-phase  model  was  better  than  the  standard 
vB  model  in  8 of  11  species  for  females  and  7 of  11  for 
males,  the  two-phase  model  did  not  perform  better  than 
the  vB  (model  1 here)  for  spiny  dogfish.  Because  Araya 
and  Cubillos  (2006)  included  only  one  spiny  dogfish 
population  (Black  Sea),  which  appears  to  have  different 


age  and  growth  characteristics  from  those  in  the  GOA, 
and  only  examined  average  length  at  age  data  (Avsar, 
2001),  we  felt  that  it  was  worth  while  to  investigate  the 
two-phase  family  of  models  in  this  study.  Braccini  et 
al.  (2007)  found  that  the  two-phase  model  was  the  best 
statistical  fit  for  the  piked  spurdog,  which  is  a species 
similar  to  spiny  dogfish;  however,  the  resultant  mod- 
els showed  some  of  the  same  characteristic  difficulties 
that  we  encountered.  Those  results  also  indicated  a de- 
crease in  length  after  transition  (Fig.  7,  Braccini  et  al. 
2007)  and  that  the  At  parameter  appears  to  change  only 
briefly  before  returning  to  its  original  value.  Braccini  et 
al.  did  not  address  these  issues  as  we  have  attempted 
here.  A more  comprehensive  examination,  which  would 
include  multiple  data  sets  from  different  regions  for 


128 


Fishery  Bulletin  108(2) 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


129 


130 


Fishery  Bulletin  108(2) 


0 10  20  30  40 

Age  (years) 

Figure  7 

Comparison  of  published  spiny  dogfish  ( Squalus  acanthias)  female  growth 
models  from  sources  listed  in  Table  4.  (A)  Growth  models  published  for  Pacific 
Ocean  spiny  dogfish:  “Alaska”  includes  the  Gulf  of  Alaska  (GOA)  model  from 
this  study  and  a Prince  William  Sound  (PWS)  model  (Vega,  2006);  “British 
Columbia  inshore”  includes  three  models  for  dogfish  sampled  within  the 
Strait  of  Georgia  and  Hecate  Strait  (Ketchen,  1975;  Saunders  and  McFar- 
lane,  1993);  “Puget  Sound  inshore”  covers  models  based  on  samples  collected 
within  the  Puget  Sound  area  south  off  British  Columbia  and  east  of  the 
Washington  coast  (Vega,  2006);  “Pacific  Coast  South”  includes  four  models 
based  on  samples  collected  off  Oregon  and  California  (Vega,  2006);  “Pacific 
Coast  North”  includes  models  based  on  samples  collected  off  of  Washington 
and  the  west  coast  of  Vancouver  Island  (Ketchen,  1975;  Jones  and  Geen, 
1977;  Vega,  2006);  (B)  The  growth  models  from  the  Atlantic  Ocean,  North 
Sea.  and  Black  Sea  (Holden  and  Meadows,  1962;  Sosinski  1978;  Nammack  et 
al.,  1985;  Fahy,  1989;  Avsar,  2001;  Henderson  et  al.,  2002;  Soldat  [footnote  1 
in  Table  6]).  Note  the  different  x-axis  scales. 


each  species,  and  a complete  sample 
of  the  size  range  may  lead  to  a more 
conclusive  determination  as  to  which 
species  exhibit  two-phase  growth. 

Disregarding  the  two-phase  mod- 
els, the  best-fit  model  was  model  2 
for  males  and  model  5c  for  females. 
In  this  situation,  given  the  lack  of 
data  and  difficulties  with  the  two- 
phase  models,  it  may  be  more  ap- 
propriate to  select  the  best  model 
not  based  on  the  AIC  criteria  alone, 
but  to  also  consider  the  biological 
soundness  of  the  models.  Model  2 
(males)  and  model  5c  (females)  are 
the  statistical  best  fit  of  the  more 
biologically  reasonable  models.  Both 
of  these  best-fit  models  require  L0  as 
an  input,  not  as  an  estimated  pa- 
rameter. The  lack  of  data  for  spiny 
dogfish  <50  cm  TLext  likely  causes 
the  models  that  estimate  L0  to  have 
difficulty  fitting  the  data  and  as  a 
result  estimate  L0  to  be  larger  than 
would  be  expected. 

In  the  majority  of  published  stud- 
ies on  spiny  dogfish  age  and  growth 
the  traditional  von  Bertalanffy 
model  is  used.  To  facilitate  a broad- 
er comparison  of  our  results  with 
growth  parameter  estimates  for  oth- 
er regions  of  the  geographic  distri- 
bution of  spiny  dogfish,  we  compared 
parameters  estimated  from  model  1 
(Table  4)  with  growth  curves  fitted 
by  using  the  traditional  vB  formula- 
tion, as  reported  in  published  stud- 
ies (Table  5,  Fig.  7).  Clear  differ- 
ences in  spiny  dogfish  growth  exist 
between  the  North  Pacific  and  North 
Atlantic  oceans.  For  instance,  we 
found  that  male  and  female  dogfish 
reach  larger  asymptotic  sizes  (87.2 
and  112.2  cm  TLext,  respectively)  in 
the  GOA  than  off  the  northeastern 
United  States  (82.5  and  100.5  cm 
TLgxt,  respectively;  Nammack  et  al., 
1985).  Indeed,  virtually  all  stud- 
ies have  found  large  differences  in 
growth  of  spiny  dogfish  between 
the  North  Pacific  and  North  Atlan- 
tic (Table  5,  Fig.  7).  Fish  from  the 
North  Atlantic  tend  to  grow  more 
rapidly,  achieve  smaller  asymptotic 
sizes,  and  have  shorter  life  spans 
than  those  from  the  Pacific.  Differ- 
ences in  growth  also  exist  within 
the  Pacific  (Table  5,  Fig.  7).  For  ex- 
ample, our  GOA  growth  estimates 
are  similar  to  those  for  spiny  dog- 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


131 


Table  6 

Summary  of  von  Bertalanffy  parameters  (model  1)  for  growth  models  for  female  spiny  dogfish  ( Squalus  acanthias ) from  the 
North  Pacific  and  North  Atlantic  oceans  and  the  North  and  Black  seas.  Parameters  are  asymptotic  length  (L„)  size  at  birth 
(L0),  growth  coefficient  (k),  and  the  theoretic  age-at-size  length  zero  (t0).  Here,  L0  was  solved  for  from  the  published  parameter 
estimates  for  the  purposes  of  comparison. 


Location 

k 

tQ 

L0 

Reference 

Alaska,  Gulf  of  Alaska 

121.4 

0.034 

-12.1 

40.9 

This  study 

Alaska,  Prince  William  Sound 

110.4 

0.038 

-11.6 

39.4 

Vega (2006) 

British  Columbia,  Hecate  Strait 

125.1 

0.031 

-10.6 

35.0 

Ketchen  ( 1975) 

British  Columbia,  Strait  of  Georgia 

129.1 

0.034 

-7.3 

28.4 

Ketchen ( 1975) 

British  Columbia,  Strait  of  Georgia 

114.9 

0.044 

-3.6 

16.8 

Saunders  and  McFarlane  (1993) 

British  Columbia,  offshore 

128.5 

0.036 

-6.9 

28.3 

Jones  and  Geen  (1977) 

U.S.,  inshore  (WA  north) 

113.5 

0.04 

-5.2 

21.3 

Vega (2006) 

U.S.,  inshore  (WA  south) 

100.4 

0.036 

-8.4 

26.2 

Vega (2006) 

U.S.,  offshore  (WA) 

123.6 

0.027 

-6.9 

21.0 

Vega (2006) 

U.S.,  offshore  (WA) 

152.9 

0.036 

-6.7 

32.8 

Ketchen  (1975) 

U.S.,  offshore  (OR) 

101.9 

0.027 

-12.7 

29.6 

Vega (2006) 

U.S.,  offshore  (OR  and  CA  combined) 

90.9 

0.031 

-13.0 

30.2 

Vega (2006) 

U.S.,  offshore  (CA  north) 

158.9 

0.009 

-25.3 

32.4 

Vega (2006) 

U.S.,  offshore  (CA  south) 

123.6 

0.027 

-6.9 

21.0 

Vega (2006) 

Northwest  Atlantic  (U.S.) 

100.5 

0.106 

-2.9 

26.6 

Nammack  et  al.  (1985) 

Northeast  Atlantic  (Ireland) 

98.8 

0.090 

-1.6 

13.3 

Fahy (1989) 

Northeast  Atlantic  (Ireland) 

112.0 

0.150 

-3.4 

44.7 

Henderson  et  al.  (2002) 

Northwest  Atlantic 

104.5 

0.095 

-3.7 

31.0 

Soldat2 

North  Sea 

137.1 

0.054 

-4.7 

30.7 

Sosinski  1978  (as  cited  in  Avsar,  2001) 

North  Sea 

101.4 

0.110 

-3.6 

33.2 

Holden  and  Meadows  (1962) 

Black  Sea 

145.0 

0.170 

-0.7 

16.3 

Avsar (2001) 

1 Soldat,  V.  T.  2002.  Spiny  dogfish  ( Squalus  acanthias  L.)  of  the  northwest  Atlantic  Ocean  (NWA).  NAFO  Sci.  Counc.  Res  Doc  02/84,  33  p. 


fish  from  offshore  Washington  State  waters  (Fig.  7)  but 
greater  than  those  caught  in  inshore  Washington  State 
waters  (Puget  Sound)  and  British  Columbia  (Ketchen, 
1975;  Jones  and  Geen,  1977;  Saunders  and  McFarlane, 
1993;  Vega,  2006).  The  age  and  growth  studies  from 
British  Columbia  were  conducted  on  spiny  dogfish  col- 
lected in  inshore  waters  (Strait  of  Georgia  and  Hecate 
Strait);  therefore  the  possibility  cannot  be  ruled  out 
that  spiny  dogfish  from  the  British  Columbia  offshore 
region  would  have  growth  estimates  similar  to  those 
of  Washington  offshore  and  GOA  spiny  dogfish.  The 
vB  growth  model  parameter  estimates  (Lx  and  k ) for 
northern  California  spiny  dogfish  (defined  as  spiny  dog- 
fish between  Point  Conception  to  the  Oregon  border; 
Vega,  2006)  were  radically  different  from  our  results 
for  the  GOA,  but  the  fits  for  California  may  have  been 
adversely  affected  by  small  sample  size. 

The  wide  variability  in  length-at-age  contributes  to 
the  lack  of  statistically  significant  differences  among 
growth  models  and  worn-band  estimation  models.  This 
variability  may  be  attributable  to  one  or  more  of  the 
following  factors:  measurement  error  in  either  length 
or  age  readings,  sampling  bias,  true  underlying  vari- 
ability in  growth  at  age,  and  misidentification  of  worn 
and  unworn  spines.  We  considered  the  potential  role  of 
each  of  these  factors. 


Measurement  error  in  the  length  measurements  alone 
is  insufficient  to  explain  the  relatively  large  variabil- 
ity in  the  size-at-age  data.  Aging  errors  may  take  two 
forms:  imprecision  and  bias.  We  found  no  bias  among 
the  three  readers  tested,  but  imprecision  of  the  band 
counts  among  readers  could  contribute  to  variability 
in  the  size-at-age  data,  especially  for  older  ages.  We 
used  the  median  band  count  (from  the  three  readers) 
to  account  for  reduced  precision  because  this  measure 
of  central  tendency  is  less  sensitive  to  outliers  than  the 
mean  for  small  sample  sizes  (Dudewicz  and  Mishra, 
1988).  A more  thorough  analysis  of  the  precision  of  age 
estimates  for  spiny  dogfish  in  the  Pacific  Ocean  revealed 
the  overall  coefficient  of  variation  for  aging  estimates 
among  four  laboratories  to  be  19%  (Rice  et  al.,  2009). 
Systematic  bias  was  found  for  two  of  the  laboratories 
(one  biased  high,  the  other  biased  low)  in  relation  to 
the  other  two,  but  relative  bias  did  not  always  result 
in  statistically  different  parameters  estimated  from  vB 
growth  curves  (Rice  et  al.,  2009). 

Age  validation  is  crucial  for  growth  studies  to  assure 
that  physical  structures  used  for  aging  are  correctly  in- 
terpreted. For  instance,  a systematic  aging  error  could 
result  if  the  periodicity  of  band  formation  is  not  an- 
nual. Annual  periodicity  of  band  deposition  on  second 
dorsal  spines  was  validated  for  spiny  dogfish  in  British 


132 


Fishery  Bulletin  108(2) 


Columbia  (Beamish  and  McFarlane,  1985;  McFarlane 
and  Beamish,  1987).  Moreover,  radioactive  carbon  iso- 
topes absorbed  into  spiny  dogfish  spines  provided  age 
estimates  that  agree  with  previous  aging  results  for  the 
British  Columbia  spiny  dogfish  (Campana  et  ah,  2006) 
and  verified  that  periodicity  is  annual,  even  at  old  ages 
(Campana,  2001).  We  assumed  that  this  annual  peri- 
odicity of  band  formation  in  spiny  dogfish,  which  was 
confirmed  for  this  species  in  British  Columbia,  also  ap- 
plies to  fish  from  the  GOA.  Because  spiny  dogfish  from 
British  Columbia  have  different  age  characteristics 
(e.g.,  worn  band  curves,  Fig.  5)  from  those  of  the  GOA, 
it  is  possible  that  the  pattern  of  band  deposition  may 
also  differ. 

Sampling  bias  was  considered  by  examining  potential 
differences  in  average  size  at  age  among  gear  type  and 
location  of  capture.  Because  there  were  no  significant 
differences  among  the  average  size  at  age  with  the 
different  gear  types  used  or  the  areas  sampled,  we  do 
not  believe  that  sampling  bias  was  a significant  factor 
affecting  our  results.  However,  the  lack  of  significant 
differences  in  our  study  should  not  be  misconstrued  to 
rule  out  considerations  of  sampling  bias  in  future  spiny 
dogfish  studies,  because  this  species  may  school  by  size 
and  sex  (Nammack  et  al.,  1985;  Ketchen,  1986). 

In  the  western  North  Atlantic  Ocean  commercial 
fisheries  target  the  largest  and  oldest  age  classes  (Rago 
et  al.,  1998).  Thus,  the  size-frequency  distributions 
determined  from  commercial  catches  may  not  be  repre- 
sentative of  the  full  size  range  of  fish  in  the  population. 
Moreover,  depletion  of  large  old  fish  from  the  population 
by  heavy  exploitation  means  that  subsequent  research 
surveys  may  not  catch  a representative  sample  of  the 
full  size  and  age  ranges  of  the  population.  In  the  GOA, 
spiny  dogfish  are  taken  as  bycatch  in  multiple  fisheries. 
In  some  cases,  dogfish  bycatch  is  largely  unaccounted 
for,  owing  to  the  lack  of  observers  on  small  (<60-ft)  ves- 
sels, such  as  those  vessels  with  salmon  gill  nets,  as  well 
as  some  longline  vessels  targeting  halibut  and  sablefish, 
resulting  in  an  unknown  level  of  total  fishing  mortal- 
ity (Courtney  et  al.,  2006).  However,  in  the  GOA,  it  is 
unlikely  that  the  fishing  mortality  has  truncated  the 
size  distribution  of  spiny  dogfish  because  spiny  dogfish 
are  not  targeted  and  recent  (2006)  estimates  of  spiny 
dogfish  biomass  are  80-100%  of  the  estimated  theoreti- 
cal population  carrying  capacity  (Rice,  2007).  Therefore, 
it  is  unlikely  that  the  fishery  has  created  size-selective 
impacts  that  would  lead  to  erroneous  selection  of  the 
two-phase  models  as  the  best-fit  models  (Braccini  et 
al.,  2007). 

One  limitation  of  our  size-frequency  distributions 
is  the  absence  of  spiny  dogfish  smaller  than  50  cm 
TLext.  The  lack  of  samples  from  smaller  spiny  dogfish  is 
likely  due  to  fishery-dependent  opportunistic  sampling 
which  apparently  occurs  in  areas  devoid  of  juvenile 
spiny  dogfish.  Examination  of  NMFS  spring  and  fall 
trawl  surveys  along  the  U.S.  east  coast  revealed  that 
in  spring  most  juveniles  were  caught  in  water  between 
50  and  150  m deep  (range:  7-390  m)  in  offshore  waters 
from  North  Carolina  to  the  eastern  edge  of  Georges 


Bank,  whereas  in  fall  most  were  caught  between  25 
and  75  m (range:  12-366  m)  in  various  locations,  such 
as  on  Georges  Bank,  Nantucket  Shoals,  and  throughout 
the  Gulf  of  Maine  (McMillan  and  Morse,  1999).  Spiny 
dogfish  smaller  than  50cm  TLext  have  been  surveyed  in 
both  Puget  Sound,  Washington  (Tribuzio  et  al.,  2009), 
and  in  the  northern  Strait  of  Georgia  (McFarlane  et 
al.,  2006)  by  using  bottom  trawl  gear.  In  this  study, 
we  made  numerous  unsuccessful  attempts  to  capture 
juvenile  dogfish  smaller  than  50  cm  TLext  in  the  GOA 
using  sport  and  longline  gear  in  Yakutat  Bay,  long- 
line  gear  with  small  (10/0  circle)  hooks  in  Southeast 
Alaska  (K.  Munk,  personal  commun.1),  and  commercial 
bottom  trawls  off  Kodiak  Island  (J.  Gauvin,  personal 
commun.2). 

A missing  size  group,  such  as  small  dogfish  in  our 
case,  may  cause  growth  models  to  overestimate  t0  or  L0, 
thus  decreasing  the  k estimate.  Further,  this  missing 
size  group  may  have  caused  the  age  of  transition,  th,  in 
the  two-phase  models  to  be  underestimated.  Also,  the 
lack  of  small  animals  may  have  limited  our  ability  to 
discriminate  among  competing  growth  models.  We  used 
band-diameter  data  and  back-calculated  lengths  derived 
from  unworn  spines  to  attempt  to  address  this  data 
gap.  The  inclusion  of  the  band-diameter  data  greatly 
improved  the  worn-band  estimation  models,  but  mini- 
mally changed  the  growth  models.  Few  of  the  estimated 
growth  model  parameters  based  on  the  back-calculated 
and  mean  back-calculated  data  were  significantly  dif- 
ferent from  those  estimated  from  the  observed  data 
alone. 

Back-calculation  methods  are  designed  to  be  used 
when  sample  sizes  are  small  or  if  sampling  has  not  oc- 
curred each  month  (Cailliet  and  Goldman,  2004),  but 
in  this  case  it  was  the  entire  smaller  end  of  the  size 
range  that  was  being  estimated.  With  the  modified 
Fraser-Lee  size-at-birth  method,  we  had  to  assume  that 
average  size  at  birth  was  known.  We  use  26.2  cm,  which 
is  based  on  data  collected  from  spiny  dogfish  inside  the 
Strait  of  Georgia,  British  Columbia  (Ketchen  1972). 
Sizes  at  birth  are  reportedly  similar  for  the  species 
across  the  northern  hemisphere,  with  ranges  of  23-30 
cm  (Ketchen  1972,  Tribuzio  et  al.  2009).  We  also  as- 
sumed that  2.45  mm  was  the  spine  diameter  at  birth, 
based  on  studies  of  British  Columbia  spiny  dogfish  (Mc- 
Farlane and  King  2009).  Because  this  is  an  average  as 
well,  it  is  likely  that  some  spines  were  classified  as  “un- 
worn” when  they  were  actually  “worn.”  Spines  that  are 
classified  as  “unworn”  can  lead  to  underestimating  the 
age,  and  in  the  case  of  the  back-calculation  resulted  in 
instances  where  20  cm  or  more  of  growth  was  predicted 
in  the  first  year.  Back-calculations  may  not  be  appropri- 
ate for  this  species  when  dorsal  fin  spines  are  used  as 
aging  structures,  but  may  work  well  if  a structure  such 
as  vertebrae  are  used. 


1 Munk,  Kristen.  2007.  Alaska  Department  of  Fish  and 
Game,  Juneau,  AK,  99801. 

2 Gauvin,  John.  2007.  Gauvin  and  Associates,  LLC.  Burien, 
WA  98166. 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


133 


The  relatively  large  variability  in  size  at  age  of  spiny 
dogfish  in  the  GOA  could  also  reflect  true  underlying 
variability  in  growth  rates.  Individuals  experiencing 
different  thermal  and  feeding  histories  are  expected  to 
have  different  growth  characteristics.  It  is  also  conceiv- 
able that  our  samples  represent  collections  of  dogfish 
from  multiple,  mixed  populations.  For  instance,  4 of 
2940  recoveries  (0.14%)  of  spiny  dogfish  tagged  in  Brit- 
ish Columbia  were  recovered  in  Alaska  (McFarlane  and 
King,  2003).  Because  the  movements  of  spiny  dogfish 
from  other  areas  to  and  from  Alaska  are  unknown, 
the  degree  of  mixing  is  uncertain.  However,  there  is 
no  evidence  of  genetic  differentiation  in  the  Northeast 
Pacific  based  on  analyses  of  eight  microsatellite  loci 
from  dogfish  sampled  from  the  Bering  Sea,  the  Gulf  of 
Alaska,  Strait  of  Georgia,  Puget  Sound,  and  the  coasts 
of  Washington,  Oregon,  and  California  (Hauser,  2009). 
Mixtures  of  spiny  dogfish  from  other  areas  with  growth 
characteristics  that  are  different  from  those  of  Alaska 
resident  dogfish  could  contribute  to  the  variability  in 
size  at  age  that  we  observed  in  the  GOA.  Nevertheless, 
the  existence  of  a statistically  significant  difference  in 
growth  rates  from  different  areas  of  the  Northeast  Pa- 
cific (Vega,  2006;  Table  4 this  document)  indicates  that 
mixing  is  incomplete. 

Our  findings  have  at  least  two  important  implications 
for  management  of  the  species.  First,  for  estimation  of 
stock  productivity  and  biological  reference  points  for 
spiny  dogfish  in  the  GOA,  it  is  important  to  use  growth 
curves  that  are  fitted  to  size-at-age  data  from  dogfish 
captured  in  the  GOA.  Although  alternative  growth  mod- 
el parameters  were  not  statistically  significantly  differ- 
ent from  one  another  in  our  study,  the  variation  among 
predicted  length  may  be  of  biological  significance.  For 
instance,  the  worn-band  estimation  curves  for  the  GOA 
and  British  Columbia  resulted  in  very  different  esti- 
mates of  ages  (Fig.  5);  use  of  growth  curves  for  British 
Columbia  would  result  in  estimated  numbers  of  worn 
bands  from  dogfish  spines  in  the  GOA  that  would  be 
biased  low.  For  example,  for  a spiny  dogfish  with  a 
1.8-mm  EBD,  the  GOA  model  would  estimate  an  age 
of  one  year,  whereas  both  of  the  British  Columbia  mod- 
els would  estimate  an  age  of  four  years.  A fish  with  a 
6-mm  EBD  would  be  estimated  to  be  30  years  old  by  the 
GOA  model  and  46  and  37  years  old  by  the  two  British 
Columbia  models.  Such  biases  in  growth  estimates  may 
lead  to  biases  in  estimates  of  biological  reference  points 
for  fishery  management. 

Second,  as  in  other  portions  of  their  range,  the  largest 
spiny  dogfish  are  the  oldest  females.  Because  commer- 
cial fisheries  for  spiny  dogfish  select  for  the  largest  in- 
dividuals, fishing  mortality  rates  are  disproportionately 
higher  for  this  reproductive  segment  of  the  population. 
In  the  Northwest  Atlantic  Ocean,  a sharp  increase  in 
landings  during  1987-1993  led  to  a fivefold  increase  in 
fishing  mortality  rates  on  females  from  0.016  to  0.26; 
and  fishing  mortality  rates  exceeding  0.10  on  large 
(a80-cm)  females  resulted  in  negative  pup  replace- 
ment, subsequently  leading  to  stock  decline  (Rago  et 
al.,  1998).  Thus,  to  sustain  spiny  dogfish  in  the  GOA, 


fishery  management  plans  should  consider  not  only 
slow  growth  rates,  low  fecundity,  and  late  maturation 
of  this  species  (King  and  McFarlane,  2003),  but  also 
the  potentially  disproportionate  number  of  removals  of 
mature  females  from  the  stock  by  commercial  fishing  by 
estimating  size-  and  sex-specific  fishing  mortality  rates 
and  biological  reference  points. 

Future  research  should  address  the  many  uncer- 
tainties remaining  about  spiny  dogfish  biology  and 
life  history  in  Alaska.  In  particular,  results  from  this 
study  indicate  several  areas  of  research  needed  to 
improve  our  understanding  of  spiny  dogfish  age  and 
growth.  First,  although  demonstrated  for  fish  captured 
off  British  Columbia  (Beamish  and  McFarlane,  1985; 
McFarlane  and  Beamish,  1987;  Campana  et  al.,  2006), 
validation  of  annual  band  formation,  as  well  as  worn- 
band  properties,  for  spiny  dogfish  collected  from  the 
GOA  should  be  conducted  to  describe  potential  sources 
of  bias  in  the  age  estimates  for  spiny  dogfish  at  this 
northern  portion  of  their  range  in  the  Pacific  Ocean. 
Second,  the  collection  of  juvenile  dogfish  (<50  cm)  is 
needed  to  provide  more  precise  estimates  of  growth 
over  their  full  life  history,  as  well  as  to  help  identify 
statistically  best-fit  growth  models.  Third,  tagging 
studies,  such  as  those  conducted  in  British  Columbia 
(King  and  McFarlane,  2003),  would  help  elucidate  the 
degree  to  which  dogfish  in  Alaska  represent  mixed 
stocks  with  different  growth  attributes;  such  tagging 
results  would  help  to  delineate  stock  boundaries  essen- 
tial for  fishery  management.  Fourth,  controlled  experi- 
ments are  necessary  to  fully  examine  the  selectivity  of 
various  fishing  gears  for  spiny  dogfish  by  size  and  sex. 
This  would  be  an  important  preliminary  step  toward 
gear  standardization,  if  long-term  sampling  programs 
are  envisioned  for  spiny  dogfish.  Finally,  continued 
sampling  of  spiny  dogfish  over  small  regional  scales 
is  necessary  to  fully  evaluate  potential  geographic 
differences  in  growth  and  resultant  parameters  (i.e., 
natural  mortality)  within  the  GOA,  as  well  as  to  more 
broadly  understand  the  life  history  of  this  species  in 
this  portion  of  its  range.  Although  our  study  would 
not  have  been  possible  without  the  diversity  of  low- 
cost  sampling  opportunities  afforded  to  us,  including 
the  valuable  assistance  of  state  and  federal  agencies 
and  sport  and  commercial  fishermen,  further  progress 
will  be  accelerated  by  a full-scale,  directed  field  pro- 
gram, which  would  be  more  successful  at  providing 
an  unbiased  sample  set  of  spiny  dogfish  in  the  waters 
off  Alaska,  and  which  would  aid  in  efforts  to  build  a 
more  detailed  stock  assessment,  and  thus  models  of 
population  dynamics. 

Acknowledgments 

We  are  grateful  for  funding  of  this  research  by  the 
North  Pacific  Research  Board  (NPRB  publication  no. 
227),  the  Rasmuson  Fisheries  Research  Center,  and  the 
Alaska  Fisheries  Science  Center’s  Population  Dynamics 
Fellowship  through  the  Cooperative  Institute  for  Arctic 


134 


Fishery  Bulletin  108(2) 


Research  (CIFAR).  We  thank  V.  Gallucci,  J.  Rice,  A. 
Andrews,  and  W.  Strasberger  for  field  and  laboratory 
assistance,  and  G.  Bargmann,  S.  Rosen,  and  J.  Topping 
at  the  Washington  Department  of  Fish  and  Wildlife 
for  reading  spines  and  training.  We  acknowledge  the 
National  Marine  Fisheries  Service;  Alaska  Department 
of  Fish  and  Game;  chartered  vessels  and  crew  of  the 
FVs  Kingfisher , Winter  King,  and  Sea  View,  commercial 
fishermen  in  Yakutat,  Cordova,  and  Kasilof;  Gauvin 
and  Associates,  LLC.,  and  Alaska  Pacific  and  Trident 
Seafoods  for  kindly  providing  sampling  opportunities. 
Finally,  we  are  grateful  to  T.  Quinn  II  and  K.  Goldman 
for  considerable  helpful  analytical  advice. 


Literature  cited 

Araya,  M.,  and  L.  A.  Cubillos 

2006.  Evidence  of  two-phase  growth  in  elasmobranchs.  En- 
viron. Biol.  Fishes  77:293-300. 

Avsar,  D. 

2001.  Age,  growth,  reproduction  and  feeding  of  the 
spurdog  ( Squalus  acanthias  Linnaeus,  1758)  in  the 
southeastern  Black  Sea.  Estuar.  Coast.  Shelf  Sci.  52: 
269-278. 

Beamish,  R.  J.,  and  G.  A.  McFarlane. 

1985.  Annulus  development  on  the  second  dorsal  spine 
of  the  spiny  dogfish  ( Squalus  acanthias)  and  its  valid- 
ity for  age  determination.  Can.  J.  Fish.  Aquat.  Sci. 
42:1799-1805. 

Beverton,  R.  J.  H.,  and  S.  J.  Holt. 

1957.  On  the  dynamics  of  exploited  fish  populations.  U.K 
Ministry  of  Agriculture  and  Fisheries,  Fisheries  Inves- 
tigations 2,  533  p. 

Boldt,  J. 

2003.  Ecosystem  considerations  for  2004.  In  Stock 
assessment  and  fishery  evaluation  report  for  the  ground- 
fish  resources  of  the  Bering  Sea/Aleutian  Islands  and 
Gulf  of  Alaska  region,  335  p.  [Available  from  North 
Pacific  Fishery  Management  Council,  605  W.  4th  Ave., 
Suite  306,  Anchorage,  AK  99501.] 

Bonfil,  R. 

2005.  The  purpose  of  stock  assessment  and  the  objectives 
of  fisheries  management.  In  Management  techniques 
for  elasmobranch  fisheries  ( J.  A.  Musick,  and  R.  Bonfil, 
eds.),  p.  6-14.  FAO  Fisheries  Tech.  Paper  474. 

Braccini,  J.  M.,  B.  M.  Gillanders,  and  T.  I.  Walker. 

2007.  Comparison  of  deterministic  growth  models  fitted 
to  length-at-age  of  the  piked  spurdog  ( Squalus  mega- 
lops) in  south-eastern  Australia.  Mar.  Freshw.  Res. 
58:24-33. 

Burnham,  K.  P.,  and  D.  R.  Anderson. 

2004.  Multimodel  inference:  understanding  AIC  and  BIC 
in  model  selection.  Soc.  Meth.  Res.  33:261—304. 

Cailliet,  G.  M.,  and  K.  J.  Goldman. 

2004.  Age  determination  and  validation  in  chondrich- 
thyan  fishes.  In  The  biology  of  sharks  and  their  rela- 
tives (J.  F.  Carrier,  J.  A.  Musick,  and  M.  R.  Heithaus, 
eds.),  p.  399-447.  CRC  Press,  Boca  Raton,  FL. 

Cailliet,  G.  M.,  W.  D.  Smith,  H.  F.  Mollet,  and  K.  J.  Goldman. 

2006.  Age  and  growth  studies  of  Chondrichthyan  fishes: 
the  need  for  consistency  in  terminology,  verification, 
validation,  and  growth  function  fitting.  Environ.  Biol. 
Fishes  77:211-228. 


Campana,  S.  E. 

1990.  How  reliable  are  growth  back-calculations  based 
on  otoliths?  Can.  J.  Fish.  Aquat.  Sci.  47:2219-2227. 

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. 

Campana,  S.  E.,  C.  Annand,  and  J.  I.  McMillan. 

1995.  Graphical  and  statistical  methods  for  determining 
the  consistency  of  age  determinations.  Trans.  Am. 
Fish.  Soc.  124:131-138. 

Campana,  S.,  C.  Jones,  G.  A.  McFarlane,  and  S.  Myklevoll. 

2006.  Bomb  dating  and  age  validation  using  the  spines 
of  spiny  dogfish  ( Squalus  acanthias).  Environ.  Biol. 
Fishes  77:327-336. 

Carlson,  J.  K.,  and  I.  E.  Baremore. 

2005.  Growth  dynamics  of  the  spinner  shark  (Carcha- 
rhinus  brevipinna)  off  the  United  States  southeast  and 
Gulf  of  Mexico  coasts:  a comparison  of  methods.  Fish. 
Bull.  103:280_291. 

Compagno,  L.  J.  V. 

1984.  Sharks  of  the  world:  an  annotated  and  illustrated 
catalogue  of  shark  species  known  to  date,  pt.  2:  Car- 
charhiniformes,  675  p.  FAO  Fish.  Synop.  125  (vol. 
4).  FAO,  Rome. 

Courtney,  D.,  C.  A.  Tribuzio,  K.  J.  Goldman,  and  J.  Rice. 

2006.  GOA  sharks.  In  Stock  assessment  and  fishery 
evaluation  report  for  the  groundfish  resources  of  the 
Gulf  of  Alaska  for  2007,  Appendix  E,  p.  481-561.  [Avail- 
able from  North  Pacific  Fishery  Management  Council, 
605  W.  4th  Ave.,  Suite  306,  Anchorage,  AK  99501.] 

Driggers,  W.  B.,  J.  K.  Carlson,  D.  Oakley,  G.  Ulrich,  B.  Cullum, 
and  J.  M.  Dean. 

2004.  Age  and  growth  of  the  blacknose  shark,  Carcharhi- 
nus  acronotus,  in  the  western  North  Atlantic  Ocean  with 
comments  on  regional  variation  in  growth.  Environ. 
Biol.  Fishes  71:171-178. 

Dudewicz,  E.  J.,  and  S.  A.  Mishra. 

1988.  Modern  mathematical  statistics,  864  p.  John 
Wiley  & Sons,  Inc.,  New  York,  NY. 

Fabens,  A.  J. 

1965.  Properties  and  fitting  of  the  von  Bertalanffy  growth 
curve.  Growth  29:265-289. 

Fahy,  E. 

1989.  The  spurdog  Squalus  acanthias  (L)  fishery  in 
south  west  Ireland.  Irish  Fish.  Invest.  Ser.  B (Mar.) 
32,  22  p. 

Francis,  R.  I.  C.  C. 

1990.  Back-calculation  of  fish  length:  a critical  review. 
J.  Fish  Biol.  36:883-902. 

Goldman,  K.  J.,  S.  Branstetter,  and  J.  A.  Musick. 

2006.  A re-examination  of  the  age  and  growth  of  sand 
tiger  sharks,  Carcharias  Taurus,  in  the  western  North 
Atlantic:  the  importance  of  ageing  protocols  and  use  of 
multiple  back-calculation  techniques.  Environ.  Biol. 
Fishes  77:241-252. 

Gulland,  J.  A. 

1969.  Manual  of  methods  for  fish  stock  assessment:  part 
1,  Fish  population  analysis.  FAO  manuals  in  fisheries 
science,  no.  4,  154  p.  FAO  Rome. 

Hauser,  L. 

2009.  The  molecular  ecology  of  dogfish  sharks  (Squalus 
acanthias) . In  Biology  and  management  of  dogfish 
sharks  (V.  F.  Gallucci,  G.  McFarlane,  G.  Bargmann, 
eds.),  p.  229-252.  Am.  Fish.  Soc.,  Bethesda,  MD. 


Tribuzio  et  al.:  Age  and  growth  of  Squalus  acanthias  in  the  Gulf  of  Alaska 


135 


Henderson,  A.  C.,  K.  Flannery,  and  J.  Dunne. 

2002.  Growth  and  reproduction  in  the  spiny  dogfish  Squa- 
lus acanthias  L.  (Elasmobranchii:  Squalidae),  from  the 
west  coast  of  Ireland.  Sarsia  87:350-361. 

Hoenig,  J.  M.,  M.  J.  Morgan,  and  C.  A.  Brown. 

1995.  Analyzing  differences  between  two  age  determi- 
nation methods  by  tests  of  symmetry.  Can.  J.  Fish. 
Aquat.  Sci.  52:364-368. 

Holden,  M.  J. 

1974.  Problems  in  the  rational  exploitation  of  elasmo- 
branch  populations  and  some  suggested  solutions.  In 
Sea  fisheries  research  (E.  H.  Jones,  ed.),  p.  187-215. 
Logos,  London. 

1977.  Elasmobranchs.  In  Fish  population  dynamics  (J. 
A.  Gulland,  ed.),  p.  187-215.  Wiley,  London. 

Holden,  M.  J.,  and  P.  S.  Meadows. 

1962.  The  structure  of  the  spine  of  the  spur  dogfish  ( Squa- 
lus acanthias  L.)  and  its  use  for  age  determination.  J. 
Mar.  Biol.  Assoc.  U K.  42:179-197. 

Jones,  B.  C.,  and  G.  H.  Geen. 

1977.  Age  and  growth  of  spiny  dogfish  ( Squalus  acanth- 
ias) in  the  Strait  of  Georgia,  British  Columbia.  Fish. 
Mar.  Serv.  Res.  Dev.,  Tech.  Rep.  699,  16  p. 

Ketchen,  K.  S. 

1972.  Size  at  maturity,  fecundity,  and  embryonic  growth  of 
the  spiny  dogfish  (Squalus  acanthias ) in  British  Colum- 
bia waters.  J.  Fish.  Res.  Board  Can.  29:1717-1723. 

1975.  Age  and  growth  of  dogfish  Squalus  acanthias  in 
British  Columbia  waters.  J.  Fish.  Res.  Board  Can. 
32:43-59. 

1986.  The  spiny  dogfish  ( Squalus  acanthias)  in  the  north- 
east Pacific  and  a history  of  its  utilization.  Can.  Spec. 
Publ.  Fish.  Aquat.  Sci.  88,  78  p. 

Ketchen,  K.S.,  N.  Bourne,  and  T.  H.  Butler. 

1983.  History  and  present  status  of  fisheries  for  marine 
fishes  and  invertebrates  in  the  Strait  of  Georgia,  British 
Columbia.  Can.  J.  Fish.  Aquat.  Sci.  48:1095-1119. 

King,  J.  R.,  and  G.  A.  McFarlane. 

2003.  Marine  fish  life  history  strategies:  applications 
to  fisheries  management.  Fish.  Manag.  Ecol.  10:249- 
264. 

McFarlane,  G.  A.,  and  R.  J.  Beamish. 

1987.  Validation  of  the  dorsal  spine  method  of  age  deter- 
mination for  spiny  dogfish.  In  Age  and  growth  in  fish 
(R.  C.  Summerfelt,  G.  E.  Hall,  eds.),  p.  287-300.  Iowa 
State  Univ.,  Ames,  IA. 

McFarlane,  G.  A.,  and  J.  R.  King. 

2006.  Migration  patterns  of  spiny  dogfish  ( Squalus 
acanthias)  in  the  North  Pacific  Ocean.  Fish.  Bull. 
101:358-367. 

2009.  Re-evaluating  the  age  determination  of  spiny  dog- 
fish (Squalus  acanthias)  using  oxytetracycline  and  fish  at 
liberty  up  to  twenty  years.  In  Biology  and  management 
of  dogfish  sharks  (V.  F.  Gallucci,  G.  McFarlane,  G.  Barg- 
mann,  eds.),  p.  77-88.  Am.  Fish.  Soc.,  Bethesda,  MD. 

McFarlane,  G.  A.,  J.  R.  King,  and  V.  R.  Hodes. 

2006.  Biological  results  of  the  Strait  of  Georgia  spiny 
dogfish  (Squalus  acanthias)  longline  survey  October 
18-31,  2005.  Can.  Data  Rep.  Fish.  Aquat.  Sci.  1182, 
24  p. 

McMillan,  D.  G.,  and  W.  W.  Morse. 

1999.  Essential  fish  habitat  source  document:  Spiny 
dogfish,  Squalus  acanthias,  life  history  and  habitat 
characteristics.  NOAA  Tech.  Memo.  NMFS-NE-150, 19  p. 


Mollet,  H.  F.,  J.  M.  Ezcurra,  and  J.  B.  O’Sullivan. 

2002.  Captive  biology  of  the  pelagic  stingray,  Dasyatis 
violacea  (Bonaparte,  1832).  Mar.  Freshw.  Res.  53:531- 
541. 

Nammack,  M.  F.,  J.  A.  Musick,  and  J.  A.  Colvocoresses. 

1985.  Life  history  of  spiny  dogfish  off  the  Northeastern 
United  States.  Trans.  Am.  Fish.  Soc.  114:367—376. 

Rago,  P.  J.,  K.  A.  Sosebee,  J.  K.  T.  Brodziak,  S.  A.  Murawski,  and 
E.  D.  Anderson. 

1998.  Implications  of  recent  increases  in  catches  on  the 
dynamics  of  Northwest  Atlantic  spiny  dogfish  ( Squalus 
acanthias).  Fish.  Res.  39:165-181. 

Rice,  J. 

2007.  Population  dynamics  of  spiny  dogfish  (Squalus 
acanthias)  in  the  Gulf  of  Alaska  with  an  emphasis  on 
the  analysis  of  bycatch  data.  M.S.  thesis,  109  p.  Univ. 
Washington,  Seattle. 

Rice,  J.,  V.  F.  Gallucci,  and  G.  H.  Kruse. 

2009.  Evaluation  of  the  precision  of  age  estimates  for 
spiny  dogfish.  In  Biology  and  management  of  dogfish 
sharks  (V.  F.  Gallucci,  G.  McFarlane,  G.  Bargmann, 
eds.),  p.  161-168.  Am.  Fish.  Soc,,  Bethesda,  MD. 

Ricker,  W.  E. 

1975.  Computation  and  interpretation  of  biological  sta- 
tistics of  fish  populations.  Bull.  Fish.  Res.  Board  Can. 
191:1-382. 

1979.  Growth  rates  and  models.  In  Fish  physiology,  vol. 
IX  (W.  S.  Hoar,  D.  J.  Randall,  and  J.  R.  Brett,  eds.),  p. 
677-743.  Academic  Press,  New  York. 

Saunders,  M.  W.,  and  G.  A.  McFarlane. 

1993.  Age  and  length  at  maturity  of  the  female  spiny 
dogfish,  Squalus  acanthias,  in  the  Strait  of  Georgia, 
British  Columbia,  Canada.  Environ.  Biol.  Fishes 
38:49-57. 

Soriano,  M.,  J.  Moreau,  J.  M.  Hoenig,  and  D.  Pauly. 

1992.  New  functions  for  the  analysis  of  two-phase  growth 
of  juvenile  and  adult  fishes,  with  application  to  Nile 
perch.  Trans.  Am.  Fish.  Soc.  121:486-493. 

Sosinski,  J. 

1978.  Characteristics  of  the  North  Sea  spurdog  (Squalus 
acanthias  L.)  stock.  Fish.  Biol.  8:9-22. 

Tribuzio,  C.  A.,  V.  F.  Gallucci,  G.  Bargman. 

2009.  Reproductive  biology  and  management  implications 
for  spiny  dogfish  (Squalus  acanthias)  in  Puget  Sound, 
WA.  In  Biology  and  management  of  dogfish  sharks 
(V.  F.  Gallucci,  G.  McFarlane,  G.  Bargmann,  eds.),  p. 
181-194.  Am.  Fish.  Soc.,  Bethesda,  MD. 

Tribuzio,  C.  A.,  C.  Rodgveller,  J.  Heifetz,  D.  Courtney,  and  K.  J. 
Goldman. 

2008.  Assessment  of  the  shark  stocks  in  the  Gulf  of 
Alaska.  In  Stock  assessment  and  fishery  evaluation 
report  for  the  groundfish  resources  of  the  Gulf  of  Alaska 
for  2009,  chapter  18,  p.  557-612.  [Available  from  North 
Pacific  Fishery  Management  Council,  605  W.  4th  Ave., 
Suite  306,  Anchorage,  AK  99501.] 

Vega,  N.  M. 

2006.  Biogeography  of  the  spiny  dogfish,  Squalus 
acanthias,  over  a latitudinal  gradient  in  the  North- 
east Pacific.  M.S.  thesis,  117  p.  Univ.  Washington, 
Seattle,  WA. 

von  Bertalanffy,  L. 

1938.  A quantitative  theory  of  organic  growth  (inquiries 
on  growth  laws  II).  Human  Biol.  10:181-213. 


136 


Effective  herding  of  flatfish  by  cables 
with  minimal  seafloor  contact 


Carwyn  F.  Hammond1 

Email  address  for  contact  author:  craig.rose@noaa.gov 

1 NOAA,  National  Marine  Fisheries  Service 

Alaska  Fisheries  Science  Center,  Conservation  Engineering  Program 
7600  Sand  Point  Way  NE 
Seattle,  Washington  98115 

2 Best  Use  Cooperative 

4241  21st  Avenue  West,  Suite  302 
Seattle,  Washington  98199 


Abstract — Otter  trawls  are  very 
effective  at  capturing  flatfish,  but 
they  can  affect  the  seafloor  ecosys- 
tems where  they  are  used.  Alaska 
flatfish  trawlers  have  very  long 
cables  (called  sweeps)  between  doors 
and  net  to  herd  fish  into  the  path 
of  the  trawl.  These  sweeps,  which 
ride  on  and  can  disturb  the  seafloor, 
account  for  most  of  the  area  affected 
by  these  trawls  and  hence  a large  pro- 
portion of  the  potential  for  damage 
to  seafloor  organisms.  We  examined 
modifications  to  otter  trawls,  such 
that  disk  clusters  were  installed  at 
9-m  intervals  to  raise  trawl  sweeps 
small  distances  above  the  seafloor, 
greatly  reducing  the  area  of  direct 
seafloor  contact.  A critical  consider- 
ation was  whether  flatfish  would  still 
be  herded  effectively  by  these  sweeps. 
We  compared  conventional  and  modi- 
fied sweeps  using  a twin  trawl  system 
and  analyzed  the  volume  and  com- 
position of  the  resulting  catches.  We 
tested  sweeps  raised  5,  7.5,  and  10 
cm  and  observed  no  significant  losses 
of  flatfish  catch  until  sweeps  were 
raised  10  cm,  and  those  losses  were 
relatively  small  (5-10%).  No  size  com- 
position changes  were  detected  in  the 
flatfish  catches.  Alaska  pollock  ( Ther - 
agra  chalcogramma ) were  captured 
at  higher  rates  with  two  versions  of 
the  modified  sweeps.  Sonar  observa- 
tions of  the  sweeps  in  operation  and 
the  seafloor  after  passage  confirmed 
that  the  area  of  direct  seafloor  contact 
was  greatly  reduced  by  the  modified 
sweeps. 


Manuscript  submitted  16  January  2009. 
Manuscript  accepted  13  November  2009. 
Fish.  Bull.  108:136-144  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


Craig  S.  Rose  (contact  author)1 
John  R.  Gauvin2 


Otter  trawling  is  one  of  the  most 
effective  methods  for  capturing  com- 
mercial quantities  of  flatfish  and  is 
the  principal  method  for  flatfish  har- 
vest in  Alaska  waters.  However,  trawl 
fisheries  have  received  increasing 
attention  for  their  potential  to  affect 
seafloor  habitats.  Changes  to  seafloor 
ecosystems  resulting  from  the  pas- 
sage of  trawl  gear  have  been  have 
described  in  a wide  range  of  studies 
(Barnes  and  Thomas,  2005;  Lokke- 
borg,  2005).  These  include  changes 
to  infaunal  (Tuck  et  al.,  1998)  and 
epifaunal  (Kaiser  et  al.,  1998;  Prena 
et  al.,  1999;  McConnaughey  et  al., 
2000)  communities,  as  well  as  indi- 
rect effects  from  changes  to  seafloor 
structure  and  resuspension  of  sedi- 
ments (Churchill,  1989).  The  most 
common  response  to  mitigate  these 
problems  has  been  closures  of  sen- 
sitive areas  to  trawling.  When  such 
areas  have  rough,  rocky  substrates, 
regulations  requiring  that  trawl  foot- 
rope  cross-sections  be  below  a certain 
size  have  been  used  to  discourage 
fishing  in  these  areas;  the  smaller 
footropes  make  nets  more  vulnerable 
to  damage  (Hannah,  2003;  Bellman 
et  al.,  2005). 

Alaskan  commercial  flatfish  fisher- 
ies, among  the  largest  in  the  world, 
are  pursued  almost  exclusively  with 
demersal  otter  trawls.  (The  excep- 
tion is  the  fishery  for  Pacific  halibut 
| Hippogossus  stenolepis ],  a large,  pi- 
scivorous species  that  is  harvested 
by  longlines.)  These  otter  trawls  gen- 


erally use  very  long  cables,  herein 
called  “sweeps,”  that  skim  the  sea- 
floor ahead  and  to  both  sides  of  the 
trawl  net.  In  Alaska  flatfish  fisher- 
ies, the  fishermen  have  used  progres- 
sively longer  sweeps  to  increase  the 
width  of  their  gear  and,  hence,  the 
area  from  which  flatfish  are  captured. 
These  sweeps  now  account  for  the 
overwhelming  majority  of  the  seafloor 
area  swept  by  these  trawlers  to  cap- 
ture flatfish.  Although  these  sweeps 
greatly  increase  flatfish  catches,  they 
also  account  for  most  of  the  negative 
effects  of  trawling  on  the  seafloor. 

Although  some  reviews  (Kaiser  et 
al.,  2007)  have  recommended  devel- 
opment of  modified  fishing  gear  to 
reduce  the  effects  of  trawling  on  sea- 
floor communities,  studies  that  test 
such  gear  are  just  beginning  to  be 
published.  He  (2007)  reviewed  such 
efforts  for  all  mobile  fishing  gears.  A 
substantial  effort  in  Europe  focused 
on  modifications  for  beam  trawling 
(van  Marlen  et  al.,  2005).  Guyonnet 
et  al.  (2008)  described  tests  of  modi- 
fied gear  that  reduce  the  contact  of 
the  cables  between  trawl  doors  and 
nets  with  the  seafloor.  Although 
their  tests  were  accomplished  with 
different  modifications  to  gear 
(dangling  chain  sections  attached 
to  neutrally  bouyant  rope)  and  in  a 
very  different  fishery,  their  concept 
is  very  similar  to  the  modifications 
we  tested. 

Ryer  (2008)  has  described  flat- 
fish behaviors  that  are  important  to 


Rose  et  al. : Effective  herding  of  flatfish  by  cables  with  minimal  seafloor  contact 


137 


their  capture  by  trawls.  Flatfish  generally  react  to 
approaching  objects  at  much  closer  ranges  (<1  m)  than 
do  roundfish  and  remain  very  close  to  the  seafloor 
when  avoiding  such  objects.  To  target  these  behaviors, 
towing  cables  (angled  toward  the  trawl  net  across  the 
sea  floor)  are  used  in  both  demersal  seines  and  otter 
trawls  to  herd  flatfish  into  the  path  of  a capture  net. 
Flatfish  avoid  the  approaching  cable  by  continuous  or 
burst-and-pause  swimming,  both  of  which  move  them 
gradually  into  the  path  of  the  capture  device.  Conven- 
tional sweep  cables  have  equal  diameters  throughout, 
and  no  structures  to  interrupt  their  contact  with  the 
seafloor.  Flere,  we  test  whether  effective  herding  re- 
sponses could  be  stimulated  if  such  cables  were  raised 
a short  distance  above  the  seafloor. 

Like  most  flatfish  fisheries,  those  in  Alaska  operate 
on  seafloors  consisting  of  unconsolidated  mixtures  of 
sand  and  mud.  The  potential  for  reducing  damage  to 
the  physical  and  biological  features  of  these  habitats 
by  raising  sweeps  a short  distance  off  the  bottom  is 
dependent  on  the  presence  of  low  vertical  relief  or  flex- 
ible structures  of  the  bottom  relief.  This  modification 
would  likely  not  prevent  damage  to  high  relief  and  rigid 
or  fragile  features  more  common  on  rockier  substrates. 
For  the  modifications  tested  here  to  be  effective,  their 
effects  on  both  catch  rates  and  seafloor  features  need 
to  be  examined. 

To  develop  practical  modifications  for  the  trawl  sys- 
tems used  in  Alaska’s  flatfish  fisheries,  we  convened  a 
series  of  meetings  with  trawler  captains  and  gear  man- 
ufacturers. For  initial  study,  they  recommended  raising 
the  sweeps  slightly  above  the  seafloor,  allowing  small 
and  flexible  animals  and  other  habitat  structure  to  pass 
beneath.  In  the  current  study  we  examine  the  proposed 
change,  focusing  on  determining  which  adjustments 


maintain  catch  rates  and  on  using  direct  observations 
to  demonstrate  reduced  seafloor  contact. 


Methods 

To  test  the  effect  of  the  modified  sweeps  on  their  ability 
to  herd  flatfish,  we  used  a twin  trawl  system  (Fig.  1).  A 
twin  trawl  system  tows  two  separate  trawls,  including 
sweeps,  simultaneously  on  parallel,  adjacent  tracks. 
Close  proximity  and  simultaneous  operation  assure  that 
both  nets  encounter  very  similar  compositions  of  fish 
species  at  similar  abundances.  Therefore  differences  in 
catch  are  principally  due  to  differences  in  the  capture 
effectiveness  of  the  two  trawls.  The  only  difference 
between  the  trawls  in  this  experiment  was  the  use  of 
the  elevating  disks  on  the  sweeps  of  the  trawls. 

Twin  trawl  tests 

Field  experiments  were  conducted  during  September 
2006  in  the  eastern  Bering  Sea  onboard  the  FV  Cape 
Horn.  The  Cape  Horn  is  a 47-m  trawler  processor, 
active  in  the  mixed  groundfish  fisheries  of  the  Bering 
Sea.  This  vessel  was  equipped  for  a twin  trawling 
system,  with  an  extra  winch  and  towing  cable.  The 
sweeps  and  trawls  were  towed  with  conventional  trawl 
doors  on  each  side  and  a weight  (clump)  in  the  middle. 
Both  doors  and  the  clump  were  towed  from  three  sepa- 
rate cables  that  were  adjusted  so  that  both  sides  fished 
evenly.  Towing  sites  were  selected  to  provide  com- 
mercial catch  rates  of  a mixture  of  the  four  principal 
flatfish  species  of  the  Bering  Sea  shelf:  yellowfin  sole 
( Limanda  aspera);  northern  rock  sole  (Lepidopsetta 
polyxystra);  flathead  sole  (Hippoglossoides  elassodon ); 


138 


Fishery  Bulletin  108(2) 


and  arrowtooth  flounder  ( Atheresthes  stomias ).  Towing 
continued  through  both  day  and  night  periods,  reflect- 
ing commercial  practice.  All  of  the  tows  were  in  areas 
with  bottom  substrates  composed  of  mixtures  of  sand 
and  mud  (McConnaughey  and  Smith,  2000). 

The  trawls  were  identical  two-seam  nets  with  200- 
mm  mesh  in  the  forward  portions  and  equipped  with 
130-mm  codends.  The  mouth  opening  of  each  net  was 
much  wider  (25  m)  than  high  (3  m).  Similar  nets  in  a 
single  trawl  configuration  are  used  to  target  flatfish  on 
the  eastern  Bering  Sea  shelf.  The  distances  between 
each  of  the  doors  and  the  central  clump  were  monitored 
for  equality  with  acoustic  measurement  systems  and 
were  each  approximately  80  m.  Both  nets  were  equipped 
with  sensors  that  indicated  the  direction  of  water  flow 
in  relation  to  the  center  of  the  headrope.  The  three 
towing  cables  were  adjusted  to  keep  that  flow  perpen- 
dicular to  the  headropes  of  both  nets  and  to  keep  their 
door-clump  openings  equal,  assuring  comparable  fishing 
characteristics  for  both  fishing  systems. 

The  sweeps  were  180-m  long  and  were  composed  of 
5-cm  (2-inch)  diameter  combination  rope  constructed  of 
steel  cable  covered  with  polyethylene  fiber.  This  is  the 
most  common  sweep  material  currently  used  in  U.S. 
Bering  Sea  flatfish  fisheries.  Sweeps  used  on  vessels  to 
target  flatfish  on  the  eastern  Bering  Sea  shelf  are  200 
to  450  m long  (C.  Rose,  unpubl.  data).  The  sweeps  of 
the  two  adjacent  trawls  had  to  be  about  half  as  long  as 
those  used  with  commercial  single  trawls,  because  the 
entire  twin  trawl  system  is  similar  in  width  to  a con- 
ventional single  trawl.  The  shorter  sweep  lengths  were 
necessary  to  assure  that  the  angle  of  the  test  sweeps 
to  the  direction  of  towing  was  similar  to  that  common 
in  the  fishery.  In  this  field  study,  clusters  of  disks  (disk 
clusters)  were  attached  over  the  experimental  sweeps 
at  9-m  (30-ft)  intervals  (Fig.  2).  The  disks  were  either 


15,  20,  or  25  cm  (6,  8,  or  10  inch)  in  diameter  attached 
to  5-cm  (2  inch)  diameter  sweeps,  creating  nominal 
clearance  between  the  cables  and  the  seafloor  of  5,  7.5, 
and  10  cm  (2,  3,  and  4 inch),  respectively.  Nominal 
clearances  are  those  immediately  adjacent  to  a disk 
when  the  disk  is  resting  on  a hard  surface.  The  press- 
ing of  disks  into  the  seafloor  and  the  sagging  of  sweeps 
between  elevating  devices  would  affect  actual  clear- 
ances. For  stability,  disk  clusters  were  approximately 
the  same  length  as  their  diameter.  These  disk  clusters 
were  fixed  in  position  with  a combination  of  clamps 
and  rope  seizings,  which  were  run  through  the  sweep 
cable  to  prevent  the  clusters  from  sliding  along  the 
cable.  Disk  clusters  were  installed  on  the  aft  90  m of 
the  modified  sweeps.  Halfway  through  each  experiment, 
the  sweeps  were  switched  between  the  two  trawl  nets, 
but  each  trawl  net  remained  in  place. 

Catches  from  each  trawl  were  kept  separate  through- 
out the  sampling  process.  As  catches  entered  the  sam- 
pling area,  they  passed  across  a motion-compensated 
flow  scale  which  provided  a total  catch  weight.  All  indi- 
viduals of  four  flatfish  species  (yellowfin  sole,  northern 
rock  sole,  flathead  sole,  and  arrowtooth  flounder)  and 
two  gadids  (Pacific  cod  [Gadus  macrocephalus ] and 
Alaska  pollock  [ Theragra  chalcogramma ])  were  sorted 
into  separate  holding  bins.  These  are  the  principal 
flatfish  and  gadid  species  harvested  from  the  eastern 
Bering  Sea  shelf.  Fish  from  each  bin  were  then  run 
across  a second  flow  scale  to  measure  the  weight  of  each 
of  these  species.  During  the  sorting  of  catch  from  each 
trawl,  50-150  fish  of  each  species  were  sampled  and 
their  fork  lengths  were  measured  to  1-cm  intervals  to 
determine  their  size  composition.  These  length  samples 
were  periodically  taken  from  the  catch  as  it  passed 
through  the  sorting  area  to  avoid  bias  in  case  fish  size 
varied  between  parts  of  each  catch.  Because  of  their 
large  size,  limited  holding  space  and  handling 
requirements  precluded  adequate  length  sampling 
of  Pacific  cod. 

Tows  were  planned  to  last  2 hours,  unless  catch 
sensors  indicated  substantial  catches  (>8  met- 
ric tons  [t]  per  net)  before  that  time.  Actual  tow 
durations  ranged  from  33  to  150  minutes.  We 
eliminated  hauls  where  debris  (e.g.,  crab  pots) 
was  large  enough  to  clog  the  net,  or  where  gear 
components  became  entangled,  because  such  con- 
ditions could  influence  gear  performance  and  the 
size  and  composition  of  the  resulting  catch.  Tow 
locations  were  selected  in  order  to  encounter  com- 
mercial concentrations  of  the  major  flatfish  species 
of  the  eastern  Bering  Sea  shelf.  Environmental 
parameters  at  the  trawl,  including  depth,  tempera- 
ture and  light  level,  were  sampled  throughout  the 
experiment  with  a Mk9  logger  (Wildlife  Comput- 
ers, Redmond,  WA)  mounted  at  the  center  of  the 
trawl’s  headrope. 

We  used  a high-resolution,  rapid-update  sonar 
(SoundMetrics  DIDSON,  Dual-frequency  IDen- 
tification  SONar,  Lake  Forest  Park,  WA)  to  ob- 
serve how  the  sweep  modifications  affected  sea- 


Schematic  diagram  of  a cluster  of  disks  (disk  cluster)  attached 
to  trawl  sweeps  to  raise  the  sweeps  above  the  seafloor  to 
test  whether  this  gear  modification  reduces  flatfish  herding. 
Rubber  disks  (A,  20  cm-diameter,  and  B,  15  cm-diameter) 
were  installed  over  the  sweep  cable,  between  clamps  (D)  that 
fix  their  location  on  the  cable.  Steel  washers  (C)  prevented 
rubber  disks  from  passing  over  clamps.  Ropes  seized  over  and 
tucked  through  cable  (E)  blocked  clamps  from  shifting. 


Rose  et  al. : Effective  herding  of  flatfish  by  cables  with  minimal  seafloor  contact 


139 


180°W  1 70°W  160°W 


Figure  3 

Fishing  locations  (•)  in  the  eastern  Bering  Sea  for  the  2006  tests  of 
the  effects  of  raised  sweeps  on  flatfish  herding.  Regions  shaded  with 
diagonal  lines  are  areas  of  trawl  closures  around  the  Pribilof  Islands 
and  in  Bristol  Bay.  Contour  lines  indicate  depths. 


floor  contact.  This  was  mounted  in 
a protective  sled,  which  was  towed 
both  behind  the  sweeps,  to  show 
interactions  between  the  sweeps 
and  the  seafloor,  and,  separate 
from  the  trawl,  across  the  track  of 
a previous  haul,  to  show  marks  left 
on  the  seafloor.  These  observations 
were  made  only  on  sweeps  with  the 
20-cm  disks.  The  sled  was  also 
towed  across  tracks  from  previous 
trawl  tows  with  conventional  and 
modified  sweeps  and  was  equipped 
with  a video  camera  for  detailed 
imagery. 

To  estimate  the  proportional 
change  in  catch  due  to  the  sweep 
modifications,  the  difference  be- 
tween the  natural  logarithms  of 
the  catch  weights  from  modified 
and  unmodified  trawls  (Log Dif) 
was  calculated  for  each  species 
from  each  twin-trawl  haul: 

Log  Dif  = In  (Catch  modlfied)  - 

In  (Catch  unmodified).  (1) 

This  statistic,  equivalent  to  the  log- 
arithm of  the  ratio  between  catches  with  modified  and 
unmodified  nets,  was  appropriate  because  absolute  catch 
sizes  were  uncontrolled  and  varied  widely.  A statistic 
based  on  subtracting  the  untransformed  trawl  catches, 
like  that  for  an  ordinary  paired  t-test,  would  have  varied 
proportionally  to  absolute  catch  rates,  whereas  catch 
ratios,  as  measured  by  Log  Dif  were  independent  of  the 
fish  densities  encountered  during  each  tow.  Averages 
and  confidence  intervals  of  Lo gDif  were  computed  for 
each  species  and  sweep  modification.  To  report  these 
results  as  ratios,  the  averages  and  confidence  intervals 
were  then  back-transformed  with  the  exponential  func- 
tion. Catch  results  were  only  used  for  species  with  more 
than  a minimal  catch  (>10  fish)  in  both  nets.  The  null 
hypothesis  that  the  sweep  modifications  did  not  affect 
catch  was  tested  with  a t-test  of  whether  average  Lo  gDif 
was  different  from  0,  equivalent  to  a paired  t-test  for 
differences  between  the  log-transformed  catches. 

To  test  whether  the  sweep  modifications  affected  the 
size-selectivity  for  different  fish  species  and  to  minimize 
variability,  we  pooled  fish  into  three  size  classes  for 
each  species,  except  for  arrowtooth  flounder,  where  a 
wide  size  range  made  four  size  classes  more  appropri- 
ate. The  size-class  boundaries  were  set  so  that  approxi- 
mately one-third  (one-fourth  for  arrowtooth  flounder)  of 
the  fish  in  the  combined  control  catches  were  in  each 
category.  To  maintain  consistency  with  the  weight- 
based  analysis  of  overall  catch,  and  because  the  Alaska 
trawl  fleet  classifies  fish  sizes  by  weight,  the  boundaries 
of  the  size  classes  were  defined  by  individual  weights 
instead  of  lengths,  and  the  catches  of  each  size  class 
were  computed  as  weights,  instead  of  numbers.  Length- 


weight  functions  from  the  annual  Bering  Sea  shelf  trawl 
survey  (NMFS,  unpubl.  data1)  were  used  to  convert  the 
sampled  lengths  to  their  corresponding  weights.  The 
catch  of  each  size  class  was  estimated  by  expanding  the 
proportion  of  that  size  class,  by  weight,  from  the  sample 
of  catch  for  that  species.  As  with  the  total  catch  data, 
averages  and  confidence  intervals  were  calculated.  We 
used  analysis  of  variance  to  test  for  differences  between 
size  classes  for  each  combination  of  species  and  for  each 
sweep  modification. 

Results 

From  6 to  23  September  2006,  61  successful  twin  trawl 
hauls  were  conducted,  including  19,  26,  and  16  hauls 
with  experimental  sweep  clearances  of  5,  7.5,  and  10  cm, 
respectively.  Depths  at  these  tow  sites  (Fig.  3)  ranged 
from  70  to  117  m,  and  bottom  temperatures  ranged  from 
2.5°  to  5.5°C. 

Sonar  imagery  during  towing  showed  that  unmodified 
sweeps  produced  a continuous  cloud  of  disturbed  sedi- 
ment due  to  contact  with  the  seafloor.  Variation  in  the 
density  of  that  cloud  appeared  to  result  from  contact 
with  high  and  low  spots  on  the  seafloor,  and  rapid  oscil- 
lation of  strong  and  weak  cloud  intensity  appeared  to 
be  due  to  vibration  of  the  sweeps.  In  contrast,  the  sedi- 
ment cloud  from  the  modified  sweep  appeared  only  di- 
rectly behind  the  disk  cluster.  The  only  clouds  from  the 


1 NMFS,  Alaska  Fisheries  Science  Center,  RACE  Division, 
7600  Sand  Point  Way  NE,  Seattle,  WA 


140 


Fishery  Bulletin  108(2) 


Figure  4 

(A)  Sonar  and  ( B ) video  imagery  of  the  seafloor  after  passage  of  the  raised  otter  trawl  sweeps.  Video  picture  was 
taken  as  the  seafloor  sled  passed  over  the  location  indicated  on  the  sonar  image.  Otter  trawl  sweeps  were  raised 
with  widely  spaced  disk  clusters,  which  caused  the  parallel  tracks  seen  in  the  sonar  image  and  the  flattened 
swath  in  the  video  image. 


sweeps  themselves  were  brief  puffs  after  contact  with 
high  spots  on  the  seafloor.  Areas  covered  by  the  modi- 
fied sweeps  showed  marks  from  the  disk  clusters  that 


were  approximately  10-cm  wide  separated  by  seafloor 
indistinguishable  from  unaffected  areas  (Fig.  4A).  This 
disk  cluster  mark  was  approximately  5%  of  the  2-m 
interval  between  marks.  This  spacing  is 
much  shorter  than  the  9-m  spacing  on  the 
cable  because  sweeps  are  sharply  angled 
to  their  direction  of  movement  (angle-of- 
attack).  Images  of  such  tracks  from  the 
video  (Fig.  4B)  showed  a flattening  of  very 
low-profile  surface  textures. 

The  use  of  15-cm  disks  on  the  sweeps 
did  not  cause  significant  differences  in 
catch  rates  (LogDif  was  not  different 
from  0)  for  any  of  the  six  species,  and 
only  the  pollock  catch  rate  changed  (12% 
increase,  P=0.007)  with  the  20-cm  disks 
(Fig.  5).  Northern  rock  sole  and  flathead 
sole  catches  both  decreased  significantly 
(-11%,  PcO.OOl,  and  -5%,  P=0.02,  respec- 
tively) when  the  25-cm  disks  were  used, 
whereas  pollock  catch  increased  again 
(+12%,  P=0.03).  Decreases  for  the  other 
two  flatfish  were  also  observed — although 
not  statistically  significant  at  the  0.05 
level  (P=0.08  for  arrowtooth  flounder  and 
P=0.07  for  yellowfin  sole).  A consistent 
decrease  in  the  mean  relative  catch  with 
increasing  disk  size  for  all  of  the  flatfish 
species,  although  only  significant  for  the 
largest  disks,  indicates  that  smaller  ef- 
fects may  have  occurred  for  the  smaller 
disks  that  could  not  be  statistically  de- 


1.30  -i 


Yellowfin 

sole 


Pacific  cod  Alaska 
pollock 


Figure  5 

Estimates  of  and  95%  confidence  intervals  for  the  ratios  of  fish  catches 
with  the  modified  trawl  sweeps  raised  to  three  different  heights  off  the 
seafloor  to  fish  catches  with  conventional  sweeps  for  the  four  principal 
flatfish  species  (yellowfin  sole  [Limanda  aspera ];  northern  rock  sole 
[Lepidopsetta  polyxystra]\  flathead  sole  I Hippoglossoides  elassodon J; 
arrowtooth  flounder  [Atheresthes  stomias ]);  and  two  principal  gadid 
species:  Pacific  cod  ( Gadus  macrocephalus)\  and  Alaska  pollock  ( Ther - 
agra  chalcogramma ) taken  in  Bering  Sea  trawl  fisheries. 


Rose  et  al.:  Effective  herding  of  flatfish  by  cables  with  minimal  seafloor  contact 


141 


Figure  6 

Size  compositions  for  flatfish  and  gadid  species  taken  during  tests  of 
whether  raised  trawl  sweeps  reduce  herding  of  fish.  Yellowfin  sole  (Limanda 
aspera);  northern  rock  sole  ( Lepidopsetta  polyxystra) ; flathead  sole  (Hippo- 
glossoides  elassodon)\  arrowtooth  flounder  ( Atheresthes  stomias);  Pacific  cod 
( Gadus  macrocephalus );  and  Alaska  pollock  ( Theragra  chaleogramma). 


tected  in  our  experiment.  Pacific  cod 
catches  did  not  change  significantly 
with  any  of  the  modifications. 

For  evaluating  the  likelihood  of 
substantial  losses  of  catch,  the  confi- 
dence intervals  provide  more  informa- 
tion than  the  basic  significance  tests 
alone.  For  example,  the  lower  confi- 
dence bounds  for  the  effects  of  20-cm 
disks  on  flatfish  catches  leave  only  a 
2.5%  (1  of  40)  probability  that  catch 
losses  would  exceed  4-6%.  Correspond- 
ing “worst  case”  losses  for  the  15-cm 
disks  were  even  smaller.  Similarly,  al- 
though none  of  the  Pacific  cod  catch 
results  passed  the  threshold  of  a 95% 
two-tailed  probability  of  being  differ- 
ent from  no  change,  all  three  of  the 
confidence  intervals  were  almost  en- 
tirely above  a value  of  1.  Therefore,  a 
trawler  could  implement  one  of  these 
modifications  with  little  expectation  of 
catching  fewer  Pacific  cod  and  with  a 
reasonable  chance  of  slight  increases 
in  Pacific  cod  catch. 

The  size  composition  of  each  spe- 
cies from  the  unmodified  nets  (Fig.  6) 
showed  truncation  at  the  lower  end  of 
the  size  distribution,  owing  to  use  of 
large  mesh  in  the  body  of  the  net  (20  cm,  stretch  mea- 
sure), intermediates  (14  cm)  and  codends  (15  cm)  that 
release  smaller  fish.  Although  the  proportions  varied 
somewhat  between  experiments,  each  study  encoun- 
tered a representative  range  of  sizes  available  to  the 
commercial  fishery. 

ANOVA  tests  for  differences  in  catch  effects  between 
major  size  classes  (thirds  or  quartiles  of  control  size 
frequencies)  revealed  no  significant  differences  for  any 
of  the  flatfish  species  (Fig.  7).  One  significant  difference 
(P=0.04)  was  detected  for  pollock  in  sweeps  with  the 
smallest  disks  (15  cm),  attributable  to  a lower  catch 
rate  of  the  smallest  pollock.  Confidence  intervals  were 
included  in  Figure  7 to  aid  comparisons  between  size 
groups  within  species  and  sweep  modification  classes. 
Confidence  intervals  were  wider  for  the  largest  and 
smallest  categories  because  few  individuals  from  these 
ranges  were  encountered  in  some  tows,  increasing  vari- 
ability, whereas  all  tows  had  substantial  numbers  of 
fish  in  the  central  ranges. 

Discussion 

Flatfish  can  be  effectively  herded  by  trawl  sweeps 
and  with  greatly  reduced  seafloor  contact.  Signifi- 
cant catch  reductions,  averaging  5%  for  flathead  sole 
and  11%  for  rock  sole,  were  only  detected  when  25-cm 
disks  were  installed  that  raised  the  sweeps  10  cm 
above  the  substrate  at  the  ends  of  each  9-m  section.  No 
detectable  catch  reductions  occurred  during  tests  with 


smaller  clearances  (5  and  7.5  cm).  Confidence  intervals 
indicated  only  a 2.5%  probability  of  catch  reductions 
greater  that  5%  with  7.5-cm  clearances.  Nor  did  sweeps 
with  such  clearances  appear  to  change  size  selectivity 
significantly. 

Flatfish  exhibit  predator  avoidance  behaviors  that 
allow  them  to  be  effectively  herded  by  the  sweeps.  In 
contrast  to  roundfish,  flatfish  cease  movement  when  a 
predator  is  detected  and  only  flee  upon  very  close  ap- 
proach (Ryer,  2008).  Therefore,  observed  flatfish  reac- 
tions to  trawl  gear  (Main  and  Sangster,  1981;  Rose, 
1996;  Ryer  and  Barnett,  2006)  mostly  occur  at  horizon- 
tal ranges  of  much  less  than  1 m.  However,  because  con- 
ventional fishing  gear  has  either  continuous  or  closely 
spaced  contact  with  the  seafloor,  there  has  been  little  or 
no  information  to  assess  the  role  of  gear  contact  or  prox- 
imity to  the  seafloor  in  either  initiating  or  sustaining 
the  flight  behaviors  that  result  in  herding.  Given  the 
cryptic  behaviors  of  flatfish,  we  could  not  assume  that 
stimuli  several  centimeters  above  the  seafloor  would  be 
as  effective  as  those  that  would  directly  contact  flatfish 
on  the  seafloor.  The  current  results  demonstrate  that 
flatfish  do  respond  with  effective  herding  behaviors  to 
sweep  cables  displaced  from  the  seafloor  by  5 to  10  cm. 
Even  the  largest  of  the  flatfish  encountered  here  would 
not  have  contacted  the  raised  sweeps  if  they  remained 
resting  on  the  seafloor.  At  the  highest  clearance  (10 
cm),  slightly  reduced  catches  indicated  that  the  flight 
response  began  to  break  down  and  some  of  the  flat- 
fish were  not  herded  as  well  as  with  the  conventional 
sweeps.  Winger  et  al.  (2004)  found  that  flatfish  size 


142 


Fishery  Bulletin  108(2) 


Yellowfin  sole 


Northern  rock  sole 


Elevating  disk  diameter 

Flathead  sole 


Elevating  disk  diameter 


□ <1000  g 
■ 1000-1500  g 
£1  >1500  g 


Elevating  disk  diameter  Elevating  disk  diameter 

Figure  7 

Estimates  of  and  95%  confidence  intervals  for  ratios  of  fish  catches  during  tests  with  modified  trawl  sweeps  raised  to 
three  different  heights  off  of  the  seafloor  to  catches  with  conventional  sweeps  for  broad  size  classes  of  four  principal  flat- 
fish species  and  a principal  gadid  species  taken  in  Bering  Sea  trawl  fisheries:  yellowfin  sole  ( Liman  da  aspera);  northern 
rock  sole  (Lepidopsetta  polyxystra)\  flathead  sole  (Hippoglossoides  elassodon)\  arrowtooth  flounder  ( Atheresthes  stomias ); 
and  Alaska  pollock  (Tlieragra  chcilcogramma) . 


affected  behavioral  responses  to  approaching  sweeps, 
including  tailbeat  frequency  and  swimming  endurance. 
Although  any  of  these  behaviors  could  affect  herding- 
related  capture  rates,  the  current  study  did  not  indicate 
behavioral  differences  between  size  classes  in  response 
to  the  elevated  sweeps. 

We  followed  commercial  practices  in  the  gear  type 
used,  weight-based  catch  metrics,  towing  durations, 
catch  handling,  and  round-the-clock  operations.  This 


procedure  was  undertaken  to  increase  the  relevance  of 
our  results  to  those  with  the  greatest  stake  in  deciding 
on  the  use  of  these  modifications:  the  fishermen  and 
fishing  companies.  Fishermen  actively  participated  in 
designing  the  gear  modifications  and  in  conducting  the 
research. 

To  examine  consequences  of  using  modified  sweeps 
in  the  fishery  and  to  improve  precision,  all  tows  were 
analyzed  together,  including  day  and  night  tows,  even 


Rose  et  al.:  Effective  herding  of  flatfish  by  cables  with  minimal  seafloor  contact 


143 


though  light  levels  affect  the  herding  process  (Ryer  and 
Barnett,  2006).  The  effects  of  light  on  flatfish  herding 
are  analyzed  and  reported  in  a separate  paper  (Ryer 
et  al.,  2010). 

Although  not  the  focus  of  this  study,  an  unexpected 
result  was  the  increase  in  pollock  catches  that  occurred 
with  two  of  the  sweep  modifications.  Pollock  herd  differ- 
ently from  flatfish,  reacting  to  stimuli  at  much  greater 
distances  (Rose,  1996).  The  forward  sections  of  the 
most  modern  pollock  trawls  have  “meshes”  that  are 
more  than  25-m  long.  Although  large  groups  of  pol- 
lock could  easily  swim  through  such  meshes,  they  still 
avoid  the  netting  and  are  eventually  herded  into  parts 
of  the  net  that  physically  restrain  them.  These  nets 
would  not  work  if  pollock  herded  only  at  short  ranges. 
Separation  of  the  sweeps  from  the  seafloor,  or  the  disk 
clusters  themselves,  could  have  increased  visibility  of 
the  sweeps,  which  may  have  enhanced  pollock  herding. 
Both  factors  would  be  reduced  at  the  smallest  disks, 
where  herding  improvement  was  not  detected. 

Sonar  observations  of  the  elevated  sweeps  showed 
that  their  interaction  with  the  seafloor  was  radically 
changed.  The  continuous  sediment  clouds  produced 
along  the  entire  length  of  the  unmodified  sweeps  were, 
for  the  modified  sweeps,  reduced  to  isolated  clouds  be- 
hind each  disk,  indicating  substantial  reductions  in  the 
area  of  direct  contact.  Therefore,  any  effects  based  on 
direct  contact,  as  well  as  resuspension  of  sediments, 
should  have  been  greatly  reduced.  The  sonar  images 
of  the  seafloor  after  passage  of  the  sweep  showed  that 
the  contact  area  of  the  disks  was  approximately  5%  of 
the  total  swept  area.  Seafloor  texture  between  the  disk 
tracks  was  indistinguishable  from  unaffected  areas, 
but  areas  covered  by  conventional  sweeps  showed  slight 
smoothing.  The  seafloor  directly  contacted  by  the  disks 
was  uniformly  smoothed.  Although  the  texture  change 
due  to  conventional  sweeps  appeared  slight,  the  resus- 
pension observed  during  fishing  indicated  some  distur- 
bance of  the  bottom  and  we  believe  that  the  substantial 
reduction  of  contact  due  to  using  the  disks  more  than 
compensates  for  any  increased  disturbance  to  the  small 
area  directly  under  the  disks. 

In  another  recent  study  (Guyonnet  et  al.,  2008),  the 
concept  of  slightly  raising  trawl  sweeps,  therein  called 
“legs,”  was  also  applied  to  reduce  their  impact  on  the 
seafloor.  Instead  of  disk  clusters,  Guyonnet  et  al.  used 
neutrally  buoyant  sweep  material  that  was  weighted 
only  by  dangling  chains  attached  every  50  cm.  They 
also  found  no  significant  effects  on  catch  composition 
or  size  selectivity  for  target  animals.  They  found  that 
damage  to  benthic  animals  was  reduced  with  the  al- 
ternative gear. 

Our  results  alone,  although  promising,  do  not  address 
the  full  potential  of  sweep  modifications  to  reduce  the 
effects  on  the  seafloor  of  trawling  for  Bering  Sea  flat- 
fish. Although  creating  several  centimeters  of  separation 
between  the  sweeps  and  the  seafloor  greatly  reduces  the 
potential  for  damage  to  infauna  and  small  epifauna, 
it  does  not  prevent  contact  with  seafloor  features  and 
animals  larger  than  that  spacing.  The  vulnerability  of 


ecosystem  features  to  trawling  operations  is  a function 
of  the  amount  of  damage  caused  by  each  trawl  exposure 
(e.g.,  the  proportion  of  a particular  species  in  the  path 
of  a trawl  that  dies  due  to  trawl  contact)  and  the  fre- 
quency and  coverage  of  the  trawling  effort.  An  analysis 
of  such  factors  for  the  Bering  Sea  shelf  highlighted 
structure-forming  animals  as  the  seafloor  feature  most 
vulnerable  to  trawling.2  The  structure-forming  animals 
of  the  eastern  Bering  Sea  shelf  are  generally  small  and 
flexible;  therefore  it  is  quite  conceivable  that  creating 
a space  below  the  sweeps  could  also  reduce  damage  to 
these  animals.  That  potential  is  being  examined  by 
the  authors  in  a subsequent  study  that  will  focus  on 
how  these  sweep  modifications  change  damage  rates  to 
structure-forming  animals  of  the  Bering  Sea  shelf. 

Successful  gear  modifications  for  reducing  trawling 
effects  on  seafloor  habitats  would  add  a habitat  pro- 
tection option  in  addition  to  area  closures  and  gear 
switching.  Closures  of  areas  to  trawling  can  move  fish- 
ing effort  from  productive  grounds,  and  therefore  can 
increase  the  total  effort  required  or  concentrate  fishing 
and  its  effects  in  the  remaining  fishing  grounds  (Fu- 
jioka,  2006).  The  list  of  alternative  gear  for  harvesting 
these  flatfish  is  quite  limited  and  none  are  without 
some  negative  effects  on  habitat.  With  beam  trawling, 
herding  sweeps  are  not  used  to  concentrate  fish  into  the 
path  of  the  capture  device.  Therefore,  the  entire  area 
from  which  fish  are  collected  is  swept  with  the  capture 
net  itself.  Studies  to  reduce  the  effects  of  beam  trawls 
on  habitat  have  focused  on  other  stimuli  to  move  fish 
from  the  seafloor  into  the  net  (van  Marlen  et  al.,  2005). 
The  capture  process  for  demersal  seines  is  similar  in 
many  ways  to  that  of  Alaska  otter  trawls  with  long 
sweeps — weighted  cables  are  pulled  across  the  seafloor 
to  herd  fish  into  the  path  of  a capture  net.  Demersal  en- 
tangling nets  depend  on  natural  movements  of  the  fish 
to  bring  them  to  the  gear,  and  therefore  they  are  effec- 
tive only  during  periods  when  fish  are  actively  moving. 
They  are  still  unlikely  to  produce  catch  rates  similar 
to  those  produced  with  trawls  unless  vast  fleets  of  nets 
are  deployed.  Such  extensive  net  deployments  would 
exacerbate  the  most  notable  problem  with  demersal 
entangling  nets — ghost  fishing  of  derelict  and  lost  gear. 
Finally,  although  longline  fishing  is  the  foundation  for 
one  of  the  most  successful  commercial  flatfish  fisheries 
(Pacific  halibut),  most  flatfish  species  are  not  of  the  size 
and  do  not  have  a predatory  diet  that  make  longlines 
particularly  effective. 

Implementing  the  trawl  gear  modifications  described 
here  would  require  some  adaptations  in  equipment  and 
handling  methods  for  fishermen.  The  volume  of  the 
elevating  devices  would  require  additional  space  on 
deployment  reels  or  net  drums,  thus  requiring  either 
that  sweep  lengths  be  shortened  to  fit  onto  the  reels 
or  larger  reels  be  installed  on  vessels.  The  disks  would 


2 Final  environmental  impact  statement  for  essential  fish 
habitat  identification  and  conservation  in  Alaska.  April 
2005  [online],  http://www.fakr.noaa.gov/habitat/seis/efheis. 
htm. 


144 


Fishery  Bulletin  108(2) 


also  complicate  deployment  and  retrieval  because  they 
do  not  wrap  as  evenly  onto  reels  as  unmodified  sweeps. 
Potential  advantages  with  the  use  of  disks  would  in- 
clude longer  usability  of  sweeps  and  reduced  drag  (im- 
proved fuel  efficiency),  both  due  to  reduced  contact  of 
the  sweeps  with  the  seafloor.  An  important  factor  in 
identifying  these  implementation  and  operational  issues 
early,  as  well  as  in  the  development  of  potential  solu- 
tions, has  been  the  direct  participation  of  the  fishing  in- 
dustry in  this  research  and  our  ability  to  conduct  these 
tests  under  conditions  identical  to  most  of  the  important 
operational  aspects  of  the  commercial  fishery. 

Acknowledgments 

The  authors  thank  K.  Hjelm,  captain  of  the  FV  Cape 
Horn , and  his  crew  for  their  tireless  work  and  creativ- 
ity in  helping  with  this  research.  We  also  appreciate 
the  support  of  numerous  Bering  Sea  trawl  captains  and 
companies  in  discussing,  motivating,  and  moving  our 
study  forward.  Scott  McEntire  developed  the  systems 
necessary  for  collection  of  video  and  sonar  data,  a criti- 
cal contribution  to  this  project.  We  are  very  grateful  to 
our  sampling  crew:  D.  Benjamin,  N.  Roberson,  E.  Acuna, 
and  H.  Kenney,  and  those  from  a pilot  study,  J.  Olsen, 
J.  Hagga,  and  C.  Shavey.  Our  thanks  are  also  extended 
to  the  many  reviewers  both  anonymous  and  within  the 
Alaska  Fisheries  Science  Center,  whose  thoughtful  com- 
ments and  suggestions  greatly  improved  this  article. 


Literature  cited 

Barnes,  P.  W.,  and  J.  P.  Thomas  (eds.). 

2005.  Benthic  habitats  and  the  effects  of  fishing.  Am. 
Fish.  Soc.  Symp.  4,  890  p. 

Bellman,  M.  A.,  S.  A.  Heppell,  and  C.  Goldfinger. 

2005.  Evaluation  of  a US  west  coast  groundfish  habi- 
tat conservation  regulation  via  analysis  of  spatial  and 
temporal  patterns  of  trawl  fishing  effort.  Can.  J.  Fish. 
Aquat.  Sci.  62:2886-2900. 

Churchill,  J.  H. 

1989.  The  effect  of  commercial  trawling  on  sediment 
resuspension  and  transport  over  the  Middle  Atlantic 
Bight  continental  shelf.  Cont.  Shelf  Res.  9:841—865. 

Fujioka,  J.  T. 

2006.  A model  for  evaluating  fishing  impacts  on  habitat 
and  comparing  fishing  closure  strategies.  Can.  J.  Fish. 
Aquat.  Sci.  63:2330-2342. 

Guyonnet,  B.,  J.  Grail,  and  B.  Vincent. 

2008.  Modified  otter  trawl  legs  to  reduce  damage  and 
mortality  of  benthic  organisms  in  North  East  Atlantic 
fisheries  (Bay  of  Biscay).  J.  Mar.  Syst.  72:2-16. 

Hannah,  R.  W. 

2003.  Spatial  changes  in  trawl  fishing  effort  in  response 
to  footrope  diameter  restrictions  in  the  U.S.  west  coast 
bottom  trawl  fishery.  N.  Am.  J.  Fish.  Manag.  23:693- 
702. 

He,  P. 

2007.  Technical  measures  to  reduce  seabed  impact  of 


mobile  fishing  gears.  In  Bycatch  reduction  in  the  world’s 
fisheries  (S.  J.  Kennelly  ed.),  p.  141-179.  Springer, 
New  York. 

Kaiser,  M.  J.,  D.  B.  Edwards,  P.  J.  Armstrong,  K.  Radford, 
N.  E.  L.  Lough,  R.  P.  Flatt,  and  H.  D.  Jones. 

1998.  Changes  in  megafaunal  benthic  communities  in 
different  habitats  after  trawling  disturbance.  ICES  J. 
Mar.  Sci.  55:353-361. 

Kaiser,  M.  J.,  N.  Graham,  C.  S.  Rose,  and  P.  H.  Weibe. 

2007.  Ecosystem-sensitive  approaches  to  fishing:  reconcil- 
ing fisheries  with  conservation  through  improvements  in 
fishing  technology.  ICES  J.  Mar.  Sci.  64:1610-1611. 

Lokkeborg,  S. 

2005.  Impacts  of  trawling  and  scallop  dredging  on  ben- 
thic habitats  and  communities.  FAO  Fish.  Tech.  Paper 
472,  58  p.  FAO,  Rome. 

Main,  J.,  and  G.  I.  Sangster. 

1981.  A study  of  the  fish  capture  process  in  a bottom 
trawl  by  direct  observations  from  a towed  underwater 
vehicle.  Scott.  Fish.  Res.  Rep.  23,  24  p.  Marine 
Scotland,  Aberdeen. 

McConnaughey,  R.  A.,  C.B.  Dew,  and  K.  Meir. 

2000.  An  examination  of  chronic  trawling  effects  on  soft 
bottom  benthos  in  the  eastern  Bering  Sea.  ICES  J.  of 
Mar.  Sci.  57:1377-1388. 

McConnaughey,  R.  A.,  and  K.R.  Smith. 

2000.  Associations  between  flatfish  abundance  and  sur- 
ficial  sediments  in  the  eastern  Bering  Sea.  Can.  J. 
Fish.  Aquat.  Sci.  57:2410-2419. 

Prena,  J.,  P.  Schwinghamer,  T.  W.  Rowell,  D.  C.  Gordon,  K.  D. 
Gilkinson,  W.  P.  Vass,  and  D.  L.  McKeown. 

1999.  Experimental  otter  trawling  on  a sandy  bottom 
ecosystem  of  the  Grand  Banks  of  Newfoundland:  analysis 
of  trawl  bycatch  and  effects  on  epifauna.  Mar.  Ecol. 
Progr.  Ser.  181:107-124. 

Rose,  C.  S. 

1996.  Behavior  of  North  Pacific  groundfish  encounter- 
ing trawls:  applications  to  reduce  bycatch.  In  Solv- 
ing bycatch:  considerations  for  today  and  tomorrow,  p. 
235-241.  Alaska  Sea  Grant  College  Program  Report 
96-03,  Univ.  Alaska,  Fairbanks. 

Ryer,  C.  H. 

2008.  A review  of  flatfish  behavior  relative  to  trawls. 
Fish.  Res.  90:138-146 

Ryer,  C.  H.,  and  L.  A.  K.  Barnett. 

2006.  Influence  of  illumination  and  temperature  upon 
flatfish  reactivity  and  herding  behavior:  Potential 
implications  for  trawl  capture  efficiency.  Fish.  Res. 
81:242-250. 

Ryer,  C.  H.,  C.  S.  Rose,  and  P.  S.  Iseri. 

2010.  Flatfish  herding  behavior:  diel  patterns  of  trawl 
sweep  efficiency  as  inferred  from  field  and  laboratory 
manipulations.  Fish.  Bull.  108:145-154. 

Tuck,  I.  D.,  S.  J.  Hall,  M.  R.  Robertson,  E.  Armstrong,  and  D.  J. 
Basford. 

1998.  Effects  of  physical  trawling  disturbance  in  a previ- 
ously unfished  sheltered  Scottish  sea  loch.  Mar.  Ecol. 
Progr.  Ser.  162:227-242. 

van  Marlen,  B,  M.  J.  N.  Bergman,  S.  Groenwold,  and  M.  Fonds. 

2005.  New  approaches  to  the  reduction  of  non-target 
mortality  in  beam  trawling.  Fish.  Res.  72:333-345. 

Winger,  P.  D.,  S.  J.  Walsh,  P.  He,  and  J.  A.  Brown. 

2004.  Simulating  trawl  herding  in  flatfish:  the  role  of  fish 
length  in  behavior  and  swimming  characteristics.  ICES 
J.  Mar.  Sci.  61:1179-1185. 


145 


Flatfish  herding  behavior  in  response 

to  trawl  sweeps:  a comparison  of  diet  responses 

to  conventional  sweeps  and  elevated  sweeps 


Email  address  for  contact  author:  cliff.ryer@noaa.gov 

1 Fisheries  Behavioral  Ecology  Program 

Resource  Assessment  and  Conservation  Engineering  Division 

Alaska  Fisheries  Science  Center,  NOAA  Fisheries 

Hatfield  Marine  Science  Center 

2030  Marine  Science  Drive 

Newport,  Oregon  97365 

2 Resource  Assessment  and  Conservation  Engineering  Division 
Alaska  Fisheries  Science  Center,  NOAA  Fisheries 

7600  Sand  Point  Way 
Seattle,  Washington  98115 


Abstract — Commercial  bottom  trawls 
often  have  sweeps  to  herd  fish  into 
the  net.  Elevation  of  the  sweeps  off 
the  seafloor  may  reduce  seafloor 
disturbance,  but  also  reduce  herd- 
ing effectiveness.  In  both  field  and 
laboratory  experiments,  we  examined 
the  behavior  of  flatfish  in  response 
to  sweeps.  We  tested  the  hypotheses 
that  1)  sweeps  are  more  effective  at 
herding  flatfish  during  the  day  than 
at  night,  when  fish  are  unable  to  see 
approaching  gear,  and  that  2)  eleva- 
tion of  sweeps  off  the  seafloor  reduces 
herding  during  the  day,  but  not  at 
night.  In  sea  trials,  day  catches  were 
greater  than  night  catches  for  four 
out  of  six  flatfish  species  examined. 
The  elevation  of  sweeps  10  cm  sig- 
nificantly decreased  catches  during 
the  day,  but  not  at  night.  Laboratory 
experiments  revealed  northern  rock 
sole  ( Lepidopsetta  polyxystra ) and 
Pacific  halibut  ( Hippoglossus  stenol- 
epis ) were  more  likely  to  be  herded 
by  the  sweep  in  the  light,  whereas  in 
the  dark  they  tended  to  pass  under  or 
over  the  sweep.  In  the  light,  elevation 
of  the  sweep  reduced  herding,  and 
more  fish  passed  under  the  sweep.  In 
contrast,  in  the  dark,  sweep  elevation 
had  little  effect  upon  the  number  of 
fish  that  exhibited  herding  behavior. 
The  results  of  both  field  and  labo- 
ratory experiments  were  consistent 
with  the  premise  that  vision  is  the 
principle  sensory  input  that  controls 
fish  behavior  and  orientation  to  trawl 
gear,  and  gear  performance  will  differ 
between  conditions  where  flatfish  can 
see,  in  contrast  to  where  they  cannot 
see,  the  approaching  gear. 


Manuscript  submitted  8 June  2009. 
Manuscript  accepted  15  December  2009. 
Fish.  Bull.  108:145-154  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


Clifford  H.  Ryer1  (contact  author) 
Craig  S.  Rose2 
Paul  J.  Iseri1 


Trawl  catches  for  many  fish  species 
exhibit  diel  patterns  (Casey  and  Myers, 
1998).  This  is  generally  viewed  as  a 
product  of  two  independent  factors: 
availability  and  catchability  of  the 
fish  species.  Many  gadids  exhibit  diel 
changes  in  availability  associated  with 
vertical  migration  (Beamish,  1965; 
Casey  and  Myers,  1998;  Schabets- 
berger  et  al.,  2000).  Gadids  aggregate 
close  to  the  bottom  during  the  day  and 
are  highly  available  to  bottom  trawls. 
At  night,  dispersal  into  the  overly- 
ing water  renders  them  less  available. 
Interestingly,  for  many  flatfish  species 
the  opposite  pattern,  higher  catches 
at  night,  has  been  observed  (Walsh, 
1991;  Walsh  and  Hickey,  1993;  Casey 
and  Myers,  1998).  Seasonal  migra- 
tions will  occasionally  take  flatfish 
into  the  water  column  (Metcalfe  et  al., 
1990;  Nichol  and  Sommerton,  2009), 
as  will  the  occasional  exploitation  of 
pelagic  prey.  However,  under  normal 
circumstances  many  flatfish  species 
appear  to  remain  on  the  bottom  and 
are  consistently  available  to  trawl 
gear,  day  or  night.  As  a consequence, 
greater  flatfish  catches  at  night  are 
thought  to  be  associated  with  higher 
catchability,  that  is,  with  a decreased 
ability  to  evade  capture  (Ryer,  2008). 

Video  cameras  mounted  on  trawls, 
as  well  as  supplemental  bag  nets  be- 
hind the  main  net,  have  documented 


extensive  flatfish  escapement  beneath 
the  footrope  during  the  day  (Main 
and  Sangster,  1981;  Walsh,  1988).  For 
obvious  technical  reasons,  behavior 
in  front  of  the  footrope,  or  sweeps,  at 
night  has  not  been  observed  in  field 
studies,  except  with  flash  photogra- 
phy (Walsh  and  Hickey,  1993).  How- 
ever, laboratory  experiments  indicate 
that  northern  rock  sole  ( Lepidopsetta 
polyxystra),  Pacific  halibut  ( Hippo- 
glossus stenolepis),  and  English  sole 
( Parophrys  vetulus)  are  more  likely 
to  rise  or  hop  into  the  water  column 
during  darkness,  than  to  herd  (Ryer 
and  Barnett,  2006).  By  moving  off  the 
bottom,  these  fish  remove  themselves 
from  the  “zone  of  influence”  of  the 
ground  gear,  and  as  they  cease  swim- 
ming they  pass  over  the  footrope  and 
into  the  net.  This  behavior  potentially 
explains  why  flatfish  are  captured  in 
greater  numbers  at  night. 

This  paradigm,  i.e.,  higher  flatfish 
catches  at  night,  stems  largely  from  a 
series  of  published  studies  (Main  and 
Sangster,  1981;  Walsh,  1988,  1991; 
Walsh  and  Hickey,  1993;  Casey  and 
Myers,  1998;  and  references  therein), 
based  on  survey  trawls.  On  survey 
trawls,  the  combined  length  of  bridles 
and  sweeps  is  typically  minimized.  In 
contrast,  on  commercial  flatfish  trawls 
lengthy  sweeps  are  used  to  herd  fish 
inward  toward  the  net  (Winger  et 


146 


Fishery  Bulletin  108(2) 


al.,  1999,  2004).  On  some  modern  flatfish  trawls  these 
sweeps  may  be  up  to  400  m in  length,  and  as  much 
as  90%  of  the  seafloor  is  subject  to  the  action  of  gear 
which  is  designed  to  affect  capture  by  manipulating 
flatfish  swimming  behavior.  But  for  the  very  reason  that 
footropes  are  more  efficient  in  the  dark,  sweeps  may 
be  less  efficient.  If  flatfish,  unable  to  see  the  approach- 
ing sweep,  rise  or  hop  into  the  water  column,  rather 
than  herding  as  happens  during  the  day,  they  will  pass 
over  the  sweep  and  be  lost  to  the  catch.  This  situation 
raises  the  possibility  that  flatfish  trawls  that  rely  upon 
sweep  herding  may  capture  more  flatfish  during  the  day 
than  during  the  night — a pattern  not  seen  with  survey 
trawls,  which  have  minimal  sweeps. 

In  this  study  we  investigated  the  performance  of 
trawls  equipped  with  sweeps  under  day  and  night  con- 
ditions, using  a combination  of  manipulative  at-sea  and 
laboratory  procedures.  For  our  at-sea  experiment,  we 
used  a data  set  acquired  during  a series  of  cruises  in 
the  eastern  Bering  Sea,  the  goal  of  which  was  to  evalu- 
ate sweeps  designed  to  reduce  damage  to  benthic  habi- 
tat (Rose  et  al.,  2010).  In  brief,  trawling  was  conducted 
with  sweeps  that  were  elevated,  to  various  degrees,  off 
the  seafloor  to  evaluate  the  trade-off  between  reductions 
in  habitat  disturbance  and  decreased  flatfish  herding 
efficiency.  Here  we  test  hypotheses  related  to  our  prin- 
ciple premise:  flatfish  behavior  initiated  by  ground-gear 
is  principally  controlled  by  ambient  light  levels.  More 
specifically,  first  we  test  the  hypothesis  that  trawls 
configured  with  control  (commercial  type)  sweeps  in 
contact  with  the  bottom,  will  catch  more  flatfish  dur- 
ing the  day  than  during  the  night.  Following  from  this, 
we  test  a second  related  hypothesis:  the  elevation  of 
sweeps  off  the  bottom  will  have  differential  effects, 
day  as  opposed  to  night.  During  the  day,  elevation  will 
reduce  sweep  efficiency,  resulting  in  lower  flatfish  catch. 
During  the  night,  because  sweeps  are  already  rela- 
tively ineffective,  elevation  of  the  sweeps  will  have  no 
influence  upon  their  efficiency,  as  reflected  by  flatfish 
catch.  Lastly,  we  conducted  comparable  experiments 
under  both  light  and  dark  conditions,  using  simulated 
ground-gear  in  the  laboratory  where  behavior  could  be 
quantified,  to  ascertain  whether  the  proposed  effects 
of  elevated  sweeps  on  catch  are  directly  attributable  to 
ambient-light-mediated  differences  in  flatfish  behavior 
in  relation  to  ground  gear. 

Methods 

At-sea  experiments 

Tows  of  paired  trawls  (control  and  elevated  sweeps) 
were  conducted  during  September  2007  in  the  eastern 
Bering  Sea  onboard  the  FV  Cape  Horn.  Details  of  gear 
and  onboard  procedures  can  be  found  in  Rose  et  al. 
(2010).  Briefly,  the  Cape  Horn  is  a 47-m  trawler  proces- 
sor, configured  so  as  to  allow  twin  trawling,  i.e.,  fishing 
with  two  identical  nets  side-by-side.  Each  net  had  a 
set  of  independent  180-m  sweeps,  being  spread  by  one 


otter  board  on  each  side  of  the  vessel,  and  connected 
in  the  middle  by  a towed  weight  (clump).  The  sweeps 
were  composed  of  5-cm  diameter  combination  rope,  con- 
structed of  steel  cable  and  covered  by  polyethylene  fiber. 
Modifying  the  sweeps  on  one  net,  while  keeping  all  other 
trawl  characteristics  consistent,  allowed  the  difference 
between  the  two  catches  to  reflect  the  effect  of  the  modi- 
fication. In  this  field  study,  disk  clusters  were  attached 
to  the  experimental  sweeps  at  9-m  intervals.  The  disks 
were  either  15,  20,  or  25  cm  in  diameter.  This  created 
a nominal  spacing  between  the  sweeps  and  the  seafloor 
of  5,  7.5,  and  10  cm,  respectively.  Test  tows  were  made 
with  modified  sweeps  on  one  net  and  unmodified  sweeps 
on  the  other.  Halfway  through  each  experiment,  the 
modified  sweeps  and  unmodified  sweeps  were  switched 
(left  to  right,  right  to  left). 

Catches  from  each  trawl  were  kept  separate  until  the 
entire  catch  had  been  sampled.  As  catches  entered  the 
sampling  area,  they  were  passed  across  a motion-com- 
pensated flow  scale  to  determine  total  catch  weight.  The 
five  or  six  most  abundant  species  were  then  completely 
sorted  into  holding  bins.  Fish  from  each  bin  were  then 
run  across  a second  flow  scale  to  measure  the  weight  of 
each  of  those  species.  To  estimate  the  weight  of  other 
species,  samples  of  the  unsorted  catch  were  taken  at 
intervals,  sorted,  and  weighed  by  species.  The  com- 
position of  these  samples  was  then  expanded  to  the 
weight  of  the  entire  catch  by  calculating  the  fraction 
of  the  sample  weight  to  the  total  catch  weight.  For  the 
species  cited  in  this  paper,  Pacific  halibut  and  Alaska 
plaice  catches  were  estimated  from  the  samples  and 
all  other  species  were  fully  weighed  on  the  second  flow 
scale.  During  the  sorting  phase,  samples  of  50-150  fish 
of  each  species  were  drawn  and  measured  to  determine 
their  length  composition.  Length  samples  were  taken 
from  throughout  the  catch  as  it  passed  through  the 
sorting  area  and  the  length  of  each  individual  in  the 
sample  was  measured 

Sixty-one  paired  hauls  were  made  over  depths  rang- 
ing from  70  to  117  m.  Ambient  light  on  the  bottom  is 
greatly  influenced  by  water  depth.  To  minimize  poten- 
tial depth  effects  upon  ambient  light,  we  limited  our 
analysis  to  hauls  where  depth  was  between  79  and 
94  m:  a 15-m  range.  In  addition,  we  eliminated  hauls 
where  large  debris  (crab  pots,  etc.)  were  encountered, 
or  where  gear  components  became  entangled,  assuming 
that  such  conditions  would  influence  gear  performance 
and  catch.  After  examining  in  situ  light  measurements 
(Wildlife  Computers,  MK9  light  meter,  Redmond,  WA) 
we  further  eliminated  daytime  hauls  where  light  levels 
fell  below  l.OxlO-4  pmol  photons/m2/s,  and  nighttime 
hauls  exceeding  1.0xl0~5  pmol  photons/m2/s.  This  step 
eliminated  hauls  made  around  dusk  or  dawn  and  set 
a clear  differentiation  between  daytime  and  nighttime 
light.  In  the  resulting  data  set  (36  hauls),  mean  tow 
depth  did  not  differ  between  nighttime  and  daytime 
tows  (day:  n= 7,  mean  [x]  = 82  m,  standard  error  [SE]  = 1; 
night:  n =19,  x=84  m,  SE  = 1;  t(34]=1.54,  P=0.133).  Tow 
durations  ranged  from  33  to  150  min,  being  somewhat 
longer  at  night  (x=115.8,  SE  = 5.9)  than  during  the  day 


Ryer  et  al.:  Flatfish  herding  behavior  in  response  to  trawl  sweeps 


147 


(x-  87.5,  SE  = 6.3,  £[34]  = 3.28,  P = 0.003).  During  long 
tows,  accumulating  catch  can  distort  meshes  and  back 
up  into  the  intermediate  portion  of  the  net,  altering 
gear  selectivity  (Herrmann,  2005).  However,  catches  in 
this  study  were  small  compared  to  net  capacity,  never 
filling  the  codend.  Hence,  we  assume  that  differences 
in  duration  between  day  and  night  did  not  influence 
net  performance  or  fish  catchability  in  a manner  that 
would  bias  our  results.  Similarly,  during  long  tows 
proportionately  more  fish  will  tire  and  fall  back  into 
the  net,  particularly  so  for  many  roundfish  species, 
which  can  swim  for  prolonged  periods  in  front  of  the 
net  (Main  and  Sangster,  1981).  However,  flatfish  typi- 
cally swim  for  less  than  1 minute  in  front  of  nets  ( Ryer, 
2008),  and  thus  this  source  of  bias  was  also  unlikely 
in  our  study. 

For  our  first  analysis,  we  compared  daytime  and 
nighttime  catches  from  the  control  nets  only;  where 
sweeps  were  in  contact  with  the  bottom  along  their 
entire  length.  Catch  per  unit  of  effort  (CPUE:  kg/min) 
was  calculated  for  total  catch  (all  species)  as  well  as  for 
six  flatfish  species:  yellowfish  sole  ( Limanda  aspera ); 
flathead  sole  ( Hippoglossoides  elassodon);  arrowtooth 
flounder  ( Atheresthes  stomias)\  rock  sole  ( Lepidopsetta 
spp. );  Alaska  plaice  ( Pleuronectes  quadrituberculatus); 
and  Pacific  halibut.  CPUE  values  were  natural  log  (In) 
transformed  and  tested  for  day  and  night  differences 
with  t-tests  (Sokal  and  Rohlf,  1969).  Where  variances 
were  heteroscedastic,  Satterthwaite’s  adjusted  degrees 
of  freedom  were  used  (Snedecor  and  Cochran,  1980).  Be- 
cause CPUE  was  based  upon  weight,  we  also  compared 
mean  total  length  between  daytime  and  nighttime  hauls 
for  each  flatfish  species. 

For  our  second  analysis,  we  used  the  subset  of  samples 
from  trawls  where  25.4-cm  disks  were  attached  to  el- 
evate sweeps  of  the  experimental  net  to  an  approximate 
height  of  10  cm  (the  distance  between  sediment  surface 
and  bottom  of  the  sweep  material).  For  this  analysis, 
catch  of  the  experimental  net  was  compared  to  that  of 
the  paired  control  net  (with  bottom  contact  sweeps)  by 
using  a paired  /-test  (Sokal  and  Rohlf,  1969).  Separate 
analyses  were  conducted  for  daytime  (rc  = 10  pairs)  and 
nighttime  (n  = 5 pairs)  hauls.  Similar  analysis  was  con- 
ducted for  flatfish  lengths. 

Laboratory  experiments 

Northern  rock  sole  and  Pacific  halibut  were  collected 
as  age-0  juveniles  by  using  a 2-m  plumb-staff  beam 
trawl  from  Chiniak  Bay,  Kodiak,  Alaska.  Fish  were 
transported  to  the  Hatfield  Marine  Science  Center  in 
Oregon  and  reared  in  2. 2-m  (diameter)  circular  tanks 
with  flow-through  seawater  (28-35%o,  9°C  [±  1°])  on  a 
diet  of  krill  and  gelatinized  food.  After  reaching  age  1, 
fish  were  transferred  to  3-m  diameter  pools  for  contin- 
ued growth. 

Simulated  sweep  exposure  took  place  in  an  elongated 
tank  (10.7x1.5x1.2  m)  filled  to  a depth  of  0.9  m.  This 
tank  was  provided  with  flow-through  seawater  (28-35%e) 
and  located  in  a light-proof  room,  allowing  for  control  of 


illumination  by  an  overhead  bank  of  fluorescent  lamps. 
The  tank  bottom  was  covered  to  a depth  of  4 cm  with 
sand,  allowing  flatfish  to  completely  bury  themselves. 
Details  of  this  apparatus  are  presented  elsewhere  (Ryer 
and  Barnett,  2006)  and  will  only  be  described  briefly 
here.  By  means  of  a moveable  carriage  a simulated 
sweep  was  propelled  down  the  length  of  the  tank.  This 
sweep  consisted  of  a piece  of  5-cm  diameter  PVC  pipe, 
painted  green  to  resemble  the  actual  sweep  used  in  the 
field  study.  It  could  be  positioned  so  that  it  ran  down 
the  tank  in  contact  with  the  bottom,  or  elevated  so  that 
it  was  approximately  10  cm  off  the  bottom. 

Fish  were  maintained  on  a 12/12  h photo  period 
during  all  experiments,  with  lights  turned  on  at  0700 
and  off  at  1900.  At  1600  on  the  day  before  the  trials, 
the  length  of  the  tank  was  subdivided  into  three  equal 
3-m  sections,  by  means  of  four  removable  partitions, 
of  which  two  of  these  partitions  prevented  fish  from 
moving  to  the  extreme  ends  of  the  tank.  Next,  fish 
were  introduced  to  each  of  the  three  main  sections  of 
the  tank.  This  sectioning  assured  that  fish  would  not 
aggregate  in  a single  area  of  the  tank.  At  0800  on  the 
day  of  trials,  the  footrope  carriage  was  lowered  into 
the  tank,  behind  one  of  the  end  partitions  and  secured 
to  its  tracks.  Then  the  lighting  was  either  turned  off 
(dark  trials)  or  kept  on  (light  trials),  and  fish  were 
allowed  2 h acclimation  before  a trial.  Illumination  at 
the  sand  surface  was  measured  once  at  the  beginning 
of  the  study.  For  light  trials,  illumination  was  approxi- 
mately 1.5  pmol  photons/m2/s  (-125  lux),  whereas,  for 
dark  trials  illumination  was  <lxl0-8  pmol  photons/m2/s 
(~10-6  lux).  Both  species  used  in  this  study  have  the 
same  light  thresholds  ( 10  5 pmol  photons/m2/s)  for  vi- 
sual discrimination  of  small  motile  prey  (Hurst  et  al., 
2007),  and  we  assumed  they  would  see  approaching 
footrope  in  the  light  trials,  but  not  in  the  dark  trials. 
Illumination  was  measured  with  a research  radiometer 
(International  Light  Inc.,  Model  IL1700,  Peabody,  MA) 
equipped  with  a 2ji  PAR  (photosynthetically  active  ra- 
diation) sensor.  Water  supply  to  the  tank  was  filtered 
through  sand,  making  it  unlikely  that  water  clarity, 
and  hence  light  levels,  changed  appreciably  from  day 
to  day.  At  1000  h,  immediately  before  a trial,  the  parti- 
tions were  removed;  for  dark  trials  red  flashlights  were 
used  during  this  process,  and  care  was  taken  to  avoid 
shining  the  lights  directly  into  the  tank.  Five  minutes 
later  the  footrope  carriage  was  pulled  from  one  end  of 
the  tank  to  the  other  at  a speed  of  1.0  m/s  (±  0.1  m/s),  a 
speed  roughly  equal  3.6  km/h  or  2 knots;  flatfish  trawls 
are  commonly  towed  at  2-5  knots.  Afterwards,  the 
lights  in  the  room,  if  turned  off,  were  turned  back  on 
and  rakes  were  used  to  herd  fish  back  into  each  of  the 
three  main  sections  of  the  tank,  after  which  the  parti- 
tions were  put  back  in  place  and  the  footrope  carriage 
was  removed  from  the  tank.  This  entire  process  was 
repeated  in  the  afternoon,  using  the  opposite  lighting 
from  that  of  the  morning:  at  1200  h,  a footrope  carriage 
was  lowered  into  the  tank  and  lighting  was  adjusted; 
at  1400  h,  partitions  were  removed  and  the  footrope 
carriage  was  pulled.  We  assume  that  this  alternation 


148 


Fishery  Bulletin  108(2) 


in  treatment  order  precluded  any  bias  attributable  to 
flatfish  habituation  or  learning. 

Positioned  behind  and  above  the  footrope  were  three 
(50W)  infrared  LED  (light  emitting  diode)  lamps,  aimed 
forward  and  down,  so  that  they  illuminated  the  footrope 
and  tank  bottom  immediately  in  front  of  the  footrope. 
The  wavelength  of  light  emitted  by  these  lamps  peaked 
at  880  nm,  and  emissions  dropped  to  0 below  760  nm. 
Most  fish  are  insensitive  to  light  at  those  wavelengths 
(Douglas  and  Hawryshyn,  1990)  and  results  from  light- 
threshold  feeding  studies  for  all  three  flatfish  species 
used  in  this  study  are  consistent  with  this  generaliza- 
tion (Hurst  et  al.,  2007).  Two  underwater  video  cameras 
(Aqua-Vu,  model  ZT-120,  Crosslake,  MN ) were  mounted 
alongside  the  lamps,  also  directed  at  the  area  in  front 
of  the  footrope.  This  arrangement  allowed  for  visual 
monitoring  out  to  1.1  m in  advance  of  the  footrope.  The 
video  footage  was  captured  from  a remote  location  by 
digital  mini-DV  recorders. 

Trials  were  conducted  with  three  age  classes  of  Pa- 
cific halibut:  age-1,  age-2,  and  age-3,  as  well  as  age- 
2 northern  rock  sole.  For  age-3  Pacific  halibut,  three 
groups  of  five  fish  each  were  examined.  Trials  took 
place  over  two  consecutive  days.  On  the  first  day  sweep 
height  was  randomly  set  to  either  the  “in  contact”  or 
“elevated”  position.  On  the  second  day  the  alternative 
position  was  used.  During  each  day,  fish  were  exposed 
to  the  simulated  sweep  approach  twice;  once  in  the  light 
and  once  in  the  dark.  The  order  of  application  of  light 
vs.  dark  trials  was  also  randomly  determined.  After  the 
second  day  fish  were  then  removed  from  the  tank,  their 
total  length  was  measured,  and  they  were  replaced  by 
a new  group.  Age-3  Pacific  halibut  ranged  from  37-52 
cm  in  total  length. 

For  age-2  Pacific  halibut,  age-1  Pacific  halibut,  and 
age-2  rock  sole,  groups  consisting  of  10  fish  each  were 
trialed  differently.  Each  group  was  trialed  for  only  a 
single  day,  at  one  sweep  height.  For  age-2  Pacific  hali- 
but, six  groups  were  trialed  at  each  sweep  height.  For 
age-1  Pacific  halibut  and  age-2  northern  rock  sole,  five 
groups  were  trialed  at  each  sweep  height.  As  before, 
the  order  of  light  and  dark  trials  was  randomized.  Age- 
2 Pacific  halibut  ranged  from  19-31  cm,  age-1  hali- 
but from  8-14  cm,  and  age-2  northern  rock  sole  from 
9-17  cm. 

Fish  behavior  was  quantified  by  using  the  slow-mo- 
tion  playback  of  digital  video.  First,  the  number  of 
fish  encountered,  i.e.,  observed,  as  the  sweep  made 
its  transit  from  one  end  of  the  tank  to  the  other,  was 
recorded  from  each  trial.  Then  the  initial  behavioral 
response  of  each  observed  fish  was  assigned  to  one 
of  four  categories:  1)  pass  under,  2)  hop,  3)  rise,  and 
4)  herd.  Fish  characterized  by  “under”  either  did  not 
react  at  all  to  the  approaching  sweep,  or  reacted  when 
contacted  by  the  sweep,  but  passed  under  the  sweep 
as  it  progressed  down  the  tank.  “Hop”  characterized 
fish  that  reacted  to  the  sweep  with  one  or  two  sinu- 
soidal body  undulations,  typically  after  being  struck 
by  the  sweep,  which  resulted  in  the  fish  “hopping”  off 
the  substrate.  However,  this  initial  startle  reaction 


< 

>. 

Q. 


10  ’ 

• 

Night 

10  2 
10  3 

o 

O O 
■ Qr-O- 

o 

oo 

o 

Day 

10  4 

o 

o c 
o 

1 

Q 

o 

10  s 

• 

10  6 

• • X 

• 

• • 

• % 

• 

• 

• — 

78 


82 


86 

Depth 


90 


94 


Figure  1 

In  situ  natural  log-transformed  light  data  for  trawl  tows 
conducted  during  day  and  night,  plotted  by  mean  depth 
over  the  course  of  each  tow.  Regression  analysis  indi- 
cated no  effect  of  depth  upon  ambient  light  over  this 
relatively  narrow  range  of  depths  and  hence,  regressions 
are  plotted  as  zero-slope  lines. 


was  not  followed  by  any  further  swimming,  such  that 
the  fish  tended  to  hang  stationary  in  the  water,  and 
passed  over  the  sweep  as  it  progressed  down  the  tank. 
“Rise”  characterized  the  motion  of  fish  that  departed 
the  bottom  with  sustained  swimming  in  an  upward 
direction,  such  that  the  distance  between  fish  and  bot- 
tom continuously  increased  as  the  fish  swam.  This  was 
in  contrast  to  fish  characterized  by  “herd”  where  fish 
maintained  a distance  of  less  than  one  body  length  be- 
tween themselves  and  the  bottom  as  they  swam  along 
in  front  of  the  sweep,  i.e.  herding  behavior.  Ryer  and 
Barnett  (2006)  investigated  whether  initial  orienta- 
tion, i.e.,  the  direction  fish  were  facing,  influenced 
behavioral  response.  No  relationship  was  observed,  and 
consequently,  no  data  on  fish  orientation  were  recorded 
in  this  study.  Categorical  data  on  behavioral  response 
were  pooled  across  replicate  groups  and  analyzed  by 
contingency  table  analysis  by  using  log-linear  models 
(Fienberg,  1980). 


Results 

At-sea  experiment 

Mean  ambient  light  on  the  seafloor  (Fig.  1)  was  greater 
during  daytime  tows  (2.0xl0~3  pmol  photons/m2/s) 
than  during  nighttime  tows  (8.4xl0~7  pmol  photons/ 
m2/s,  F(1  33]=352.76,  P<0.001).  However,  over  the  rela- 
tively narrow  range  of  tow  depths  used  in  this  analy- 
sis, depth  had  no  influence  upon  bottom  ambient  light 
level  (Fu  33]=0.27,  P=0.607).  Mean  total  catch  (CPUE) 
in  terms  of  weight  (kg/min)  was  greater  during  the 
day  than  at  night  (Table  1,  day:  x=100.6  kg,  SE  = 9.61; 
night:  x=53.07  kg,  SE  = 6.14).  This  pattern  of  diurnally 


Ryer  et  al.:  Flatfish  herding  behavior  in  response  to  trawl  sweeps 


149 


LU 


LU 


C Arrowtooth  flounder 


LU 

q_  E Pacific  halibut 


Figure  2 

Mean  catch  per  unit  of  effort  (CPUE)  ±1  standard  error  (SE)  for  daytime  and  nighttime 
hauls  for  each  of  six  flatfish  species  from  control  nets  where  the  sweep  was  in  contact  with 
the  seafloor:  (A)  yellowfin  sole  (Limanda  aspera);  (B)  flathead  sole  ( Hippoglossoides  elas- 
sodon)\  (C)  arrowtooth  flounder  (Atheresthes  stomias );  (D)  rock  sole  (Lepidopsetta  spp.); 
(E)  Alaska  plaice  ( Pleuronectes  quadrituberculatus);  and  (F)  Pacific  halibut  ( Hippoglossus 
stenolepis ). 


Table  1 

Statistics  for  a comparison  of  day  and  night  trawl  catches,  by  total  catch,  and  catch  of  six  individual  species  of  flatfish.  For  both 
day  and  night  tows,  trawl  nets  were  equipped  with  control  sweeps  (that  had  contact  with  the  bottom).  Where  needed,  Satterth- 
waite’s  adjusted  degrees  of  freedom  were  used  to  mitigate  for  nonhomogeneity  of  variance. 

Species 

t-test  statistic 

df 

P value 

Total  catch 

4.85 

31.3 

<0.001 

Yellowfin  sole  (Limanda  aspera ) 

1.71 

30.6 

0.097 

Flathead  sole  (Hippoglossoides  elassodon ) 

-7.44 

34 

<0.001 

Arrowtooth  flounder  (Atheresthes  stomias) 

-3.26 

34 

0.003 

Rock  sole  (Lepidopsetta  spp.) 

-2.38 

29.3 

0.024 

Alaska  plaice  (Pleuronectes  quadrituberculatus) 

-3.74 

26.4 

0.001 

Pacific  halibut  (Hippoglossus  stenolepis) 

1.58 

34 

0.126 

larger  catches  was  also  exhibited  by  four  out  of  six 
flatfish  species  examined  (Table  1,  Fig.  2).  Flathead 
sole,  arrowtooth  flounder,  rock  sole,  and  Alaska  plaice 
were  all  characterized  by  higher  CPUE  during  the  day. 
Yellowfin  sole  and  Pacific  halibut  exhibited  no  signifi- 
cant differences  in  catch  between  day  and  night.  Of  the 
four  species  for  which  fish  total  length  was  measured 
in  catch  subsamples  (i.e.,  yellowfin  sole,  flathead  sole, 
arrowtooth  flounder,  and  rock  sole),  fish  tended  to  be 
slightly  larger  at  night.  This  was  only  statistically  sig- 
nificant for  yellowfin  sole  (f[24  4]=3.93,  P=0.001),  where 
fish  averaged  1 cm  longer  during  the  night  (x=32.8  cm, 
SE  = 0.2)  than  during  the  day  (x=31.8  cm,  SE  = 0.1),  and 
rock  sole  (f[32]=2.91,  P=0.006),  where  fish  averaged  0.9 


cm  longer  during  the  night  (*=33.3  cm,  SE  = 0.2)  than 
during  the  day  (5;=32.4  cm,  SE  = 0.2). 

The  effect  of  elevating  sweeps  10  cm  off  the  bottom 
differed,  depending  upon  whether  tows  were  made  dur- 
ing the  day  or  night  (Fig.  3).  During  the  day,  total  catch 
tended  to  decrease  when  sweeps  were  elevated  (Table 
2,  elevated:  5c=93.4,  SE  = 8.7;  control:  3c=100.6,  SE  = 9.6). 
However,  during  the  night,  elevation  of  sweeps  had  little 
influence  upon  catch  (elevated:  x=55.1,  SE  = 6.8;  control: 
x=53.1,  SE  = 6.1).  This  same  pattern  was  evident  for  four 
out  of  six  flatfish  species  examined.  Species  for  which 
daytime  elevation  of  sweeps  decreased  catch  included 
flathead  sole,  arrowtooth  flounder,  rock  sole,  and  Alaska 
plaice.  Sweep  configuration  had  no  significant  effect  on 


150 


Fishery  Bulletin  108(2) 


A Yellowfin  sole  B Flathead  sole  C Arrowtooth  flounder 


E 

2: 

wj 


LU 

3 

Q. 

o 


Figure  3 

Mean  catch  per  unit  of  effort  (CPUE)  ±1  standard  error  (SE)  for  daytime  and  nighttime 
catches  of  each  of  six  flatfish  species  from  both  control  nets,  where  the  sweep  was  in  contact 
with  the  seafloor,  as  well  as  experimental  nets  where  the  sweep  was  elevated  10.2  cm  off 
the  seafloor:  (A)  yellowfin  sole  ( Limanda  aspera);  (B)  flathead  sole  ( Hippoglossoides  elas- 
sodon)\  (C)  arrowtooth  flounder  ( Atlieresthes  stomias);  (D)  rock  sole  ( Lepidopsetta  spp.); 
(E)  Alaska  plaice  ( Pleuronectes  quadrituberculatus)\  and  (F)  Pacific  halibut  (Hippoglossus 
stenolepis). 


Table  2 

Statistics  for  comparison  of  total  catch  and  catch  of  six  individual  species  of  flatfish  between  trawl  nets  equipped  with  control 
(bottom  contact)  and  those  equipped  with  elevated  (10  cm  off  bottom)  sweeps,  from  both  day  and  night  tows. 

Species 

Paired  t-test  statistic 

df 

P value 

Total  catch 

Day 

2.11 

9 

0.064 

Night 

-0.22 

4 

0.834 

Yellowfin  sole  ( Limanda  aspera ) 

Day 

1.84 

9 

0.099 

Night 

0.09 

4 

0.935 

Flathead  sole  (Hippoglossoides  elassodon) 

Day 

2.33 

9 

0.045 

Night 

-0.78 

4 

0.481 

Arrowtooth  flounder  (Atheresthes  stomias ) 

Day 

4.35 

9 

0.002 

Night 

-0.71 

4 

0.519 

Rock  sole  (Lepidopsetta  spp.) 

Day 

5.42 

9 

<0.001 

Night 

0.23 

4 

0.830 

Alaska  plaice  (Pleuronectes  quadrituberculatus) 

Day 

2.39 

9 

0.041 

Night 

-0.67 

4 

0.539 

Pacific  halibut  (Hippoglossus  stenolepis) 

Day 

-0.59 

9 

0.753 

Night 

0.29 

4 

0.785 

daytime  catches  of  yellowfin  sole  or  Pacific  halibut.  In 
contrast  to  daytime  results,  elevated  sweeps  had  no  ef- 
fect upon  nighttime  catches  for  any  species.  Of  the  four 
species  that  were  measured,  fish  lengths  did  not  differ 
between  tows  with  elevated  sweeps  and  control  tows, 
regardless  of  time  of  day  (P>0.05  for  each  species,  day 
and  night). 


Laboratory  experiment 

Overall,  28%  of  fish  initiated  herding  behavior  in 
response  to  simulated  sweep  disturbance.  Herding  was 
most  prevalent  in  the  light,  and  tended  to  be  replaced 
by  fish  passing  under  the  sweep,  as  well  as  hopping  or 
rising  off  the  bottom  in  the  dark  (Fig.  4).  There  was  also 


Ryer  et  al.:  Flatfish  herding  behavior  in  response  to  trawl  sweeps 


151 


A Age-3  Pacific  halibut — light 


Age-3  Pacific  halibut — dark 


B Age-2  Pacific  halibut — light 


Age-2  Pacific  halibut — dark 


C Age-1  Pacific  halibut — light 


50 

40 

30 

20 

10 

0 


Age-1  Pacific  halibut — dark 


D Age-2  Northrock  sole — light  Age-2  Northrock  sole — dark 


Figure  4 

Behavioral  response  of  flatfish,  under  light  and  dark  conditions,  with  the  simulated  sweep  both 
in  contact  (control)  and  elevated  10  cm  off  the  bottom:  (A)  age-3  Pacific  halibut  ( Hippoglossus 
stenolepis)\  (B)  age-2  Pacific  halibut;  (C)  age-1  Pacific  halibut;  and  (D)  age-2  northern  rock  sole 
( Lepidopsetta  polyxystra).  “Pass  under”  represents  fish  that  either  did  not  react  to  the  sweep, 
or  reacted  late,  such  that  they  passed  under  the  sweep  as  it  progressed  down  the  tank.  “Hop” 
characterized  fish  that  reacted  to  the  sweep  with  one  or  two  body  undulations,  but  almost 
immediately  pass  over  the  sweep.  “Rise”  characterized  fish  in  which  the  initial  jump  off  the 
bottom  was  followed  by  sustained  swimming  in  an  upward  direction,  such  that  the  distance 
between  fish  and  bottom  continuously  increased  as  the  fish  swam.  “Herd”  characterized  fish 
which,  after  reacting  to  the  gear,  swam  along  in  front  of  the  sweep,  close  to  the  bottom,  typically 
maintaining  a distance  of  less  than  one  body  length  between  themselves  and  the  bottom. 


a tendency  for  herding  in  the  light  to  decrease  when 
the  sweep  was  elevated.  These  observations  are  sup- 
ported by  results  of  log-linear  model  analysis,  in  which 
ambient  light  (light,  dark)  mediated  the  influence  of 


sweep  height  upon  behavioral  response  (G[3|  = 9.96, 
P=0.019).  All  three  age  classes  of  Pacific  halibut,  and 
northern  rock  sole,  behaved  comparably;  there  were 
no  significant  effects  of  species  or  age  on  the  type  of 


152 


Fishery  Bulletin  108(2) 


response  displayed,  or  interactions  with  light  level  or 
sweep  height  (P>0.05  for  all).  Examination  of  Figure  4 
could  lead  one  to  conclude  that  age-3  halibut  behaved 
somewhat  differently  than  the  other  species  and  age 
groups.  However,  the  number  of  age-3  halibut  tested 
(;z=15)  was  small  compared  to  each  of  the  other  species 
and  age  groups  (n  >50  for  each),  and  as  a consequence, 
had  little  influence  upon  our  statistical  model.  We  pooled 
data  across  species  and  collapsed  response  categories 
down  to  those  fish  that  herded  in  contrast  to  those  that 
did  not  (pass  under,  hop,  and  rise  combined),  so  as  to 
render  the  data  into  a form  most  similar  to  our  at-sea 
trawl-catch  experiments.  Again,  ambient  light  (light 
or  dark)  mediated  the  influence  of  sweep  height  upon 
behavioral  response  (Gtlj  = 5.75,  P=0.017).  In  Figure  5 
we  have  simplified  this  relationship  by  graphing  the 
percentage  of  fish  herding  under  the  two  light  and  sweep 
height  treatments.  In  addition  to  a conspicuous  decrease 
in  herding  in  the  dark,  elevation  of  the  sweep  decreased 
herding  in  the  light  but  had  little  influence  in  the  dark- 
ness— results  consistent  with  those  observed  in  the  at- 
sea  experiment. 

Discussion 

Ambient  illumination  controls  many  aspects  of  fish 
behavior,  from  feeding  and  habitat  use  (Janssen,  1978; 
Helfman  and  Schultz,  1984;  Ryer  and  Olla,  1999;  De 
Robertis  et  ah,  2003;  Petrie  and  Ryer,  2006)  to  social  and 
antipredator  behavior  (Shaw,  1961;  Ryer  and  Olla,  1998). 
Similarly,  light  has  a pervasive  influence  upon  interac- 
tions between  fish  and  trawls.  In  this  study,  field  data 
were  largely  consistent  with  our  principal  hypothesis; 
that  trawls  configured  with  sweeps  that  are  in  contact 
with  the  seafloor  would  catch  more  flatfish  during  the 


60  r 


50  - 

CD 

40 

0 

.c 

E 30  ' 

c 

0 

9 20  - 

(D 

CL 

10  - 


0 L-1 

Light  Dark 

Figure  5 

Percentage  of  fish  that  herded  in  response  to  simulated 
trawl  sweep  disturbance  under  both  light  and  dark  con- 
ditions, with  the  sweep  both  in  contact  (control)  and 
elevated  10  cm  off  the  bottom.  Data  were  pooled  across 
species  and  age  classes. 


day  than  during  the  night.  This  pattern  was  observed 
for  four  out  of  six  flatfish  species  examined:  flathead  sole, 
arrowtooth  flounder,  rock  sole,  and  Alaska  plaice.  Herd- 
ing, as  seen  in  both  roundfish  and  flatfish,  is  an  ordered 
behavioral  response  in  which  fish  move  away  from  an 
approaching  threat,  i.e.,  the  doors,  sweeps,  bridles,  and 
wings  of  the  net.  Through  either  continuous  swimming, 
or  sudden  swimming  bursts,  interspersed  with  rests  on 
the  bottom  (Winger  et  al.,  1999,  2004),  fish  then  funnel 
to  the  center  of  the  gear,  where  they  concentrate  before 
tiring  and  falling  back  into  net.  Several  studies  have 
demonstrated  that  both  roundfish  (Olla  et  al.,  2000;  Ryer 
and  Olla,  2000)  and  flatfish  (Ryer  and  Barnett,  2006) 
lose  the  ability  to  orient  themselves  in  relation  to  gear 
and  initiate  herding  when  ambient  light  falls  below  the 
threshold  for  visual  perception  of  the  gear  (Kim  and 
Wardle,  1998a,  1998b). 

Given  the  brief  evolutionary  time  during  which  fish 
have  interacted  with  towed  fishing  gear,  approximate- 
ly 100  years,  it  is  unlikely  that  specific  gear  avoid- 
ance behavior  has  evolved.  Rather,  we  consider  it  most 
parsimonious  to  assume  gear  avoidance  is  rooted  in 
antipredator  behavior.  Although  flatfish  may  initially 
erupt  from  the  seafloor  upon  being  disturbed  by  trawl 
ground-gear,  as  when  attacked  by  a predator,  subse- 
quent herding  behavior  is  consistent  with  “distance 
keeping”  behavior,  during  which  the  fish  attempts  to 
maintain  a safe  distance  between  itself  and  a slowly 
pursuing  predator.  Scuba  and  skin  divers  who  have  at- 
tempted to  follow  fish  along  the  seafloor  are  certainly 
familiar  with  this  behavior.  For  flatfish,  movement  in 
the  vertical  dimension  also  plays  a critical  role  during 
herding.  It  has  been  observed  that  flatfish  remain  close 
to  the  bottom  during  herding,  usually  less  than  half  a 
body  length  (Ryer,  2008).  Staying  close  to  the  bottom 
reduces  drag,  lessening  thrust  requirements  to  achieve 
a given  speed — the  ground  effect  (Videler,  1993;  Gib- 
son, 2005).  Rising  off  the  bottom  makes  flatfish  more 
conspicuous,  and  due  to  the  location  of  a flatfish’s  eyes, 
also  interferes  with  visual  tracking  of  a pursuing  preda- 
tor, in  this  case,  the  trawl  ground-gear.  Although  they 
herd  close  to  the  bottom  in  the  light,  Pacific  halibut 
and  northern  rock  sole  respond  differently  to  ground- 
gear  in  the  darkness,  as  demonstrated  by  laboratory 
experiments  (Ryer  and  Barnett,  2006).  Unable  to  see, 
the  fish  respond  to  contact  with  the  ground-gear  ini- 
tially by  hopping  or  swimming  upward  and  away  from 
the  bottom.  Similarly,  in  this  study  the  percentage  of 
fish  moving  off  the  bottom  increased  from  4%  in  the 
light  to  21%  in  darkness,  for  all  species  and  bar  heights 
combined.  Moving  off  the  bottom  in  darkness  probably 
functions  as  an  antipredator  tactic,  making  the  flatfish 
more  difficult  to  follow  and  may  simply  be  the  flatfish 
version  of  the  Mauthner-cell  triggered  (lateral  line) 
startle  response  (Eaton  and  Hackett,  1984). 

Our  second  hypothesis,  that  elevation  of  sweeps  off 
the  bottom,  10  cm  in  this  case,  would  decrease  catch 
during  daylight,  but  not  at  night,  was  also  partial- 
ly supported  by  our  field  experiment.  Again,  four  of 
six  flatfish  species  examined  displayed  the  predicted 


Control 

Elevated 


Ryer  et  al.:  Flatfish  herding  behavior  in  response  to  trawl  sweeps 


153 


catch  pattern.  Arguably,  our  analysis  is  based  upon 
a small  set  of  paired  tows,  particularly  at  night  (n  = 5 
pairs).  Taken  alone,  these  at-sea  trials  might  not  be 
convincing.  However,  these  results  were  mirrored  by 
our  laboratory  experiments,  where  the  elevation  of 
sweeps  decreased  herding  to  a greater  extent  in  the 
light,  compared  to  darkness.  The  elevation  of  sweeps 
had  several  consequences,  all  of  which  were  likely  to 
have  influenced  flatfish  behavior.  First,  because  most 
flatfish  react  to  ground  gear  at  a very  short  distance, 
often  only  after  being  struck,  the  likelihood  that  fish 
would  simply  not  react  and  be  passed  over  by  sweeps 
was  probably  increased  by  sweep  elevation.  Further, 
part  of  the  visual  stimulus  to  herd  that  is  associated 
with  ground  gear  is  the  sand  and  mud  cloud  that  is 
kicked  up  by  the  gear.  This  visual  stimulus  would  be 
absent  or  greatly  diminished  by  sweep  elevation,  fur- 
ther decreasing  the  likelihood  of  flatfish  response.  Our 
laboratory  experiments  with  rock  sole  exhibited  a pat- 
tern of  response  nearly  identical  to  that  seen  in  the 
field  and  indicated  that  passage  under  or  over  the  gear 
was  probably  responsible  for  the  decline  in  herding  as- 
sociated with  sweep  elevation  during  the  day;  in  the 
light,  fish  passing  beneath  the  sweep  increased  by  24% 
when  the  sweep  was  elevated.  Lastly,  even  when  herd- 
ing is  initiated,  it  must  be  maintained.  Flatfish  will 
sometimes  dive  under  ground  gear  when  they  perceive 
a gap  between  the  gear  and  the  bottom — a trait  that 
has  been  used  to  reduce  flatfish  bycatch  (DeAlteris  et 
al.,  1997).  Sweep  elevation  probably  facilitated  such 
escape.  Unfortunately,  our  laboratory  data  were  of  little 
aid  in  evaluating  this  possibility.  Because  of  the  physi- 
cal limitations  of  our  apparatus,  we  characterized  only 
the  initial  behavioral  response  of  fish — not  prolonged 
behavioral  sequences  that  would  characterize  such 
deliberate  escape  tactics. 

Our  field  data  indicate  that  Pacific  halibut  could  have 
a different  pattern  of  availability  or  catchability,  com- 
pared to  that  of  the  other  flatfish  species  we  examined. 
By  virtue  of  size,  Pacific  halibut  stand  apart  from  most 
other  flatfish.  Beyond  three  or  four  years  of  age,  their 
size  likely  renders  them  immune  to  most  predators.  This 
may  make  them  more  likely  to  venture  from  the  bottom, 
as  may  their  piscivorous  diet.  Consequently,  they  may 
be  more  likely  than  other  species  to  rise  off  the  bottom 
and  swim  back  over  sweeps.  If  so,  it  follows  that  most 
of  the  fish  captured  are  those  directly  in  the  path  of 
the  net,  excluding  the  area  swept  by  the  sweeps.  Our 
trawling  operations  tended  to  produce  larger,  albeit  not 
significant,  Pacific  halibut  catches  at  night — a trend  re- 
ported by  commercial  fishermen  as  well.  It  may  be  that 
with  their  greater  speed  and  endurance,  many  halibut 
escape  trawls  during  the  day,  but  at  night  cannot  see 
the  gear  to  coordinate  their  escape.  In  contrast  to  the 
halibut  results,  the  nonsignificant  differences  for  yel- 
lowfin  sole  were  similar  in  direction  and  magnitude  to 
the  significant  differences  detected  for  the  other  small 
flatfishes.  This  finding  opens  the  possibility  that  these 
flatfishes  had  similar  reactions,  but  our  experiment  just 
did  not  have  the  statistical  power  to  detect  them. 


Diel  patterns  of  catch  in  trawl  fisheries  and  surveys 
reflect  not  only  patterns  in  fish  availability,  but  gear- 
specific  behavioral  influences  upon  catchability  that 
are  directly  controlled  by  ambient  illumination.  Results 
of  our  laboratory  experiments,  along  with  earlier  ex- 
periments (Ryer  and  Barnett,  2006),  indicate  that  trawl 
footropes  are  likely  to  be  more  efficient  at  displacing 
flatfish  from  the  bottom  and  rapidly  transitioning  them 
to  the  net  under  conditions  of  darkness  (Ryer,  2008).  In 
contrast,  sweeps  are  probably  more  effective  at  herding 
flatfish  inwards  to  the  path  of  the  net  under  daylight 
conditions.  This  disparity  is  probably  responsible  for 
the  observed  pattern  of  higher  flatfish  catches  at  night 
with  survey  nets,  where  bridles  and  sweeps  are  kept  to 
minimal  length,  as  compared  to  higher  daytime  catch- 
es with  commercial  flatfish  nets  and  lengthy  sweeps  . 
These  differences,  as  explained  by  the  results  of  this 
work,  highlight  the  importance  of  fish  behavior  for  fish 
capture  technology. 

Acknowledgments 

We  wish  to  thank  C.  Hammond  and  J.  Gauvin  for  assis- 
tance with  the  at-sea  portion  of  this  project,  as  well  as 
the  captain  and  crew  of  the  FV  Cape  Horn.  M.  Ottmar, 
and  S.  Haines  assisted  with  animal  husbandry  and  labo- 
ratory experiments.  A.  Stoner,  M.  Davis,  B.  Laurel,  and 
T.  Hurst  provided  helpful  comments  and  discussion  of 
ideas  explored  in  this  research,  and  R.  Hannah  and  W. 
Wakefield  provided  helpful  critiques  of  an  early  draft  of 
this  manuscript.  C.  Sweitzer  assisted  with  manuscript 
preparation. 


Literature  cited 

Beamish,  F.  W.  H. 

1965.  Vertical  migration  by  demersal  fish  in  the  North- 
west Atlantic.  J.  Fish.  Res.  Board  Can.  23:109-139. 

Casey,  J.  M.,  and  R.  A.  Myers. 

1998.  Diel  variation  in  trawl  catchability:  is  it  as  clear 
as  day  and  night?  Can.  J.  Fish.  Aquat.  Sci.  55:2329- 
2340. 

DeAlteris,  J.,  H.  Milliken,  and  D.  Morse. 

1997.  Bycatch  reduction  in  the  Northwest  Atlantic 
small-mesh  bottom-trawl  fishery  for  silver  hake  ( Merluc - 
cius  bilinearis).  In  Developing  and  sustaining  world  fish- 
eries: the  state  of  science  and  management.  2nd  World 
Fisheries  Congress  (D.  A.  Hancock,  D.  C.  Smith,  A. 
Grant,  and  J.  P.  Beumer,  eds.),  p.  568-573.  CSIRO, 
Collingwood,  Australia. 

De  Robertis,  A.,  C.  H.  Ryer,  A.  Veloza,  and  R.  D.  Brodeur. 

2003.  Differential  effects  of  turbidity  on  prey  consump- 
tion of  piscivorous  and  planktivorous  fish.  Can.  J.  Fish. 
Aquat.  Sci.  60:1517-1526. 

Douglas,  R.  H.,  and  C.  W.  Hawryshyn. 

1990.  Behavioural  studies  of  fish  vision.  In  The  visual 
system  of  fish  (Douglas,  R.  H.,  and  M.  B.  A.  Djamgoz, 
eds.),  p.  373-418.  Chapman  and  Hall,  London. 

Eaton,  R.  C.,  and  J.  T.  Hackett. 

1984.  The  role  of  the  Mauthner  cell  in  fast-starts  involv- 


154 


Fishery  Bulletin  108(2) 


ing  escapes  in  teleost  fishes.  In  Neural  mechanisms  of 
startle  behavior  (Eaton,  R.C.,  ed.),  p.  213-266.  Plenum 
Publ.  Corp.,  New  York. 

Fienberg,  S.  E. 

1980.  The  analysis  of  cross  classified  categorical 
data.  MIT  Press,  Cambridge,  MA. 

Gibson,  R.  N. 

2005.  The  behaviour  of  flatfishes.  In  Flatfishes:  biology 
and  exploitation  (Gibson,  R.  N.,  ed.),  p.  213-239.  Black- 
well  Science,  Oxford. 

Helfman,  G.  S.,  and  E.  T.  Schultz. 

1984.  Social  transmission  of  behavioural  traditions  in 
a coral  reef  fish.  Anim.  Behav.  2:379-384. 

Herrmann,  B. 

2005.  Effect  of  catch  size  and  shape  on  the  selectivity  of 
diamond  mesh  cod-ends  I.  Model  development.  Fish. 
Res.  71:1-13. 

Hurst,  T.  P.,  C.  H.  Ryer,  J.  M.  Ramsey,  and  S.  A.  Haines. 

2007.  Divergent  foraging  strategies  of  three  co-occurring 
north  Pacific  flatfishes.  Mar.  Biol.  151:1087-1098. 

Janssen,  J. 

1978.  Will  alewives  ( Alosa  pseudoharengus ) feed  in  the 
dark?  Environ.  Biol.  Fish.  3:239-240. 

Kim,  Y.  -H„  and  C.  S.  Wardle. 

1998a.  Modelling  the  visual  stimulus  of  towed  fishing 
gear.  Fish.  Res.  34:165-177. 

1998b.  Measuring  the  brightness  contrast  of  fishing 
gear,  the  visual  stimulus  for  fish  capture.  Fish.  Res. 
34:151-164. 

Main,  J.,  and  G.  I.  Sangster. 

1981.  A study  of  the  fish  capture  process  in  a bottom 
trawl  by  direct  observations  from  a towed  underwater 
vehicle.  Scott.  Fish.  Res.  Rep.  23. 

Metcalfe,  J.  E.,  G.  P.  Arnold,  and  P.  E.  Webb. 

1990.  Energetics  of  migration  by  selective  tidal  stream 
transport:  an  analysis  for  plaice  tracked  in  the  southern 
North  Sea.  J.  Mar.  Biol.  Assoc.,  U.K.  70:149-162. 

Nichol,  D.,  and  D.  A.  Somerton. 

2009.  Evidence  of  the  selection  of  tidal  streams  by 
northern  rock  sole  (Lepidosetta  polyxystra)  for  trans- 
port in  the  eastern  Bering  Sea.  Fish.  Bull.  107:221- 
234. 

Olla,  B.  L.  M.  W.  Davis,  and  C.  Rose. 

2000.  Differences  in  orientation  and  swimming  of 
walleye  pollock  Theragra  chalcogramma  in  a trawl  net 
during  light  and  dark  conditions:  concordance  between 
field  and  laboratory  observations.  Fish.  Res.  44:261- 
266. 

Petrie,  M.  E.,  and  C.  H.  Ryer. 

2006.  Hunger,  light  level  and  body  size  affect  refuge 
use  by  post-settlement  lingcod  Ophiodon  elongatus.  J. 
Fish.  Biol.  69:957-969. 

Rose,  C.  S.,  J.  R.  Gauvin,  and  C.  F.  Hammond. 

2010.  Effective  herding  of  flatfish  by  cables  with  minimal 
seafloor  contact.  Fish.  Bull.  108:136-144 


Ryer,  C.  H. 

2008.  A review  of  flatfish  behavior  relative  to  trawls. 
Fish.  Res.  90:138-246. 

Ryer,  C.  H.,  and  L.  A.  K.  Barnett. 

2006.  Influence  of  illumination  and  temperature  upon 
flatfish  reactivity  and  herding  behavior:  Potential 
implications  for  trawl  capture  efficiency.  Fish.  Res. 
81:242-250. 

Ryer,  C.  H.,  and  B.  L.  Olla. 

1998.  Effect  of  light  on  juvenile  walleye  pollock  shoaling 
and  their  interaction  with  predators.  Mar.  Ecol.  Prog. 
Ser.  167:215-226. 

1999.  Light-induced  changes  in  the  prey  consumption 
and  behavior  of  two  juvenile  planktivorous  fish.  Mar. 
Ecol.  Prog.  Ser.  181:412-51. 

2000.  Avoidance  of  an  approaching  net  by  juvenile  walleye 
pollock  Theragra  chalcogramma  in  the  laboratory:  the 
influence  of  light  intensity.  Fish.  Res.  45:195-199. 

Schabetsberger,  R.,  R.  D.  Brodeur,  L.  Ciannelli,  J.  M.  Napp,  and 
G.  L.  Swartzman. 

2000.  Diel  vertical  migration  and  interaction  of  zoo- 
plankton and  juvenile  walleye  pollock  (Theragra  chal- 
cogramma) at  a frontal  region  near  the  Pribilof  Islands, 
Bering  Sea.  ICES  J.  Mar.  Sci.  57:1283-1295. 

Shaw,  E. 

1961.  Minimal  light  intensity  and  the  dispersal  of  school- 
ing fish.  Bull.  Inst.  Oceanogr.  1213:1-8. 

Snedecor,  G.  W.,  and  W.  G.  Cochran. 

1980.  Statistical  methods,  593  p.  Iowa  State  Univ. 
Press,  Ames,  IA. 

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

1969.  Biometry,  776  p.  W.H.  Freeman  and  Co.,  San 
Francisco. 

Videler,  J.  J. 

1993.  Fish  swimming,  260  p.  Chapman  & Hall, 
London. 

Walsh,  S.  J. 

1988.  Diel  variability  in  trawl  catches  of  juvenile  and 
adult  yellowtail  flounder  on  the  Grand  Banks  and  the 
effect  on  resource  assessment.  N.  Am.  J.  Fish.  Manag. 
8:373-381. 

1991.  Diel  variation  in  availability  and  vulnerability  of 
fish  to  a survey  trawl.  J.  Appl.  Ichthyol.  7:147-159. 

Walsh,  S.  J.,  and  W.  M.  Hickey. 

1993.  Behavioural  reactions  of  demersal  fish  to  bottom 
trawls  at  various  light  conditions.  ICES  J.  Mar.  Sci. 
Symp.  196:68-76. 

Winger,  P.  D.,  P.  He,  and  S.  J.  Walsh. 

1999.  Swimming  endurance  of  American  plaice  ( Hippoglos - 
soides  platessoides ) and  its  role  in  fish  capture.  ICES 
J.  Mar.  Sci.  56:252-265. 

Winger,  P.  D.,  S.  J.  Walsh,  P.  He,  and  J.  A.  Brown. 

2004.  Simulating  trawl  herding  in  flatfish:  the  role  of  fish 
length  in  behaviour  and  swimming  characteristics.  ICES 
J.  Mar.  Sci.  61:1179-1185. 


155 


Spatial  and  temporal  variation  in  otolith  chemistry 
for  tautog  ( Tautoga  onitis ) in  Narragansett  Bay 
and  Rhode  Island  coastal  ponds 

Ivan  Mateo  (contact  author)1 
Edward  G.  Durbin2 
David  A.  Bengtson1 
Richard  Kingsley2 
Peter  K.  Swart3 
Daisy  Durant4 

Email  address  for  contact  author:  imateo32@hotmail.com 

1 University  of  Rhode  Island 

Department  of  Fisheries,  Animal  and  Veterinary  Sciences 
Kingston,  Rhode  Island  02881 

2 University  of  Rhode  Island 
Graduate  School  of  Oceanography 
Narragansett,  Rhode  Island  02882 

3 Division  of  Marine  Geology  and  Physics 

Rosenstiel  School  of  Marine  and  Atmospheric  Sciences 
University  of  Miami 
Miami,  Florida  33149 

4 Narragansett  Bay  National  Estuarine  Research  Reserve 
P.O.  Box  151 

Prudence  Island,  Rhode  Island  02872 


Abstract — The  elemental  composi- 
tion of  otoliths  may  provide  valuable 
information  for  establishing  connec- 
tivity between  fish  nursery  grounds 
and  adult  fish  populations.  Concen- 
trations of  Rb,  Mg,  Ca,  Mn,  Sr,  Na, 
K,  Sr,  Pb,  and  Ba  were  determined 
by  using  solution-based  inductively 
coupled  plasma  mass  spectrometry  in 
otoliths  of  young-of-the  year  tautog 
(Tautoga  onitis ) captured  in  nursery 
areas  along  the  Rhode  Island  coast 
during  two  consecutive  years.  Stable 
oxygen  (6180)  and  carbon  (<513C)  iso- 
topic ratios  in  young-of-the  year  oto- 
liths were  also  analyzed  with  isotope 
ratio  mass  spectrometry.  Chemical 
signatures  differed  significantly 
among  the  distinct  nurseries  within 
Narragansett  Bay  and  the  coastal 
ponds  across  years.  Significant  dif- 
ferences were  also  observed  within 
nurseries  from  year  to  year.  Classi- 
fication accuracy  to  each  of  the  five 
tautog  nursery  areas  ranged  from  85% 
to  92%  across  years.  Because  accu- 
rate classification  of  juvenile  tautog 
nursery  sites  was  achieved,  otolith 
chemistry  can  potentially  be  used  as 
a natural  habitat  tag. 


Manuscript  submitted  16  June  2009. 
Manuscript  accepted  14  December  2009. 
Fish.  Bull.  108:155-161  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


The  dependence  of  fish  production 
and  population  dynamics  on  disper- 
sal and  migration  among  multiple 
habitats,  referred  to  as  “connectiv- 
ity,” is  a critical  property  of  marine 
populations.  Connectivity  rates  deter- 
mine colonization  patterns  for  new 
habitats,  the  resiliency  of  popula- 
tions to  harvest,  and  can  be  used  in 
the  design  of  marine  protected  areas 
(MPAs).  Quantifying  connectivity 
rates  in  marine  organisms  is,  how- 
ever, extremely  difficult  because  the 
natal  and  nursery  origins  of  adults  are 
almost  unknown.  Recently,  tagging 
techniques  with  natural  isotopic  and 
elemental  markers  have  been  devel- 
oped for  species  that  were  not  able  to 
be  tagged  or  recaptured  by  conven- 
tional approaches.  Chemical  natural 
habitat  tags  in  the  otoliths  of  juvenile 
fish  have  been  used  to  differentiate 
individuals  from  different  estuarine 
and  riverine  systems  (Thorrold  et  al., 
1998a;  Thorrold  et  al.,  1998b;  Gilland- 
ers  and  Kingsford,  2000;  Gillanders, 
2002b)  and  other  types  of  nearshore 
habitats,  such  as  estuary  as  opposed 
to  rocky  reef  (Gillanders  and  Kings- 


ford, 1996)  and  estuary  as  opposed  to 
exposed  coastal  habitats  (Yamashita 
et  al.,  2000;  Forrester  and  Swearer, 
2002).  In  addition,  through  chemical 
analysis  of  the  juvenile  core  region  of 
adult  otoliths,  natural  habitat  tags 
have  been  used  to  determine  the  pro- 
portion of  the  adult  population  that 
resided  in  different  juvenile  habitats 
(Yamashita  et  al.,  2000;  Thorrold  et 
al.,  2001;  Gillanders,  2002a). 

The  tautog  (Tautoga  onitis ) is  an 
economically  and  ecologically  impor- 
tant species  found  in  the  waters  of 
eastern  North  America  from  the  Gulf 
of  Maine  to  North  Carolina.  Juvenile 
tautog  are  known  to  depend  on  shal- 
low water  habitats  where  they  are 
safe  from  high  levels  of  predation 
and  can  find  necessary  food  resourc- 
es (Dorf  and  Powell,  1997;  Arendt, 
1999).  However,  the  relative  impor- 
tance of  open  coastline  and  enclosed 
bays  and  lagoons  as  nursery  habitat 
for  tautog  is  still  poorly  understood 
(Sogard  et  al.,  1992).  In  light  of  the 
fact  that  the  northeastern  coast  of 
the  United  States  has  experienced  a 
major  loss  of  its  estuarine  habitats 


156 


Fishery  Bulletin  108(2) 


because  of  human  alteration  of  the  coastal  zone  (Brom- 
berg and  Bertness,  2005),  data  are  needed  to  quantify 
the  importance  of  specific  coastal  habitat  types  in  sus- 
taining tautog  populations. 

Our  long-term  goal  is  to  investigate  the  utility  of 
naturally  occurring  habitat  tags  to  determine  habitat 
linkages  in  Narragansett  Bay  and  other  nearby  es- 
tuarine systems  by  juvenile  tautog.  This  is  an  initial 
crucial  step  to  quantify  the  relative  contribution  of 
estuarine  habitats  for  the  population  connectivity  of 
adult  tautog. 

Materials  and  methods 

Sampling  of  juvenile  fish 

In  Rhode  Island,  young-of-the-year  (YOY)  tautog  of  45- 
64  mm  fork  length  (FL)  were  sampled  from  three  sites 
in  Narragansett  Bay:  Mt.  Hope  Bay  (MH),  Gaspee  Point 
(GP),  and  Rose  Island  (RS);  and  from  two  sites  from  the 
coastal  ponds  along  the  Rhode  Island  southern  shore: 
Point  Judith,  lower  pond  (PJ),  and  Charlestown  Pond 
(CP)  (Fig.  1).  The  samples  were  obtained  in  coopera- 
tion with  Rhode  Island  Department  of  Environmental 
Management,  Division  of  Marine  Fisheries  (RIDEM), 
during  August  and  September  of  2005  and  2006.  The 
sampling  stations  were  selected  to  include  different 
nursery  areas  and  possibly  different  chemical  back- 
grounds and  according  to  information  on  juvenile  tautog 
abundance  from  RIDEM.  Average  monthly  surface  tem- 
peratures and  salinities  at  Gaspee  Point  for  2005  were 
22°C  and  24.9%e,  and  for  2006  were  20.6°C  and  22.5%e. 
For  Mount  Hope  Bay,  average  surface  temperatures 
and  salinities  were  21.7°C  and  27.0 %c,  and  for  2006 
were  20.5°C  and  24.9%c.  Data  from  the  closest  point 
to  Rose  Island  showed  average  surface  temperatures 
and  salinities  for  2006  were  17.4°C  and  30.8 %c.  Twenty 
juveniles  per  site  per  year  were  captured  for  analysis. 
Sampled  fish  were  kept  frozen  until  dissection  for  the 
removal  of  otoliths. 

Laboratory  processing  of  samples 

Before  dissection,  each  fish  was  weighed  (wet  weight 
to  the  nearest  0.1  g)  and  measured  (FL  and  standard 
length  [SL]  to  the  nearest  0.1  mm).  Both  sagittal  oto- 
liths were  removed  from  each  fish,  cleaned  of  adhering 
tissue,  rinsed  3x  with  Milli-Q-filtered  (Millipore  Corp., 
Billerica,  MA)  water,  and  allowed  to  dry  in  a class-100 
laminar-flow  hood.  The  left  sagittal  otolith  was  used 
for  trace  metal  analysis  and  the  right  otolith  was  used 
for  stable  isotope  analysis.  A total  of  164  otoliths  were 
prepared  for  trace  metal  analysis.  Each  otolith  was 
weighed  on  a Thermo  Cahn  microbalance  (±  0.01  mg) 
(Thermo  Fisher  Scientific,  Waltham,  MA).  Samples  were 
then  placed  in  acid-washed  2.5-mL  snap-cap  polypropyl- 
ene containers.  The  otolith  weights  ranged  from  0.08 
to  0.34  mg  and  averaged  0.18  mg.  Otoliths  for  trace 
metal  analysis  were  transferred  to  5-mL  clean  polypro- 


pylene tubes  and  0.2  mL  of  triple-distilled  17%  HN03 
was  added  to  insure  complete  dissolution  (in  about  30 
seconds).  An  internal  thulium  single-element  standard 
spike  was  added  (to  correct  for  variable  matrix  effects 
during  the  inductively  coupled  plasma  mass  spectrom- 
etry analyses)  and  then  the  solution  was  diluted  to  1.8 
mL  with  triple-distilled  water.  This  dilution  resulted 
in  a Ca  concentration  of  approximately  40  ppm  in  the 
analyzed  otolith  solution. 

Otolith  chemistry 

Elemental  concentrations  of  YOY  otoliths  were  deter- 
mined through  solution-based  ICPMS  at  the  University 
of  Rhode  Island  Graduate  School  of  Oceanography.  All 
measurements  were  carried  out  on  a Finnigan  ele- 
ment high-resolution  inductively  coupled  plasma  mass 
spectrometer  (HR-ICPMS)  (Thermo  Fisher  Scientific, 
Waltham,  MA).  A procedural  blank  was  prepared  in  the 
same  manner  as  had  been  used  for  the  other  samples, 
but  with  no  otolith  present.  The  procedural  blank  was 
compared  to  the  system  blank  to  determine  if  contami- 
nation occurred  during  processing.  System  blanks  were 
made  from  the  same  acid  used  for  sample  dissolution 
and  were  run  every  four  samples.  A drift-correction 
standard  was  prepared  by  gravimetrically  spiking  a 
CaC03  standard  solution  with  the  appropriate  concen- 
trations of  Na,  K,  Rb,  Mg,  Ca,  Mn,  Ni,  Cu,  Zn,  Sr,  Ba, 
Co,  and  Pb  to  match  the  typical  elemental  composition 
of  the  otoliths.  This  drift-correction  standard  was  ana- 
lyzed every  four  samples  to  track  and  correct  for  varia- 
tions in  instrument  sensitivity  during  each  analytical 
time  period.  The  choice  of  these  thirteen  elements  for 
our  study  was  based  on  previous  studies  of  elemental 
composition  of  juvenile  fish  otoliths.  Analytical  results 
were  expressed  as  absolute  concentrations  of  elemental 
molar  ratios  with  respect  to  calcium:  Element:Ca  ratios, 
expressed  as  units  of  mmol/mol  or  pmol/mol. 

The  elements  that  were  always  above  detection  lim- 
its (Rb,  Mg,  Ca,  Sr,  and  Ba)  were  used  for  subsequent 
analysis.  The  average  relative  standard  deviations 
were  as  follows:  Rb  (3%),  Mg  (10%),  Ca  (1%),  Sr  (1%), 
and  Ba  (5%).  The  limits  of  detection  were  as  follows 
(values  in  ppm):  Rb  (0.007),  Mg  (0.02),  Sr  (0.077),  and 
Ba  (0.014).  The  detection  limits  for  the  whole  otolith 
dissolution-solution-based  method  were  calculated  as 
three  times  the  standard  deviation  of  the  counts  per 
second  (cps)  of  the  isotope  of  interest  in  acid  blanks 
divided  by  the  sensitivity  in  cps/ppm  of  the  CRM22 
carbonate  standard.  For  every  isotope,  these  were  in 
the  sub-ppm  range — a result  that  compares  with  the 
3 to  2000  ppm  range  of  the  elements  of  interest  in  the 
sample  otoliths. 

Stable  carbon  and  oxygen  isotopes  of  these  otolith 
samples  were  determined  at  Rosenstiel  School  of  Marine 
and  Atmospheric  Sciences,  University  of  Miami,  by  us- 
ing an  automated  carbonate  device  (Kiel  III)  attached 
to  a thermo  Finnigan  delta-plus  stable  isotope  mass 
spectrometer  (Thermo  Fisher  Scientific,  Waltham,  MA). 
Data  were  expressed  by  using  conventional  d notation 


Mateo  et  al.:  Otolith  chemistry  for  Tautoga  onitis  in  Narragansett  Bay  and  Rhode  Island  coastal  ponds 


157 


71°25'W  71°15'W 


71  °40'W  71  °30W 


(A)  Map  of  tautog  (Tautoga  onitis)  sampling  stations  in  Narragansett  Bay  for  2005 
and  2006  that  were  surveyed  for  juvenile  otolith  element  concentrations  and  isotopic 
signatures.  Sampling  stations  are  shown  in  arrows.  (B)  Map  of  the  south  coast  of  Rhode 
Island  showing  two  coastal  ponds  (Charlestown,  Point  Judith)  that  were  surveyed  for 
tautog  for  years  2005  and  2006  to  determine  otolith  element  concentrations  and  isotopic 
signatures.  Sampling  stations  are  shown  in  arrows. 


in  relation  to  V-PDB  (Vienna  Peedee  Belemnite).  Data 
were  corrected  for  the  usual  isobaric  interferences.  The 
external  precision  (calculated  from  replicate  analyses  of 
an  internal  laboratory  calcite  standard)  was  0.04%  for 
513C  and  0.08%  for  d180. 


Statistical  analysis 

Two-way  analysis  of  variance  (ANOVA)  was  used  to  test 
for  differences  in  fish  body  length  among  stations  and 
years.  We  also  examined  relationships  between  otolith 


158 


Fishery  Bulletin  108(2) 


Table  1 

Average  size  distribution  of  tautog  ( Tautoga  onitis)  collected  in  Rhode  Island  for  analysis  of  otolith  elemental  concentrations  and 
stable  isotopic  signatures.  The  numbers  of  fish  measured  at  each  station  ( n ) to  obtain  average  fork  lengths  (FL  in  mm)  in  each 
year  are  shown.  Numbers  in  parentheses  are  standard  errors. 

Station 

2005 

2006 

n 

FL 

n 

FL 

Gaspee  Point  (GP) 

17 

59.6(1.2) 

Mount  Hope  Bay  (MH) 

17 

59.1  (2.1) 

21 

63.0(2.9) 

Rose  Island  (RS) 

20 

52.1  (1.2) 

17 

45.3(3.5) 

Point  Judith,  lower  pond  (PJ) 

18 

49.4  (2.1) 

19 

57.3  (1.5) 

Charlestown  Pond  (CP) 

17 

50.2(3.0) 

18 

54.9(2.7) 

weight  and  otolith  elemental  composition  and  isotopic 
signatures  with  analysis  of  covariance  (ANCOVA).  If 
there  was  a significant  relationship,  we  removed  the 
effect  of  size  (otolith  weight  used  as  a proxy  for  fish  size) 
to  ensure  that  differences  in  fish  size  among  samples  did 
not  confound  any  site-specific  differences  in  otolith  chem- 
istry. Concentrations  of  elements  were  weight-detrended 
by  subtraction  of  the  product  of  the  common  within- 
group  linear  slope  multiplied  by  the  otolith  weight  from 
the  observed  concentration  (Campana  et  al.,  2000). 

To  detect  differences  in  the  concentrations  of  par- 
ticular elements  and  multi-element  fingerprints  among 
stations  and  between  years,  we  performed  ANOVA  and 
multivariate  analyses  of  variance  (M ANOVA).  Pillai’s 
trace  statistic  was  chosen  as  the  multivariate  test  sta- 
tistic because  it  is  more  robust  than  other  multivari- 
ate statistics  (Wilkes’s  lamda,  Hotelling’s  T-test)  to 
small  sample  sizes,  unequal  cell  sizes,  and  situations  in 
which  covariances  are  not  homogeneous.  Tukey’s  HSD 
test  was  used  to  detect  a posteriori  differences  among 
means  (a=0.05).  Before  statistical  testing,  residuals 
were  examined  for  normality  and  homogeneity  among 
stations.  To  meet  model  assumptions,  all  analyses  were 
performed  on  natural  log-transformed  data.  We  also 
used  linear  discriminant  function  analyses  (DFAs)  on 
tautog  juvenile  data  to  visualize  spatial  differences  in 
juvenile  otolith  chemistry  data  within  sites  and  to  ex- 
amine classification  success  for  juveniles  from  different 
sites  or  regions.  Classification  success  is  the  percentage 
of  fish  that  are  correctly  assigned  to  their  actual  region 
given  the  information  on  location  where  the  fish  was 
collected  and  the  chemical  signature  of  each  fish.  Cross 
validations  were  performed  by  using  jackknife  (“leave 
one  out”)  procedures  in  SYSTAT  (vers.  11,  Systat  Soft- 
ware, Inc.,  Chicago,  IL). 

Results 

Size  distribution 

Mean  (FL)  of  juvenile  tautog  at  stations  in  Rhode  Island 
ranged  from  45  to  63  mm  (Table  1).  There  were  sig- 


nificant differences  in  mean  length  among  stations 
(ANOVA,  PcO.OOl)  and  between  years  (ANOVA,  P<0.05) 
within  Rhode  Island  stations.  There  were  no  significant 
differences  in  mean  FL  among  stations  within  Narra- 
gansett  Bay.  However,  in  2005,  mean  FL  from  all  sta- 
tions within  Narragansett  Bay  were  significantly  longer 
than  that  for  individuals  caught  in  the  coastal  ponds 
(Point  Judith,  lower  pond,  Charlestown  Pond)  (Tukey 
test,  P<0.05).  In  2006;  only  Mount  Hope  Bay  had  fish 
significantly  longer  than  those  from  Rose  Island  (Tukey 
test,  P<0.05). 

Otolith  chemistry 

Results  of  MANOVA  showed  that  the  chemical  signa- 
tures of  trace  metals  and  stable  isotopes  combined  in 
tautog  otoliths  differed  significantly  among  stations 
(MANOVA,  F18  384  = 20.72,  PcO.OOl)  and  years  (MANOVA, 
P6  126  = 9.05,  P<0.001)  within  Rhode  Island.  Signifi- 
cant interaction  between  station  and  year  (MANOVA, 
F18  3g4=5.18,  P<0.001)  implied  that  chemical  signatures 
differed  between  years  depending  on  the  station  studied. 
Classification  success  for  tautog  by  using  both  trace 
metals  and  stable  isotopes  for  stations  within  Rhode 
Island  for  each  of  the  two  years  ranged  from  85%  to 
92%  (Table  2). 

Individual  elemental  concentrations 

In  Rhode  Island,  one  trace  element  (Rb)  and  one  stable 
isotope  (d13C)  showed  significant  relationships  with  the 
covariable  otolith  weight  in  the  ANCOVA  (P<0.001) 
and  therefore  required  the  effect  of  otolith  weight  be 
removed  for  subsequent  ANOVA  analysis.  The  ele- 
mental concentrations  and  isotope  signatures  varied 
significantly  among  stations  (ANOVA,  P<0.001),  and 
between  years  (ANOVA,  P<0.001)  (Fig.  2).  Significant 
interaction  between  station  and  year  (ANOVA,  PcO.OOl) 
indicated  that  concentration  of  individual  elements  dif- 
fered between  years  depending  on  the  station  studied. 
In  Rhode  Island,  elemental  concentrations  of  Sr,  Ba, 
Mg,  Rb,  and  the  stable  isotopes  d13C  and  <5180  varied 
significantly  among  stations  in  2005,  whereas  only  Ba 


Mateo  et  al.:  Otolith  chemistry  for  Tautoga  onitis  in  Narragansett  Bay  and  Rhode  Island  coastal  ponds 


159 


Table  2 

Classification  success  (as  a percentage)  results  determined  by  jack-knife  cross  validation  procedure  for  linear  discriminant  func- 
tion analysis  of  chemical  concentrations  in  tautog  (Tautoga  onitis)  otoliths  collected  at  Rhode  Island  stations  in  2005  and  2006, 
with  the  use  of  solution-based  inductively  coupled  plasma  mass  spectrometry  for  the  combined  trace  metals  (Sr,  Ba,  Mg,  Rb): 
[Sr/Ca],  [Ba/Ca],  [Rb/Ca],  [Mg/Ca])  and  for  d13C  and  stable  isotopes.  Names  of  the  stations  are  Gaspee  Point  (GP),  Mount 
Hope  Bay  (MH),  Rose  Island  (RS),  Point  Judith,  lower  pond  (PJ),  Charlestown  Pond  (CP). 


GP 

MH 

RS 

PJ 

CP 

Classification  success  {%) 

2005 

GP 

14 

0 

0 

0 

1 

93 

MH 

0 

13 

0 

2 

1 

81 

RS 

0 

0 

19 

1 

0 

95 

PJ 

0 

0 

0 

16 

1 

94 

CP 

0 

1 

0 

0 

16 

94 

Total 

14 

14 

19 

19 

19 

92 

2006 

GP 

0 

0 

0 

0 

0 

MH 

0 

13 

0 

2 

1 

81 

RS 

0 

0 

14 

2 

1 

82 

PJ 

0 

2 

0 

15 

0 

88 

CP 

0 

0 

0 

2 

15 

88 

Total 

15 

14 

21 

17 

85 

and  6180  varied  significantly  among  stations  in  2006 
(ANOVA,  P<0.001)  (Fig.  2).  For  example,  6180  was  high- 
est at  Rose  Island  at  the  mouth  of  Narragansett  Bay, 
whereas  <513C  magnitudes  were  similar  across  years  for 
all  Narragansett  Bay  stations.  Sr  concentrations  within 
Narragansett  Bay  and  the  coastal  ponds  also  remained 
similar  in  magnitude  throughout  the  years  of  study. 

Discussion 

The  elemental  composition  of  juvenile  tautog  otoliths 
varied  considerably  within  and  among  estuaries  and 
between  years.  We  found  very  strong  differences  in  the 
concentrations  of  Mg,  Sr,  Ba,  and  Rb,  as  well  as  in  the 
stable  isotopic  signatures  of  d 13C  and  6 180,  among  sta- 
tions within  RI.  High  classification  success  rates  (gener- 
ally >85%)  of  the  discriminant  functions  derived  from 
trace  element  and  stable  isotope  signatures  together 
confirmed  their  use  as  an  effective  natural  tag  of  the 
estuarine  nursery  area  of  juvenile  tautog.  Although 
most  of  the  variance  in  trace  element  signatures  was 
concentrated  among  estuaries,  we  also  found  signifi- 
cant differences  in  elemental  fingerprints  and  stable 
isotopes  in  tautog  otoliths  among  sites  about  10  to  25 
km2  apart  within  Narragansett  Bay  resulting  in  100% 
classification  success  within  that  water  body.  These 
data  indicate  that  the  physicochemical  characteristics 
of  specific  sections  of  the  estuaries  may  vary  enough  to 
generate  the  differences  in  otolith  chemistry  that  we 
observed  within  each  estuary. 

Elemental  fingerprints,  however,  should  not  be  regard- 
ed as  permanent  markers  of  actual  estuarine  habitat  or 
environment  (Forrester  and  Swearer,  2002;  Swearer  et 


ah,  2003).  Estuarine  habitats  are  very  dynamic;  seawa- 
ter properties  and  composition  at  a particular  location 
can  vary  over  tidal  to  annual  time  scales  (Peters,  1999). 
As  a result,  it  may  be  expected  that  the  magnitude  of 
variations  in  elemental  fingerprints  in  otoliths  among 
estuaries  will  not  remain  constant  over  time.  The  sig- 
nificant interannual  differences  we  report  among  year 
classes  in  age-0  tautog  otolith  elemental  signatures  is 
similar  to  interannual  differences  in  otolith  chemis- 
try reported  for  other  marine  fishes  (Gillanders  and 
Kingsford,  2000;  Gillanders,  2005).  Thus,  interannual 
differences  indicate  that  age-0  tautog  elemental  signa- 
tures must  be  analyzed  on  a year-class-specific  basis 
because  there  were  stations  where  concentrations  were 
not  consistent  between  years. 

It  is  not  surprising  to  see  such  clear  differences 
in  otolith  chemical  signatures  among  the  stations 
sampled  in  Narragansett  Bay.  Data  from  RIDEM  show 
that  there  were  also  significant  differences  in  salinity 
regimes  in  these  regions  during  the  late  spring  and 
summer  of  2005  and  2006  (H.  Stoffel,  and  J.  McNa- 
mee,  unpubl.  data1).  The  proximity  of  Rose  Island 
station  to  the  mouth  of  Narragansett  Bay  meant  that 
high  salinities  (up  to  30 %c)  would  be  observed.  On 
the  other  hand,  the  lower-salinity  stations  within  the 
upper  region  of  Narragansett  Bay  are  located  much 
closer  to  the  industrial  area  and  watershed  and  there- 
fore potentially  more  prone  to  terrestrial  influences 
from  freshwater  runoff  resulting  in  reduced  salinities 
(20-25%*). 


1 Stoffel,  H.,  and  J.  McNamee.  2008.  Rhode  Island  Dept. 
Environmental  Management  (RIDEM),  Jamestown,  RI 
02879. 


160 


Fishery  Bulletin  108(2) 


Sampling  year 


Figure  2 

Variation  in  trace  elements  and  stable  isotopes  concentrations  measured  in  otoliths  of  young-of- 
the-year  tautog  ( Tautoga  onitis ) collected  in  Rhode  Island  in  2005  and  2006.  All  trace  element 
data  (element/CaxlO6)  are  ln(x+l)  transformed.  Rhode  Island  station  codes  are  GP=Gaspee  Point, 
MH  = Mount  Hope  Bay,  RS  = Rose  Island,  PJ  = Point  Judith,  lower  pond,  and  CP=  Charlestown 
Pond. 


Successful  discrimination  between  estuarine  nurs- 
eries in  the  present  study  was  accomplished  through 
otolith  elemental  fingerprints,  fulfilling  one  of  the  re- 
quirements for  their  possible  use  as  natural  tags  (Cam- 
pana  et  al.,  2000).  The  estuarine  nursery  origin  of 
juvenile  tautog  was  accurately  identified  based  on  oto- 
lith elemental  fingerprints  and  stable  isotopes.  Several 
methods  based  on  laser  ablation  (Thorrold  et  al.,  2001; 
Gillanders,  2002a)  or  micromilling  techniques  (Gil- 


landers  and  Kingsford,  1996;  Gillanders,  2005;  Brown, 
2006)  could  be  used  to  determine  elemental  fingerprints 
found  in  the  otolith  cores  of  adult  tautog  for  comparison 
with  the  juvenile  estuarine  fingerprints  that  we  have 
established.  We  believe  solution-based  techniques  are 
more  suitable  than  microprobe  techniques  for  analysis 
of  tautog  otolith  elemental  concentrations  because  1) 
solution-based  techniques  tend  to  have  higher  sensitiv- 
ity, accuracy,  and  precision  compared  to  microprobe 


Mateo  et  al.:  Otolith  chemistry  for  Tautoga  onitis  in  Narragansett  Bay  and  Rhode  Island  coastal  ponds 


161 


techniques  (Campana,  1999;  Campana  et  al.,  2000);  and 
2)  solution-based  techniques  can  provide  more  precise 
natural  tags  on  fish  with  limited  movement  within  habi- 
tats during  their  first  year  of  life.  For  example,  tautog 
have  a short  larval  period  of  15  to  20  days  (Sogard  et 
al.,  1992;  Dorf  and  Powell  1997)  and  once  larvae  have 
settled,  they  have  small  home  range  of  approximately 
20  meters  (Able  et  al.,  2005)  during  their  first  year  of 
life.  Thus,  juvenile  cores  samples  from  age  classes  rep- 
resenting fish  born  in  2005  and  2006  could  be  extracted 
by  micromilling  procedures  and  their  chemical  elements 
can  be  analyzed  by  solution  ICPMS.  Present  results  are 
a step  towards  establishing  juvenile  movement  to  adult 
habitats,  which  must  be  examined  in  nursery  studies 
(Beck  et  al.,  2001).  Identifying  links  between  juvenile 
and  adult  habitats,  and  understanding  connectivity 
between  estuarine  nurseries  and  adult  populations, 
has  the  potential  to  aid  fishery  managers  and  aid  in 
the  management  and  conservation  of  estuarine  fish 
nursery  habitats. 

Acknowledgments 

We  would  like  to  thank  C.  Powell,  M.  Burnett,  and  B. 
Murphy  from  RIDEM;  as  well  as  P.  Stout  from  Camp 
Fuller,  and  R.  Dickau  from  Pond  Shore  Association  for 
helping  to  collect  fish.  Special  thanks  go  to  B.  Taplin,  R. 
Pruell  and  the  late  L.  Meng  from  U.S.  Environmental 
Protection  Agency,  and  to  K.  Castro  from  University  of 
Rhode  Island  Sea  Grant  Fisheries  Extension  for  support 
and  inspiration  for  this  project.  This  study  was  funded 
through  University  of  Rhode  Island  Sea  Grant  Program 
and  the  Nature  Conservancy  Global  Marine  Initiative. 


Literature  cited 

Able,  K.  W„  L.  S.  Hales,  and  S.  M.  Hagan 

2005.  Movement  and  growth  of  juvenile  (age  0 and  1+) 
tautog  ( Tautoga  onitis  [ L . ] ) and  cunner  ( Tautogola - 
brus  adspersus  [Walbaum])  in  a southern  New  Jersey 
estuary.  J.  Exp.  Mar.  Biol.  Ecol.  327:22-35. 

Arendt,  M.  D. 

1999.  Seasonal  residence,  movement,  and  activity  of  adult 
tautog  ( Tautoga  onitis)  in  lower  Chesapeake  Bay.  M. 
S.  thesis,  104  p.  School  of  Marine  Science,  College  of 
William  and  Mary,  Gloucester  Point,  VA. 

Beck,  M.  W„  K.  L.  Heck,  Jr.,  K.  W.  Able,  D.  L.  Childers,  D.  B.  Egg- 
leston, B.  M.  Gillanders,  B.  Halpern,  C.  G.  Hays,  K.  Hoshino, 
T.  J.  Minello,  R.  J.  Orth,  P.  F.  Sheridan,  and  M.  P.  Weinstein. 

2001.  The  identification,  conservation,  and  manage- 
ment of  estuarine  and  marine  nurseries  for  fish  and 
invertebrates.  BioScience  51:633-641. 

Bromberg,  K.  D,  and  M.  D.  Bertness. 

2005.  Reconstructing  New  England  salt  marsh  losses 
using  historical  maps.  Estuaries  28:823-832. 

Brown,  J.  A. 

2006.  Using  the  chemical  composition  of  otoliths  to 
evaluate  the  nursery  role  of  estuaries  for  English  sole 
Pleuronectes  vetulus  populations.  Mar.  Ecol.  Prog.  Ser. 
306:269-281. 


Campana,  S.  E. 

1999.  Chemistry  and  composition  of  fish  otoliths:  path- 
ways, mechanisms  and  applications.  Mar.  Ecol.  Prog. 
Ser.  188:263-297. 

Campana,  S.  E.,  G.  A.  Chouinard,  J.  M.  Hanson,  A.  Frechet,  and 
J.  Brattey. 

2000.  Otolith  elemental  fingerprints  as  biological  tracers 
of  fish  stocks.  Fish.  Res.  46:343-357. 

Dorf,  B.  A.,  and  J.  C.  Powell. 

1997.  Distribution,  abundance,  and  habitat  characteris- 
tics of  juvenile  tautog  (Tautoga  onitis , Family  Labridae) 
in  Narragansett  Bay,  Rhode  Island,  1988-1992.  Estu- 
aries 20:589-600. 

Forrester,  G.  E.,  and  S.  E.  Swearer. 

2002.  Trace  elements  in  otoliths  indicate  the  use  of  open- 
coast  versus  bay  nursery  habitats  by  juvenile  California 
halibut.  Mar.  Ecol.  Prog.  Ser.  241:201-213. 

Gillanders,  B.  M. 

2002a.  Connectivity  between  juvenile  and  adult  fish 
populations:  do  adults  remain  near  their  recruitment 
estuaries?  Mar.  Ecol.  Prog.  Ser.  240:215-223. 

2002b.  Temporal  and  spatial  variability  in  elemental 
composition  of  otoliths:  Implications  for  determining 
stock  identity  and  connectivity  of  populations.  Can. 
J.  Fish.  Aquat.  Sci.  59:669-679. 

2005.  Using  elemental  chemistry  of  fish  otoliths  to 
determine  connectivity  between  estuarine  and  coastal 
habitats.  Estuar.  Coast.  Shelf  Sci.  64:47-57. 

Gillanders,  B.  M.,  and  M.  J.  Kingsford. 

1996.  Elements  in  otoliths  may  elucidate  the  contribu- 
tion of  estuarine  recruitment  to  sustaining  coastal  reef 
populations  of  a temperate  reef  fish.  Mar.  Ecol.  Prog. 
Ser.  141:13-20. 

2000.  Elemental  fingerprints  of  otoliths  of  fish  may 
distinguish  estuarine  ‘nursery’  habitats.  Mar.  Ecol. 
Prog.  Ser.  201:273-286. 

Peters,  H. 

1999.  Spatial  and  temporal  variability  of  turbulent 
mixing  in  an  estuary.  J.  Mar.  Res.  57:805-845. 

Sogard,  S.  M.,  K.  W.  Able,  and  M.  P.  Fahay. 

1992.  Early  life  history  of  the  tautog,  Tautoga  onitis,  in 
the  Mid-Atlantic  Bight.  Fish.  Bull.  90:529-539. 

Swearer,  S.  E.,  G.  E.  Forrester,  M.  A.  Steele,  A.  J.  Brooks,  and 
D.  W.  Lea. 

2003.  Spatio-temporal  and  interspecific  variation  in  oto- 
lith trace-elemental  fingerprints  in  a temperate  estua- 
rine fish  assemblage.  Estuar.  Coast.  Shelf.  Sci.  56: 
1111-1123. 

Thorrold,  S.  R.,  C.  M.  Jones,  S.  E.  Campana,  J.  W.  McLaren,  and 
J.  W.  H.  Lam. 

1998a.  Trace  element  signatures  in  otoliths  record  natal 
river  of  juvenile  American  shad  ( Alosa  sapidissima). 
Limnol.  Oceanogr.  43:1826-1835. 

Thorrold,  S.  R.,  C.  M.  Jones,  P.  K.  Swart,  and  T.  E.  Targett. 

1998b.  Accurate  classification  of  juvenile  weakfish 
Cynoscion  regalis  to  estuarine  nursery  areas  based  on 
chemical  signatures  in  otoliths.  Mar.  Ecol.  Prog.  Ser. 
173:253-265. 

Thorrold,  S.  R.,  C.  Latkoczy,  P.  K.  Swart,  and  C.  M.  Jones. 

2001.  Natal  homing  in  a marine  fish  metapopulation. 
Science  291:297-299. 

Yamashita,  Y.,  T.  Otake,  and  H.  Yamada. 

2000.  Relative  contributions  from  exposed  inshore  and 
estuarine  nursery  grounds  to  the  recruitment  of  stone 
flounder,  Platichthys  bicoloratus,  estimated  using  otolith 
Sr:Ca  ratios.  Fish.  Oceanogr.  9:316-327. 


162 


Fish  assemblages  associated 
with  three  types  of  artificial  reefs: 
density  of  assemblages 
and  possible  impacts 
on  adjacent  fish  abundance 

Reiji  Masuda  (contact  author)1  Yoshiaki  Kai1 

Masami  Shiba2  Asami  Nakanishi1 

Yoh  Yamashita1  Masaru  Torikoshi1 

Masahiro  Ueno1  Masaru  Tanaka3 

Email  address  for  contact  author:  reiji@kais. kyoto-u.ac.jp 

1 Maizuru  Fisheries  Research  Station 
Kyoto  University 

Nagahama,  Maizuru 
Kyoto  625-0086,  Japan 

2 Ashiu  Forest  Research  Station 
Kyoto  University 

Miyama,  Nantan 
Kyoto  601-0703,  Japan 

3 University  of  Malaysia  Sabah 
Locked  Bag  No.  2073 
88999,  Kota  Kinabalu 
Sabah,  Malaysia 


Abstract— We  evaluated  the  effective- 
ness of  wooden  artificial  reefs  (ARs) 
as  fish  habitat.  Three  types  of  ARs, 
made  of  cedar  logs,  broadleaf  tree 
logs,  and  PVC  pipes,  respectively, 
were  deployed  in  triplicate  at  8-m 
depth  off  Maizuru,  Kyoto  Prefec- 
ture, Sea  of  Japan,  in  May  2004.  Fish 
assemblages  associated  with  each  of 
the  nine  ARs  were  observed  by  using 
SCUBA  twice  a month  for  four  years. 
Fish  assemblages  in  the  adjacent 
habitat  were  also  monitored  for  two 
years  before  and  four  years  after  reef 
deployment.  In  the  surveyed  areas 
(ca.  10  m2)  associated  with  each  of  the 
cedar,  broadleaf,  and  PVC  ARs,  the 
average  number  of  fish  species  was 
4.14,  3.49,  and  3.00,  and  the  average 
number  of  individuals  was  40.7,  27.9, 
and  20.3,  respectively.  The  estimated 
biomass  was  also  more  greater  when 
associated  with  the  cedar  ARs  than 
with  other  ARs.  Visual  censuses  of  the 
habitat  adjacent  to  the  ARs  revealed 
that  the  number  of  fish  species  and 
the  density  of  individuals  were  not 
affected  by  the  deployment  of  the  ARs. 
Our  results  support  the  superiority 
of  cedar  as  an  AR  material  and  indi- 
cate that  deployment  of  wooden  ARs 
causes  no  reduction  of  fish  abundance 
in  adjacent  natural  reefs. 


Manuscript  submitted  29  January  2009. 
Manuscript  accepted  14  December  2009. 
Fish.  Bull.  108:162-173  (2019). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


Habitat  complexity  plays  a major  role 
in  the  survival  of  young  demersal 
fishes  by  providing  a refuge  from  pre- 
dation (Ferreira  et  al.,  2001;  Scharf 
et  al.,  2006;  Hamilton  and  Konar, 
2007).  Fish  species  richness  is  highly 
dependent  on  the  rugosity  and  vari- 
ety of  growth  forms  in  the  habitat, 
whereas  the  height  of  vertical  struc- 
tures is  an  important  predictor  of 
total  fish  abundance  (Gratwicke  and 
Speight,  2005).  In  this  respect,  arti- 
ficial reefs  (ARs)  are  often  deployed 
to  improve  the  quality  of  habitat 
(Gorham  and  Alevizon,  1989).  In 
addition  to  their  role  as  refuges,  ARs 
host  encrusting  invertebrates  that 
can  be  consumed  as  prey  by  fishes 
(Seaman  and  Jensen,  2000).  Fish  are 
often  more  abundant  at  ARs  than  at 
natural  reefs,  probably  because  the 
vertical  structures  potentially  allow 
more  varied  refuges  for  fish  settle- 
ment and  recruitment  than  the  usual 
more  moderately  sloped  bottoms  of 
natural  reefs  (Rilov  and  Benayahu, 
2000;  Reed  et  al.,  2006). 


Although  the  deployment  of  struc- 
tures functioning  as  ARs  may  well 
have  started  long  ago  by  fishermen 
in  various  localities  around  the  globe, 
research  on  this  subject  is  relatively 
recent  (Seaman  and  Sprague,  1991). 
Two  countries,  United  States  and  Ja- 
pan, have  relatively  long  histories  of 
nationwide  projects  on  ARs.  In  the 
case  of  the  United  States,  the  main 
goal  of  deploying  ARs  has  been  to 
improve  catch  for  recreational  fish- 
ermen. Common  materials  used  for 
these  ARs  have  been  waste  products, 
such  as  automobiles,  tires,  and  oil 
and  gas  platforms.  The  use  of  such 
products  has  caused  environmental 
concerns,  resulting  in  a shift  toward 
the  construction  of  ARs  with  concrete 
(Collins  et  al.,  2002).  In  contrast,  the 
purpose  of  Japanese  deployments  of 
ARs  have  primarily  been  to  improve 
commercial  fishery  production,  and 
governmental  agencies  have  invested 
heavily  in  the  construction  of  large 
ARs  made  of  concrete  and  steel  to  be 
deployed  in  coastal  areas. 


Masuda  et  al.:  Fish  assemblages  associated  with  three  types  of  artificial  reefs 


163 


The  recent  trend  for  ARs  in  Japan  has  shifted 
from  concrete  to  wooden  construction.  This  has 
been  partly  due  to  funding  shortages,  but  also 
because  fishermen  have  found  that  wooden  ARs 
attract  fish  more  rapidly  than  those  made  of  con- 
crete or  steel.  Indeed,  most  coastal  prefectures 
in  Japan  deploy  wooden  ARs  with  or  without 
governmental  subsidies  under  the  supervision 
of  local  fishermen’s  cooperatives.  The  materials 
and  shape  of  wooden  ARs  differ  depending  on 
each  fishery  cooperative.  As  much  as  70%  of  the 
land  area  in  Japan  is  forested,  half  of  which  is 
plantation  forests  of  conifers,  such  as  Japanese 
cedar  (Cryptomeria  japonica)  and  hinoki  cypress 
( Chamaecyparis  obtusa).  Although  these  forests 
require  occasional  thinning,  many  of  them  lack 
such  maintenance  because  of  the  decline  in  the 
market  price  of  timber.  Therefore,  the  construc- 
tion of  wooden  ARs  also  has  the  socioeconomic 
potential  to  stimulate  the  demand  for  forestry 
materials. 

The  primary  goal  of  the  present  study  was  to 
confirm  the  efficacy  of  wooden  ARs,  especially 
those  made  of  cedar  tree  logs  as  fish  habitat.  For 
this  purpose,  fish  assemblages  associated  with 
ARs  made  from  cedar  trees  were  compared  to 
those  made  from  broadleaf  trees  and  those  made 
with  polyvinyl  chloride  (PVC)  pipes.  There  is  a 
debate  whether  ARs  merely  attract  fishes  from  adjacent 
areas  or  whether  they  do  improve  fishery  productivity 
(Grossman  et  al.,  1997;  Pickering  and  Whitmarsh, 
1997).  We  therefore  tested  the  possibility  that  ARs  at- 
tract fishes  from  adjacent  areas  and  thus  concentrate 
fish  abundance  at  the  ARs,  rather  than  fish  abundance 
is  spread  over  the  fishing  ground  as  a whole.  A visual 
census  had  been  conducted  twice  a month  for  more 
than  two  years  before  the  deployment  of  these  ARs  in 
adjacent  areas;  hence  the  fish  fauna  was  compared  in 
the  area  before  and  after  the  deployment  of  ARs. 

Materials  and  methods 

Deployment  and  visual  census  of  artificial  reefs 

Three  types  of  ARs  were  prepared.  The  design  of  the  ARs 
was  modified  from  that  designed  by  the  Atake  Forestry 
Association,  Yamaguchi,  Japan  (http://www.geocities.jp/ 
abu_kikori/katsudou/gyosyou/gyosyou2.html,  accessed 
on  December  2003;  also  see  Fig.  1).  The  first  type  of  AR 
(cedar  AR)  was  constructed  of  16  log  sections  (1.5  m long, 
6.9-18.4  cm  diameter)  of  Japanese  cedar  ( Cryptomeria 
japonica)  arranged  in  a parallel  cross  formation.  Each 
corner  was  tied  with  rope  and  fixed  with  a stainless 
steel  rod.  Diagonal  wires  helped  maintain  the  rectan- 
gular shape.  The  second  type  of  AR  (broadleaf  AR)  was 
constructed  from  six  species  of  broadleaf  trees  harvested 
from  the  Ashiu  Forest  Research  Station,  Kyoto  Univer- 
sity, and  assembled  with  the  same  dimensions  as  those 
used  for  the  cedar  AR.  The  broadleaf  tree  species  used 


were  Japanese  cherry  birch  ( Betula  grossa),  hornbeam 
( Carpinus  laxiflora),  Japanese  beech  ( Fagus  crenata), 
Chinese  chestnut  (Castanea  crenata),  redvein  maple 
( Acer  rufinerve),  and  macropoda  holly  ( Ilex  macropoda). 
The  diameter  of  broadleaf  and  cedar  logs  ranged  from 
7.5  to  19.2  cm.  The  third  type  of  AR  (PVC  AR)  was  made 
of  hollow  PVC  pipes  (11.8  cm  diameter,  3 mm  thickness) 
and  was  assembled  in  the  same  manner  as  that  used  for 
the  other  two  types  of  ARs. 

These  three  types  of  ARs  were  constructed  in  trip- 
licate and  deployed  at  a depth  of  8 m off  the  Maizuru 
Fisheries  Research  Station  (MFRS),  Nagahama,  Maiz- 
uru, Kyoto  (35°29'N  lat.  and  135°22'E  long.)  on  21  May 
2004  (Fig.  2).  The  shore  in  this  area  is  a concrete  bank 
and  its  subtidal  zone  consists  of  natural  rocks,  concrete 
blocks,  both  partly  covered  by  live  oyster  ( Crassostrea 
gigas)  and  their  dead  shells,  and  sandy  silt  with  some 
macroalgal  vegetation.  The  substrate  in  the  research 
area  consisted  of  muddy  silt  with  no  macroalgae  veg- 
etation. Each  AR  was  sunk  with  240  kg  of  sand  bags 
(60  kg  attached  to  each  corner  of  the  AR).  ARs  were 
set  15  m apart. 

Twice  monthly  visual  censuses  of  fish  assemblages  as- 
sociated with  each  AR  were  conducted  for  four  consecu- 
tive years  after  AR  deployment.  All  census  observations 
were  made  by  the  first  author  with  SCUBA  equipment. 
The  area  in  and  around  each  AR  was  observed  for  about 
three  minutes  and  the  species,  size,  and  number  of 
fish  were  recorded.  A census  commenced  from  one  of 
the  lateral  sides  of  an  AR  and  extended  out  to  about 
1 m from  each  side.  The  observer  then  swam  around 
and  above  the  AR,  and  the  fish  inside  the  AR  were 


164 


Fishery  Bulletin  108(2) 


Figure  2 

Map  of  study  area  for  artificial  reef  deployment  off  Maizuru,  Kyoto,  in  2004.  Upper-left 
map  shows  location  of  Wakasa  Bay  (in  box)  along  the  Sea  of  Japan.  The  arrow  in  the 
upper-right  map  represents  the  location  of  the  research  area  in  Maizuru  Bay.  Lower 
map  shows  the  research  area  off  the  Maizuru  Fisheries  Research  Station  (MFRS), 
Kyoto  University,  with  the  nine  artificial  reefs  (three  typesxthree  replicates)  deployed 
in  a line.  Observations  were  conducted  after  the  visual  census  of  the  adjacent  habitat 
(transects  1-3).  Census  lines  are  expressed  by  thick  dotted  lines,  and  -2  m,  -5  m,  and 
-10  m isobaths  are  expressed  by  thin  dotted  lines. 


recorded.  Fish  were  considered  as  associating  with  an 
AR  if  they  were  swimming  or  dwelling  within  1 m of 
the  AR  (Sherman  et  ah,  2002),  and  thus  fish  in  an 
area  of  about  10  m2  were  counted  for  each  AR.  Fish 
standard  length  (SL)  was  estimated  with  the  help  of  a 
scale  marked  on  a clipboard  and  was  recorded.  Length 


estimates  were  occasionally  calibrated  by  capturing 
and  measuring  fish.  These  calibrations  revealed  that 
visual  SL  estimates  were  within  10%  error  of  the  actual 
measured  SL.  Water  temperature  and  visibility  during 
observations  ranged  from  10.1°  to  28.8°C  and  from  1 
to  5 m,  respectively.  Biomass  calculation  for  each  AR 


Masuda  et  al.:  Fish  assemblages  associated  with  three  types  of  artificial  reefs 


165 


was  conducted  according  to  the  method  of  San- 
tos et  al.  (2005)  and  Friedlander  et  al.  (2007). 

The  estimated  average  length  of  each  species  for 
each  sample  was  converted  to  mass  by  using  the 
length-mass  relationship 

M=aSL6, 

where  a and  b - constants  for  allometric  growth; 

SL  = standard  length;  and 
M = mass. 

Length-mass  parameters  were  obtained  from  Fish- 
Base  (www.fishbase.org,  accessed  on  July  2008) 
and  calibration  was  based  on  our  own  samples. 

The  number  of  fish  species  (species  richness), 
total  number  of  fish  individuals  (abundance),  to- 
tal fish  biomass,  and  number  of  individuals  of 
each  fish  species  associated  with  each  type  of  AR 
were  compared  among  the  three  types  of  ARs  by 
repeated  measures  ANOVA  followed  by  Tukey’s 
HSD  test.  Data  for  the  number  of  fish  individuals 
and  their  biomass  were  log  (x+l)  transformed  to 
obtain  homoscedasticity. 

Estimation  of  the  impact  of  AR  deployment 
on  fish  abundance  in  the  adjacent  area 

Fish  assemblages  in  the  area  surrounding  the  ARs 
were  compared  before  and  after  AR  deployment. 

Data  from  the  twice  monthly  visual  censuses  in 
each  area  were  used  for  this  purpose  (Masuda, 

2008;  Fig.  2).  The  number  and  size  of  fish  of  each  spe- 
cies found  along  three  400-m2  belt  transects  have  been 
recorded  twice  a month  since  1 January  2002.  One 
transect  was  close  to  the  location  of  the  ARs  that  we 
deployed  in  the  present  study  (transect  1),  and  the  other 
two  were  relatively  distant  (transects  2 and  3).  There- 
fore, species  richness  and  fish  abundance  in  transect  1 
would  decline  after  AR  deployment  if  fish  were  simply 
attracted  from  the  adjacent  natural  reef  to  these  ARs. 
Each  of  the  three  transects  included  areas  of  rocky  reef, 
live  oysters  and  their  dead  shells,  a sandy  or  muddy 
silt  bottom,  and  an  artificial  vertical  structure  made 
of  concrete  blocks  that  had  been  deployed  more  than 
20  years  earlier.  The  size  (length  x width  x height)  of 
the  concrete  structures  along  transects  1,  2,  and  3 were 
0. 5x3x2. 4 m,  1. 8x3x1  m,  and  2. 5x2. 5x2  m,  respectively. 
Data  from  23  May  2002  to  15  May  2004,  and  those  from 
29  May  2004  to  8 May  2008  were  used  to  compare  the 
fish  assemblages  before  and  after  deployment  of  the 
ARs.  Analyses  of  covariance  (ANCOVA)  was  used  to 
compare  species  richness  and  fish  abundance  in  each 
transect  before  and  after  deploying  the  wooden  or  PVC 
ARs,  and  bottom  water  temperature  was  used  as  a 
covariant  because  fish  species  richness  and  abundance 
increase  almost  linearly  with  the  increase  of  bottom 
water  temperature  in  this  habitat  (Masuda,  2008).  The 
number  of  individuals  of  each  species  was  also  compared 
by  ANCOVA  before  and  after  deployment  of  the  ARs.  All 


Table  1 

The  mean  (±  standard  error)  number  of  species,  individuals,  and 
estimated  biomass  of  fish  attracted  to  the  cedar,  broadleaf,  and 
PVC  artificial  reefs  over  the  entire  observation  period  (2004-08) 
and  for  each  of  the  four  years  (n  = 3 ARs  per  type).  Different  let- 
ters represent  significant  differences  among  AR  types  (P<0.01, 
Tukey’s  HSD  test). 


Cedar  ARs 

Broadleaf  ARs 

PVC  ARs 

No.  of  species 

Whole  period 

4.14  ±0.138° 

3.49  ±0.1076 

3.00 

±0.113° 

1st  year 

5.14  ±0.332° 

3.44  ±0.2456 

2.51 

±0.201° 

2nd  year 

4.10±0.289° 

3.49  ±0.244h 

2.83 

±0.232° 

3rd  year 

3.93  ±0.226° 

3.63±0.225°6 

3.28 

±0.2176 

4th  year 

3.38  ±0.196 

3.40  ±0.193 

3.38 

±0.195 

No.  of  individuals 

Whole  period 

40.7  ±4.43° 

27.9  ±2.88fe 

20.3 

±2.18° 

1st  year 

84.5  ±12.9° 

36.8  ±5.86h 

29.6 

±6.23° 

2nd  year 

24.1  ±5.00° 

28.0  ±5.88“ 

10.9 

±2.25h 

3rd  year 

32.1  ±8.31° 

24.7  ±6.80ft 

19.0 

±4.00b 

4th  year 

22.0  ±4.59 

22.1  ±4.15 

21.9 

±3.82 

Fish  biomass  (grams) 

Whole  period  284  ±34.7° 

143  ±19.1* 

157 

±40. 76 

1st  year 

498  ±89.8° 

113  ±24.4ft 

243 

±1576 

2nd  year 

222  ±51.6 

134  ±38.7 

89.1 

±19.3 

3rd  year 

310  ±82.0 

179  ±44.8 

141 

±28.4 

4th  year 

108  ±29.9fc 

148  ±41.8ft 

155 

±28.2° 

statistical  analyses  were  conducted  with  the  software 
JMP  (vers.  5.0. 1J,  SAS  Institute,  Inc.,  Cary,  NC)  with 
an  alpha  level  of  0.01. 

Results 

Fish  assemblages  associated  with  the  ARs 

Both  species  richness  and  fish  abundance  were  high- 
est associated  with  the  cedar  ARs,  intermediate  with 
the  broadleaf  ARs,  and  lowest  with  the  PVC  ARs  when 
compared  over  the  entire  sampling  period  (Table  1). 
These  differences  were  significant  among  the  three  AR 
types  in  both  of  these  measurements  (repeated  mea- 
sures ANOVA  followed  by  Tukey’s  HSD  test:  PcO.Ol). 
The  greater  effectiveness  of  the  cedar  ARs  was  promi- 
nent in  the  first  year  after  deployment  but  decreased 
with  time  and  became  nonsignificant  in  the  fourth 
year  (Table  1;  Fig.  3).  Fish  biomass  was  greatest  in 
the  cedar  and  PVC  ARs  in  the  first  and  fourth  year, 
respectively,  but  did  not  differ  significantly  in  the 
second  and  third  years. 

A total  of  62  fish  species  were  observed  in  96  dives 
on  these  nine  ARs,  among  which  six  species  were 
found  most  frequently  in  the  cedar  ARs,  two  in  the 
broadleaf  ARs,  and  two  in  the  PVC  ARs  (Table  2). 
Five  most  commonly  observed  fish  species  in  the  ARs 


166 


Fishery  Bulletin  108(2) 


were  black  rockfish  ( Sebastes  inermis),  jack  mackerel 
( Trachurus  japonicus),  bambooleaf  wrasse  (Pseudola- 
brus  sieboldi),  chameleon  goby  (Tridentiger  trigono- 
cephalus)  and  whitespotted  pigmy  filefish  ( Rudarius 
ercodes ) (Fig.  4);  the  former  three  species  are  tar- 
geted in  commercial  fisheries,  whereas  the  latter  two 
are  prey  species  of  other  commercial  species.  Jack 
mackerel  is  pelagic  and  migratory,  and  the  other  four 
species  are  demersal  and  relatively  sedentary.  The 
typical  fishes  showing  high  preference  for  the  cedar 
ARs  were  black  rockfish,  sunrise  sculpin  (Pseudo- 
blennius  cottoides),  black  sea  bream  (Acanthopagrus 
schlegelii),  whitespotted  pigmy  filefish,  thread-sail 
filefish  ( Stephanolepis  cirrhifer),  and  finepatterned 
puffer  ( Takifugu  poecilonotus).  Two  species  of  goby  (7s- 
tigobius  hoshinonis  and  T.  trigonocephalus)  were  most 
abundant  in  the  broadleaf  ARs  (Fig.  4).  Redspotted 
grouper  ( Epinephelus  akaara)  and  barface  cardinalfish 
( Apogon  semilineatus)  were  most  abundant  in  the  PVC 


ARs.  Jack  mackerel  and  bambooleaf  wrasse  were  the 
most  abundant  species  during  the  entire  census  period 
(Table  2),  but  they  did  not  show  any  clear  preference 
for  a particular  type  of  AR. 

Maximum,  minimum,  and  average  body  length  in 
two  highly  abundant  and  commercially  important  spe- 
cies, black  rockfish  and  jack  mackerel,  are  plotted  for 
each  type  of  artificial  reef  in  Figure  5.  Black  rockfish 
generally  had  a wide  range  (1.5-16  cm)  of  body  length, 
whereas  jack  mackerel  had  a smaller  body  size  range 
(4-12  cm).  This  was  prominent  in  cedar  ARs,  especially 
shortly  after  the  deployment  of  the  AR  (Fig.  5A). 

A bryozoan  community  was  established  within  two 
to  three  months  of  deploying  the  cedar  ARs.  Other  en- 
crusting epibenthic  assemblages,  such  as  Porifera,  Cni- 
daria,  Mollusca,  and  Annelida,  gradually  formed  on  the 
broadleaf  and  PVC  ARs  after  one  year.  The  upper  sec- 
tions of  the  ARs  attracted  these  encrusting  organisms 
more  rapidly  than  the  lower  sections.  In  the  fourth  year, 
some  of  the  upper  sections  of  the  cedar  and 
broadleaf  ARs  began  to  decay  because  of  foul- 
ing by  encrusting  organisms,  particularly  wood 
boring  piddock  (Martesia  striata).  Crabs  ( Cha - 
rybdis  japonica)  and  sea  cucumbers  ( Stichopus 
japonicus)  were  common  in  all  types  of  ARs. 
At  least  four  fish  species,  black  sea  bream, 
Temminck’s  surfperch  (Ditrema  temmincki), 
whitespotted  pigmy  filefish,  and  thread-sail 
filefish,  were  observed  feeding  on  the  encrust- 
ing organisms  on  and  around  the  cedar  ARs. 
Conger  eel  ( Conger  myriaster),  two  species  of 
groupers,  and  large  individuals  of  bambooleaf 
wrasse  resided  inside  the  PVC  pipes.  Some  fish, 
such  as  thread-sail  filefish  and  redfin  velvetfish 
( Paracentropogon  rubripinnis),  overwintered, 
showing  minimal  movement  in  the  cedar  ARs 
through  the  winter. 

Fish  assemblages  in  the  adjacent  habitat 

Visual  censuses  of  the  areas  adjacent  to  the 
ARs  revealed  that  both  fish  species  richness 
and  abundance  showed  clear  seasonal  changes 
corresponding  to  variations  in  sea  bottom  water 
temperature  (Fig.  6).  A total  of  73,922  fish  indi- 
viduals from  90  species  were  recorded  from 
23  May  2002  to  8 May  2008  in  transects  1-3. 
There  was  no  significant  change  in  fish  species 
richness  or  abundance  along  any  of  the  three 
transects  after  the  deployment  of  ARs  (P>0.5, 
ANCOVA;  Table  3).  Species-to-species  analy- 
sis revealed  that  although  there  were  several 
cases  of  increases  or  decreases  in  abundance 
after  deployment,  there  was  no  evidence  of  a 
systematic  decrease  in  species  richness  along 
transect  1,  in  which  one  species  decreased  and 
four  species  increased  after  the  deployment  (see 
far-right  column  in  Table  2).  The  average  (±SE) 
number  of  individuals  in  the  entire  census  area 
of  the  adjacent  habitat  was  171  ±12.6  per  400  m2. 


_CO 

§ 100 
"D 
> 

T3 

C 

d 10 
z 


1 

1000 
3 

« 100 
C/5 
03 

E 

0 

1 10 

U5 

u_ 

1 

Jul  Oct  Jan  Apr  Jul  Oct  Jan  Apr  Jul  Oct  Jan  Apr  Jul  Oct  Jan  Apr 
2004  2004  2004  2004 

Figure  3 

Species  richness,  fish  abundance,  and  fish  biomass  associated 
with  each  type  of  artificial  reef  on  each  observation  day  between 
May  2004  and  April  2008.  Plotted  data  are  averages  of  the  two 
monthly  observations  carried  out  at  each  triplicate  artificial 
reef.  Note  log  scale  for  individuals  and  biomass  plots. 


Masuda  et  al.:  Fish  assemblages  associated  with  three  types  of  artificial  reefs 


167 


168 


Fishery  Bulletin  108(2) 


-H> 

1 

+ 

+ 

+ 

O) 

to 

£ 

1— 1 

CM 

+ 

i—l 

CM 

+ 

i—l 

CO 

+ 

a 

u 

H 

-O 

03 

03 

LO 

o 

-ch 

00 

CD 

-o 

-o 

1—1 

CO 

O 

o 

o 

O 

03 

LO 

CO 

i—i 

CM 

CM 

CM 

O 

o 

o 

O 

CM 

CM 

to 

o 

o 

o 

O 

PVC 

O 

d 

d 

o 

o 

o 

d 

d 

o 

o 

d 

+1 

+1 

+i 

+i 

+i 

+i 

+i 

+i 

+1 

+i 

+i 

03 

03 

o 

i— H 

to 

00 

o 

CM 

03 

CD 

CD 

1—1 

CM 

CM 

rH 

03 

i—l 

±SE 

O 

O 

o 

o 

d 

o 

o 

O 

o 

o 

d 

o 

o 

d 

o 

o 

o 

£ 

o3 

<v 

C+-. 

o 

03 

»o 

-o 

£ 

JD 

o 

CM 

CD 

CD 

oo 

to 

CD 

T— 1 

03 

i— 1 

o 

o 

T— 1 

CD 

CM 

CM 

rH 

o 

CO 

O 

o 

o 

to 

O 

o 

O 

o 

a 

o 

o 

o 

o 

d 

d 

o 

d 

d 

o 

d 

k 

+i 

+1 

+i 

+i 

+1 

+i 

+i 

+i 

+i 

+i 

+i 

'-a 

pq 

o 

CO 

oo 

o 

i—i 

i—i 

CD 

03 

to 

03 

i— 1 

tO 

co 

1— 1 

1—1 1 

LO 

to 

to 

i—l 

i—i 

CO 

O 

d 

o 

o 

p 

o 

d 

d 

o 

o 

d 

o 

d 

d 

d 

o 

-o 

6 

03 

o 

o 

c- 

o 

CJ 

CD 

CD 

o 

oo 

i> 

co 

to 

O 

o 

o 

to 

03 

o 

CD 

CO 

i—i 

T3 

i—i 

o 

o 

o 

CO 

to 

o 

O 

o 

i— i 

03 

i—i 

T— 1 

d 

d 

o 

d 

o 

i — i 

d 

o 

d 

o 

o 

+i 

+ 1 

+i 

+1 

+i 

+1 

+i 

+1 

+i 

+1 

+1 

+i 

LO 

CO 

03 

o 

o 

03 

o 

CD 

h- 

CD 

CD 

i— 1 

1—1 

CD 

O 

1—1 

o 

i-H 

o 

CM 

00 

d 

00 

rH 

o 

CM 

d 

CM 

o 

i—i 

d 

d 

d 

d 

o 

o 

> 

T— 1 

o 

CM 

co 

CO 

LO 

CM 

o 

00 

o 

CM 

CM 

CM 

1— 1 

r— i 

"C 

<D 

3 

Cti 

E 

o 

< 1 ) 

03 

rH 

O 

co 

LO 

o 

CD 

LO 

CM 

CD 

E 

03 

!— 1 

to 

CM 

CO 

O 

cr 

O 

i— 1 

rH 

U 

<u 

PQ 

(N 

0) 

k 

S3 

03 

CM 

CM 

co 

CM 

i — 1 

CO 

to 

O 

CD 

LO 

oo 

CO 

LO 

i— i 

s 

o 

03 

£ 

00 

i—i 

00 

o 

CJ 

1 

to 

1 

03 

| 

t> 

1 

to 

| 

to 

1 

r— 1 

| 

»o 

1 

i—i 

I 

1—1 

1 

CM 

| 

oo 

1 

03 

CM 

co 

CD 

CM 

r— I 

rH 1 

o 

rH 

LO 

CM 

CD 

CO 

-H> 

£h 

rH 

bJD 

£ 

03 

£ 

03 

CO 

00 

00 

o 

00 

03 

CM 

[> 

CO 

o 

pq 

§ 

CO 

CO 

to 

UO 

CO 

CM 

CO 

i-H 

CM 

CD 

LO 

03 

CD 

co 

3 

£ 

co 

•K. 

~a 

a 

CO 

3 

CO 

O 

a 

Cj 

a 

<33 

O 

.co 

O 

S3 

2 

»o 

s 

a 

£ 

<33 

CJ 

o 

o 

co 

a 

co 

co 

<33 

"a 

o 

S 

& 

2 

k 

k 

co 

a 

o 

a 

o 

co 

<33 

2 

ty) 

a 

a, 

a. 

UjO 

~£ 

CJ 

o 

(D 

CO 

o 

CO 

S3 

O 

co 

co 

-a 

S 

03 

a 

2 

co 

a 

O 

*a 

£ 

CJ 

.a 

2 

k 

<33 

P, 

£ 

O 

k 

<33 

a 

a 

o 

^<3> 

<33 

O 

Sh 

.g, 

a 

£ 

CO 

O 

S' 

2 

a 

a 

2 

S3 

<33 

£ 

CO 

•2 

Cj 

£ 

o 

a 

a 

a 

tuo 

O) 

-o 

k 

£ 

k 

£ 

f a 

fa 

o 

o 

<33 

£ 

£ 

<D 

C3 

k 

<33 

• 2? 

S3 

'•? 

"e 

£ 

<33 

CL 

►5 

cj 

k 

£ 

£ 

_£ 

_£ 

m 

o 

Eh 

co 

lx 

co 

E-h 

Eh 

03 

03 

03 

03 

T3 

03 

T3 

2 

4— < 

4— > 

03 

03 

03 

Jh 

>> 

r£ 

£ 

£ 

>> 

a 

03 

03 

<33 

03 

£ 

o 

T3 

03 

Sh 

03 

O 

o 

Ph 

m 

i 

03 

Discussion 

The  greater  effectiveness  of  cedar  ARs 

We  found  that  ARs  made  from  logs  of  cedar  trees  had  a 
higher  fish  species  richness  and  abundance  than  those 
made  of  broadleaf  trees  or  PVC  pipes.  The  greater  effec- 
tiveness of  the  cedar  ARs  can  be  attributed  to  the  direct 
or  indirect  effects  of  cedar  wood  as  an  AR  material. 
Qualitative  observations  support  the  latter  because  we 
observed  some  fish  feeding  on  encrusting  organisms  on 
the  cedar  ARs.  Cedar  emits  volatile  compounds  that 
repel  terrestrial  invertebrates  to  protect  the  living  tree 
(Morisawa  et  al.,  2002),  but  such  chemicals  might  not  be 
effective  as  repellants  in  seawater,  making  it  a suitable 
habitat  for  fouling  marine  organisms.  The  rapid  growth 
of  cedar  trees  results  in  relatively  soft  tissues  that  can 
further  make  the  wood  a suitable  substrate  for  fouling 
organisms.  A comparison  of  the  abundance  of  epibenthic 
assemblages  between  cedar  and  broadleaf  logs  will  be 
required  to  confirm  this  hypothesis. 

Redspotted  grouper  was  significantly  more  abundant 
in  PVC  ARs  than  in  the  other  two  types  of  ARs.  The 
body  length  of  this  species  was  an  average  of  14  cm 
and  ranged  from  10  to  19  cm  (Table  2),  and  the  inner 
diameter  of  the  PVC  pipes  was  11  cm.  ARs  with  holes 
are  expected  to  host  more  fish  (Kellison  and  Sedberry, 
1998),  especially  large  predators  (Hixon  and  Beets, 
1989).  Indeed,  PVC  pipes,  because  of  their  size,  pro- 
vided a suitable  shelter  for  redspotted  groupers.  Yel- 
lowspotted  grouper  ( E . awoara),  conger  eel,  and  some 
large  individuals  of  bambooleaf  wrasse  also  used  the 
cavities  of  the  PVC  pipes. 

Two  species  of  goby  were  more  abundant  in  the  broad- 
leaf ARs  than  in  the  other  two  ARs.  Most  of  these  go- 
bies ranged  from  1 to  5 cm.  Predation  pressure  by  the 
abundant  sunrise  sculpin  and  black  rockfish  in  the 
cedar  ARs,  and  groupers  in  the  PVC  ARs,  may  have 
reduced  the  survival  of  gobies  in  these  two  types  of 
ARs,  resulting  in  the  relatively  higher  abundance  of 
gobies  in  the  broadleaf  ARs. 

Black  rockfish  associated  with  cedar  ARs  ranged  from 
1.5  to  16  cm  SL.  Black  rockfish  is  a viviparous  fish 
and  matures  at  12  cm  BL  in  1-2  years  after  birth,  and 
1.5  cm  and  16  cm  SL  individuals  represent  1.5-month 
and  4-5  year-old  individuals,  respectively  (Hisada  et 
al.,  2000).  Whitespotted  pigmy  filefish  associated  with 
cedar  ARs  ranged  from  1 to  5 cm  SL.  Whitespotted 
pigmy  filefish  mature  at  3 cm  SL  (Ishida  and  Tanaka, 
1983).  Therefore  these  species  use  ARs  as  settlement 
sites,  nurseries,  and  adult  habitats.  Jack  mackerel  as- 
sociated with  ARs  ranged  from  4 to  12  cm  SL.  Jack 
mackerel  mature  at  14  cm  SL  (Ochiai  et  al.,  1983)  and 
attain  4 cm  in  2 months  (Xie  et  al.,  2005).  Therefore 
they  use  ARs  mainly  as  nursery  habitat  and  are  loosely 
associated  with  ARs.  This  finding  is  in  agreement  with 
that  of  Rooker  et  al.  (1997)  who  reported  that  some  mid- 
water pelagic  fishes,  such  as  carangids  and  scombrids, 
were  transient  members  of  the  AR  fish  assemblages. 
Considering  that  there  are  both  pelagic  predators,  such 


Masuda  et  al.:  Fish  assemblages  associated  with  three  types  of  artificial  reefs 


169 


2004  2005  2006  2007  2008 

Figure  4 

The  monthly  average  of  individuals  of  black  rockfish  iSebastes  inermis),  jack  mackerel  ( Trachurus  japonicus ),  bambooleaf 
wrasse  ( Pseudolabrus  sieboldi),  chameleon  goby  ( Tridentiger  trigonocephalus),  and  whitespotted  pigmy  filefish  (Rudarius 
ercodes)  associated  with  each  type  of  artificial  reef  installed  off  Maizuru,  Kyoto,  in  2004. 


Table  3 

The  number  of  species  and  number  of  individuals  of  fish  recorded  during 
observations  along  transects  1,  2,  and  3 before  and  after  the  deployment  of 
the  artificial  reefs,  expressed  as  the  mean  ±standard  error  (n  = 48  and  96 
observations  for  before  and  after  deployment,  respectively). 

Transect  1 

Transect  2 

Transect  3 

No.  of  species 

Before  deployment 

9.69  ±0.61 

9.67  ±0.60 

8.88  ±0.62 

After  deployment 

9.40  ±0.43 

9.44  ±0.47 

8.40  ±0.42 

No.  of  individuals 

Before  deployment 

116.9  ±21.5 

237.6  ±44.6 

178.0  ±37.4 

After  deployment 

165.7  ±22.7 

225.9  ±34.8 

171.1  ±28.9 

as  Japanese  seabass  (Lateolabrax  ja- 
ponicus), and  benthic  predators,  such 
as  Japanese  flounder  ( Paralichthys  oli- 
vaceus ),  in  this  area  (Masuda,  2008), 
these  ARs  may  well  be  used  as  refuges 
from  predators. 

Because  the  size  of  ARs  was  1.5x1. 5 
m and  fish  were  counted  within  a dis- 
tance of  1 m,  the  survey  area  repre- 
sented about  10  m2  for  each  AR.  The 
density  of  fish  associated  with  the  AR 
was  estimated  as  4.07,  2.79,  and  2.03 
fish  per  m2  in  and  around  the  cedar, 
broadleaf,  and  PVC  ARs,  respectively 
(Table  1).  Santos  et  al.  (2005)  stud- 
ied fish  assemblages  associated  with 
ARs  made  of  concrete  blocks  located 
at  a similar  latitude  but  deeper  depth 
(17-22  m)  in  south  Portugal  (37°00'N  lat.,  7°45'  and 
8°00'E  long.),  and  estimated  the  mean  fish  density  as 
2.01  ±0.74  fish  per  m2  and  fish  biomass  as  123.6  ±77.4  g 
per  m2.  Fish  density  on  our  cedar  ARs  was  about  twice 


as  much  but  the  biomass  was  much  less  than  the  value 
reported  by  Santos  et  al.  This  finding  was  probably  the 
result  of  the  cedar  ARs  hosting  more  recruited  juveniles 
than  adults. 


170 


Fishery  Bulletin  108(2) 


Dq  Cedar 


Trachurus  japonicus 


: Jan  Apr  Jul  Oct  Jan  Apr  Jul  Oct  Jan  Apr  Jul  Oct  Jan  Apr 

1 2005  2006  2007  2008 


2004  2005 


onr\7  200ft 


Figure  5 

Monthly  maximum  (rectangles),  minimum  (triangles),  and  average  (circles)  body  length  of  black 
rockfish  ( Sebastes  inermis)  found  associated  with  (A)  cedar,  (B)  broadleaf,  and  (C)  PVC  artificial 
reefs,  and  those  of  jack  mackerel  (Trachurus  japonicus)  found  associated  with  (D)  cedar,  (E) 
broadleaf,  and  (F)  PVC  artificial  reefs. 


The  cedar  ARs  hosted  fish  assemblages  within  the 
first  two  to  three  months  of  deployment.  These  recruits 
may  have  come  from  the  adjacent  coastal  habitat  or 
from  offshore.  Rapid  colonization  of  ARs  was  also  re- 
ported by  Bohnsack  et  al.  (1994)  who  observed  that  fish 
species,  number  of  individuals,  and  biomass  reached 
peak  levels  within  two  months  of  deploying  concrete 
ARs  in  Florida. 

There  was  only  one  species,  Acentrogobius  pflaumii, 
that  decreased  in  abundance  in  transect  1 after  the  de- 
ployment of  ARs.  This  goby  is  the  fifth  most  frequently 
observed  fish  in  the  adjacent  natural  reef  (Masuda, 
2008),  but  relatively  few  were  associated  with  ARs. 
Therefore  it  is  unlikely  that  the  attraction  to  ARs  in- 
duced the  decline  in  the  population  along  transect  1. 
The  relative  stability  of  fish  species  and  abundance 
observed  among  the  three  transects  supports  the  con- 
cept of  an  inshore  migration  and  is  in  agreement  with 
data  of  Connell  (1997)  who  found  that  the  number  of 


recruits  did  not  differ  between  ARs  located  close  to  and 
far  from  a natural  reef.  Sanchez-Jerez  and  Ramos-Espla 
(2000)  also  confirmed  that  antitrawling  reefs  deployed 
in  a seagrass  habitat  had  little  effect  on  seagrass  fish 
assemblages  in  the  surrounding  area.  We  therefore 
conclude  that  the  three  types  of  ARs  deployed  in  this 
study  provided  additional  habitat  for  young  fish  without 
any  significant  depletion  of  numbers  in  the  existing  fish 
community. 

The  average  number  of  fish  in  the  adjacent  habitat 
was  171  individuals  per  400  m2,  or  0.43  individuals  per 
m2.  Fish  density  on  the  cedar  reef  was  thus  10  times 
larger  than  that  of  the  adjacent  area.  Bohnsack  et  al. 
(1991)  reviewed  experimental  studies,  where  fish  densi- 
ties at  natural  reefs  were  compared  with  those  at  arti- 
ficial reefs,  and  found  that  in  some  cases  the  latter  can 
host  densities  of  more  than  10  times  that  of  the  former. 
Therefore,  our  results  of  fish  density  on  cedar  ARs  are 
within  the  range  of  previously  reported  ARs. 


Masuda  et  al.:  Fish  assemblages  associated  with  three  types  of  artificial  reefs 


171 


Deployment  of  wooden  ARs  as  a tool 
for  ecosystem-based  fishery  management 

The  major  anthropogenic  impacts  on  coastal  ecosys- 
tems include  overfishing,  loss  of  physical  complexity 
induced  by  construction  or  trawling,  and  eutrophica- 
tion induced  by  water  discharge.  ARs  made  of  cedar 
and  other  materials  have  the  potential  to  attenuate  at 
least  some  of  these  problems.  ARs  are  useful  in  that 
they  preclude  trawling,  protect  juveniles  in  nursery 
grounds,  and  provide  fishing  sites  for  artisanal  fisher- 
men (Polovina,  1991).  Our  study  site  had  also  been  a 
trawl  fishing  ground  for  bivalves  and  sea  cucumbers, 
but  fishermen  could  not  trawl  at  our  ARs.  The  preven- 
tion of  trawling  resulted  in  the  accumulation  of  rela- 
tively large  individuals  of  sea  cucumber  in  our  ARs 
(R.  Masuda,  unpubl.  data).  Habitat  complexity,  such 
as  vertical  relief  and  holes,  can  be  a positive  factor 
for  the  survival  of  juvenile  fish.  For  instance,  Gorham 
and  Alevizon  (1989)  showed  that  the  attachment  of 
polypropylene  rope  to  ARs  significantly  increases  the 
abundance  of  juvenile  fish.  Wooden  ARs  not  only  pro- 
vide vertical  relief  but  also  provide  a porous  substrate 
for  boring  and  attachment  by  encrusting  organisms, 
such  as  boring  sponges,  oysters,  and  wood  boring  pid- 
dock.  Some  demersal  fishes,  such  as  black  rockfish, 
wrasses,  and  gobies  might  well  use  these  encrusting 
organisms  for  both  refuge  and  as  prey. 

Most  of  the  encrusting  organisms  on  ARs  are 
plankton  feeders  that  can  use  a wide  size  range  of 
phytoplankton  and  zooplankton.  For  example,  a sin- 
gle oyster  filters  several  liters  of  sea  water  per  day 
and  produces  pseudofeces  that  contain  about  half 
of  the  organic  content  of  that  trapped  on  the  gills 
(Deslous-Paoli  et  al.,  1992).  Most  juvenile  and  young 
demersal  fish  feed  on  benthic  organisms  in  addition 
to  relatively  large  zooplankton.  Therefore,  encrusting 
organisms  on  ARs  can  transform  phytoplankton  and 
microzooplankton  to  a usable  energy  source  for  fish- 
es. Fabi  et  al.  (2006)  demonstrated  that  ARs  provide 
the  main  food  source  (e.g.,  encrusted  organisms  and 
crustaceans)  for  the  three  major  fish  species  ( Sciaena 
umbra,  Diplodus  annularis,  and  Lithognathus  mor- 
rnyrus)  they  studied.  Furthermore,  improved  water 
clarity  due  to  the  filtering  function  of  the  encrusting 
organisms  is  likely  to  result  in  the  better  growth  of 
primary  producers,  such  as  macroalgae.  The  use  of  fish 
reefs  as  biofilters  for  nutrient  removal  has  also  been 
proposed  by  Seaman  and  Jensen  (2000). 

The  efficacy  of  wooden  ARs  is  of  a short  duration  (up 
to  3-5  years)  compared  to  those  made  of  concrete,  which 
can  last  decades  (Yabe,  1995).  However,  fishermen  have 
observed  that  wooden  ARs  attract  fish  sooner  than  other 
types  of  AR.  Although  wooden  ARs  biodegrade  sooner 
than  concrete  ARs,  from  an  ecological  point  of  view  of 
providing  immediate  refuge,  habitat,  and  a source  of 
food,  they  have  long-term  effects  on  the  marine  environ- 
ment. Simple  wooden  ARs  that  combine  logs  and  con- 
crete blocks  sink  easily  in  a muddy  substrate,  and  their 
life  as  an  effective  AR  can  be  as  short  as  one  year  (R. 


Figure  6 

Seasonal  changes  in  the  mean  (±standard  error)  number 
of  fish  species  and  individuals  per  transect  in  the  area 
immediately  adjacent  to  the  artificial  reefs,  and  the  surface 
and  bottom  water  temperatures  measured  in  those  areas 
during  the  surveys  from  January  2002  to  June  2008,  at 
Nagahama,  Maizuru,  Japan.  Vertical  arrows  represent  the 
date  (21  May  2004)  of  artificial  reef  deployment. 


Masuda,  personal  observ. ).  The  shape  of  wooden  ARs 
presented  in  this  article,  with  a double-cross  formation 
(Fig.  1),  provides  an  open  and  stable  vertical  relief  that 
can  attract  more  fish  recruits.  This  formation  can  also 
act  as  a stable  substrate  for  encrusting  organisms  that 
can  function  as  powerful  biofilters,  and  has  a longer 
durability  than  other  wooden  constructs. 

The  recruitment  of  reef  fishes  is  often  limited  by 
the  availability  of  suitable  nearshore  nursery  habitats, 
which  tend  to  be  vulnerable  to  anthropogenic  impacts. 
The  decrease  of  reef  fish  populations  is  therefore  partly 
attributable  to  the  loss  of  nursery  habitats,  such  as 
natural  rocky  reefs  and  seagrass  beds.  The  deployment 
of  wooden  ARs  may  provide  an  opportunity  to  mitigate 
this  trend  of  decline  in  nursery  quality  and  because 
they  are  highly  biodegradable,  the  risks  of  unexpect- 


172 


Fishery  Bulletin  108(2) 


ed  negative  impacts  on  the  environment  are  minimal. 
Stock  enhancement,  defined  as  the  release  of  cultured 
juveniles  into  wild  populations  to  augment  harvest,  has 
been  used  as  a strategy  to  reconstruct  depleted  fisheries 
resources  (Bell  et  al.,  2008).  We  suggest  that  the  release 
of  reef-associating  fish  juveniles,  such  as  black  rockfish, 
combined  with  the  deployment  of  wooden  ARs  would 
be  an  efficient  approach  for  the  recovery  of  depleted 
coastal  fisheries. 

A major  problem  of  deploying  ARs  is  that  they  at- 
tract fishermen  as  well  as  fishes.  There  is  always  the 
possibility  that  fishermen  will  catch  more  fish  than  the 
increase  of  production  because  fish  attracted  to  ARs 
are  generally  more  easily  exploitable  than  those  spread 
over  natural  reefs  (Powers  et  al.,  2003).  Indeed,  we  of- 
ten observed  local  anglers  fishing  at  our  experimental 
reefs.  Therefore,  a management  strategy  is  critically 
important  in  controlling  the  harvesting  pressure  at  AR 
sites  (Pickering  and  Whitmarsh,  1997).  As  our  long- 
term goal  is  to  improve  the  productivity  of  local  inshore 
fishing  grounds,  we  would  suggest  that  part  of  the  ar- 
eas to  be  enhanced  should  have  ARs  distributed  within 
them  and  be  managed  as  marine  protected  areas. 

Acknowledgments 

We  are  grateful  to  H.  Fujii  and  other  technical  staff 
in  the  Ashiu  Forest  Research  Station  for  providing  the 
materials  and  construction  for  the  wooden  ARs,  and 
I.  Shiga,  K.  Sato,  and  graduate  students  at  Maizuru 
Fisheries  Research  Station  (MFRS)  for  help  in  deploy- 
ing the  ARs.  D.  Robert  of  MFRS,  W.  Seaman  of  Univer- 
sity of  Florida,  and  three  anonymous  reviewers  kindly 
provided  constructive  and  insightful  comments  on  the 
manuscript.  This  research  was  partly  supported  by  a 
Grant-in-Aid  for  Scientific  Research  from  the  Japan 
Society  for  the  Promotion  of  Science. 

Literature  cited 

Bell,  J.  D.,  K.  M.  Leber,  H.  L.  Blankenship,  N.  R.  Loneragan, 
and  R.  Masuda. 

2008.  A new  era  for  restocking,  stock  enhancement  and 
sea  ranching  of  coastal  fisheries  resources.  Rev.  Fish. 
Sci.  16:1-9. 

Bohnsack,  J.  A.,  D.  L.  Johnson,  and  R.  F.  Ambrose. 

1991.  Ecology  of  artificial  reef  habitats  and  fishes.  In 
Artificial  habitats  for  marine  and  freshwater  fisher- 
ies (W.  Seaman  Jr.,  and  L.  M.  Sprague,  eds.),  p.  61- 
107.  Academic  Press,  San  Diego,  CA. 

Bohnsack,  J.  A.,  D.  E.  Harper,  D.  B.  McClellan,  and  M.  Hulsbeck. 
1994.  Effects  of  reef  size  on  colonization  and  assemblage 
structure  of  fishes  at  artificial  reefs  off  southeastern 
Florida,  U.S.A.  Bull.  Mar.  Sci.  55:796-823. 

Collins,  K.  J.,  A.  C.  Jensen,  J.  J.  Mallinson,  V.  Roenelle,  and 
I.  P.  Smith. 

2002.  Environmental  impact  assessment  of  a scrap  tyre 
artificial  reef.  ICES  J.  Mar.  Sci.  59:S243-S249. 
Connell,  S.  D. 

1997.  The  relationship  between  large  predatory  fish  and 


recruitment  and  mortality  of  juvenile  coral  reef-fish 
on  artificial  reefs.  J.  Exp.  Mar.  Biol.  Ecol.  209:261- 
278. 

Deslous-Paoli,  J.  — M.,  A.  — M.  Lannou,  P.  Geairon,  S.  Bougrier, 
O.  Raillard,  and  M.  Heral. 

1992.  Effects  of  the  feeding  behaviour  of  Crassostrea 
gigas  (bivalve  Molluscs)  on  biosedimentation  of  natural 
particulate  matter.  Hydrobiologia  231:85-91. 

Fabi,  G.,  S.  Manoukian,  and  A.  Spagnolo. 

2006.  Feeding  behavior  of  three  common  fishes  at  an 
artificial  reef  in  the  north  Adriatic  Sea.  Bull.  Mar. 
Sci.  78:39-56. 

Ferreira,  C.  E.  L.,  J.  E.  A.  Confalves,  and  R.,  Coutinho. 

2001.  Community  structure  of  fishes  and  habitat  com- 
plexity on  a tropical  rocky  shore.  Environ.  Biol.  Fish. 
61:353-369. 

Friedlander,  A.  M.,  E.  K.  Brown,  and  M.  E.  Monaco. 

2007.  Coupling  ecology  and  GIS  to  evaluate  efficacy 
of  marine  protected  areas  in  Hawaii.  Ecol.  Appl. 
17:715-730. 

Gorham,  J.  C.,  and  W.  S.  Alevizon. 

1989.  Habitat  complexity  and  the  abundance  of  juvenile 
fishes  residing  on  small  scale  artificial  reefs.  Bull. 
Mar.  Sci.  44:662-665. 

Gratwicke,  B.,  and  M.  R.  Speight. 

2005.  The  relationship  between  fish  species  richness, 
abundance  and  habitat  complexity  in  a range  of  shallow 
tropical  marine  habitats.  J.  Fish  Biol.  66:650-667. 

Grossman,  G.  D.,  G.  P.  Jones,  and  W.  J.  Seaman  Jr. 

1997.  Do  artificial  reefs  increase  regional  fish  production? 
A review  of  existing  data.  Fisheries  22(4):  17 — 2 3 . 

Hamilton,  J.,  and  B.  Konar. 

2007.  Implications  of  substrate  complexity  and  kelp 
variability  for  south-central  Alaskan  nearshore  fish 
communities.  Fish.  Bull.  105:189-196. 

Hisada,  T.,  T.  Inoue,  and  Y.  Hamanaka. 

2000.  Age,  growth  and  maturity  of  a black  rockfish  in 
the  western  Wakasa  Bay.  Bull.  Kyoto  Inst.  Ocean. 
Fish.  Sci.  22:44-49. 

Hixon,  M.  A.,  and  J.  P.  Beets. 

1989.  Shelter  characteristics  and  Caribbean  fish  assem- 
blages: experiments  with  artificial  reefs.  Bull.  Mar. 
Sci.  44:666-680. 

Ishida,  Y.,  and  S.  Tanaka. 

1983.  Growth  and  maturation  of  the  small  filefish  Rudar- 
ius  ercodes  in  Odawa  Bay.  Bull.  Japan.  Soc.  Sci.  Fish. 
49:547-553. 

Kellison,  G.  T.,  and  G.  R.  Sedberry. 

1998.  The  effects  of  artificial  reef  vertical  profile  and 
hole  diameter  on  fishes  off  South  Carolina.  Bull.  Mar. 
Sci.  62:763-780. 

Masuda,  R. 

2008.  Seasonal  and  interannual  variation  of  subtidal 
fish  assemblages  in  Wakasa  Bay  with  reference  to  the 
warming  trend  in  the  Sea  of  Japan.  Environ.  Biol. 
Fish.  82:387-399. 

Morisawa,  J.,  C.-S.  Kim,  T.  Kashiwagi,  S.  Tebayashi,  and 
M.  Horiike. 

2002.  Repellents  in  the  Japanese  cedar,  Cryptome- 
ria  japonica,  against  the  pill-bug,  Armadillidium 
vulgare.  Biosci.  Biotechnol.  Biochem.  66:2424-2428. 

Ochiai,  A.,  K.  Mutsutani,  and  S.  Umeda. 

1983.  On  the  first  year’s  growth,  maturity  and  artificial 
spawning  of  cultured  jack  mackerel.  Bull.  Japan.  Soc. 
Sci.  Fish.  49:541—545. 


Masuda  et  al.:  Fish  assemblages  associated  with  three  types  of  artificial  reefs 


173 


Pickering,  H.,  and  D.  Whitmarsh. 

1997.  Artificial  reefs  and  fisheries  exploitation:  a review 
of  the  ‘attraction  versus  production’  debate,  the  influ- 
ence of  design  and  its  significance  for  policy.  Fish. 
Res.  31:39-59. 

Polovina,  J.  J. 

1991.  Fisheries  applications  and  biological  impacts  of 
artificial  reefs.  In  Artificial  habitats  for  marine  and 
freshwater  fisheries  ( W.  Seaman  Jr.,  and  L.  M.  Sprague, 
eds.),  p.  153-176.  Academic  Press,  San  Diego,  CA. 

Powers,  S.  P.,  J.  H.  Grabowski,  C.  H.  Peterson,  and  W.  J. 

Lindberg. 

2003.  Estimating  enhancement  of  fish  production  by  off- 
shore artificial  reefs:  uncertainty  exhibited  by  divergent 
scenarios.  Mar.  Ecol.  Prog.  Ser.  264:265-277. 

Reed,  D.  C.,  S.  C.  Schroeter,  D.  Huang,  T.  W.  Anderson,  and  R. 

F.  Ambrose. 

2006.  Quantitative  assessment  of  different  artificial  reef 
designs  in  mitigating  losses  to  kelp  forest  fishes.  Bull. 
Mar.  Sci.  78:133-150. 

Rilov,  G.,  and  Y.  Benayahu. 

2000.  Fish  assemblage  on  natural  versus  artificial  reefs: 
the  rehabilitation  perspective.  Mar.  Biol.  136:931-942. 

Rooker,  J.  R.,  Q.  R.  Dokken,  C.  V.  Pattengill,  and  G.  J.  Holt. 

1997.  Fish  assemblages  on  artificial  and  natural  reefs  in 
the  Flower  Garden  Banks  National  Marine  Sanctuary, 
USA.  Coral  Reefs  16:83-92. 

Sanchez-Jerez,  P.,  and  A.  Ramos-Espla. 

2000.  Changes  in  fish  assemblages  associated  with 
the  deployment  of  an  antitrawling  reef  in  seagrass 
meadows.  Trans.  Am.  Fish.  Soc.  129:1150-1159. 


Santos,  M.,  N.,  C.  C.  Monteiro,  and  G.  Lasserre. 

2005.  Observations  and  trends  on  the  intra-annual  varia- 
tion of  the  fish  assemblages  on  two  artificial  reefs  in 
Algarve  coastal  waters  (southern  Portugal).  Sci.  Mar. 
69:415-426. 

Scharf,  F.  S.,  J.  P.  Manderson,  and  M.  C.  Fabrizio. 

2006.  The  effects  of  seafloor  habitat  complexity  on 
survival  of  juvenile  fishes:  Species-specific  interac- 
tions with  structural  refuge.  J.  Exp.  Mar.  Biol.  Ecol. 
335:167-176. 

Seaman,  W.,  Jr.,  and  A.  C.  Jensen. 

2000.  Purposes  and  practices  of  artificial  reef 
evaluation.  In  Artificial  reef  evaluation  (W.  Seaman 
Jr.,  ed),  p.  1-19.  CRC  Press,  Boca  Raton,  FL. 

Seaman,  W.,  Jr.,  and  L.  M.  Sprague. 

1991.  Artificial  habitat  practices  in  aquatic  systems.  In 
Artificial  habitats  for  marine  and  freshwater  fisheries 
( W.  Seaman  Jr.,  and  L.  M.  Sprague,  eds.),  p.  1-29.  Aca- 
demic Press,  San  Diego,  CA. 

Sherman,  R.  L.,  D.  S.  Gilliam,  and  R.  E.  Spieler. 

2002.  Artificial  reef  design:  void  space,  complexity,  and 
attractants.  ICES  J.  Mar.  Sci.  59:S196-S200. 

Xie,  S.,  Y.  Watanabe,  T.  Saruwatari,  R.  Masuda,  Y.  Yamashita, 
C.  Sassa,  and  Y.  Konishi. 

2005.  Growth  and  morphological  development  of  sag- 
ittal otoliths  of  larval  and  early  juvenile  Trachurus 
japonicus.  J.  Fish  Biol.  66:1704-1719. 

Yabe, K. 

1995.  The  note  of  evaluation  of  artificial  fish  reefs  on 
the  sand  beaches  at  Haboro,  Hokkaido.  Bull.  Hokkaido 
Tokai  Univ.  8:101-108. 


Abstract — The  Pacific  sardine  ( Sar - 
dinops  sagax)  is  distributed  along  the 
west  coast  of  North  America  from  Baja 
California  to  British  Columbia.  This 
article  presents  estimates  of  biomass, 
spawning  biomass,  and  related  biolog- 
ical parameters  based  on  four  trawl- 
ichthyoplankton  surveys  conducted 
during  July  2003-March  2005  off 
Oregon  and  Washington.  The  trawl- 
based  biomass  estimates,  serving  as 
relative  abundance,  were  198,600  t 
(coefficient  of  variation  [CV]  = 0.51) 
in  July  2003,  20,100  t (0.8)  in  March 
2004,  77,900  t (0.34)  in  July  2004, 
and  30,100  t (0.72)  in  March  2005 
over  an  area  close  to  200,000  km2. 
The  biomass  estimates,  high  in  July 
and  low  in  March,  are  a strong  indi- 
cation of  migration  in  and  out  of  this 
area.  Sardine  spawn  in  July  off  the 
Pacific  Northwest  (PNW)  coast  and 
none  of  the  sampled  fish  had  spawned 
in  March.  The  estimated  spawn- 
ing biomass  for  July  2003  and  July 
2004  was  39,184  t (0.57)  and  84,120  t 
(0.93),  respectively.  The  average  active 
female  sardine  in  the  PNW  spawned 
every  20-40  days  compared  to  every 
6-8  days  off  California.  The  spawning 
habitat  was  located  in  the  southeast- 
ern area  off  the  PNW  coast,  a shift 
from  the  northwest  area  off  the  PNW 
coast  in  the  1990s.  Egg  production  in 
off  the  PNW  for  2003-04  was  lower 
than  that  off  California  and  that  in 
the  1990s.  Because  the  biomass  of 
Pacific  sardine  off  the  PNW  appears 
to  be  supported  heavily  by  migratory 
fish  from  California,  the  sustainabil- 
ity of  the  local  PNW  population  relies 
on  the  stability  of  the  population  off 
California,  and  on  local  oceanographic 
conditions  for  local  residence. 


Manuscript  submitted  22  October  2008. 
Manuscript  accepted  29  December  2009. 
Fish.  Bull.  108:174-192  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


Biomass  and  reproduction  of 

Pacific  sardine  ( Sardinops  sagax ) 

off  the  Pacific  northwestern  United  States, 

2003-2005 


Nancy  C.  H.  Lo  (contact  author) 

Beverly  J.  Macewicz 
David  A.  Griffith 

Email  address  for  contact  author:  Nancy.Lo@noaa.gov 

Southwest  Fisheries  Science  Center 
8604  La  Jolla  Shores  Dr. 

La  Jolla,  California  92037 


Pacific  sardine  ( Sardinops  sagax ; here- 
after “sardine”)  are  distributed  widely 
off  the  west  coast  of  North  America 
from  Baja  California,  Mexico,  to  Brit- 
ish Columbia,  Canada;  the  majority  of 
the  population  is  located  off  California 
(Felin,  1954;  Murphy,  1966;  Emmett 
et  ah,  2005;  McFarlane  et  ah,  2005; 
Smith,  2005).  Tagging  studies  have 
shown  that  sardine  migrate  along  the 
west  coast  (Janssen,  1938;  Clark  and 
Janssen,  1945).  The  sardine  popu- 
lation reached  a peak  in  the  early 
1930s  at  3.5  million  metric  tons  (t) 
and  declined  rapidly  in  the  mid-1950s 
(Marr,  1950).  The  sardine  fishery  off 
California  and  British  Columbia  dates 
from  1916  (Fig.  1).  Pacific  sardine 
was  one  of  the  economically  impor- 
tant species  off  California  and  British 
Columbia  in  the  1930s  when  a fishery 
began  off  Oregon  and  Washington  (the 
Pacific  Northwest:  PNW).  The  PNW 
catch  peaked  at  nearly  50,000  t in 
1938  (Marr,  1950;  Mosher  and  Eckles, 
1954;  Murphy,  1966).  In  the  1960s, 
however,  a moratorium  on  sardine 
fishing  was  established  in  U.S.  waters 
because  of  low  catches  (Murphy,  1966; 
MacCall,  1976).  In  the  mid  1980s,  sar- 
dine became  common  as  bycatch  in 
fisheries  off  Baja  California  and  Cali- 
fornia state  (Wolf,  1992;  Deriso  et  al., 
1996)  and  reappeared  from  Oregon 
to  British  Columbia  in  1992  (Emmett 
et  al.,  2005;  McFarlane  et  al.,  2005), 
apparently  in  response  to  the  1992-93 
El  Nino  event.  The  sardine  population 
now  supports  a relatively  large  fish- 
ery with  annual  catches  over  50,000 
t in  recent  years  (Fig.  1).  Sardine 


also  serve  as  important  food  for  tuna, 
salmon,  marlin,  mackerel,  sharks,  and 
some  groundfish  species,  as  well  as 
many  seabirds,  seals,  sea  lions,  dol- 
phins, and  whales  (Snodgrass  and 
Lowry,  personal  commun.1)  (Preti  et 
al.,  2001,  2004;  Emmett  et  al.,  2005). 
The  reappearance  of  sardine  popula- 
tions in  the  north  California  Current 
ecosystem  adds  another  forage  base  for 
predators  and  an  emerging  resource  of 
consumer  interest  to  the  ecosystem. 

Pacific  sardine  off  the  PNW  are 
considered  to  be  a part  of  the  north- 
ern subpopulation,  the  majority  of 
which  is  distributed  off  the  western 
United  States  and  Canada  (Smith, 
2005),  as  determined  from  historical 
tagging  studies  (Clark  and  Janssen, 
1945),  size  at  age,  and  other  biological 
characters.  Historical  tagging  studies 
indicated  that  some  large  sardine  mi- 
grate from  California  to  the  PNW  in 
late  spring  and  early  summer  to  feed, 
and  that  the  majority  of  the  large 
sardine  off  the  PNW  move  south  to 
California  in  the  winter  to  spawn  in 
the  spring  (Clark  and  Janssen,  1945). 
The  major  spawning  area  of  this 
northern  subpopulation  was  believed 
to  be  located  off  southern  California 
before  the  1960s  (Ahlstrom,  1948; 
Marr,  1960;  Smith,  2005).  Spawning 
also  may  have  occurred  in  the  PNW 
because  young  fish  were  caught  by 
commercial  boats  in  Canadian  wa- 
ters in  1940  (Hart,  1943).  However, 


1 Snodgrass,  Owyn.  2009.  Southwest 
Fisheries  Science  Center,  La  Jolla,  CA. 
Lowry,  Mark.  2009.  Southwest  Fisher- 

ies Science  Center,  La  Jolla,  CA. 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardmops  sagax  off  the  Pacific  coast. 


175 


the  importance  of  the  PNW  as  a spawning  area 
has  not  been  studied.  After  the  resurgence  of 
Pacific  sardine  off  California,  ichthyoplankton 
and  fishery-independent  trawl  surveys  have  been 
conducted  off  California  to  assess  the  biological 
characteristics  of  the  sardine  population  since  the 
mid  1980s,  when  the  estimated  sardine  biomass 
approached  20,000  t (Wolf,  1992;  Lo  et  al.,  2005). 
Beginning  in  the  mid  1990s,  sardine  abundance, 
distribution,  and  ecological  relationships  off  the 
PNW  and  Canada  were  analyzed  with  data  from 
salmon  surface-rope  trawl  surveys  off  the  PNW 
and  trawl  surveys  off  Vancouver  Island,  Canada 
(Bentley  et  al.,  1996;  Emmett  et  al.,  2005;  Mc- 
Farlane  et.  al.,  2005);  however,  very  few  of  those 
surveys  were  designed  specifically  to  assess  the 
biological  characteristics  of  Pacific  sardine. 

Four  trawl  surveys  off  the  PNW  were  conducted 
in  July  2003,  March  and  July  2004,  and  March 
2005  to  provide  fishery-independent  measures  of 
biological  characteristics  of  sardine  in  this  area, 
and  to  answer  the  following  questions:  1)  Do  sar- 
dine migrate  between  the  PNW  and  California? 

2)  To  what  extent  does  Pacific  sardine  spawning 
in  the  PNW  depend  on  the  sardine  population  off 
California?  and  3)  How  much  does  the  Pacific  sar- 
dine egg  production  in  the  PNW  contribute  to  that  of  the 
whole  population?  To  answer  these  questions,  we  esti- 
mated spring  and  summer  biomasses  with  length  distri- 
butions to  serve  as  signals  of  migration;  the  location  and 
spatial  extent  of  spawning  habitat  to  examine  the  fol- 
lowing: the  effect  of  the  reduction  of  the  spawning  area 
in  the  PNW  to  the  local  population;  daily  egg  production 
and  its  contribution  to  the  total  egg  production;  adult 
reproductive  parameters  to  estimate  rates  of  spawning, 
fecundity  and  maturity;  and  spawning  biomass.  These 
measurements  were  compared  with  available  PNW  mea- 
surements from  the  mid-1990s  and  those  off  California 
in  the  same  time  period,  to  facilitate  our  understanding 
of  the  population  dynamics  of  the  Pacific  sardine  off 
the  northern  west  coast  of  the  North  American  con- 
tinent, and  to  better  manage  the  entire  population. 

Materials  and  methods 

Survey 

In  order  to  obtain  unbiased  estimates  of  the  biologi- 
cal characteristics  of  Pacific  sardine  off  the  PNW,  the 
Fisheries  Resources  Division  of  the  Southwest  Fisheries 
Science  Center,  conducted  four  surveys  in  July  2003, 
March  and  July  2004,  and  March  2005  aboard  the  FV 
Frosti.  Multiple  gear  types  were  used:  a surface  trawl  to 
collect  adult  samples,  the  CalVET  plankton  net  (Califor- 
nia Cooperative  Oceanic  Fisheries  Investigation  verti- 
cal-egg-tow  net;  Smith  et  al.,  1985),  and  the  continuous 
underway  fish  egg  sampler  (CUFES;  Checkley  et  al., 
1997)  to  collect  ichthyoplankton  samples  and  record 
hydroacoustics.  The  survey  region  encompassed  the  area 


of  the  northeast  Pacific  Ocean  from  42°  to  48°N  latitude 
and  from  inshore  out  to  128°W  longitude. 

The  basic  survey  pattern  comprised  seven  transect 
lines  oriented  on  the  parallels  at  a spacing  of  60  nauti- 
cal miles  (111  km).  Stations  were  spaced  at  30  nautical 
miles  (55.5  km)  along  each  transect  measured  from 
the  offshore  station.  Forty-two  predetermined  stations 
were  sampled  by  trawl  and  CalVET  tow  during  each 
survey.  For  the  July  2003  survey,  the  primary  goal  was 
to  estimate  the  spawning  biomass  of  Pacific  sardine.  In 
the  offshore  area,  few  trawls  were  undertaken  because 
both  acoustics  and  CUFES  samples  showed  little  sign 
of  sardine  schools  and  eggs.  The  inshore  sampling  was 
discontinued  close  to  the  100-m  isobath  during  July 
2003  to  avoid  net  damage  in  shallow  water.  All  fish- 
ing was  conducted  at  night,  when  Pacific  sardine  are 
distributed  in  the  upper  50  m of  the  water  column  and 
oceanographic  conditions  at  depths  greater  than  50  m 
would  have  little  influence  on  the  spatial  and  vertical 
distributions  of  sardine  schools.  Moreover,  within  60  km 
from  the  shore,  the  densities  of  fish  were  not  related  to 
the  distance  from  shore  (Emmett  et  al.,  2005).  Therefore 
we  expected  little  bias  introduced  from  sampling  along 
the  100-m  isobath.  With  more  experience,  we  found  that 
we  could  tow  the  net  at  a shallower  depth  than  initially 
expected,  and  during  subsequent  surveys  we  occasion- 
ally fished  inshore  at  shallower  depths  (see  below,  Figs. 
2-5).  For  the  remaining  three  surveys,  most  trawls 
were  evenly  distributed  along  the  transect  line  and  be- 
tween transect  lines  in  the  inshore  area.  Occasionally, 
trawls  were  made  during  transit  between  transect  lines. 

Both  trawl  and  CalVET  samples  were  collected  dur- 
ing all  four  surveys  and  CUFES  samples  were  collected 
during  July  surveys  only  (Figs.  2-5).  Trawl-related 


176 


Fishery  Bulletin  108(2) 


48°N 


46°N 


44°N 


42°N 


40°N 


48°N 


46°N 


44°N 


42°N 


40°N 


• Positive  CalVET 
o Negative  CalVET 
▲ 0<egg  density  from  CUFES<0.5  eggs/min! 
J i Egg  density  from  CUFES>0  5 eggs/min 
|~j  High  density  spawning  region  j 
Temperature  contours  are  °C 


California 


100 


200 


128°W 


126°W 


124°W 


122°W 


128°W 


126°W 


124°W 


122“W 


Figure  2 

(A)  Locations  of  trawls  (stars)  used  for  the  estimation  of  biomass  of  Pacific  sardine  (Sardinops  sagax),  excluding  added 
trawls  in  the  inshore  area  (Table  1),  and  (B)  locations  of  California  vertical  egg  tows  (CalVET:  circles)  and  where  con- 
tinuous underway  egg  samples  were  taken  (measured  in  eggs  per  minute  (CUFES:  triangles).  Contours  are  sea  surface 
temperatures  (°C).  The  dashed  vertical  line  is  the  125°W  longitude  divider  of  the  two  sampling  strata.  The  offshore 
shaded  area  in  (B)  is  the  major  spawning  habitat.  Positive  tows  were  those  tows  during  which  sardine  were  caught. 
Negative  tows  were  tows  when  sardine  were  not  caught. 


station  activities  were  performed  between  twilight  and 
dawn,  whereas  CalVET  and  CUFES  samples  were  col- 
lected throughout  all  24  hours.  At  each  station,  a Cal- 
VET sample  was  collected  and  sea  surface  temperature 
(SST)  was  recorded,  whereas  between  stations,  CUFES 
samples  and  water  temperature  were  taken  at  a fixed 
3-m  depth  (Figs.  2 and  4).  The  CUFES  data  were  used 
primarily  to  map  the  spawning  area  based  on  the  den- 
sity of  sardine  eggs. 

A Nordic  264  trawl  (NET  Systems,  Bainbridge  Is- 
land, WA),  with  a vertical  opening  of  20  m,  a mouth 
area  of  approximately  360  m2,  and  a 7-mm  codend 
mesh  (Emmett  et  ah,  2005),  was  towed  to  sample  the 


upper  18-20  m of  the  water  column.  The  distance  trav- 
eled by  each  trawl  was  recorded  and  was  later  con- 
verted to  the  volume  sampled.  The  swept  area  (m2) 
is  the  volume  (m3)  divided  by  20  m.  During  the  July 
2003  survey,  few  trawls  were  taken  in  the  offshore 
area.  Additional  trawls  were  taken  inshore  to  collect 
extra  samples  to  determine  reproductive  parameters 
in  areas  of  sardine  spawning  activity  identified  by 
sardine  egg  densities  in  CUFES  samples  or  the  pres- 
ence of  schools  as  indicated  by  acoustic  signals  (Fig.  2). 
Similarly,  during  July  2004,  trawls  were  taken  in  the 
southern  spawning  area  off  Port  Orford,  OR  (Fig.  4), 
in  addition  to  the  prepositioned  and  between-transect 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardinops  sagax  off  the  Pacific  coast. 


177 


trawls.  Data  from  the  added  trawls 
were  excluded  in  estimating  the  total 
biomass  to  avoid  bias.  For  the  two 
March  surveys,  all  locations  (fixed 
stations  and  between-transects)  were 
trawled  regardless  of  spawning  or 
acoustic  signals  (Figs.  3 and  5).  The 
total  number  of  trawls  for  each  sur- 
vey was  close  to  50  (Table  1). 

For  each  trawl,  the  total  weight 
(kg)  of  the  Pacific  sardine  catch  was 
recorded  and  up  to  50  Pacific  sardine 
were  randomly  sampled  from  each 
trawl  where  sardine  were  caught 
(hereafter  referred  to  as  a “positive 
trawl”).  Sex  was  determined  for  each 
fish,  and  standard  length  (SL)  and 
weight  were  measured.  For  the  female 
fish,  the  ovaries  were  first  examined 
for  torpedo  shape  and  or  development 
of  visible  oocytes  (yolking  or  hydrat- 
ing). When  oocytes  were  not  visible 
and  the  ovary  was  small,  clear,  and 
torpedo  shaped,  the  ovary  was  re- 
corded as  code  1 (clearly  immature). 
Otherwise,  the  additional  ovarian 
codes  2 (intermediate),  3 (active),  or 
4 (hydrated)  (Table  2)  were  used  to 
identify  potentially  mature  females — 
because  only  histological  analysis  can 
verify  sardine  maturity  with  certain- 
ty (Macewicz  et  al.,  1996).  All  ovaries 
were  removed  and  preserved  in  10% 
neutral  buffered  formalin.  If  a 50- 
fish  subsample  did  not  have  25  po- 
tentially mature  females  (ovary  codes 
2-4),  more  females  were  sampled  to 
attain  25  per  trawl  for  estimation 
of  reproductive  parameters  used  for 
computing  spawning  biomass.  Addi- 
tional females  were  also  processed  to 
estimate  batch  fecundity,  but  were 
not  included  in  the  original  random 
subsample  for  length  distributions. 
We  also  obtained  length  distribu- 
tions based  on  data  from  commercial 
purse  seine  catches  off  the  PNW  in 
the  summer  seasons  and  from  a test 
purse  seine  set  in  March  2005. 


48°N 


46°N 


44°N 


42°N 


29  February-19  March  2004 


hf. 


*&::  ■ : ■ 


Astoria . WASHINGTON 
OREGON 


V 


/Coos  Bay 


OREGON 

CALIFORNIA 


40>^ 
38^4? 


36°N 


34°N 


32°N 


30°N 


N 


# Positive  CalVET 
Negative  CalVET 

a 0<egg  density  from  CUFES<1  eggs/min 
& Egg  density  from  CUFES>1  eggs/min 

■ j High  density  spawning  regi 

★ Positive  sardine  trawls 
☆ Negative  sardine  trawls 

Temperature  contours  are  °C 
I 1 

128°W  126°W  124°W  122°W  120°W  118°W  116°W 


Seasonal  biomass 

A swept-area  method  was  used  to 
estimate  the  total  biomass  of  Pacific 
sardine  in  summer  and  spring  based 
on  July  and  March  trawl  data,  respec- 
tively. Because  the  efficiency  of  the 
trawl  catch  has  not  been  evaluated, 
the  biomass  estimates  must  be  con- 
sidered as  relative  and  minimum 


Figure  3 

Locations  of  trawl  (stars)  and  California  vertical  egg  tows  (CalVET:  circles), 
for  2004  March  ichthyoplankton-trawl  survey  off  the  Pacific  Northwest 
(top  map),  and  locations  of  trawls,  CalVET  tows  (circles),  and  continuous 
underway  egg  sampling  (CUFES:  triangles)  for  the  March-April  2004 
California  Cooperative  Oceanic  Fisheries  Investigations  (CalCOFI)  daily 
egg  production  survey  (bottom  map).  Solid  symbols  indicate  that  Pacific 
sardine  ( Sardinops  sagax)  were  captured  in  the  sample  at  that  site.  Con- 
tours are  sea  surface  temperatures  (°C).  The  dashed  vertical  line  at  125°W 
longitude  (seen  in  top  map)  is  the  divider  of  the  two  sampling  strata.  The 
shaded  area  on  the  bottom  is  the  identified  spawning  habitat. 


178 


Fishery  Bulletin  108(2) 


Figure  4 

(A)  Locations  of  trawls  (stars)  used  for  biomass  estimation  of  Pacific  sardine  ( Sardinops  sagax),  excluding  added  trawls 
in  the  inshore  area  (Table  1),  and  (B)  California  vertical  egg  tows  (CalVET:  circles)  and  continuous  underway  egg 
sampling  in  eggs/minute  (CUFES:  triangles)  for  2004  July  trawl-ichthyoplankton  survey  off  the  Pacific  Northwest. 
Contours  are  sea  surface  temperatures  (°C).  The  dashed  vertical  line  is  the  125°W  longitude  divider  of  the  two  sam- 
pling strata.  The  shaded  area  is  the  major  spawning  habitat.  Positive  tows  were  those  tows  during  which  sardine  were 
caught.  Negative  tows  were  tows  when  sardine  were  not  caught. 


abundances.  A stratified  sampling  design  was  used 
to  estimate  biomass  and  spawning  biomass,  because 
more  stations  were  assigned  close  to  the  shore  than 
offshore.  Otherwise,  estimates  would  be  biased  toward 
the  inshore  area  (Holt  and  Smith,  1979).  The  survey 
area  was  divided  into  an  inshore  area  (stratum  1)  and 
an  offshore  area  (stratum  2)  with  125°W  longitude  as 
the  dividing  line.  For  the  July  2003  survey,  we  excluded 
the  nonpredetermined  trawls  (i.e.,  those  trawls  locations 
of  which  were  not  determined  before  the  survey)  taken 
in  the  vicinity  of  positive  trawls  to  prevent  an  overesti- 
mate of  the  total  biomass.  The  catch  for  each  tow  was 
expressed  as  kg/m2  ( = catch  [kg]/swept  areaT  m2]=catch 
[kg]/volume  of  water  [m3]/depth  20  m),  where  the  volume 


of  water  filtered  was  computed  as  the  distance  covered 
by  each  tow  multiplied  by  the  area  of  the  vertical  trawl 
mouth  opening  of  approximately  360  m2  (with  20  m as 
diameter).  We  estimated  relative  total  biomass  (B)  and 
its  standard  error  (SE)  for  each  survey  as  follows: 

S = ^ ( A,  106 ) / 1000,  (1) 

i 

SE(B)  = ( ^ ( var( Xl ) (A,  106  )2 )1/2/ 1000  ( 2 ) 

i 

where  B = the  estimate  of  the  total  biomass  (t); 

Xi  = the  mean  catch  (kg/m2);  and 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardmops  sagax  off  the  Pacific  coast. 


179 


A(  = the  area  (km2)  in  stratum 
i,  i=l  (inshore)  and  2 (off- 
shore). 

Note:  the  coefficient  of  variance  (CV) 
of  the  estimate  is  CV(B)  - SE (B)/B. 
Bootstrap  simulation  was  used  to  esti- 
mate the  bias  of  the  estimate  (Eq.  1), 
and  the  bias-corrected  estimate  (Bc) 
as  Bc=B-(Bb-B),  where  B is  computed 
from  Equation  1,  Bh  is  the  estimate 
from  the  bootstrap  simulation,  and  the 
mean  square  error  (MSE -variance  + 
bias2)  of  the  biomass  estimates  (Eq.  2). 

We  also  computed  a crude  esti- 
mate of  the  recruit  biomass  (age- 
zero  year  or  incoming  year  class)  as 
ancillary  information  for  compara- 
tive purposes  for  spring  in  2004  and 
2005,  based  on  the  biomass  of  fish 
<120  mm  SL  because  120  mm  was 
the  break  point  for  the  length-fre- 
quency distribution  in  March  surveys 
from  this  study  (Fig.  6)  and  it  was 
reported  that  age-0  sardine  in  the 
PNW  were  <110  mm  (measured  by 
fork  length)  (Emmett  et  al.,  2005). 
Recruit  biomass  (BR)  was  estimated 
by  using  Equations  1 and  2,  where 
Xt=  the  mean  catch  (kg/m2)  of  fish 
<120  mm  SL  in  the  tth  stratum  ( XR  ■ ). 
The  catch  of  recruits  for  each  trawl 
would  be  obtained  as 

XR,ij=Xi*Uij,  length  <120mm’  where  Xij  = 

the  total  catch  from  the  yth  trawl, 
and  U lengths  120  mm  =the  Weight  of  fish 

<120  mm  SL  divided  by  the  total  fish 
weight  based  on  our  random  samples 
with  a maximum  of  50  fish  from  each 
tow. 

Spawning  habitat 

The  spawning  habitat  was  defined  as 
the  area  of  relatively  high  egg  densi- 
ties during  early  summer,  because 
June-July  was  the  peak  spawning 
time  for  Pacific  sardine  off  the  PNW 
as  determined  from  egg  and  larval 
data  collected  in  the  mid-1990s  (Bent- 
ley et  al.,  1996).  Because  the  number 
of  positive  CalVET  tows  was  low  (four 
of  54  tows  during  July  2003  and  3 of 
48  tows  during  July  2004),  we  chose  to 
use  data  from  CUFES  sampling.  The 
spawning  habitat  area  was  defined 
as  the  area  where  the  majority  of  egg 
densities  exceeded  a threshold  of  0.5 
eggs/min  because  the  egg  densities 
were  generally  low.  Off  California, 


48°N  - 


46°N 


44°N  - 


42°N 


36°N 


34°N  - 


32°N 


128°W 


126°W 


124°W 


122°W 


120°W 


118°W 


116°W 


40'X, 

38°^ 


30°N 


Temperature  contours  are  °C 


• Positive  CalVET 
Negative  CalVET 

* 0<egg  density  from  CUFES<1  eggs/min 
± Egg  density  from  CUFES>1  eggs/min 

fH  High  density  spawning  region 

★ Positive  sardine  trawls 
☆ Negative  sardine  trawls 


) 

1 5 April— 


1 May  2005 

Long  Beach 


2-21  March  2005 


Figure  5 

Locations  of  trawl  (stars)  and  California  vertical  egg  tows  (CalVET: 
circles)  for  the  2005  March  ichthyoplankton-trawl  survey  off  the  Pacific 
Northwest  (top)  and  trawls,  CalVET  tows  (circles),  and  continuous 
underway  egg  samples  (CUFES:  triangles)  for  the  April-May  2005 
California  Cooperative  Oceanic  Fisheries  Investigations  (CalCOFI) 
daily  egg  production  survey  (bottom).  Solid  symbols  indicate  that  Pacific 
sardine  (Sardinops  sagax)  were  captured  in  the  sample  at  that  site. 
Contours  are  sea  surface  temperatures  (°C).  The  dashed  vertical  line 
is  the  125°W  longitude  divider  of  the  two  sampling  strata.  The  shaded 
area  is  the  spawning  habitat. 


180 


Fishery  Bulletin  108(2) 


Table  1 

Estimates  of  biomass  of  Pacific  sardine  ( Sardinops  sagax),  biomass-related  parameters  for  each  survey  (stratum  1,  stratum  2 
[with  latitude  125°W  as  the  dividing  line  between  them],  and  the  entire  survey  area),  and  confidence  intervals  for  biomass  and 
recruits:  incoming  year  class  (fish  <120  mm  standard  length)  in  tons  (t).  Either  the  coefficient  of  variation  (CV)  or  number  of 
positive  (pos)  trawls  is  in  parentheses. 


Stratum  1 

Stratum  2 

Entire  survey  area 

July  2003 

Mean  density  (kg/m3) 

2.41e-°04 

2.2e-°06 

4.85e-°05 

Biomass  (t)  (CV) 

192,801(0.57) 

7207(0.85) 

200,008(0.56) 

No  of  trawls1  (pos) 

38(34) 

10(2) 

48(36) 

No  of  trawls  used  for  biomass  (positive) 

17(14) 

5(2) 

22(16) 

Survey  area  (km2)  (%  of  entire  survey  area) 

40,043(19) 

166,333(81) 

206,377(100) 

Bootstrap  results 

Biomass  (t)  (CV) 

193,946(0.53) 

7406(0.82) 

201,360(0.51) 

Mean  square  error  (MSE)1/2 

102,174 

6094 

102,378 

Bias-corrected 
Confidence  interval  (t) 

191,656 

7009 

198,656 

44,286-421,321 

March  2004 

Mean  density(kg/m3) 

2.6e-°05 

4 6e-°08 

5.2e-°06 

Biomass  (t)  (CV) 

21,243(0.83) 

155(0.7) 

21,398(0.82) 

Recruits  (t)(  CV ) 

21,030(0.83) 

69(1.0) 

21,099(0.83) 

No  of  trawls  (pos) 

25(7) 

34(2) 

59(9) 

Survey  area  (km2)  (%) 

40,043(19) 

166,334(81) 

206,377(100) 

Bootstrap  results 

Biomass  (t)  (CV) 

22,494(0.81) 

155(0.69) 

22,650(0.8) 

MSE1/2 

18,260 

106 

18260 

Bias-corrected 
Confidence  interval  (t) 

19,992 

155 

20,147 

0-63,017 

Recruits  (t)(CV) 

21,629(0.81) 

69(1.02) 

21,698(0.81) 

MSE1/2 

17,576 

70 

17,578 

Bias-corrected 
Confidence  interval  (t) 

20,432 

69 

20,501 

4749-45,808 

July  2004 

Mean  density  (kg/m3) 

9.0e-005 

2.3e-°06 

2 ie-°°5 

Biomass  (t)  (CV) 

72,206(0.405) 

6989(0.992) 

79,194(0.379) 

No  of  trawls1  (pos) 

20(16) 

38(11) 

58(27) 

No  of  trawls  used  for  biomass  (pos) 

17(15) 

30(14) 

47(19) 

Survey  area  (km2)  (%) 

40,043(21) 

150,932(79) 

190,975(100) 

Bootstrap  results 

Biomasst  t)  (CV) 

73,186(0.41) 

7299(0.95) 

80,485(0.38) 

MSE1/2 

29,723 

6928 

30,605 

Bias-corrected 
Confidence  interval  (t) 

71,226 

6678 

77,903 

30,474-146,176 

March  2005 

Mean  density  (kg/m3) 

3.7e-°05 

2.3e-007 

7.9e~006 

Biomass  (t)(CV) 

29,488(0.69) 

700(0.57) 

30,188(0.68) 

Recruits  (t)  (CV) 

55(1.0) 

0(0) 

54.80(1.0) 

No  of  trawls  (pos) 

15(11) 

34(9) 

49(20) 

Survey  area  (km2)  (%) 

40,043(21) 

150,932(79) 

190,976(100) 

Bootstrap  results 

Biomass  (t)(CV) 

29,573(0.73) 

705(0.57) 

30,278(0.72) 

MSE1/2 

21,713 

402 

21,714 

Bias-corrected 
Confidence  interval  (t) 

29,403 

695 

30,098 

1800-86,035 

Recruits  (t)  (CV) 

56.6(0.98) 

0(0) 

57(0.98) 

MSE1/2 

56 

0 

56 

Bias-corrected 
Confidence  interval  (t) 

53 

0 

53 

70-1640 

1 During  the  July  2003  cruise,  data  from  22  out  of  48  trawls  were  used  for  biomass  computation.  The  total  48  trawls  included  38  (34  with  sardine) 
in  stratum  1 and  10  (2  with  sardine)  in  stratum  2.  During  the  July  2004  cruise,  only  data  from  the  first  47  trawls  out  of  58  trawls  were  used.  The 
total  58  trawls  included  20  trawls  (16  with  sardine)  in  stratum  1 and  38  trawls  (11  with  sardine)  in  stratum  2. 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardmops  sagax  off  the  Pacific  coast. 


181 


Table  2 

Gross  anatomical  classification  of  female  and  male  Pacific  sardine  ( Sardinops  sagax)  gonads. 

Gonad  code 

Female:  Ovary  description 

1 

Clearly  immature : Oocytes  are  not  visible.  Ovary  is  very  small,  translucent  or  clear,  and  thin,  but  with 
rounded  edges  (torpedo  shaped). 

2 

Intermediate'.  Individual  oocytes  are  not  visible  to  the  unaided  eye  (no  visible  yolk  or  hydrate  oocytes  in 
the  ovaries),  but  ovary  is  not  clearly  immature.  Includes  possible  maturing  and  regressed  ovaries. 

3 

Active : Yolked  oocytes  in  ovaries  visible  to  the  unaided  eye  in  any  size  or  amount,  including  the  smaller 
opaque  oocytes  (around  0.4-0. 5 mm)  to  the  large  yellowish  oocytes  (about  0.6-0. 8mm). 

4 

Hydrated : Hydrated  oocytes  are  present,  yolked  oocytes  may  also  be  seen.  Hydrated  oocytes  (large  and 
transparent),  from  few  to  many,  or  even  if  loose  or  “oozing”  or  “running”  from  ovary,  qualify  for  this 
class 

Male: 

Testis  description 

1 

Clearly  immature'.  Testis  is  very  small,  knife  shaped,  translucent  or  clear,  and  thin  with  a flat  ventral 
edge. 

2 

Intermediate'.  No  milt  is  evident  and  testis  is  not  clearly  immature  (includes  maturing  or  regressed 
testes). 

3 

Active:  Milt  is  present  either  oozing  from  the  gonopore,  in  the  duct,  or  in  the  testis  (observed  when  the 
testis  was  cut). 

the  threshold  was  one  egg/min.  We  obtained  the  SST 
for  CUFES  samples  with  >0.5  eggs/min  as  a proxy  for 
the  oceanographic  conditions.  No  biological  variables 
such  as  zooplankton  volume  (Lynn,  2003)  were  collected 
during  these  surveys. 

Daily  egg  production 

The  daily  egg  production  (P0)  is  defined  as  the  newly 
spawned  eggs  produced  per  0.05  m2  per  day,  where  0.05 
m2  was  the  surface  area  covered  by  the  CalVET  net  tow 
The  daily  rate  of  egg  production  and  the  daily  specific 
fecundity  rate  from  adult  parameters  (Lasker,  1985) 
are  needed  to  compute  spawning  biomass.  In  California 
waters,  sardine  egg  data  from  CalVET  tows  and  yolksac 
larval  data  from  both  CalVET  tows  and  bongo  nets,  and 
sardine  ages  were  used  to  model  the  embryonic  mortal- 
ity curve,  a negative  exponential  curve  (Lo  et  al.,  1996, 
2005): 

Pl  = P0e^zt\  (3) 

where  Pt  = the  daily  production  rate  at  age  t (days); 

2 = the  daily  instantaneous  embryonic  mortality 
rate;  and 

P 0 = the  intercept,  is  the  daily  egg  production  at 
age  zero. 

Because  few  eggs  were  caught  during  CalVET  net  tows 
in  July  surveys  and  no  eggs  were  caught  in  March 
surveys  (Fig.  2-5,  Table  3),  no  attempt  was  made  to 
estimate  egg  production  for  the  March  surveys.  For  July 
surveys,  it  was  impossible  to  model  the  egg  mortality 
curve  because  the  mortality  curve  requires  sufficient 


data  on  egg  abundance  for  each  egg  stage  and  age. 
Instead,  we  used  an  alternative  algorithm  to  estimate 
P0,  an  integral  method  (P0 1)  based  on  the  standing  stock 
of  eggs  from  CalVET  tows. 

The  estimate  of  P0  (P0  7)  was  based  on  the  relation- 
ship between  the  mean  catch  of  eggs  from  CalVET 
tows  (Y)  and  egg  production  (P0)  through  the  inte- 
gral of  Pt  over  the  period  from  spawning  to  hatching 
(th).  The  mean  catch  of  eggs  from  CalVET  tows  was 
a weighted  average  with  the  area  in  each  stratum  as 
weight.  This  method  requires  prior  knowledge  of  the 
egg  mortality  rate  and  the  temperature-dependent 
hatching  time: 

*h  th 

Y=^Ptdt  = ^ P0e~ztdt.  (4) 

o o 

Integrating  the  above  equation  yields  the  estimate  of 
P0  as  a function  of  the  mean  egg  density,  Y,  incubation 
time,  th  , and  the  daily  instantaneous  mortality  rate,  2: 


with  variance  calculated  by  using  the  delta  method: 

<3Pa  J 9 d Pv  1 9 

var(  P0  7 ) = ( J r var(  2)  + ( - )"  var(  Y), 

az  oY 

Y[1  - exp(-2^  )(1  + zth )]  2 , , 

= ( - 0 ) var(2) 

[1-exp  {-zth)Y 

+( )2  var(Y). 

1 - exp(  -zth ) 


182 


Fishery  Bulletin  108(2) 


The  2 value  was  the  estimate  from  the  daily  egg  pro- 
duction method  (DEPM)  surveys  off  California  in  2003 
(0.48  [CV=0.08])  and  2004  (0.25  [CV=0.04])  (Lo  et  ah, 
2005)  because  of  the  lack  of  sufficient  data  to  estimate 
2 off  the  PNW.  Age  at  hatching  (in  days)  was  2.5  days 
computed  from  the  temperature-dependent  sardine 
egg  development  model  for  stage  XII  given  in  Lo  et  al. 
(1996):  ^ = 30.65*  exp(-0.145*temp-0.037*12)*121  41/24, 
where  temp  is  the  average  temperature  from  positive 
CUFES  collections  during  the  July  surveys,  and  equals 


16.4°C  and  16.3°C  for  2003  and  2004,  respectively.  This 
integral  estimate  is  biased  upward  on  the  basis  of  a com- 
parison of  P0  j and  the  P0  from  the  nonlinear  regression 
from  four  California  daily  egg  production  surveys  and 
a simple  theoretical  population.  Both  cases  indicated 
that  the  relative  bias  (rb  = ( P0I  — P0 )/  P0I)  was  close 
to  20%  of  P0  j Thus  the  bias-corrected  egg  production 
(P o,f)  would  be  P0c=P0I  (l-rb)=_P0  7(0.8). 

The  mean  density  of  eggs  (Y-)  (eggs/0.05  m2)  was 
estimated  for  each  of  two  strata  (i=l,2),  with  125°W 
latitude  as  the  dividing  line. 
The  overall  mean  density  (Y) 
for  the  whole  survey  area  was  a 
weighted  average  with  the  area 
in  each  stratum  as  the  weight 
and  was  used  to  estimate  the 
daily  egg  production.  No  esti- 
mate of  egg  production  for  each 
stratum  was  obtained  because 
of  the  small  sample  sizes. 

To  understand  the  relative 
contribution  of  egg  production 
from  the  PNW  area,  we  com- 
puted a ratio  of  the  total  egg 
production  in  the  PNW  to  the 
total  egg  production  in  the 
whole  area  (PNW  and  Califor- 
nia) as  P0  j Aj/^P0  j Aj  , where 
P0  ■ is  the  daily  egg  production 
during  the  peak  spawning  time 
in  the  survey  area  Ay,  j= 1 re- 
fers to  the  PNW  area  in  July 
and  j- 2 refers  to  California  in 
April. 

Adult  reproductive  state 
and  parameters 

For  all  four  surveys,  we  used 
histological  analysis  of  all  ovar- 
ian tissues,  along  with  trawl  and 
female  data,  to  provide  accurate 
assessment  of  adult  parameters 
and  reproductive  state  such  as 
maturity,  spawning  period, 
recent  spawning  activity,  post- 
spawning condition,  or  iden- 
tification of  advanced  oocyte 
development  for  a selection  of 
females  for  batch  fecundity  esti- 
mation. In  the  laboratory,  each 
preserved  ovary  was  blotted  and 
weighed  to  the  nearest  mg.  A 
piece  of  each  ovary  was  removed, 
a histological  slide  was  pre- 
pared, and  the  tissue  sections 
were  stained  with  hematoxylin 
and  eosin.  We  analyzed  oocyte 
development,  atresia,  and  post- 
ovulatory follicle  age  to  assign 


Washington-Oregon 


California 


-p 

03 


0.35 
0.30 
0 25 
0.20 
0 15 
0.10 
0.05 
0.00 


March  2004 


-M- 


0.35 
0 30 
0.25 
0.20 
0 15 
0.10 
0.05 
0.00 


March-April  2004 


80  100  120  140  160  180  200  220  240  260  280 


80  100  120  140  160  180  200  220  240  260  280 


80  100  120  140  160  180  200  220  240  260  280 


Standard  length  class  (mm) 

Figure  6 

Length  frequency  distribution  of  Pacific  sardine  ( Sardinops  sagax)  off  Washington 
and  Oregon,  and  California  during  2003,  2004,  and  2005  from  fishery  independent 
trawl  surveys  (black  bars)  and  port  sampling  of  commercial  purse  seine  catches 
(gray  bars).  Catch  data  were  provided  by  California  Dept,  of  Fish  and  Game,  Oregon 
Dept,  of  Fish  and  Wildlife,  and  Washington  Dept,  of  Fish  and  Wildlife. 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardinops  sagax  off  the  Pacific  coast. 


183 


Table  3 

Estimated  Pacific  sardine  (Sardinops  sagax ) egg  densities,  egg  production  ( P0  c,  Eq.  5)  with  coefficient  of  variation  (CV)  in  paren- 
theses, and  number  of  collections  with  positive  collections  in  parentheses  from  California  Cooperative  Oceanic  Fisheries  Inves- 
tigation vertical  egg  tow  net  (CalVET)  and  continuous  underway  fish  egg  sampler  (CUFES)  samples  in  two  strata  with  dividing 
latitude  of  125°W  and  the  entire  survey  area  for  the  July  2003  and  July  2004  surveys.  Dashes  indicate  where  statistics  were  not 
computed  because  of  small  or  zero  catches. 


Stratum  1 

Stratum  2 

Entire  survey  area 

July  2003 
CalVET 

Egg  density  (eggs/0.05  m2)(CV) 

0.388(0.51) 

0 

0.073(0.51) 

P0  c (Egg  production  /0.05  m2/day)(CV) 

— 

— 

0.04(0.51) 

No.  of  CalVET  tows  (positive) 

18(4) 

36(0) 

54(4) 

CUFES 

Eggs/min  (CV) 

0.148(0.61) 

0.05(0.74) 

0.069(0.49) 

No  of  CUFES  samples  (positive) 

316(102) 

166(15) 

482(117) 

Survey  area  (km2)(%) 

40,043(19) 

166,334(81) 

206,377(100) 

July  2004 
CalVET 

Egg  density  (eggs/0.05  m2)(CV) 

0.0(— ) 

0.088(0.56) 

0.070(0.56) 

P0  c (Egg  production  /0.05  m2/day)(CV) 

— 

— 

0.037(0.58) 

No.  of  CalVET  tows  (positive) 

14(0) 

34(3) 

48(3) 

CUFES 

Eggs/min  (CV) 

0.11(0.42) 

0.097(0.67) 

0.1(0.53) 

No  of  CUFES  samples  (positive) 

197(65) 

450(64) 

647(129) 

Survey  area  (km2)(%) 

40,043(21) 

150,932(79) 

190,975(100) 

Table  4 

Percentage  and  average  size  in  each  maturity  class  of  Pacific  sardine  ( Sardinops  sagax ) females  in  the  random  samples  from 
trawls  conducted  during  four  research  surveys  in  2003-05  off  Oregon  and  Washington.  Maturity  was  based  on  histological 
analysis  of  ovaries. 


Survey  dates 
(n  females) 

Maturity  class 

Percentage 
of  females 

Mean  standard 
length  (mm) 

Mean  whole 
body  weight  (g) 

6-25  July  2003 

Immature 

0.7 

204 

124 

(690) 

Mature 

98.3 

238 

194 

29  Feburary-19  March  2004 

Immature 

97.2 

108 

14 

(108) 

Mature 

2.8 

207 

105 

6-25  July  2004 

Immature 

62.2 

147 

43 

(410) 

Mature 

37.8 

240 

200 

2-21March  2005 

Immature 

89.2 

161 

51 

(241) 

Mature 

10.8 

195 

87 

female  maturity  and  reproductive  state  (Macewicz  et 
al.,  1996;  Lo  et  al.,  2005). 

Sufficient  numbers  of  immature  and  mature  females 
in  the  random  50-fish  subsample  of  a positive  trawl 
for  estimation  of  the  length  at  which  50%  were  ma- 
ture were  collected  during  July  2004  and  March  2005 
(Table=4).  Females  were  grouped  into  10-mm  length 
classes  and  the  length  at  which  50%  were  mature  was 
estimated  by  logistic  regression:  y=l/(l  + e~<a+bL)),  where 
y = the  proportion  of  mature  female  sardine  and  L = the 


standard  length  in  mm.  The  length-specific  maturation 
relationships  were  compared  to  those  off  California  in 
April  1994,  2004,  and  2005  (Macewicz  et  al.,  1996;  Lo 
et  al.,  2005). 

Because  the  spawning  season  occurs  in  early  summer, 
we  used  the  two  sets  of  July  survey  data  to  estimate 
the  following  adult  reproductive  parameters,  which  were 
used  in  the  spawning  biomass  computation  based  on 
the  daily  egg  production  method  (Lasker, 1985;  Parker, 
1985;  Lo  et  al.,  2005):  the  daily  spawning  fraction  (S) 


184 


Fishery  Bulletin  108(2) 


or  the  fraction  of  mature  females  spawning  per  day;  the 
average  batch  fecundity  (number  of  eggs  per  spawning 
per  mature  female:  F );  the  fraction  of  mature  fish  that 
were  female  by  weight  (sex  ratio:  R );  and  the  average 
weight  of  mature  females  (g)  (WJ.  The  reproductive 
parameters  were  estimated  from  the  data  on  the  first 
25  mature  females  per  trawl  or  all  mature  females  if 
there  were  <25  by  following  the  methods  in  Macewicz 
et  al.  (1996).  Females  with  ovaries  histologically  iden- 
tified as  containing  hydrated  oocytes  (hydrated  ovary) 
have  temporarily  inflated  ovary  weights.  For  each  July 
survey,  the  relation  between  wet  weight  (y)  and  ovary- 
free  wet  weight  (x)  from  mature  females  lacking  hy- 
drated oocytes  was  determined  as  y=-9.0998  + 1.0758x 
in  2003  and  y=-6.316  + 1.05608x  in  2004.  Thus,  the 
observed  female  weight  was  adjusted  downward  for  fe- 
males with  hydrated  ovaries  when  calculating  average 
mature  female  weight  (Wf)  for  each  collection  by  year. 
During  March  of  2004  and  2005,  adjustments  were  not 
necessary  and  fecundity  was  not  estimated  from  mature 
females  caught  because  none  of  them  had  ovaries  with 
oocytes  in  the  migratory-nucleus  or  hydrated  stages. 
Mean  batch  fecundity  was  estimated  by  the  gravimet- 
ric method  for  54  females  from  21  trawls  from  the  July 
surveys.  The  relationship  of  batch  fecundity  to  female 
weight  (without  ovary)  was  then  determined. 

Reproductive  adult  parameters  were  summarized  for 
each  trawl.  Population  values  were  estimated  by  meth- 
ods in  Picquelle  and  Stauffer  (1985),  where  estimation 
of  each  adult  parameter  (S,  F,  W,  R)  was  based  on  a 
ratio  estimator  (Picquelle  and  Stauffer,  1985;  Lo  et  ah, 
1996)  and  used  to  calculate  spawning  biomass  and  its 
covariance  for  the  July  2003  and  July  2004  surveys. 

Spawning  biomass 


The  denominator  (. RSF/W* ) is  referred  to  as  the  daily 
specific  fecundity  (number  of  eggs/population  weight 
[g]/day). 

The  variance  of  the  spawning  biomass  estimate  (Bs) 
was  computed  from  the  Taylor  expansion  in  terms  of 
the  coefficient  of  variation  (CV)  for  each  parameter 
estimate  and  covariance  for  adult  parameter  estimates 
(Parker,  1985;  Picquelle  and  Stauffer,  1985;  Lo  et  al., 
1996;  2005): 


var(bs)=bs 


cv(p0f +cv(w/.)2 +cv(s)2  + 
cv  (r)2 + CV  (f)2  + 2 covs 


(8) 


The  last  term,  involving  the  covariance  term,  on  the 
right-hand  side  is 


COVS  = 

i i<j 


COV  ( Xj , Xj  j 
xtx- 


(9) 


where  x;=the  z'th  adult  parameter  estimate,  e.g.,  xt=F 
and  Xj=Wf.  The  sign  of  any  two  terms  is  positive  if  they 
are  both  in  the  numerator  of  Bs  or  denominator  of  Bs 
(Eq.  7);  otherwise,  the  sign  is  negative.  The  covariance 
term  is 


COV(  Xj  Xj  ) 


[n  I (n-l)\^mk{xlk-  x^g^Xj  k-  Xj) 
* , (10) 


( \ 

( \ 

x** 

V k , 

\ k J 

The  DEPM  is  a well-accepted  method  used  for  estimating 
spawning  biomass  for  fish  with  indeterminate  fecundity, 
i.e.  multiple  spawners  (Hunter  and  Lo,  1993;  Stratouda- 
kis  et  al.,  2006)  and  was  used  to  estimate  the  spawning 
biomass  of  Pacific  sardine  in  this  area  in  1994  (Bentley 
et  al.,  1996).  The  spawning  biomass  was  computed  with 
the  following  equation: 


where  k = kth  tow,  and  k=l,  ...  , n\ 
mk  and  gk  = sample  sizes;  and 
x£  k and  Xj  k = sample  means  from  the  kth  tow  for  x;  and 
Xj  respectively. 

Results 


, PqAC 
>s  RSF/Wf 


(7) 


where  P0  = 

A = 
C = 

R = 

S = 

F = 


the  daily  egg  production/0.05  m2  at  hatch- 
ing; 

the  survey  area  in  units  of  0.05  m2; 

the  conversion  factor  from  grams  (g)  to 

metric  tons  (t); 

the  fraction  of  mature  fish  that  is  female,  by 
weight  (sex  ratio); 

the  daily  spawning  fraction:  fraction  of 

mature  females  spawning  per  day; 

the  average  batch  fecundity  (number  of  eggs 

per  spawn  per  mature  female);  and 

the  average  weight  of  mature  females  (g). 


Seasonal  biomass 

The  relative  abundance  of  Pacific  sardine  was  higher 
in  summer  than  in  the  following  spring  off  the  PNW. 
The  bias-corrected  seasonal  biomass  estimates  were 
198,600  t (CV=0.51)  for  July  2003,  20,100  t (CV=0.80) 
for  March  2004,  77,900  t (CV=0.38)  for  July  2004,  and 
30,100  t (CV=0.72)  for  March  2005  over  an  area  close 
to  200,000  km2  (Table  1).  The  inshore  stratum  1 made 
up  20%  of  the  survey  area.  Yet,  for  all  years  stratum  1 
had  over  80%  of  the  biomass.  The  recruit  biomasses  (fish 
<120  mm  SL)  in  spring  of  2004  and  2005  were  quite  dif- 
ferent: 20,500  t (CV=0.81)  for  the  2003  year  class  and 
53  t (CV=0.72)  for  the  2004  year  class,  respectively.  The 
2004  point  estimate  of  the  recruit  biomass,  20,500  t, 
was  greater  than  that  of  the  total  biomass  of  20,100  t 


Lo  et  al  Biomass  and  reproductive  status  of  Sardinops  sagax  off  the  Pacific  coast. 


185 


but  this  was  primarily  due  to  the  bias  correction  based 
on  the  bootstrap  simulation  and  the  difference  was  not 
statistically  significant. 

The  relatively  large  2003  year  class  constituted  a 
major  proportion  of  the  total  biomass  in  March  2004, 
whereas  the  2004  year  class  constituted  a very  small 
proportion  of  the  fish  in  2005  (Fig.  6).  Therefore,  the 
relative  abundance  of  Pacific  sardine  in  the  spring  of 
2004  and  2005  was  primarily  supported  by  the  strong 
year  class  of  2003. 

Spawning  habitat 

The  spawning  habitat  was  located  east  of  125. 5°W  longi- 
tude in  July  2003  and  2004  (Figs.  2 and  4),  and  between 
43°  and  44.5°N  latitude  in  July  2003,  and  between  42° 
and  44.5°N  latitude  in  July  2004.  The  location  of  the 
spawning  center,  computed  as  the  weighted  latitude  and 
longitude  with  the  eggs/min  (>0.5)  as  the  weight,  was 
124. 7°W  and  43.7°N  in  2003  and  125.13°W  and  42.9°N 
for  2004.  Therefore,  the  spawning  habitat  shifted  south- 
westward  from  2003  to  2004.  Because  the  eggs  from  the 
CUFES  samples  were  distributed  more  to  the  west,  the 
size  of  the  spawning  habitat  was  10,716  km2  for  2003  and 
14,260  km2  for  2004.  The  spawning  habitat,  determined 
from  CUFES  data,  crossed  the  dividing  line  of  125°W 
between  two  strata  based  on  trawl  allocation.  For  both 
July  cruises,  the  range  of  SST  in  the  spawning  habitat 
was  13.4-18.5°C  with  a mean  close  to  16°C  (15.7°C  and 
16.0°C  for  2003  and  2004).  Note  that  the  overall  mean 
SST  for  July  2003  was  16.2°C  (range  9.4-25.3°C)  and 
the  mean  temperature  was  16.8°C  (range  9.7-19.9°C) 
in  July  2004.  The  number  of  positive  CUFES  samples 
was  117  out  of  482  in  July  2003  and  129  out  of  647  in 
July  2004.  Therefore,  the  proportion  of  positive  samples 
(24%  in  2003,  19%  in  2004)  was  similar  during  these 
two  years. 

Daily  egg  production 

The  mean  density  of  eggs  was  0.388  eggs/0.05  m2 
(CV=0.51)  in  stratum  1 and  no  eggs  were  caught  by 
CalVET  net  tows  in  stratum  2 during  the  July  2003 
survey.  The  opposite  was  true  for  the  July  2004  survey: 
no  eggs  were  caught  in  stratum  1 and  the  mean  density 
in  stratum  2 was  0.088  eggs/0.05  m2  (CV=0.56)  (Table 
3).  The  overall  mean  densities  were  0.073  eggs/0.05  m2 
(0.51)  and  0.07eggs/0.05  m2  (0.49)  for  2003  and  2004, 
respectively.  The  bias-corrected  estimates  of  the  daily 
egg  production  from  the  integral  method  (P0c)  (Eqs. 
5 and  6)  in  July  were  0.04  eggs  produced/0.05  m2/ 
day(CV=0.51)  for  2003  and  0.037  eggs  produced/0.05 
m2  /day  (CV=0.58)  for  2004.  The  mean  egg  capture 
rates  from  CUFES  samples  for  2003  and  2004  were 
0.069  eggs/min  (CV=0.49)  and  0.1  eggs/min  (CV=0.53) 
(Table  3). 

The  ratio  of  the  total  egg  production  in  the  PNW  to 
the  total  egg  production  off  the  U.S.  west  coast  (PNW 
and  California)  was  1.46%  and  2.2%  for  2003  and  2004 
and  therefore  Pacific  sardine  off  the  PNW  contrib- 


uted approximately  to  1.8%  of  the  total  egg  production 
(Table  5). 

Adult  sardine  reproductive  parameters 
and  spawning  biomass 

During  the  four  surveys,  92  of  the  214  trawls  (Figs.  2-5, 
Table  1)  captured  adults  or  subadults.  In  the  random 
subsamples  from  these  trawls,  2862  sardine  were  mea- 
sured (Fig.  6);  standard  length  ranged  from  99-289  mm 
for  females,  106-281  mm  for  males,  and  75-146  mm  for 
individuals  of  indeterminate  sex  (where  it  was  difficult 
to  accurately  determine  sex  without  microscopic  exami- 
nation). Nearly  all  females  were  mature  in  July  2003  and 
nearly  all  were  immature  in  March  2004  (Table  4).  Using 
logistic  regression  we  computed  the  standard  length  at 
which  50%  were  mature  as  195.1  mm  and  199.8  mm  for 
July  2004  and  March  2005,  respectively  (Fig.  7). 

Mean  batch  fecundity  was  estimated  for  35  females 
caught  in  July  2003  and  19  from  July  2004  (Fig.  8). 
Analysis  of  covariance  showed  no  differences  in  the 
relationship  between  female  weight  (without  ovary,  W0f) 
and  batch  fecundity  (Fb)  among  years  (P=0.531).  Com- 
bining the  data  from  July  2003  and  2004,  we  found 
that  the  relationship  between  female  weight  and  batch 
fecundity,  as  determined  by  simple  linear  regression, 
was  Fft  = -16755  + 372.1Wo/  with  the  r2  = 0.47.  Because 
the  intercept  did  not  differ  from  zero  (P=0.165),  we 
chose  the  regression  without  the  intercept,  which  yield- 
ed the  relationship  Fb  = 295.83Wo ^ , where  W ’<■  ranged 
from  111-322  g (Fig.  8).  The  latter  equation  was  used 
to  calculate  batch  fecundity  for  each  mature  Pacific 
sardine  female  in  the  July  trawl  samples. 

The  population  sex  ratio  ( R ) for  mature  fish  was 
0.534  female  (CV=0.04)  in  July  2003  and  0.568  female 
(CV=0.05)  in  July  2004  (Table  5).  The  657  mature  fe- 
male Pacific  sardine  analyzed  from  July  2003  and  196 
from  July  2004  were  considered  a random  sample  of 
the  population  in  the  area  trawled.  Population-level 
estimates  of  the  other  adult  reproductive  parameters 
were  as  follows:  average  batch  fecundity  {F)=  55,986 
eggs/spawning  event  (CV=0.04)  in  July  2003  and 
55,883  eggs/spawning  (CV=0.06)  in  July  2004;  daily 
spawning  fraction  (S)  = 0.027  (CV=0.31)  in  2003  and 
S = 0.010  (CV=0.74)  in  2004;  and  mean  mature  female 
fish  weight  (W^= 194.36  g (CV=0.02)  in  2003,  and  193.16 
g (CV=0.03)  in  2004  (Table  5).  The  daily  specific  fecun- 
dity was  calculated  as  4.21  and  1.68  eggs/gm/day  in 
2003  and  2004,  respectively  (Table  5).  The  proportion 
of  active  females  spawning  was  0.05  and  0.025  for  July 
2003  and  2004,  respectively,  which  meant  that  the  av- 
erage female  was  spawning  roughly  once  every  20  to 
40  days.  None  of  the  three  mature  females  caught  in 
March  2004  or  the  37  mature  females  caught  in  March 
2005  had  histological  evidence  of  imminent  or  recent 
spawning  (hydrating  oocytes  or  postovulatory  follicles), 
and  thus  S = 0;  hence,  spawning  biomass  was  not  esti- 
mated for  either  March  (Table  5). 

The  estimated  spawning  biomass  based  on  biased 
corrected  egg  production  from  the  integral  method  (P0  c) 


186 


Fishery  Bulletin  108(2) 


and  the  adult  reproductive  parameters  for  July  2003 
and  July  2004  (Eq.  7,  Table  5)  was  39,184  t (CV=0.57) 
and  84,120  t (CV=0.93),  respectively,  for  an  area  close 
to  200,000  km2  from  42°N  to  48°N  off  Oregon  and 
Washington. 

Discussion 

Dynamics  of  biomass 

Off  the  PNW,  the  seasonal  relative  abundances  of  Pacific 
sardine  based  on  the  swept  area  method  are  nonsta- 
tionary (i.e,  not  static):  high  in  summer  and  low  in 
spring.  Fish  residing  in  the  PNW  in  spring  are  those 


over-wintering,  and  in  the  summer  the  majority  of  fish 
>190  mm  SL  are  likely  those  migrating  from  Califor- 
nia. The  spatial  distribution  of  the  Pacific  sardine  was 
similar  between  summer  and  spring:  high  in  the  inshore 
area  and  low  in  the  offshore  area,  except  during  March 
2005  when  small  numbers  of  sardine  were  caught  in 
the  northern  offshore  area  (Fig.  5).  This  distribution  is 
quite  different  from  that  off  California  where  the  spatial 
distribution  varied  among  years  (Lo  et  al.,  2005).  The 
PNW  biomass  estimates,  high  in  July  and  low  in  March, 
together  with  the  differential  length  distributions  are 
consistent  with  the  conceptual  migration  schedule  of 
Pacific  sardine  (a  migration  route  that  appears  to  be 
similar  to  that  of  Pacific  hake,  Merluccius  productus), 
namely  of  movement  to  the  PNW  from  California  before 


Table  5 

Trawl  information,  estimated  female  adult  parameters,  egg  production,  and  spawning  biomass  (estimated  by  the  daily  egg  pro- 
duction method  (DEPM))  for  Pacific  sardine  ( Sardinops  sagax)  from  July  and  March  surveys  conducted  from  2003  through  2005 
off  Washington  and  Oregon  (Pacific  Northwest)  and  from  April  surveys  conducted  from  2003  through  2005  off  California  and 
in  1994  off  California  and  Mexico.  Either  the  coefficient  of  variation  (CV)  or  number  of  positive  trawls  is  in  parentheses.  na=not 
available. 


Pacific  Northwest 

California 

2003 

July 

2004 

March 

2004 

July 

2005 

March 

1994 

April 

2003 

April 

2004 

April 

2005 

April 

No.  trawls  (positive) 

48(36) 

59(9) 

58(27) 

49(20) 

79(43) 

0 

25(17) 

19(14) 

Ave.  surface  temperature  (°C) 
at  sardine  locations 

15.4 

10.4 

15.6 

10.4 

14.36 

13.59 

14.18 

Fraction  of  females  by  weight 

R 

0.534 

0.568 

0.538 

0.618 

0.469 

Ave.  mature  female  weight 
(g)  with  ovary 
(g)  without  ovary 
Average  batch  fecundity0 

wf 

Wof 

F 

194.36 

189.25 

55,986 

105 

102.7 

193.16 

188.90 

55,883 

102.5 

100.2 

82.53 

79.33 

24,283 

166.99 

156.29 

55,711 

65.34 

63.11 

17,662 

Relative  batch  fecundity  (oocytes/g) 

288 

289 

294 

334 

270 

No.  mature  females  analyzed 

657 

3 

196 

37 

583 

290 

175 

No.  active  mature  females 

374 

1 

81 

11 

327 

290 

148 

Fraction  of  mature  females6 
spawning  per  day  (CV) 

S 

0.027 

(0.31) 

0 

0.010 

(0.74) 

0 

0.074 

(0.23) 

0.131 

(0.17) 

0.124 

(0.31) 

Fraction  of  active  females0 

sa 

0.050 

0 

0.025 

0 

0.131 

0.131 

0.155 

spawning  per  day 
Daily  specific  fecundity 

RSF 

W 

Po 

4.21 

na 

1.68 

na 

11.7 

27.04 

15.67 

Egg  production/0.05  m2/day 
(CV)  (Eq.  5) 

0.04d 

(0.51) 

0.037d 

(0.58) 

0.193 

(0.21) 

1.520 

(0.18) 

0.960 

(0.24) 

1.916 

(0.42) 

Survey  area  (km2) 

A 

206,037 

190,975 

380,175 

365,906 

320,620 

253,620 

Spawning  biomass  (t)  (CV) 

Bs 

39,184 

(0.57) 

na 

84,120 

(0.93) 

na 

127,102 

(0.32) 

485,121 

(0.36) 

281,639 

(0.30) 

621,657 

0.54 

Eggs/min  from  CUFES  sample  (CV) 

0.069 

(0.49) 

0.1 

(0.53) 

na 

1.57 

(0.27) 

0.78 

(0.11) 

0.62 

(0.15) 

a Mature  females:  1994  estimate  was  calculated  with  ^=-10858  + 439.53  W0/-  (Macewicz  et  al.,  1996),  in  2004  with  Fb=  356.46Wo/:(Lo  et  al.,  2005), 
in  2005  with  F^-6085  + 376.28  W,^,  and  for  Pacific  Northwest  in  2003  and  2004  with  Fh= 295.83  Wa ^ 
b Mature  females  included  females  that  were  active  and  those  that  were  postbreeding  (incapable  of  further  spawning  during  the  season). 
c Active  mature  females  were  capable  of  spawning  and  had  oocytes  with  yolk  or  postovulatory  follicles  less  than  60  hours  old. 
d Calculated  by  the  integral  method  and  corrected  for  bias  (P0  c). 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardinops  sagax  off  the  Pacific  coast. 


187 


summer  to  feed,  and  a return  to  the  south  before 
spring  to  spawn  (Clark  and  Janssen,  1945;  Dorn, 

1995;  Emmett  et  al.,  2005;  Smith,  2005). 

The  U.S.  stock  biomass  of  age  1+  Pacific  sardine 
increased  from  1981  to  a peak  of  one  million  tons 
in  2000  and,  according  to  the  stock  assessment, 
began  to  decline  in  2003  (Hill  et  al.,  2007).  The 
high  biomass  off  the  PNW  in  2003  was  most  likely 
due  to  the  accumulation  of  migrant  survivors  from 
1999  through  2002,  when  the  stock  assessment 
reported  that  biomasses  were  high.  The  PNW  sar- 
dine biomass,  estimated  from  surface  rope-trawl 
surveys  for  salmon  off  the  Columbia  River,  has 
been  decreasing  since  2003  (R.  Emmett,  personal 
commun.2).  This  decrease  is  likely  due  to  1)  the 
decline  of  migratory  fish  as  a result  of  the  de- 
creasing biomass  since  2003  off  California,  2)  a 
decline  in  successful  spawning  off  the  PNW,  or  3) 
the  continued  sardine  movement  northward  into 
Canadian  waters,  or  a combination  of  the  three 
events. 

The  July  2003  survey  indicated  that  the  major- 
ity of  fish  were  large  (>190  mm  SL),  whereas  the 
July  2004  survey  showed  the  opposite  because 
most  of  the  small  fish  were  from  the  strong  2003 
year  class.  The  presence  of  large  sardine  off  Or- 
egon in  July  2003  and  California  in  March-April 
2004  is  consistent  with  the  concept  of  the  migra- 
tion of  large  fish  from  the  PNW  to  California 
before  spawning.  However,  the  large  sardine  off 
Oregon  in  July  2004  did  not  show  up  off  either 
California  or  the  PNW  during  March-April  2005 
(Fig. 6).  This  finding  may  have  been  due  to  a 
lower  total  biomass  and  a smaller  proportion  of 
large  fish  off  the  PNW  in  July  2004  (Table  1,  Fig. 

6),  or  because  during  the  2005  California  survey, 
few  trawls  were  taken  north  of  34°N  latitude 
where  most  migrants  had  resided  according  to 
the  2004  DEPM  survey  off  California,  or  it  could 
have  been  due  to  a combination  of  both  factors 
(Fig.  5). 

Although  the  summer  PNW  biomass  estimates 
were  different  between  years,  the  spring  biomass 
estimates  were  stable.  March  surveys  clearly  revealed 
the  relative  magnitude  of  the  migratory  and  the  local 
PNW  stocks  during  the  survey  years.  The  change  in 
biomass  off  the  PNW  among  years  can  be  due  to  mul- 
tiple reasons:  a change  in  the  biomass  of  the  resident 
PNW  fish,  or  a change  in  the  biomass  off  California,  or 
a change  in  the  migration  pattern  due  to  food  availabil- 
ity and  oceanographic  conditions,  or  both  (MacFarlane 
et  al.,  2005).  To  better  understand  the  dynamics  of  the 
Pacific  sardine  off  the  west  coast  of  North  America, 
spring  and  summer  synoptic  surveys  from  Baja  Cali- 
fornia, Mexico,  to  British  Columbia,  Canada,  and  from 
tagging  studies  are  necessary. 


2 Emmett,  Robert.  2009.  Northwest  Fisheries  Science  Center, 
Newport,  OR. 


Figure  7 

Fraction  of  Pacific  sardine  ( Sardinops  sagax)  females  that 
were  sexually  mature  (y)  as  a function  of  standard  length  (L) 
fitted  to  logistic  curves  for  Oregon  and  Washington  in  July 
2004  and  March  2005,  and  for  California  in  April  of  1994, 
2004,  and  2005.  Symbols  represent  the  actual  fraction  mature 
within  10-mm  length  classes. 


Spawning  habitat  and  daily  egg  production 

The  spawning  habitats  off  the  PNW  in  the  summer  of 
2003  and  2004  were  similar  in  size  between  42-44. 5°N 
and  east  of  125. 4°W.  The  spawning  area  occupied  5-7% 
of  the  survey  area,  much  smaller  than  that  off  California 
(20-25%  of  the  survey  area  in  2003-04).  The  spawning 
habitat  in  the  mid-2000s  (2003  through  2005)  seemed  to 
contract  southward  and  shoreward  compared  to  the  mid- 
1990s  (1994  through  1998)  when  it  extended  to  46°N  and 
close  to  126°W  (Emmett  et  al.,  2005).  The  temperature 
range  in  the  offshore  spawning  habitat  in  the  1990s 
(14-16°C)  was  similar  to  that  in  the  2003-04  inshore 
area  (13— 18°C);  therefore,  the  change  of  oceanographic 
conditions  may  have  caused  the  apparent  contraction  of 
spawning  habitat  between  the  mid-1990s  and  mid-2000s 
off  the  PNW.  Because  no  adult  samples  were  taken  in 
the  mid-1990s,  we  were  unable  to  compare  the  adult 


188 


Fishery  Bulletin  108(2) 


120000- 

+ July  2003  + 

O July  2004 

Number  of  oocytes 

"vl 

o 

o 

o 

o 

i i 1 1 i i 

+0  o 

. 0+  «+  + 
+ +0 

++  ++ 

+ 

20000  - 

o 

I 1 1 1 1 1 1 1 1 1 1 1 

100  200  300 

Body  weight  (gm  without  ovary) 

Figure  8 

Batch  fecundity  ( Fb ),  the  estimated  number  of  hydrated  or  migra- 
tory-nucleus oocytes,  of  Pacific  sardine  ( Sardinops  sagax)  as  a 
function  of  ovary-free  fish  weight  (WJ.  Pacific  sardine  females  were 
collected  from  trawl  samples  off  Oregon  and  Washington  during  July 
2003  and  2004.  Equation  for  regression  line  is  F6  = 295.83W0^. 

spawning  characteristics  during  these  two  periods.  The 
spawning  habitats  of  sardine  off  the  PNW  in  2003  and 
2004  were  similar,  whereas  the  spawning  habitats  off 
California  were  quite  different:  concentrated  off  central 
California  in  2004  and  distributed  through  the  whole 
survey  area  in  2005.  Note,  no  eggs  were  caught  during 
the  July  California  Cooperative  Oceanic  Fisheries  Inves- 
tigations (CalCOFI)  surveys  off  California  in  either 
CalVET  or  bongo  net  tows. 

The  daily  egg  production  off  the  PNW  was  low  in 
both  2003  and  2004  (0.04  and  0.037eggs  produced/0.05 
m2/day,  respectively),  lower  than  that  in  1994  (0.50 
eggs  produced/0.05  m2/day)  (Bentley  et  al.,  1996),  and 
lower  than  those  off  California  (1.52  and  0.96  eggs 
produced  /0.05  m2/day)  in  2003,  2004,  and  other  years. 
This  low  egg  production  in  the  PNW  contributed  only 
1.8%  of  the  total  U.S.  west  coast  egg  production  in 
2003-04.  The  low  PNW  egg  production  estimates  could 
be  the  result  of  the  July  surveys  occurring  after  the 
spawning  peak,  possibly  in  June,  as  the  SST  was  high 
(Emmett  et  al.,  2005),  or  the  result  of  the  egg  mortal- 
ity of  Pacific  sardine  off  the  PNW  being  different  from 
that  off  California,  or  both.  Future  ichthyoplankton 
surveys  with  large  sample  sizes  are  needed  to  obtain 
direct  estimates  of  the  daily  egg  production  and  egg 
mortality  off  the  PNW. 

The  egg  production  estimates  from  July  2003  and 
2004  were  very  similar  even  though  the  relative 
abundances  were  quite  different.  With  similar  egg 
production  in  two  years,  one  might  expect  that  the 
biomass  of  recruits  would  be  similar.  However,  the 
2003  year  class  was  much  stronger  than  that  of  2004. 
This  difference  would  be  most  likely  due  to  the  more 


favorable  environmental  conditions  in  2003  than  in 
2004. 

One  interesting  question  to  ask  is  what  effect  a 
reduction  of  the  spawning  habitat  or  egg  production 
would  have  on  the  PNW  Pacific  sardine  population.  The 
sustainability  of  the  Pacific  sardine  population  off  the 
PNW  depends  greatly  on  the  Pacific  sardine  population 
off  California,  oceanographic  conditions,  and  food  avail- 
ability (MacFarlane  et  al.,  2005)  because  most  of  the 
spawners  (>190  mm  SL)  off  the  PNW  in  the  summer 
are  migrants  from  California.  As  long  as  the  Pacific 
sardine  population  off  California  is  large  enough  to  al- 
low adequate  migration  to  the  PNW  in  the  summer  to 
spawn,  the  population  off  the  PNW  will  be  sustained. 
Of  course,  if  environmental  conditions  are  unfavorable, 
the  proportion  of  spawners  may  be  reduced,  affecting 
both  the  recruits  to  the  local  population  and  the  size 
of  the  population  in  the  following  spring.  If  the  popu- 
lation off  California  decreases  to  the  level  of  collapse, 
the  population  off  the  PNW  may  have  been  diminished 
well  before  the  collapse  off  California.  This  status  of  the 
PNW  population  was  evident  from  the  history  of  land- 
ings in  the  last  Pacific  sardine  collapse  (Fig.  1).  Dur- 
ing the  waning  years  of  sardine  population,  the  PNW 
commercial  landings  ended  in  1949,  16  years  before  the 
California  catch  ended  in  1965.  The  sardine  population 
began  recovering  in  the  late  1970s-early  1980s  off  Cali- 
fornia. Incidental  landings  off  California  began  in  1981, 
11  years  before  an  incidental  catch  of  sardine  off  the 
PNW  in  1992  due  to  the  favorable  El  Nino  conditions, 
and  17  years  before  directed  landings  in  1998. 

Application  of  proper  management  strategies  to  pre- 
serve the  population  off  California,  and  thus  the  mi- 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardmops  sagax  off  the  Pacific  coast. 


189 


grants,  is  essential  because  most  of  the  migrants  are 
mature  fish  and  the  leaders  of  migration  imprints:  older 
fish  lead  younger  fish  to  migrate.  This  recent  entrain- 
ment hypothesis  (Petitgas  et  al.,  2006)  is  a step  for- 
ward from  the  theory  that  fish  population  life  cycles  are 
controlled  only  by  physical  conditions  (Sinclair,  1988). 
The  entrainment  hypothesis  implies  that  the  older  fish 
are  essential  to  ensure  the  sustainability  of  popula- 
tion and  fisheries  of  Pacific  sardine  off  the  PNW  and 
thus  along  the  entire  west  coast  of  the  North  American 
continent. 

Adult  reproductive  parameters  and  spawning  biomass 

Pacific  sardine  spawn  off  the  PNW,  contrary  to  beliefs 
in  the  1930s  and  1940s  that  they  spawn  only  off  Cali- 
fornia. Although  sardine  eggs,  larvae,  and  adults  have 
been  caught  in  surveys  off  the  PNW  since  1994  (Bentley 
et  al.,  1996;  Emmett  et  al.,  2005),  only  with  the  four 
surveys  during  2003-05  were  the  reproductive  param- 
eters for  female  Pacific  sardine  off  the  PNW  examined 
in  detail. 

The  spawning  season  of  Pacific  sardine  off  the  PNW 
apparently  occurs  primarily  in  the  early  summer,  al- 
though a few  fish  possibly  spawn  in  spring.  If  July 
is  the  spawning  peak  off  the  PNW,  then  spawning  is 
less  intense  than  during  the  peak  off  California  in 
April.  The  daily  spawning  fraction  of  mature  females 
(S=0.027  and  0.01)  was  much  lower  than  that  off  Cali- 
fornia (0.07-0.17).  Previous  work  has  indicated  that  ac- 
tive mature  females  of  Sardinops  spp.  worldwide  spawn 
once  every  eight  days  (Macewicz  et  al.,  1996).  Recent 
results  off  California  are  similar  (once  every  6-8  days), 
where  as  active  mature  Pacific  sardine  females  off  the 
PNW  spawned  much  less  frequently  (only  once  every 
20-40  days).  In  addition,  females  in  July  produced 
about  288  eggs  per  gram  of  female  weight  (relative 
batch  fecundity)  off  the  PNW — few  eggs  than  similar 
females  off  California  that  spawned  334  eggs  per  gram 
of  female  weight  in  April  2004  (Table  5).  According  to 
the  April  2004  DEPM  sardine  survey  off  California, 
the  large  mature  females,  in  particular  those  >200  mm 
SL,  were  spawning  very  vigorously  (S= 0.131)  and  these 
migratory  females  may  not  have  recovered  sufficient- 
ly to  spawn  at  higher  rates  off  the  PNW  during  July 
2004,  a phenomenon  similar  to  that  which  occurred 
with  Pacific  sardine  off  Chile,  which  were  less  active 
during  a second  annual  spawning  period  (Tascheri  and 
Claramunt,  1996).  The  presence  of  a high  percentage 
of  inactive  mature  females  off  the  PNW  in  July  (43% 
in  2003  and  59%  in  2004)  indicates  two  other  possible 
explanations  for  the  low  level  of  spawning:  July  is  not 
the  peak  spawning  time  for  sardine  off  the  PNW  be- 
cause they  may  be  similar  to  northern  anchovy  where 
ovaries  with  high  levels  of  atresia  (indicating  cessa- 
tion of  reproductive  activity)  are  common  at  the  end 
of  the  spawning  season  (Hunter  and  Macewicz,  1985); 
or,  Pacific  sardine  in  the  PNW  may  behave  like  chub 
mackerel  (Scomber  japonicus)  whose  individuals  spawn 
only  for  a short  period  and  inactive  mature  females  are 


common  throughout  the  spawning  season  (Dickerson  et 
al.,  1992).  If  so,  it  may  be  necessary  in  future  surveys 
to  analyze  reproductive  samples  collected  over  a longer 
time  to  better  define  the  peak  spawning  period,  and  to 
determine  whether  the  peak  spawning  fraction  is  simi- 
lar to  the  rate  off  California  (about  0.13  spawning  per 
day)  or  whether  it  remains  low  (<0.03). 

Few  mature  Pacific  sardine  females  were  caught  off 
the  PNW  during  March  and  it  seems  that  they  may 
have  followed  warmer  water  south.  The  majority  of  the 
40  mature  females  were  inactive  (postbreeding  or  rest- 
ing) and  none  had  spawned.  It  was  surprising  that  we 
caught  12  females  of  202-260  mm  SL  that  were  active 
(their  ovaries  contained  some  oocytes  with  yolk)  and 
were  potentially  capable  of  spawning  in  the  near  future 
(3-30  days).  We  examined  the  locations  where  females 
were  caught  and  their  associated  water  temperatures. 
The  average  SST  of  trawls  during  March  was  10.4°C. 
During  March  2004,  the  three  mature  females  (one 
active)  were  caught  farthest  south  (42°N)  in  11.1°C 
water.  One  inactive  mature  female  was  caught  near 
Astoria,  OR,  in  11.1°C  water  during  March  2005,  and 
the  other  36  (11  active)  mature  females  were  caught 
inshore,  south  of  44.5°N  in  11.5°C  (10.7-12.5°C)  wa- 
ter. Immature  female  Pacific  sardine  were  generally 
found  north  of  44.5°N  in  cooler  water;  on  average  10.2°C 
(9.6-10.7°C)  in  March  2004  and  10.3°C  (9.0-12. 5°C)  in 
March  2005.  Thus,  in  the  winter,  the  older  fish  were 
able  to  move  south  following  the  warmer  water,  while 
the  younger  fish,  due  to  a lack  of  stored  energy  for  long 
distance  swimming,  remained  in  the  cold  water.  Over- 
wintering immature  females  seem  to  tolerate  water  as 
cold  as  9.0°C.  The  PNW  generally  has  warmer  coastal 
SSTs  in  the  winter  (from  downwelling)  than  in  summer. 
However  temperatures  in  the  estuaries  can  be  very  cold 
and  die  offs  of  age-0  sardine  in  the  Columbia  River  and 
other  estuaries  have  been  observed  during  the  winter 
(E.  Dorval,  personal  commun.3). 

Female  Pacific  sardine  in  the  PNW  mature  at  lengths 
greater  than  those  off  California.  Fifty  percent  of  the 
females  caught  off  the  PNW  matured  at  around  195 
mm  and  > 90%  off  California  were  mature  at  the  size 
of  the  smallest  mature  PNW  female  (182  mm).  A major- 
ity of  sardine  > 200  mm  off  the  PNW  migrate  during 
fall-winter  (Clark  and  Janssen,  1945;  Nottestad  et  al., 
1999).  During  the  April  2004  DEPM  survey,  Pacific 
sardine  were  collected  off  central  California  between 
34.8°N  and  37.3°N  and  a majority  were  the  large,  mi- 
gratory size  (those  >200  mm),  whereas  in  2005,  the 
majority  of  positive  adult  samples  were  collected  in  the 
inshore  area  of  Southern  California  between  32°N  and 
36°N  and  most  sardines  were  <200  mm.  The  length  of 
females  at  50%  maturity  off  the  PNW  was  similar  to 
the  length  estimate  (193  mm)  in  April  2004  off  Cali- 
fornia which  indicated  that  the  large  Pacific  sardines 
off  central  California  likely  were  winter  migratory  fish. 
This  conclusion  is  consistent  with  the  historical  tagging 


3 Dorval,  Emanis.  2008.  Librairie  La  Lumiere,  Rue  Baussan, 
# 34,  Turgeau,  Port-au-Prince,  Haiti,  W.I. 


190 


Fishery  Bulletin  108(2) 


results,  which  indicated  that  the  majority  of  the  tags 
released  off  the  PNW  were  recovered  off  central  Cali- 
fornia (Clark  and  Janssen,  1945). 

The  point  estimates  of  spawning  biomass  of  Pacific 
sardine  off  the  PNW  differed,  but  were  not  statistically 
different  because  of  a large  coefficient  of  variation: 
39,184  t and  84,120  t for  July  2003  and  2004,  respec- 
tively. They  were  close  to  50,000  t in  1994  (Bentley  et 
al.,  1996).  Theoretically,  the  spawning  biomass  should 
constitute  a good  proportion  of  the  total  biomass,  which 
was  not  so  for  July  2003.  This  could  be  due  to  an  under- 
estimate of  P0,  to  an  overestimate  of  the  spawning  frac- 
tion, or  both.  The  overestimate  of  the  spawning  fraction 
could  be  due  to  the  movement  of  the  postspawners  out  of 
the  spawning  area.  A DEPM  study  is  needed  to  evaluate 
such  effects  and  model  the  effects  of  fish  movement  on 
estimates  of  spawning  rate.  The  effect  of  the  timing  of 
the  survey  in  relation  to  spawning  and  movement  cycles 
needs  to  be  studied  with  new  data  and  modeling. 

The  difference  between  the  spawning  biomass  es- 
timates in  2003  and  2004  was  primarily  due  to  the 
difference  in  the  estimated  spawning  fractions  (0.027 
in  contrast  to  0.01),  because  the  estimates  of  daily  egg 
production  (P0)  were  similar.  The  large  coefficients  of 
variation  of  spawning  biomass  estimates  were  mainly  a 
result  of  the  uncertainty  in  estimates  of  P0  and  the  dai- 
ly spawning  fraction  (S)  in  July  2004.  For  low  values  of 
P0  and  S,  the  number  of  samples  has  to  be  substantially 
increased  to  obtain  a more  precise  estimate  (Picquelle 
and  Stauffer,  1985).  Estimated  spawning  biomass  for  off 
the  PNW  in  July  was  much  smaller  than  estimates  for 
off  California  during  April  in  recent  years.  The  smaller 
fish  length  at  50%  maturity  off  California  means  that 
the  more  numerous  smaller  resident  Pacific  sardine  are 
able  to  participate  in  local  spawning  at  the  same  time 
as  the  larger  migratory  sardine. 

Future  work 

The  Pacific  sardine  spawning  habitat  and  season  in  the 
PNW  are  loosely  defined  in  this  study  and  the  magnitude 
and  scope  of  the  coastal  migration  are  not  fully  explored. 
To  better  characterize  these,  we  need  to  conduct  syn- 
optic trawl-ichthyoplankon-acoustic  surveys  from  Baja 
California,  Mexico,  to  British  Columbia,  Canada,  during 
spring  and  early  summer  at  three  to  five  year  intervals. 
To  better  characterize  the  spawning  habitats  in  this 
area,  we  need  to  obtain  physical  and  biological  oceano- 
graphic data  (Lynn,  2003;  Emmett  et  al.,  2005;  Reiss  et 
al.,  2008)  and  demographic  data  of  Pacific  sardine  over 
a broader  geographic  range  because  the  Pacific  sardine 
is  a migratory  species. 

For  trawl  swept-area-based  biomass  estimates,  the 
efficiency  of  the  trawl  needs  to  be  calibrated.  Biomass 
estimates  from  acoustic  surveys  would  be  another  fish- 
ery-independent source  of  relative  abundance.  Because 
the  coefficients  of  variation  of  all  estimates  are  large, 
the  number  of  trawls  needs  to  be  increased  or  other 
statistical  estimation  procedures  should  be  explored,  or 


both,  to  improve  the  precision  of  estimates.  To  obtain 
a representative  length  distribution  of  the  population, 
fishery-independent  surveys  covering  the  whole  west 
coast  area  are  essential,  and  length  data  from  com- 
mercial vessels  should  be  used  with  caution  for  both  the 
PNW  and  California.  For  spawning  biomass,  we  need  to 
understand  the  maturation  schedules  of  females  and  the 
spawning  season  off  Oregon  and  Washington.  Numerous 
plankton  net  tows  are  needed  to  obtain  direct  estimates 
of  the  daily  egg  production  and  egg  mortality  rates  in 
early  summer.  Currently,  only  the  spawning  biomass 
of  Pacific  sardine  off  California  is  estimated  from  the 
annual  April  DEPM  survey.  Because  mature  females 
were  caught  during  two  March  surveys  off  the  PNW, 
efforts  should  be  made  to  obtain  trawl  data  off  the 
PNW  in  April.  Data  for  mature  females  collected  off  the 
PNW  could  then  be  combined  with  the  April  data  set 
off  California  to  estimate  reproductive  parameters  and 
the  spawning  biomass  of  Pacific  sardine  off  the  whole 
west  coast  of  the  United  States.  To  better  understand 
the  relationship  between  the  sardine  populations  off 
California  and  the  PNW,  we  need  to  examine  migration 
characteristics  (i.e,  migration  range,  pattern  and  sched- 
ule) and  the  effect  of  fishing  pressure  on  the  migratory 
fish  because  most  of  these  fish  are  mature  and  leaders 
of  migration  imprints.  We  need  a long  time  series  of 
abundance  for  all  regions  together,  along  with  ocean- 
ographic and  biological  data,  to  enhance  our  under- 
standing of  the  dynamics  of  the  entire  Pacific  sardine 
population  to  provide  information  for  the  development 
of  future  strategies  to  sustain  the  population. 

Acknowledgments 

We  thank  two  anonymous  reviewers  for  their  construc- 
tive comments.  We  thank  the  captain  and  crew  members 
of  the  FV  Frosti  and  the  support  for  the  charter  provided 
by  NMFS  Cooperative  Research  Program.  These  surveys 
would  not  have  been  possible  without  the  cooperation 
of  the  Northwest  Fisheries  Science  Center,  NOAA,  the 
Washington  Department  of  Fish  and  Wildlife  (WDFW), 
Oregon  Department  of  Fish  and  Wildlife  (ODFW), 
and  the  Pacific  Fishery  Management  Council  (PFMC). 
We  thank  all  those  who  participated  in  the  surveys: 

D.  Waldeck  (PFMC),  Todd  Miller  (Oregon  State  Univer- 
sity), J.  McCrae  (ODFW),  A.  Thurman  (WDFW),  and  E. 
Acuna  and  N.  Bowlin  of  Southwest  Fisheries  Science 
Center.  We  thank  J.  Hunter,  W.  Watson,  S.  Picquelle,  E. 
Dorval,  K.  Hill,  A.  Takasuka,  S.  McClatchie,  A.  MacCall, 

E.  Weber,  and  R.  Emmett  for  reviewing  the  manuscript 
and  R.  Sanford  for  organizing  the  manuscript. 


Literature  cited 

Ahlstrom,  E.  H. 

1948.  A record  of  pilchard  eggs  and  larvae  collected 
during  surveys  made  in  1939  to  1941.  U.S.  Fish  Wildl. 
Serv.,  Spec.  Sci.  Rep.  no.  54,  76  p. 


Lo  et  al.  Biomass  and  reproductive  status  of  Sardmops  sagax  off  the  Pacific  coast. 


191 


Bentley,  J.  R.,  R.  L.  Emmett,  N.  C.  H.  Lo,  and  H.  G.  Moser. 

1996.  Egg  production  of  the  Pacific  sardine  ( Sardinops 
sagax ) off  Oregon  in  1994.  Calif.  Coop.  Oceanic  Fish. 
Invest.  Rep.  37:193—200. 

Checkley,  D.  M.,  Jr.,  P.  B.  Ortner,  L.  R.  Settle,  and  S.  R.  Cummings. 

1997.  A continuous,  underway  fish  egg  sampler.  Fish. 
Oceanogr.  6(21:58-73. 

Clark,  F.  N.,  and  J.  F.  Janssen  Jr. 

1945.  Movements  and  abundance  of  the  sardine  as  mea- 
sured by  tag  returns.  Fish  Bull.  Calif.  Dep.  Fish  Game 
61:7-42. 

Deriso,  R.  B.,  J.  T.  Barnes,  L.  D.  Jacobson,  and  P.  R.  Arenas. 
1996.  Catch-at-age  analysis  for  Pacific  sardine  ( Sardinops 
sagax),  1983-1995.  Calif.  Coop.  Oceanic  Fish.  Invest. 
Rep.  37:175-187. 

Dickerson,  T.  L.,  B.  J.  Macewicz,  and  J.  R.  Hunter. 

1992.  Spawning  frequency  and  batch  fecundity  of  chub 
mackerel,  Scomber  japonicus,  during  1985.  Calif.  Coop. 
Oceanic  Fish.  Invest.  Rep.  33:130-140. 

Dorn,  M.  W. 

1995.  The  effects  of  age  composition  and  oceanographic 
conditions  on  the  annual  migration  of  Pacific  whiting, 
Merluccius  produetus.  Calif.  Coop.  Oceanic  Fish.  Invest. 
Rep.  36:97-105. 

Emmett,  R.  L.,  R.  D.  Brodeur,  T.  W.  Miller,  S.  S.  Pool,  P.  J. 
Bentley,  G.  K.  Krutzikowsky,  and  J.  McCrae. 

2005.  Pacific  sardine  ( Sardinops  sagax)  abundance,  dis- 
tribution, and  ecological  relationships  in  the  Pacific 
northwest.  Calif.  Coop.  Oceanic  Fish.  Invest.  Rep. 
46:122-143. 

Felin,  F.  E. 

1954.  Population  heterogenity  in  Pacific  pilchard.  Fish. 
Bull.  54:201-225. 

Hart,  J.  L. 

1943.  The  pilchard,  Sardinops  caerulea  (Girard)  on 
Canadian  fishing  grounds  with  special  reference  to 
an  unusual  abundance  of  young  fish.  Trans.  R.  Soc. 
Can.,  ser.  3,  37(51:55-73. 

Hill,  K.  T.,  E.  Dorval,  N.  C.  H.  Lo,  B.  J.  Macewicz,  C.  Show,  and 
R.  Felix-Uraga. 

2007.  Assessment  of  Pacific  sardine  resource  in  2007 
for  U.S.  management  in  2008.  NOAA  Tech.  Memo. 
NMFS-SWFSC-413,  157  p. 

Holt,  D.,  and  T.M.  F.  Smith. 

1979.  Post  stratification.  J.  R.  Stat.  Soc.  A.  142:33- 
46. 

Hunter,  J.  R.,  and  N.  C.  H.  Lo. 

1993.  Icthyoplankton  methods  for  estimating  fish  bio- 
mass introduction  and  terminology.  Bull.  Mar.  Sci. 
53(21:723-727. 

Hunter,  J.  R.,  and  B.  J.  Macewicz. 

1985.  Rates  of  atresia  in  the  ovary  of  captive  and  wild 
northern  anchovy,  Engraulis  mordax.  Fish.  Bull. 
83:115-136. 

Janssen,  J.  F.,  Jr. 

1938.  Second  report  of  sardine  tagging  in  California.  Fish 
Bull.  Calif.  Dep.  Fish  Game  24:376-389. 

Lasker,  R.,  ed. 

1985.  An  egg  production  method  for  estimating  spawn- 
ing biomass  of  pelagic  fish:  application  to  the  northern 
anchovy,  Engraulis  mordax.  NOAA  Tech.  Rep.  NMFS 
36,  99  p. 

Lo,  N.  C.  H.,  Y.  A.  Green  Ruiz,  M.  J.  Cervantes,  H.  G.  Moser,  and 
R.  J.  Lynn. 

1996.  Egg  production  and  spawning  biomass  of  Pacific 
sardine  ( Sardinops  sagax),  in  1994,  determined  by  the 


daily  egg  production  method.  Calif.  Coop.  Oceanic  Fish. 
Invest.  Rep.  37:160-174. 

Lo,  N.  C.  H.,  B.  J.  Macewicz,  and  D.  A.  Griffith. 

2005.  Spawning  biomass  of  Pacific  sardine  ( Sardinops 
sagax),  from  1994-2004  off  California.  Calif.  Coop. 
Oceanic  Fish.  Invest.  Rep.  46:93-112. 

Lynn,  R.  J. 

2003.  Variability  in  the  spawning  habitat  of  Pacific 
sardine  (Sardinops  sagax)  off  southern  and  central 
California.  Fish.  Oceanogr.  12(31:1-13. 

MacCall,  A.  D. 

1976.  Density  dependence  of  catchability  coefficient  in 
the  California  Pacific  sardine,  Sardinops  sagax  caerulea, 
purse  seine  fishery.  Calif.  Coop.  Oceanic  Fish.  Invest. 
Rep.  18:136-148. 

Macewicz,  B.  J.,  J.  J.  Castro-Gonzalez,  C.  E.  Cotero  Altamirano, 
and  J.  R.  Hunter. 

1996.  Adult  reproductive  parameters  of  Pacific  sardine 
(Sardinops  sagax)  during  1994.  Calif.  Coop.  Oceanic 
Fish.  Invest.  Rep.  37:140-151. 

Marr,  J.  C. 

1950.  Apparent  abundance  of  the  pilchard  ( Sardinops 
caerulea)  off  Oregon  and  Washington,  1935-43,  as  mea- 
sured by  the  catch  per  boat.  Fish.  Bull.  51:385-394. 

1960.  The  causes  of  major  variations  in  the  catch  of  the 
Pacific  sardine,  Sardinops  caerulea  (Girard).  In  Pro- 
ceedings of  the  world  scientific  meeting  on  the  biology 
of  sardine  and  related  species,  vol.  3 (Rosa,  H.,  and  G. 
I.  Murphy,  eds.),  p.  667-791.  FAO,  Rome. 

McFarlane,  G.  A.,  J.  Schweigert,  L.  MacDougall,  and  C.  Hrabok. 

2005.  Distribution  and  biology  of  Pacific  sardine 
(Sardinops  sagax 1 off  British  Columbia,  Canada.  Calif. 
Coop.  Oceanic  Fish.  Invest.  Rep.  46:144-160. 

Mosher,  K.  H.  and  H.  H.  Eckles. 

1954.  Age  determination  of  Pacific  sardine  from 
otoliths.  U.S.  Dept.  Int.,  Fish  Wild.  Serv.  Res.  Rep. 
37,  40  p. 

Murphy,  G.  L. 

1966.  Population  biology  of  the  Pacific  sardine  (Sardinops 
caerulea).  Proc.  Calif.  Acad.  Sci.  series  4,  34(11:1-84. 

Nottestad,  L.,  J.  Giske,  J.  C.  Hoist,  and  G.  Huse. 

1999.  A length-based  hypothesis  for  feeding  migrations 
in  pelagic  fish.  Can.  J.  Fish.  Aquat.  Sci.  56  (suppl. 
11:26-34. 

Parker,  K. 

1985.  Biomass  model  for  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.  5-6.  NOAA  Tech. 
Rep.  NMFS  36. 

Petitgas,  P.,  D.  Reid  , B.  Planque,  E.  Nogueira,  B.  O’Hea,  and 
U.  Cotano. 

2006.  The  entrainment  hypothesis:  an  explanation  for 
the  persistence  and  innovation  in  spawning  migrations 
and  life  cycle  spatial  patterns.  ICESCM:2006/B:07,  9 
p.  [Available  at  http://www.ices.dk/products/cmdoc- 
sindex.asp] 

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.  Lasker,  ed.1,  p.  7-16.  NOAA  Tech.  Rep.  NMFS  36. 

Preti,  A.,  S.  E.  Smith,  and  D.  A.  Ramon. 

2001.  Feeding  habitats  of  the  common  thresher  shark 
( Alopias  vulpinus)  sampled  from  the  California-based 


192 


Fishery  Bulletin  108(2) 


drift  gill  net  fishery,  1998-1999.  Calif.  Coop.  Oceanic 
Fish.  Invest.  Rep.  42:145-152. 

2004.  Diet  differences  in  the  thresher  shark  (Alopias 
vulpinus)  during  transition  from  a warm-water  regime 
to  a cool-water  regime  off  California-Oregon,  1998- 
2000.  Calif.  Coop.  Oceanic  Fish.  Invest.  Rep.  45:118-125. 

Reiss,  C.  S.,  D.  M.  Checkley  Jr.,  and  S.  J.  Bograd. 

2008.  Remotely  sensed  spawning  habitat  of  Pacific  sar- 
dine ( Sardinops  sagax)  and  Northern  anchovy  ( Engru - 
aulis  mordax)  within  the  California  Current.  Fish. 
Oceanogr.  17:126-136 

Sinclair,  M. 

1988.  Marine  populations:  an  essay  on  population  regu- 
lation and  speciation,  252  p.  Univ.  Washington  Press, 
Seattle,  WA. 

Smith,  P.  E. 

2005.  A history  of  proposals  for  subpopulation  struc- 
ture in  Pacific  sardine  ( Sardinops  sagax ) population 
off  western  North  America.  Calif.  Coop.  Oceanic  Fish. 
Invest.  Rep.  46:75-82. 


Smith,  P.  E.,  W.  Flerx,  and  R.  P.  Hewitt. 

1985.  The  CalCOFI  vertical  egg  tow  (CalVET)  net.  In  An 
egg  production  method  for  estimating  spawning  biomass 
of  pelagic  fish:  application  to  the  northern  anchovy, 
Engraulis  mordax,  (R.  Lasker,  ed.),  p.  27-32.  NOAA 
Tech.  Rep.  NMFS  36. 

Stratoudakis,  Y.,  M.  Bernal,  K.  Ganias,  and  A.  Uriarte. 

2006.  The  daily  egg  production  method:  recent  advances, 
current  applications  and  future  challenges.  Fish  Fish. 
7:35-57. 

Tascheri,  R.  and  G.  Claramunt. 

1996.  Aproximacion  a los  cambios  intra-annuales  en  el 
contenido  de  energia  del  ovario  de  sardina  ( Sardinops 
sagax  Jenyns,  1842)  en  el  norte  de  Chile.  Invest.  Mar. 
24:51-66.  [In  Spanish.] 

Wolf,  P. 

1992.  Recovery  of  the  Pacific  sardine  and  the  Califor- 
nia sardine  fishery.  Calif.  Coop.  Oceanic  Fish.  Invest. 
Rep.,  vol.  33:76-86. 


193 


Seasonal  variability 

in  ichthyoplankton  abundance 

and  assemblage  composition 

in  the  northern  Gulf  of  Mexico  off  Alabama 


Email  address  for  contact  author:  fhernandez@disl.org 

1 Dauphin  Island  Sea  Laboratory 
101  Bienville  Boulevard 
Dauphin  Island,  Alabama  36528 

2 Department  of  Marine  Sciences 
University  of  South  Alabama 

307  University  Boulevard,  LSCB  Rm  25 
Mobile,  Alabama  36688 


Abstract — Multiyear  ichthyoplankton 
surveys  used  to  monitor  larval  fish 
seasonality,  abundance,  and  assem- 
blage structure  can  provide  early  indi- 
cators of  regional  ecosystem  changes. 
Numerous  ichthyoplankton  surveys 
have  been  conducted  in  the  north- 
ern Gulf  of  Mexico,  but  few  have  had 
high  levels  of  temporal  resolution  and 
sample  replication.  In  this  study,  ich- 
thyoplankton samples  were  collected 
monthly  (October  2004-October  2006) 
at  a single  station  off  the  coast  of 
Alabama  as  part  of  a long-term  bio- 
logical survey.  Four  seasonal  periods 
were  identified  from  observed  and 
historic  water  temperatures,  includ- 
ing a relatively  long  ( June-October) 
“summer”  period  (water  tempera- 
ture >26°C).  Fish  egg  abundance, 
total  larval  abundance,  and  larval 
taxonomic  diversity  were  significantly 
related  to  water  temperature  (but  not 
salinity),  with  peaks  in  the  spring, 
spring-summer,  and  summer  periods, 
respectively.  Larvae  collected  during 
the  survey  represented  58  different 
families,  of  which  engraulids,  sciae- 
nids,  carangids,  and  clupeids  were 
the  most  prominent.  The  most  abun- 
dant taxa  collected  were  unidenti- 
fied engraulids  (50%),  sand  seatrout 
(Cynoscion  arenarius,  7.5%),  Atlantic 
bumper  ( Chloroscombrus  chrysurus, 
5.4%),  Atlantic  croaker  ( Micropogo - 
nias  undulatus,  4.4%),  Gulf  menha- 
den ( Brevoortia  patronus,  3.8%),  and 
unidentified  gobiids  (3.6%).  Larval 
concentrations  for  dominant  taxa  were 
highly  variable  between  years,  but 
the  timing  of  seasonal  occurrence  for 
these  taxa  was  relatively  consistent. 
Documented  increases  in  sea  surface 
temperature  on  the  Alabama  shelf 
may  have  various  implications  for 
larval  fish  dynamics,  as  indicated  by 
the  presence  of  tropical  larval  forms 
(e.g.,  fistularids,  labrids,  scarids,  and 
acanthurids)  in  our  ichthyoplankton 
collections  and  in  recent  juvenile  sur- 
veys of  Alabama  and  northern  Gulf 
of  Mexico  seagrass  habitats. 


Manuscript  submitted  8 October  2009. 
Manuscript  accepted  8 January  2010. 
Fish.  Bull.  108:193-207  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National 
Marine  Fisheries  Service,  NOAA. 


Frank  J.  Hernandez  Jr.  (contact  author)’ 
Sean  P.  Powers'  2 
William  M.  Graham1 


Ichthyoplankton  surveys  provide  fish- 
eries-independent  information  that  is 
inherently  “ecosystem-based”;  entire 
larval  fish  assemblages  are  collected 
(i.e.,  early  stages  of  both  exploited  and 
unexploited  finfish  species)  along  with 
zooplankton  predators  and  prey,  and 
often  with  a suite  of  environmental 
observations  (e.g.,  salinity,  tempera- 
ture). At  the  ecosystem  level,  infor- 
mation on  larval  assemblages  can  be 
used  to  detect  changes  in  marine  fish 
community  composition  and  abun- 
dances over  time  (Sherman  et  ah, 
1984).  Previous  studies  have  indicated 
that  larval  assemblages  are  the  result 
of  convergent  spawning  strategies  by 
multiple  species  taking  advantage  of 
favorable  environmental  conditions  for 
larval  fish  survival  (Doyle  et  ah,  1993; 
Sherman  et  al.,  1984).  The  composi- 
tion of  larval  fish  assemblages  varies 
spatially  and  temporally  because  of 
the  behaviors  of  the  larvae  (Gray  and 
Miskiewicz,  2000;  Hare  and  Govoni, 
2005)  and  the  spawning  adults  (Sher- 
man et  al.,  1984;  Hernandez-Miranda 
et  al.,  2003),  as  well  as  oceanographic 
transport  and  mixing  processes  (Auth, 
2008;  Muhling  et  al.,  2008).  Variabil- 
ity in  any  of  these  factors,  therefore, 
may  result  in  a different  structure 
of  larval  fish  assemblages.  Because 
larval  fish  survival  is  closely  tied  with 
primary  and  secondary  productivity 


in  coastal  oceans,  changes  in  larval 
fish  assemblage  structure  (over  larger 
time  scales)  can  be  an  early  indica- 
tor of  climate-related  environmental 
shifts  (Auth,  2008;  Brodeur  et  al., 
2008). 

Despite  the  importance  of  the  re- 
gion to  fisheries,  seasonal  variabil- 
ity in  larval  fish  assemblages  in  the 
northern  Gulf  of  Mexico  has  been 
examined  in  relatively  few  studies. 
Much  of  the  previous  ichthyoplankton 
research  has  focused  on  estuarine  as- 
semblages (Raynie  and  Shaw,  1994; 
Tolan  et  al.,  1997)  or  on  relatively 
short-term  interactions  between  as- 
semblages and  specific  oceanograph- 
ic features,  such  as  the  Mississippi 
River  plume  (Sogard  et  al.,  1987;  Go- 
voni et  al.,  1989)  or  the  Loop  Current 
(Richards  et  al.,  1993).  Other  studies 
have  used  ichthyoplankton  survey 
data  from  the  National  Marine  Fish- 
eries Service’s  (NMFS’s)  gulf-wide 
Southeast  Monitoring  and  Assess- 
ment Program  (SEAMAP),  but  these 
studies  are  typically  focused  on  a sin- 
gle species  (Scott  et  al.,  1993;  Lycz- 
kowski-Shultz  and  Ingram,  2003;  Ly- 
czkowski-Shultz  and  Hanisko,  2007). 
Ditty  et  al.  (1988)  summarized  the 
available  ichthyoplankton  literature 
at  the  time  to  provide  information  on 
larval  fish  seasonality  for  the  entire 
northern  Gulf  of  Mexico,  and  more 


194 


Fishery  Bulletin  108(2) 


Figure  I 

Location  of  the  sampling  station  used  during  the  October  2004-October  2006  ichthyo- 
plankton  monitoring  survey  (star  symbol)  and  the  NOAA  National  Data  Buoy  Center 
oceanographic  data  buoy  (NDBC  42007)  used  to  determine  the  10-year  (1993-2003) 
mean  monthly  water  temperature  estimates  for  the  region  (diamond  symbol). 


recently,  Lyczkowski-Shultz  et  al.1  reported  on  larval 
fish  seasonality  and  distribution  for  the  northeastern 
Gulf  of  Mexico. 

Although  these  latter  studies  provided  information 
on  multiple  species,  no  analyses  of  larval  fish  assem- 
blages and  environmental  variability  were  presented. 
Here  we  report  on  the  seasonality  and  concentrations 
of  larval  fishes  in  relation  to  water  temperature  based 
on  data  collected  during  an  intensive  two  year  (October 
2004-October  2006)  ichthyoplankton  survey  conducted 
off  the  coast  of  Alabama.  The  objectives  of  this  study 
were  1)  to  examine  the  seasonal  variability  in  ichthyo- 
plankton diversity  and  taxon-specific  abundances  off 
the  coast  of  Alabama;  and  2)  to  examine  variability 
in  the  relationship  between  larval  fish  assemblages 
and  seasonal  changes  in  water  temperature.  These 
objectives  would  contribute  to  our  overall  goal  of  un- 
derstanding the  oceanographic  factors  that  maintain 
larval  fish  assemblages. 


1 Lyczkowski-Shultz,  J.,  D.  S.  Hanisko,  K.  J.  Sulak,  and  G. 
D.  Dennis  III.  2004.  Characterization  of  ichthyoplankton 

within  the  U.S.  Geological  Survey’s  northeastern  Gulf  of 
Mexico  study  area — based  on  analysis  of  Southeast  Area 
Monitoring  and  Assessment  Program  (SEAMAP)  sampling 
surveys,  1982-1999,  136  p.  NEGOM  Ichthyoplankton  Synop- 
sis Final  Report,  U.S.  Dep.  Interior,  U.S.  Geological  Survey, 
USGS  SIR-2004-5059. 


Materials  and  methods 

Data  collection 

The  sampling  station  was  located  on  the  inner  continen- 
tal shelf  of  the  northern  Gulf  of  Mexico,  approximately 
18  km  south  of  Dauphin  Island,  Alabama,  at  a water 
depth  of  approximately  20  m (Fig.  1).  Ichthyoplank- 
ton sampling  was  conducted  during  monthly  day-time 
surveys  (n=26)  and  quarterly  diel  surveys  (n- 8)  from 
October  2004  to  October  2006  (Table  1).  All  samples 
were  collected  with  a Bedford  Institute  of  Oceanography 
Net  Environmental  Sampling  System  (BIONESS)  (Open 
Seas  Instrumentation,  Inc.,  Musquodoboit  Harbour, 
Nova  Scotia,  Canada),  with  a 0.25-m2  mouth  opening 
fitted  with  seven  (during  quarterly  surveys)  or  eight 
(during  monthly  surveys)  plankton  nets.  During  monthly 
surveys,  six  depth-discrete  samples  (18-15  m,  15-12 
m,  12-9  m,  9-6  m,  6-3  m,  and  3-1  m)  and  one  oblique 
sample  (18-1  m)  were  collected  during  eight  replicate 
tows  at  the  study  site  with  202-pm  mesh  nets.  An  addi- 
tional oblique  sample  was  collected  during  each  tow  with 
a 333-pm  mesh  net  for  a nominal  total  of  64  samples  per 
monthly  cruise.  All  eight  replicate  tows  were  collected 
during  daylight  hours,  generally  during  a single  day. 
During  the  quarterly  surveys,  a set  of  six  depth-discrete 
samples  (same  depth  bins  as  monthly  survey)  and  one 


Hernandez  et  al.:  Variability  in  ichthyoplankton  abundance  and  composition  in  the  northern  Gulf  of  Mexico 


195 


Table  1 

Station  data  for  ichthyoplankton  samples  collected  during  a larval  fish  monitoring  survey  at  a site  located  approximately  18  km 
south  of  Dauphin  Island,  Alabama  (October  2004-October  2006).  Seasonal  classification  is  based  on  historic  (10-year  average) 
and  observed  monthly  mean  temperatures  for  the  region  (see  Fig.  2). 

Year 

Cruise  date 

Survey  type 

Seasonal  classification 

Number  of  samples 

2004 

22  Oct 

monthly 

Summer 

54 

2004 

16-17  Nov 

diel 

Fall 

41 

2004 

29  Nov 

monthly 

Fall 

47 

2004 

08  Dec 

monthly 

Fall 

47 

2005 

06-07  Jan 

monthly 

Winter 

48 

2005 

18-21  Jan 

diel 

Winter 

76 

2005 

16  Feb 

monthly 

Winter 

50 

2005 

29  Mar 

monthly 

Spring 

23 

2005 

05  Apr 

monthly 

Spring 

18 

2005 

19  Apr 

monthly 

Spring 

47 

2005 

09-13  May 

diel 

Spring 

72 

2005 

17  May 

monthly 

Spring 

48 

2005 

09  Jun 

monthly 

Summer 

47 

2005 

13  Jul 

monthly 

Summer 

48 

2005 

09  Aug 

monthly 

Summer 

46 

2005 

14  Sep 

monthly 

Summer 

48 

2005 

27-29  Sep 

diel 

Summer 

72 

2005 

11  Oct 

monthly 

Summer 

31 

2005 

09  Nov 

monthly 

Fall 

32 

2005 

29  Nov-02  Dec 

diel 

Winter 

71 

2005 

16  Dec 

monthly 

Winter 

40 

2006 

12  Jan 

monthly 

Winter 

44 

2006 

07-10  Feb 

diel 

Winter 

60 

2006 

17  Feb 

monthly 

Winter 

43 

2006 

16  Mar 

monthly 

Spring 

39 

2006 

12-13  Apr 

monthly 

Spring 

38 

2006 

01-04  May 

diel 

Spring 

70 

2006 

17  May 

monthly 

Spring 

43 

2006 

15  Jun 

monthly 

Summer 

42 

2006 

05  Jul 

monthly 

Summer 

46 

2006 

10  Aug 

monthly 

Summer 

46 

2006 

08  Sep 

monthly 

Summer 

46 

2006 

19-22  Sep 

diel 

Summer 

66 

2006 

12  Oct 

monthly 

Summer 

47 

oblique  sample  were  collected  with  202-pm  mesh  nets 
at  dawn,  noon,  dusk,  and  midnight  (local  time)  over 
the  course  of  three  diel  periods  for  a nominal  total  of 
84  samples  per  quarterly  cruise.  Contents  of  nets  were 
rinsed  with  seawater,  sieved,  and  preserved  in  4%  forma- 
lin for  48  hours  before  being  transferred  to  70%  ethanol. 
A conductivity-temperature-depth  probe  (CTD)  (SBE19, 
Sea-Bird  Electronics,  Inc.,  Bellevue,  WA)  was  integrated 
into  the  BIONESS  system  and  provided  temperature, 
salinity,  and  depth  profiles  for  each  plankton  tow.  A flow- 
meter (General  Oceanics,  Miami,  FL)  mounted  within 
the  BIONESS  frame  estimated  the  volume  of  water 
filtered  for  each  sample.  Filtered  volume  estimates  for 
each  sample  were  compared  with  measurements  from  a 
second,  externally  mounted  flowmeter  to  estimate  filtra- 


tion efficiency.  In  all,  1634  ichthyoplankton  samples  were 
processed  and  used  in  subsequent  analyses.  Although  all 
fish  larvae  were  collected  from  a single  station,  Alabama 
has  a relatively  short  coastline  (<85  km),  thus  the  larval 
fishes  collected  likely  represent  the  ichthyofauna  of  the 
entire  Alabama  inner  shelf  region. 

Preparation  of  environmental  data 

CTD  data  were  processed  using  the  manufacturer’s 
software  (SEASOFT,  Seabird  Electronics,  Inc.,  Bellevue, 
WA)  and  averaged  into  0.5-m  bins.  Seasonal  patterns 
in  water  temperature  were  examined  using  depth-inte- 
grated monthly  mean  temperatures  recorded  during  each 
sampling  month.  For  historic  comparisons,  the  10-year 


196 


Fishery  Bulletin  108(2) 


average  for  water  temperature  was  calculated  for  each 
month  with  data  from  a coastal  observing  buoy  (NOAA 
National  Data  Buoy  Center  Station  42007)  located 
approximately  54  km  west  of  our  sampling  station  at  a 
water  depth  of  approximately  15  m (Fig.  1).  Although  the 
temperature  values  from  the  buoy  were  measured  near 
the  surface  (0.6-m  depth),  these  observations  serve  as 
good  indicators  of  seasonal  shifts  in  water-column  ther- 
mal structure,  as  indicated  by  our  own  CTD  comparisons 
of  sea  surface  temperature  and  depth-integrated  tem- 
perature (correlation  coefficient,  r2=0.98;  slope,  m=0.90; 
P<0.0001).  Together,  these  data  were  used  to  define 
ecologically  relevant  “seasons”  (rather  than  calendar 
date)  for  multivariate  analyses. 

Preparation  of  ichthyoplankton  data 

Ichthyoplankton  samples  were  sorted  and  larval  fish 
were  identified  to  the  lowest  possible  taxonomic  level 
at  the  Plankton  Sorting  and  Identification  Center  (Szc- 
zecin, Poland)  and  at  the  Dauphin  Island  Sea  Laboratory 
(Dauphin  Island,  Alabama).  Many  larval  fishes  were  not 
identified  to  the  species  level,  owing  to  the  relatively 
small  sizes  of  larvae  collected  in  the  202-pm  mesh  nets 
and  the  overall  diversity  of  larval  forms  present  in  the 
western  central  Atlantic  region,  which  includes  the 
Gulf  of  Mexico  (Marancik  et  ah,  2005).  Most  identifica- 
tions were  at  the  family  level  (52%),  followed  by  species 
(22%),  order  (14%),  and  genus  (7%)  level  identifications. 
Five  percent  of  the  larvae  collected  were  damaged  or 
unidentified. 

Unidentified  clupeiforms  (engraulids  and  clupeids) 
were  excluded  from  further  analyses  because  their  ex- 
treme concentrations  and  taxonomic  ambiguity  can 
often  mask  abundance  and  assemblage  trends  (Tolan  et 
al.,  1997;  Hernandez  et  al.,  2003).  Order-level  taxa  and 
unidentified  larvae  were  removed  from  consideration  for 
similar  reasons.  Further  taxonomic  analyses,  therefore, 
were  limited  to  taxa  that  represented  at  least  1%  of  the 
total  catch  during  any  individual  sampling  event,  where 
the  proportion  of  the  total  catch  for  each  taxonomic 
group  was  determined  after  removing  unidentified  lar- 
vae, order-level  larvae,  and  all  unidentified  clupeiforms. 
Following  Marancik  et  al.  (2005),  we  further  modified 
the  data  sets  to  exclude  genus-level  groupings  in  in- 
stances where  many  congeners  could  potentially  mask 
any  seasonal  trends.  The  following  genus-level  group- 
ings were  retained  because  each  represented  relatively 
few  congeners  with  likely  similar  early  life  histories  in 
the  northern  Gulf  of  Mexico:  Auxis  spp.  (A.  rochei  and 
A.  thazard),  Centropristis  spp.  (C.  philadelphica,  C. 
ocyurus,  and  C.  striata),  Diplectrum  spp.  ( D . bivattatum 
and  D.  formosum),  Microdesmus  spp.  (M.  lanceolatus 
and  M.  longipinnis),  and  Paralichthys  spp.  (P.  albigutta, 
P.  lethostigma,  and  P.  squamilentus) . Similarly,  all  fam- 
ily-level groups  were  removed  except  Gerreidae  (most 
likely  Eucinostomus  gula  or  E.  argentus)  and  Labridae 
(most  likely  Xyrichtys  novacula ).  In  all,  30  taxa  were 
considered  for  analyses  (Table  2).  Because  the  objective 
of  this  study  was  to  examine  the  seasonal  variability  of 


larval  fish  occurrence  and  relative  larval  fish  concentra- 
tions and  not  size-selectivity  or  vertical  distribution, 
our  analyses  included  ichthyoplankton  data  collected 
from  all  surveys  (monthly  and  quarterly  diel),  mesh 
sizes  (202  pm  and  333  pm),  and  depth  bins.  Depth 
stratification  and  gear  selectivity  will  be  addressed  in 
separate  analyses  in  forthcoming  publications. 

Analyses 

All  fish  egg  and  larval  fish  abundances  were  standard- 
ized by  the  volume  filtered  to  determine  concentration 
estimates  (no./m3).  Taxonomic  diversity  was  calculated 
for  each  sample  by  taking  the  exponential  of  Shannon 
entropy,  exp (H),  following  the  method  of  Jost  (2006). 
Monthly  mean  observations  of  total  fish  eggs,  total  fish 
larvae,  and  taxonomic  diversity  were  compared  to  mean 
temperature  and  salinity  data  by  using  least  squares 
regressions.  Two  approaches  were  used  to  examine 
larval  fish  seasonality.  First,  monthly  mean  concentra- 
tions (no./lOO  m3)  were  calculated  for  the  dominant 
taxa  to  examine  monthly  trends  in  abundance.  Second, 
observed  and  historic  water  temperature  observations 
were  used  to  define  distinct  seasons  for  the  sampling 
region.  Seasonality  in  fish  egg  concentrations,  total 
larval  fish  concentrations,  and  taxonomic  diversity  was 
examined  (after  log+1  transformation)  by  using  one-way 
ANOVAs  with  season  as  a factor  and  Tukey’s  honesty 
significant  difference  (HSD)  tests.  Lastly,  larval  con- 
centrations for  dominant  taxa  were  square-root  trans- 
formed and  analyzed  by  using  Bray  Curtis  similarity 
and  cluster  analysis  with  the  PRIMER  statistical  pack- 
age (PRIMER,  vers.  6,  Plymouth  Marine  Laboratory, 
Plymouth,  U.K.). 

Results 

Mean  monthly  water  temperature  varied  seasonally 
over  the  two  year  period,  with  a low  of  16.5°C  (January 
2005)  and  a high  of  30.2°C  (August  2006)  (Fig.  2).  The 
general  pattern  of  our  monthly  temperature  observations 
was  similar  (±2°C)  to  that  of  recent  historical  values 
(Fig.  3).  Notable  deviations  were  relatively  cooler  tem- 
perature observations  in  May  during  our  study  (mean 
differences  of  3.2°C  and  2.4°C  during  2005  and  2006, 
respectively)  and  warmer  temperatures  in  October  (mean 
differences  of  2.6°C  and  3.0°C  during  2005  and  2006, 
respectively)  and  December  (mean  difference  of  3.0°C 
in  2004).  Even  with  these  disparities,  both  data  sets 
were  in  agreement  to  define  seasonal  breaks  in  water 
temperature.  (Fig.  3).  Sampling  periods  with  mean  water 
temperature  values  <18°C  were  classified  as  winter,  and 
those  with  mean  water  temperatures  above  26°C  were 
classified  as  summer.  The  transitional  periods  of  spring 
and  fall  had  mean  water  temperatures  between  18°C 
and  26°C.  In  general,  the  observed  seasonal  pattern 
comprised  three-month  winter  (December-February) 
and  spring  (March-May)  seasons,  a relatively  long  five- 
month  summer  period  ( July-October),  and  a relatively 


Hernandez  et  at:  Variability  in  ichthyoplankton  abundance  and  composition  in  the  northern  Gulf  of  Mexico 


197 


Table  2 

Seasonal  (X)  and  peak  (*)  occurrence  of  the  dominant  larval  fish  taxa  collected  in  plankton  samples  (n=1634)  off  the  coast  of 
Alabama  from  October  2004  to  October  2006.  Seasonal  classification  is  based  on  historic  (10-year  average)  and  observed  monthly 
mean  temperatures  for  the  region,  (see  Fig.  2). 

Family 

Taxon 

Winter 

Season 

Spring  Summer 

Fall 

Elopidae 

El  ops saurus 

* 

X 

X 

Ophichthidae 

Myrophis  punctatus 

X 

X 

* 

Clupeidae 

Brevoortia  patronus 

* 

X 

X 

X 

Etrumeus  teres 

X 

* 

X 

Harengula  jaguana 

X 

* 

Opisthonema  oglinum 

* 

X 

Serranidae 

Centropristis  spp. 

X 

X 

* 

Diplectrum  spp. 

X 

* 

Serraniculus  pumilio 

X 

* 

Carangidae 

Chloroscombrus  chrysurus 

X 

* 

Decapterus  punctatus 

X 

* 

Lutjanidae 

Lutjanus  campechanus 

* 

Gerreidae 

Unidentified 

X 

* 

Sciaenidae 

Cynoscion  arenarius 

X 

X 

* 

Cynoscion  nothus 

X 

* 

X 

Larimus  fasciatus 

X 

X 

* 

X 

Leiostomus  xanthurus 

X 

X 

X 

* 

Micropogonias  undulatus 

X 

X 

* 

X 

Sciaenops  ocellatus 

X 

Labridae 

Unidentified 

X 

* 

Microdesmidae 

Microdesmus  spp. 

X 

* 

Scombridae 

Auxis  spp. 

X 

* 

Euthynnus  alletteratus 

X 

* 

Scomberomorus  maculatus 

X 

* 

Stromateidae 

Peprilus  alepidotus 

X 

* 

Peprilus  burti 

X 

X 

* 

X 

Paralichthyidae 

Citharichthys  spilopterus 

* 

X 

X 

X 

Etropus  crossotus 

* 

X 

Paralichthys  spp. 

X 

X 

* 

X 

Syacium  papillosum 

* 

X 

short  one-month  fall  period  (November).  In  one  instance, 
the  interannual  variability  in  water  temperature  at 
our  sampling  site  allowed  for  the  same  month  to  be 
designated  as  a different  season  during  different  years 
(December  was  classified  as  “fall”  in  2004  and  “winter” 
in  2005)  (Table  1). 

No  seasonal  pattern  in  salinity  was  observed  at  the 
sampling  station  (Fig.  3).  Salinity  observations  were 
generally  lower  and  more  variable  during  the  first  year 
of  the  study,  with  values  fluctuating  between  30.4  and 
34.6  between  October  2004  and  September  2005.  Sa- 
linity was  generally  higher  and  less  variable  between 
October  2005  and  October  2006,  with  values  ranging 
between  33.0  and  34.8. 

A total  of  504,478  fish  eggs  and  311,970  fish  larvae 
were  collected  over  the  course  of  the  survey.  Total  fish 
egg  concentrations  during  the  survey  ranged  from  0.16 
to  48.3  eggs/m3  (Fig.  3).  Egg  concentrations  were  sig- 


nificantly higher  in  the  spring  than  in  other  seasons 
(F=271.3,  P<0.0001,  spring>summer>fall>winter).  Total 
fish  larvae  concentrations  ranged  from  0.15  to  35.0  lar- 
vae/m3 (Fig.  3).  Larval  concentrations  were  significantly 
higher  during  summer  and  spring  seasons  (F=206.1, 
P<0.0001,  spring=summer>fall>winter).  The  diversity 
of  ichthyoplankton  assemblages,  exp (H),  ranged  from 
1.32  to  9.48  and  was  also  highest  during  the  summer 
seasons  (F=299.3,  P<0.0001,  summer>spring>fall>w 
inter)  (Fig.  3).  Species  diversity  was  significantly  re- 
lated to  temperature  as  determined  by  a least  squares 
regression  (F=34.7,  P<0.001,  r2  = 0.60).  Although  also 
significantly  correlated,  the  relationships  between  tem- 
perature and  fish  egg  concentrations  (F=4.4,  P<0.05, 
r2=0.16)  and  total  larval  concentrations  (F=6.9,  P<0.05, 
r2= 0.23)  were  not  as  strong.  No  significant  relationships 
were  observed  between  salinity  and  fish  eggs  (F=0.22, 
P=0.64,  r2=0.01),  total  fish  larvae  (F<0.01,  P=0.94, 


198 


Fishery  Bulletin  108(2) 


r2<0.01),  and  taxonomic  diversity  (,F=0.16,  P=  0.69, 
r2  = 0.01). 

Excluding  order-level  larvae  and  unidentified  larvae, 
unidentified  engraulids  dominated  our  collections  and 
represented  approximately  50%  of  the  total  (overall) 
catch  (Table  3).  Engraulid  larvae  were  present  year- 
round  and  likely  comprised  several  commonly  occurring 
species  in  the  region,  including  Anchoa  hepsetus,  A.  na- 
suta,  A.  mitchilli,  and  Engraulis  eurystole.  No  attempt 
was  made  to  examine  these  fishes  beyond  the  family 
level  because  many  were  relatively  small  (<10  mm)  and 
damaged,  and  engraulid  identifications  are  problem- 
atic in  our  region  (Farooqi  et  al.,  2006a).  Other  taxa 
that  represented  over  1%  of  the  overall  catch  included 
Cynoscioti  arenarius  (7.5%),  Chloroscombrus  chrysurus 
(5.4%),  Micropogonias  undulatus  (4.4%),  Brevoortia  pa- 
tronus  (3.8%),  unidentified  Gobiidae  (3.6%),  unidentified 
Sciaenidae  (2.8%),  unidentified  Ophidiidae  (2.5%),  Sym- 
phurus  spp.  (2.1%),  Menticirrhus  spp.  (1.2%),  unidenti- 
fied Clupeidae  (1.2%),  Syacium  spp.  (1.2%),  and  Etropus 
crossotus  (1.0%). 

Larval  fish  specimens  collected  during  the  survey 
represented  58  different  families.  Larvae  belonging  to 
22  of  these  families  could  not  be  identified  beyond  the 
family  level,  usually  because  published  descriptions  of 


representative  species  in  our  region  are  either  lacking 
or  are  insufficient  to  discern  between  different  species 
within  the  family  (e.g.,  Gerreidae,  Sparidae,  Haemu- 
lidae,  Echeneidae,  Labridae,  Scorpaenidae).  Several 
families  were  well  represented  with  numerous  species 
or  genera,  including  Ophichthidae  (11  identified  spe- 
cies), Sciaenidae  (9  species),  Carangidae  (7  species), 
Myctophidae  (6  genera),  Paralichthyidae  (5  genera),  and 
Clupeidae  (5  species).  Overall,  the  dominant  families 
collected  during  our  survey  (e.g.,  Engraulidae,  Sciaeni- 
dae, Carangidae,  and  Clupeidae)  are  the  same  as  those 
from  previous  surveys  in  the  general  vicinity  (Table 
3).  In  general,  the  taxonomic  richness  observed  in  our 
survey  falls  between  that  found  in  surveys  of  shorter 
duration  and  in  limited  spatial-scale  surveys  (e.g.,  Wil- 
liams,1983;  Rakocinski  et  al.,  1996)  and  from  SEAMAP 
surveys  that  encompass  a larger  area  and  longer  (20 
years)  time  scales  (ENTRIX,  2006). 

Seasonal  patterns  were  observed  for  most  of  the  domi- 
nant taxa  collected  (Fig.  4).  Lutjanus  campechanus  and 
Chloroscombrus  chrysurus  were  collected  only  during  the 
summer  periods  ( June-October).  Similarly,  Sciaenops 
ocellatus  larvae  were  collected  only  during  late  summer 
(September-October).  In  contrast,  Citharichthys  spilop- 
terus  was  collected  in  almost  every  sampling  event,  in- 
dicating year-round  spawning  or  extended 
pelagic  larval  durations.  Although  sev- 
eral species  had  winter  peaks,  none  were 
present  exclusively  during  winter  months. 
Brevoortia  patronus  and  Paralichthys  spp., 
for  example,  peaked  in  concentration  dur- 
ing November-December,  but  were  also 
collected  in  fall-spring.  Similar  patterns 
were  observed  for  Elops  saurus  and  Micro- 
pogonias undulatus  (late  summer-winter) 
and  Peprilus  burti  and  Leiostomus  xan- 
thurus  (late  summer-spring).  Etrumeus 
teres  differed  in  that  larvae  were  collected 
during  winter-spring  periods.  Most  of  the 
dominant  taxa,  however,  were  collected 
primarily  during  the  late  spring-late 
summer  months  (May-October),  such  as 
Myrophis  punctatus,  Harengula  jaguana, 
Opisthonema  oglinum,  Centropristis  spp., 
Diplectrum  spp.,  Serraniculus  pumilio,  De- 
capterus  punctatus , Auxis  spp.,  Euthynnus 
alletteratus,  Scomberomorus  maculatus, 
Peprilus  alepidotus,  Syacium  spp.,  ger- 
reids,  and  microdesmids.  The  remaining 
taxa  ( Cynoscion  arenarius,  C.  nothus, 
Larimus  fasciatus,  labrids,  and  Etropus 
crossotus)  were  collected  during  the  same 
period,  but  inclusive  of  the  early  spring 
months  (March-April). 

Larval  concentrations  among  the  domi- 
nant taxa  varied  widely  throughout  the 
survey  period  (Fig.  4).  Several  taxa  were 
present  in  low  numbers  throughout  the 
survey.  For  example,  mean  densities  of  E. 
saurus,  O.  oglinum,  Diplectrum  spp.,  S. 


-*-1993-2003  Mean 

♦ 2004 
■ 2005 

• 2006 


JAN  FEB  MAR  APR  MAY  JUN  JUL  AUG  SEP  OCT  NOV  DEC 
Month 

Figure  2 

Mean  monthly  temperature  observations  (depth-integrated)  at  the 
ichthyoplankton  sampling  station  and  the  10-year  average  temperature 
(1993-2003).  Sampling  station  means  are  derived  from  temperature 
profile  observations  recorded  by  the  Bedford  Institute  of  Oceanography 
Net  Environmental  Sampling  System  (BIONESS).  The  10-year  mean 
was  determined  from  near-surface  (0.6  m depth)  temperature  observa- 
tions (T  ) recorded  by  an  oceanographic  buoy  located  approximately 
54  km  west  of  the  sampling  station.  The  plotted  depth-integrated 
temperature  estimates  (T-)  were  calculated  through  the  relation  ship 
T.  = 0.90*77  + 2.37. 


Hernandez  et  al.:  Variability  in  ichthyoplankton  abundance  and  composition  in  the  northern  Gulf  of  Mexico 


199 


o 


35 


30 


25 


20 


15 


-sir 


ONDJ  FMAMJ  JASONDJ  FMAMJ  JASO 


60 

Total  fish  eggs  j 

40 

r£ 

20 

pi 

r3rj 

-i . .1  . 

oD 

r“n 

i — i 

ONDJ  FMAMJ  J ASONDJ  FMAMJ  J ASO 
40  r Total  fish  larvae 


30  - 


20  - 


c 10  - 
a) 
a 


li 


iQ. 


.rinn 


nti 


jffl 


x 

Q.  5 

X 

0 


ONDJ  FMAMJ  JASONDJ  FMAMJ  JASO 
10  f Larval  fish  assemblage 

rj|j 


nfln 


* 


r3ri 


nnnn 


* 


ONDJ  FMAMJ  J ASONDJ  FMAMJ  J ASO 
Month  (2004-2006) 

Figure  3 

Mean  temperature  and  salinity,  fish  egg  and  larval  fish  concentrations, 
and  diversity  indices  for  larval  fish  assemblages  for  October  2004-October 
2006.  Temperature  and  salinity  are  depth-integrated  mean  values  for  each 
month.  Egg  and  larval  fish  concentrations  are  standardized  by  volume  of 
water  filtered  (error  bars  denote  ±1  standard  error).  Calculation  of  diversity 
follows  Jost  (2006)  and  depicts  the  exponential  function  of  Shannon  entropy, 
H (error  bars  denote  ±1  standard  error). 


pumilio,  L.  campechanus,  Gerreidae, 

S.  ocellatus,  Labridae,  Auxis  spp.,  E. 
alletteratus,  P.  burti,  C.  spilopter- 
us,  Paralichthys  spp.,  and  Syacium 
spp.  did  not  exceed  10  larvae/100 
m3  during  any  sampling  event. 

Other  taxa  were  characterized  by 
relatively  high  concentrations,  either 
during  a single  sampling  event  (e.g., 

E.  teres,  C.  chrysurus,  C.  arenarius, 

L.  xanthurus,  Microdesmus  spp.,  S. 
maculatus,  P.  alepidotus ) or  dur- 
ing a single  year  (e.g.,  H.  jaguana). 

The  remaining  taxa  ( M . punctatus, 

B.  patronus,  Centropristis  spp.,  D. 
punctatus,  C.  nothus,  L.  fasciatus, 

M.  undulatus,  E.  crossotus)  were 
present  during  multiple  years  in 
relatively  similar  concentrations. 

Results  from  the  cluster  analysis 
were  largely  in  agreement  with  the 
observed  seasonal  patterns  previ- 
ously defined  by  water  temperature 
(Fig.  5).  Taxonomic  assemblages 
from  fall  and  winter  periods  were 
clustered  separately  from  spring 
and  summer  periods.  All  summer 
months  ( June-October)  were  clus- 
tered together  with  the  exception 
of  August  2005  and  October  2004. 

Larval  collections  in  August  2005 
were  characterized  by  atypically 
high  concentrations  of  a few  spe- 
cies, most  notably  C.  chrysurus  and 

C.  arenarius,  which  were  present 
in  concentrations  exceeding  >500 
larvae/100  m3  (Fig.  4),  resulting 
in  relatively  low  species  diversity 
(Fig.  2)  for  the  summer  period.  The 
October  2004  sampling  event  was 
included  in  the  summer  period,  al- 
though the  mean  temperature  was 
marginally  below  the  26°C  criterion 
for  the  summer  period  (Fig.  3)  and 
indicative  of  a seasonal  transitional 
period.  Similarly,  the  assemblages 
from  the  May  sampling  events  were 
relatively  distinct  from  the  ear- 
lier spring  period  sampling  events 
(March  and  April). 

Discussion 

Although  numerous  ichthyoplankton  surveys  have 
been  conducted  in  the  northern  Gulf  of  Mexico,  most 
have  been  conducted  off  the  coasts  of  Texas,  Louisi- 
ana, and  Florida  (Ditty  et  al.,  1988),  and  few  have  been 
conducted  with  a high  level  of  temporal  resolution  and 
sample  replication.  The  Alabama  shelf  region,  although 


relatively  small,  is  unique  in  that  it  is  bounded  by  two 
major  topographic  features  (Mississippi  River  Delta  to 
the  west  and  DeSoto  Canyon  to  the  east)  that  poten- 
tially inhibit  alongshore  transport  of  larvae  (Johnson 
et  al.,  2009).  In  addition,  the  Alabama  continental  shelf 
receives  freshwater  outflow  from  the  Mobile  River  system, 
which  drains  the  fourth  largest  watershed  in  the  United 
States  and  has  the  sixth  largest  freshwater  discharge 
on  the  North  American  continent  (Park  et  al.,  2007). 
As  a result,  the  inner  shelf  environment  off  Alabama 
is  a highly  productive  region  that  supports  valuable 


200 


Fishery  Bulletin  108(2) 


Table  3 

Summary  (90%  cumulative  percentage  and  abundance  ranking)  of  the  dominant  family  groups  collected  during  the  2004-2006 
ichthyoplankton  survey  in  the  northern  Gulf  of  Mexico  off  the  coast  of  Alabama  and  from  other  ichthyoplankton  surveys  in  the 
general  vicinity. 


Family 

This  study 

ENTRIX  (2006b 

Rakocinski  et  al.  (1996)2 

Williams  (1983)3 

% (Rank) 

% (Rank) 

% (Rank) 

% (Rank) 

Engraulidae 

50.5  (1) 

32.3  (1) 

82.0(1) 

69.3(1) 

Sciaenidae 

15.9(2) 

10.2  (3) 

4.0(3) 

14.0(2) 

Carangidae 

5.4(3) 

2.7  (8) 

5.0(2) 

2.8(4) 

Clupeidae 

5.0(4) 

15.5  (2) 

4.3  (3) 

Paralichthyidae 

3.9(5) 

8.5(4) 

Gobiidae 

3.6(6) 

4.1  (6) 

Ophidiidae 

2.5(7) 

3.6(7) 

Cynoglossidae 

2.1  (8) 

5.6(5) 

Synodontidae 

0.9(9) 

1.9(9) 

Triglidae 

0.8  (10) 

0.8(13) 

Serranidae 

1.9(10) 

Bregmacerotidae 

1.6(11) 

Labridae 

1.0(12) 

Callionymidae 

0.7(14) 

Stromateidae 

0.4(15) 

Scombridae 

0.3  (16) 

Lutjanidae 

0.2(17) 

Congridae 

0.2(18) 

Ophichthidae 

0.2  (19) 

Tetraodontidae 

0.2(20) 

Cumulative  % 

90.6 

91.9 

91.0 

90.4 

1 Samples  (oblique)  were  collected  as  part  of  the  SEAMAP  ichthyoplankton  survey  (Rester  el  al.,  2000)  during  the  months  of  June-November  from 
1982  to  2002  by  using  a 61-cm  bongo  net  fitted  with  333-pm  mesh.  Sample  stations  were  limited  to  the  Mississippi  and  Alabama  inner-shelf 
region. 

2 Samples  (upper  and  lower  water  column)  were  collected  monthly  from  November  1979  to  October  1980  in  Mississippi  Sound  by  using  a 1-m 
diameter  opening-closing  conical-ring  plankton  net  with  335-pm  mesh. 

3 Samples  (surface  and  demersal)  were  collected  monthly  from  March  1979  to  February  1980  in  lower  Mobile  Bay  by  using  a 1x0. 5-m  rectangular 
opening  plankton  net  with  505-pm  mesh. 


fisheries  resources  (Shipp,  1992).  The  establishment  of 
our  survey  is  the  first  to  specifically  target  larval  fish 
assemblages  in  Alabama  shelf  waters  and  is  the  only 
survey  from  the  northern  Gulf  of  Mexico  to  combine 
frequent  sampling  effort  (monthly)  with  high  temporal 
replication  (64+  samples/month)  for  a relatively  long 
duration  (25  months).  Few  ichthyoplankton  surveys  have 
been  conducted  near  our  sampling  location,  including  a 
one-year  survey  of  lower  Mobile  Bay  (Williams,  1983),  a 
one  year  survey  of  Mississippi  Sound  (Rakocinski  et  al., 
1996),  and  a summary  of  SEAMAP  ichthyoplankton  data 
collected  on  the  Mississippi  and  Alabama  shelf  during 
1982-2002  (ENTRIX,  2006).  The  fisheries-independent 
data  collected  during  our  survey,  therefore,  provide  a 
baseline  for  future  comparisons  with  respect  to  vari- 
ability in  local  oceanographic  and  climatic  features  (e.g., 
warming  water  temperatures),  water  and  land  usage 
(e.g.,  Mobile  Bay  nutrient  loading  and  water  outflow), 
and  habitat  modifications  (e.g.,  artificial  reef  programs). 


A comparison  of  results  among  multiple  ichthyoplank- 
ton surveys  is  complicated  because  the  motives  for  sur- 
veys often  differ,  resulting  in  survey-specific  protocols 
and  sampling  biases.  For  example,  the  summary  of 
larval  fish  seasonality  reported  by  Ditty  et  al.  (1988)  for 
the  northern  Gulf  of  Mexico  included  over  30  separate 
surveys  covering  a wide  range  of  spatial  extent  (Gulf- 
wide to  individual  bays  and  passes),  sampling  depths 
(neuston  to  200  m),  mesh  sizes  (0.086-1.05  mm),  gear 
types  (eight  different  samplers),  sampling  frequency 
(biweekly  to  quarterly),  and  survey  duration  (weeks 
to  years).  In  addition,  the  taxonomic  level  to  which 
ichthyoplankton  are  identified  and  at  which  they  are 
reported  varies  with  larval  fish  size,  condition  after 
capture,  and  availability  of  adequate  descriptions.  Our 
decision  to  use  a 202-pm  mesh  size  (as  opposed  to  more 
standard  sizes,  e.g.,  >333  pm)  is  the  factor  that  most 
likely  biases  our  survey  results  when  compared  with 
previous  studies.  The  effect  of  mesh  size  on  the  reten- 


Hernandez  et  al.:  Variability  in  ichthyoplankton  abundance  and  composition  in  the  northern  Gulf  of  Mexico 


201 


co 

E 


o 

o 


0 

03 


03 


C 

o 

5 

c 

0 

O 

c 

o 

O 


0.5 


S.  pumilio 


t I 


A 


4—H 


in 


ONDJ  FMAMJ  J ASONDJ  FMAMJ  J ASO 


600 

500 

400 

300 

200 

100 


C.  chrysurus 

-t-  i.  .i  i.  . | -t-nfal.. 

_ cfaA  i 1 ■ _l_ 

ONDJ  FMAMJ  JASONDJ  FMAMJ  JASO 


Month  2004-2006 

Figure  4 

Mean  larval  concentrations  (no./lOO  m3)  of  dominant  taxa  for  each  month  during  the  ichthyoplank- 
ton survey  (October  2004-October  2006).  Error  bars  denote  ±1  standard  error.  Figure  panels  are 
presented  in  taxonomic  order,  as  listed  in  Table  2. 


tion  of  larvae  has  been  documented  in  numerous  stud- 
ies, with  the  general  conclusion  that  larger  mesh  sizes 
may  efficiently  collect  the  late  larval  stages  but  under- 
estimate the  smaller  size  classes  because  of  extrusion 
(Houde  and  Lovdal,  1984;  Leslie  and  Timmins,  1989). 
Conversely,  smaller  mesh  nets  may  collect  smaller  size 


classes  of  larvae,  but  are  prone  to  clogging,  thus  reduc- 
ing their  effectiveness  in  sampling  ichthyoplankton, 
particularly  late-stage  fish  larvae  (Smith  et  al.,  1968; 
Tranter  and  Smith,  1968).  In  our  study  the  smaller 
mesh  size  enabled  us  to  achieve  better  estimates  of 
fish  egg  and  preflexion  larval  fish  concentrations,  which 


202 


Fishery  Bulletin  108(2) 


Month  2004-2006 

Figure  4 (continued) 


are  indicative  of  nearby  adult  spawning  activity.  The 
tradeoff,  however,  was  that  many  of  the  larvae  were 
too  small  to  identify  to  the  genus  or  species  level.  As 
a result,  most  fish  larvae  collected  in  this  survey  were 
identified  to  the  order  and  family  level  only  (14%  and 
52%,  respectively). 

Fifty-eight  different  families  of  fishes  were  collected 
in  our  ichthyoplankton  collections,  the  adult  forms  of 


which  represent  diverse  zoogeographic  regions  and 
ecological  niches.  As  expected,  larvae  of  nearshore 
and  inner  shelf  species  were  the  most  dominant,  such 
as  coastal  pelagic  (e.g.,  engraulids,  carangids,  clupe- 
ids,  stromateids,  gerreids)  and  coastal  demersal  (e.g., 
sciaenids,  paralichthyids,  gobiids,  cynoglossids,  syn- 
odontids)  species.  Unidentified  engraulids  were  the 
most  abundant  larval  fish  group  in  our  survey  (ap- 


Hernandez  et  al.:  Variability  in  ichthyoplankton  abundance  and  composition  in  the  northern  Gulf  of  Mexico 


203 


10 


Microdesmus 
spp. 


0 1 I I I I I "H-H 


1111 


.7(7..  I 


1 — 1 — 1 — 1 — 1 — i I 1 r 

ON  n .1  F M AM  .1  .1  A 


Month  2004-2006 


Figure  4 (continued) 


proximately  50%)  and  in  the  aforementioned  regional 
surveys  (Table  3).  Engraulid  larvae  appear  to  be  more 
abundant  in  protected  coastal  waters,  as  indicated  by 
their  higher  dominance  in  the  surveys  of  Mobile  Bay 
(82%)  and  Mississippi  Sound  (69%),  both  of  which  are 
shallow  estuarine  regions.  On  the  basis  of  identification 
of  larger  specimens,  most  of  the  engraulids  collected 
in  Mobile  Bay  and  Mississippi  Sound  were  Anchoa 


mitchilli  and  A.  hepsetus  (Williams,  1983;  Rakocinski 
et  al.,  1996),  whereas  our  collections  contained  these 
species  as  well  as  the  coastal  species  A.  nasuta  and 
Engrciulis  eurystole.  The  inner  shelf  taxa  Brevoortia 
patronus,  Cynoscion  arenarius,  Micropogonias  undula- 
tus,  Chloroscombrus  chrysurus,  and  unidentified  gobies 
were  among  the  most  dominant  ichthyoplankton  in  all 
surveys,  including  ours.  As  adults,  these  fishes  are  ex- 


204 


Fishery  Bulletin  108(2) 


Similarity 


Figure  5 

Dendrogram  depicting  relationships  (based  on  Bray  Curtis  similarities)  of 
the  dominant  taxonomic  assemblages  between  months.  Larval  concentrations 
for  dominant  taxa  were  square-root  transformed  before  analyses. 


tremely  abundant  in  estuarine  and  inner  shelf  waters 
and  serve  important  ecological  roles  as  forage  fishes 
(e.g.,  B.  patronus,  C.  chrysu/'us)  and  as  predators  link- 
ing primary  consumers  to  higher  trophic  levels  (e.g.,  M. 
undulatus,  C.  ai'enarius ) (Naughton  and  Saloman,  1981; 
Overstreet  and  Heard,  1982;  Sheridan  et  al.,  1984; 
Franks  et  al.,  2003).  The  larvae  of  these  relatively  few 
taxa  often  comprise  the  majority  of  ichthyoplankton  in 
surveys  throughout  the  northern  Gulf  of  Mexico  (Ditty, 
1986;  Cowan  and  Shaw,  1988;  Tolan  et  al.,  1997). 

Flatfish  larvae  (e.g.,  paralichthyids  and  cynoglossids) 
represented  another  dominant  coastal  group.  Cynoglos- 
sids (Sytnphurus  spp.)  were  common  year-round  in  our 
study,  which  indicates  that  our  collections  contained 
multiple  species.  These  fishes  are  commonly  reported  in 
ichthyoplankton  surveys  throughout  the  Gulf  of  Mexico, 
but  identification  of  larvae  (and  adults)  is  problematic 
owing  to  high  species  richness  and  overlapping  mer- 
istics  (Farooqi  et  al.,  2006b).  Similarly,  Citharichthys 
spp.  were  abundant  year-round,  as  were  C.  spilopterus. 
Again,  identification  down  to  species  is  problematic 
because  five  species  (C.  arctifrons,  C.  cornotus,  C.  gym- 
norhinus,  C.  macrops,  and  C.  spilopterus)  are  found  in 
the  study  region  (Lyczkowski-Shultz  and  Bond,  2006). 
Although  efforts  were  made  to  identify  larvae  conser- 
vatively, some  of  our  C.  spilopterus  may  have  included 
congeners.  This  issue  of  questionable  identification  ap- 
pears less  likely  for  the  Etropus  species  complex,  which 
was  also  abundant,  primarily  E.  crossotus  and  E.  mi- 
crostomus. 

Equally  notable  in  our  survey  was  the  absence  (or 
rarity)  of  larvae  from  taxa  that  are  common  in  our 
sampling  region  as  adults.  For  example,  serranine 
(seabasses)  serranid  larvae  were  collected,  but  epi- 


nepheline  (grouper)  larvae  were  not.  Similarly  absent 
(or  rare)  were  larvae  from  other  recreational  and  com- 
mercially important  species  such  Coryphaena  hippurus 
(Coryphaenidae),  Rachycentron  canadum  (Rachycentri- 
dae),  Balistes  eapriscus  ( Balistidae),  Lobotes  surina- 
mensis  (Lobotidae),  Chaetodipterus  faber  (Ephippidae), 
and  Mugil  cephalus  (Mugilidae),  all  of  which  spawn  in 
coastal  or  offshore  waters  of  Alabama.  The  fact  that 
we  did  not  collect  some  of  these  taxa  is  not  surprising 
(e.g.,  B.  eapriscus,  M.  cephalus)  because  they  are  more 
commonly  collected  in  the  neuston  (which  we  did  not 
sample).  The  absence  of  grouper  larvae  is  perplex- 
ing, even  though  the  rarity  of  epinepheline  larvae  has 
been  documented  in  the  northern  Gulf  of  Mexico.  For 
example,  only  37  grouper  larvae  were  collected  in  gulf- 
wide SEAMAP  ichthyoplankton  surveys  between  1982 
and  1999  (>7000  samples)  (Lyczkowski-Shultz  et  al.1). 
Most  of  the  grouper  larvae  were  collected  at  offshore 
SEAMAP  sampling  stations,  which  indicates  that  their 
occurrence  in  nearshore  environments  may  be  rare. 
It  is  possible  that  the  limited  spatial  extent  of  our 
survey  (i.e.,  a single  station)  may  have  influenced  our 
estimates  of  larval  fish  concentrations  and  variability, 
because  coastal  marine  processes  that  influence  larval 
fish  dynamics  are  often  site-specific  (e.g.,  local  wind 
regimes,  tidal  flows,  river  discharge),  but  the  overall 
seasonal  supply  of  larvae  available  at  our  sampling 
station  is  likely  representative  of  the  ichthyofauna  from 
a larger  northcentral  Gulf  of  Mexico  region  between 
the  87°W  and  89°W  longitude  (Boschung,  1992). 

The  main  objective  of  this  study  was  to  describe 
taxon-specific  seasonality  for  larval  fishes  collected  in 
the  survey  region.  For  several  reasons,  we  limited  our 
seasonal  analyses  to  water  temperature,  as  opposed  to 


Hernandez  et  al.:  Variability  in  ichthyoplankton  abundance  and  composition  in  the  northern  Gulf  of  Mexico 


205 


a suite  of  environmental  parameters.  First,  tempera- 
ture has  long  been  proposed  as  an  important  factor  in 
the  initiation  of  spawning  for  marine  fishes  (Orton, 
1920),  and  numerous  field  and  laboratory  (primarily 
aquaculture-related)  studies  have  provided  support  for 
temperature  as  a primary  influence  (Arnold  et  al.,  2002; 
Sheaves,  2006).  Second,  water  temperature  varies  pre- 
dictably at  seasonal  scales  (e.g.,  months),  as  opposed 
to  other  factors  that  vary  at  shorter  time  scales.  Our 
salinity  data  (Fig.  3),  for  example,  showed  no  seasonal 
trends  and  were  not  correlated  with  egg  or  larval  fish 
concentrations.  The  monthly  mean  salinity  values  cal- 
culated during  each  cruise  likely  reflect  short-term 
variability  related  to  tidal  flow,  riverine  outflow,  local 
wind  conditions,  and  related  factors  that  affect  salinity 
at  our  sampling  station.  In  addition,  salinity,  although 
an  important  factor  for  many  estuarine-spawning  spe- 
cies, is  generally  considered  less  important  than  tem- 
perature to  the  timing  of  marine  fish  spawning  (Bye, 
1984;  Sheaves,  2006). 

Defining  seasonality  in  terms  of  water  temperature 
also  provides  a framework  for  monitoring  fisheries  dy- 
namics with  respect  to  anticipated  rises  in  sea  tem- 
perature due  to  global  climate  change.  Our  monthly 
observed  depth-integrated  temperatures  were  relatively 
consistent  with  those  for  the  previous  ten-year  aver- 
age for  the  region,  although  winter  (December- Janu- 
ary) and  late  summer  (August-October)  values  were 
generally  higher  (Fig.  2).  Fodrie  et  al.  (2009)  noted  a 
significant  increase  in  sea  surface  temperature  near 
the  mouth  of  Mobile  Bay  over  a 20-year  period  (1987- 
2007).  The  authors  also  noted  a concurrent  increase  in 
the  number  and  occurrence  of  juvenile  subtropical  and 
tropical  fishes  collected  in  seagrass  meadows  along  the 
northern  Gulf  of  Mexico.  For  example,  in  2006-2007 
surveys,  juveniles  of  tropical  species  such  as  Chaetodon 
ocellatus  (Chaetodontidae),  Fistularia  tabaccu'ia  (Fistu- 
laridae),  Ocyurus  chrysurus  (Lutjanidae),  Thalassoma 
bifasciatum  (Labridae),  Sparisoma  viride  (Scaridae), 
and  unidentified  acanthurids  were  collected  in  coastal 
habitats  where  they  were  not  collected  during  previous 
surveys  (1971-79)  (Livingston,  1985).  Notably,  in  our 
ichthyoplankton  survey  larvae  from  all  of  these  fami- 
lies, except  Chaetodontidae,  were  collected  but  regret- 
tably, comparable  ichthyoplankton  data  from  the  1970s 
were  not  available  and  our  identifications  were  made 
only  to  the  family  level. 

Conclusions 

Increases  in  regional  water  temperatures  may  have  sig- 
nificant impacts  on  the  reproductive  success  of  marine 
fishes  and  the  subsequent  survival  of  early  life  stages, 
including  early  gonad  maturation  and  spawning  in 
adults,  altered  larval  transport  pathways,  extended 
pelagic  larval  durations,  changes  in  larval  assemblage 
structure,  and  mismatched  timing  of  larval  fish  occur- 
rence with  food  resources  and  physiological  optima, 
among  other  effects  (Sheaves,  2006;  O’Conner  et  al., 


2007;  Genner  et  al.,  2009).  Establishment  of  long-term 
baseline  surveys  provides  a means  of  monitoring  larval 
fish  assemblages  and  the  factors  that  influence  larval 
fish  dynamics  in  order  to  provide  early  indicators  of 
ecosystem  changes  due  to  environmental  perturbations. 
The  ichthyoplankton  survey  efforts  described  here  for 
the  October  2004-October  2005  period  have  since  con- 
tinued and  expanded  to  include  near  monthly  (depth- 
discrete)  ichthyoplankton  sampling  at  five  stations  along 
a cross-shelf  transect  from  inside  Mobile  Bay  extending 
offshore  to  a station  approximately  54  km  south  of  Dau- 
phin Island.  The  expanded  survey  program  (Fisheries 
Oceanography  of  Coastal  Alabama,  or  FOCAL)  will  allow 
us  to  estimate  and  monitor  the  variability  in  ichthyo- 
plankton seasonality,  abundance,  assemblage  structure, 
and  vertical  distribution  over  multiple  temporal  and 
spatial  scales. 

Acknowledgments 

We  would  like  to  thank  the  technicians  and  graduate 
students  that  participated  on  our  research  cruises: 
A.  Beck,  S.  Bosarge,  L.  Chiaverano,  T.  Clardy,  D.  del 
Valle,  N.  Geraldi,  J.  Goff,  E.  Goldman,  J.  E.  Herrmann, 
J.  M.  Herrmann,  J.  Higgins,  L.  Kramer,  B.  Lacour,  C. 
Martin,  M.  Miller,  S.  Muffelman,  C.  Newton,  C.  Pabody, 
D.  Ploetz,  C.  Schobernd,  Z.  Schobernd,  R.  Shiplett,  and 
D.  Vivian.  We  especially  thank  the  captains  and  crew  of 
the  RV  Verrill  and  RV  E.O.  Wilson  (R.  Collier,  T.  Guoba, 
C.  Lollar,  and  R.  Wilson)  and  the  Dauphin  Island  Sea 
Laboratory  technical  support  team  (M.  Dardeau,  A. 
Gunter,  and  K.  Weiss).  We  also  thank  M.  Konieczna 
and  the  scientific  staff  at  the  Plankton  Sorting  and 
Identification  Center  in  Szczecin,  Poland,  for  larval 
fish  identifications.  K.  Park  provided  assistance  with 
the  CTD  data.  H.  Fletcher  and  L.  Hu  provided  data- 
base management  support.  S.  Bosarge  produced  the 
station  map  (Fig.  1).  Valuable  comments  and  guidance 
throughout  the  course  of  our  survey  were  provided  by  J. 
Lyczkowski-Shultz  (NOAA/NMFS/SEFSC,  Pascagoula 
Laboratory,  MS)  and  S.  Heath  (Alabama  Department  of 
Conservation  and  Natural  Resources,  Marine  Resources 
Division,  Dauphin  Island,  AL).  We  also  thank  R.  Bro- 
deur  and  three  anonymous  reviewers  for  comments 
on  a previous  version  of  this  manuscript.  A portion  of 
these  data  were  collected  as  part  of  contract  # 2004- 
GPS-MSA-NC-0085  from  ConocoPhillips  Corporation, 
Houston,  TX. 


Literature  cited 

Arnold,  C.  R.,  J.  B.  Kaiser,  and  G.  J.  Holt. 

2002.  Spawning  of  cobia  Rachycentron  canadum  in 
captivity.  J.  World  Aquacult.  Soc.  33:205-208. 

Auth,  T.D. 

2008.  Distribution  and  community  structure  of  ichthyo- 
plankton from  the  northern  and  central  California  Cur- 
rent in  May  2004-06.  Fish.  Oceanogr.  17:316-331. 


206 


Fishery  Bulletin  108(2) 


Boschung,  H.  T. 

1992.  Introduction:  geographical  area  of  coverage.  In 
Catalog  of  freshwater  and  marine  fishes  of  Alabama 
(H.  T.  Boschung,  ed.),  p.  3-5.  Alabama  Mus.  Nat.  Hist. 
14,  Univ.  Alabama,  Tuscaloosa,  AL. 

Brodeur,  R.  D.,  W.  T.  Peterson,  T.  D.  Auth,  H.  L.  Soulen,  M.  M. 
Parnel,  and  A.  A.  Emerson. 

2008.  Abundance  and  diversity  of  coastal  fish  larvae 
as  indicators  of  recent  changes  in  ocean  and  climate 
conditions  in  the  Oregon  upwelling  zone.  Mar.  Ecol. 
Prog.  Ser.  366:187-202. 

Bye,  V.  J. 

1984.  The  role  of  environmental  factors  in  the  timing 
of  reproductive  cycles.  In  Fish  reproduction:  strate- 
gies and  tactics  (G.  W.  Potts  and  R.J  . Wooton,  eds.), 
p.  132-148.  Academic  Press,  London. 

Cowan,  J.  H.,  and  R.  F.  Shaw. 

1988.  The  distribution,  abundance,  and  transport  of  larval 
sciaenids  collected  during  winter  and  early  spring  from 
the  continental  shelf  waters  off  west  Louisiana.  Fish. 
Bull.  86:129-142. 

Ditty,  J.  G. 

1986.  Ichthyoplankton  in  neritic  waters  of  the  northern 
Gulf  of  Mexico  off  Louisiana:  Composition,  relative  abun- 
dance, and  seasonality.  Fish.  Bull.  84:935-  946. 

Ditty,  J.  G.,  G.  G.  Zieske,  and  R.  F.  Shaw. 

1988.  Seasonality  and  depth  distribution  of  larval  fishes 
in  the  northern  Gulf  of  Mexico  above  latitude  26°00'N. 
Fish.  Bull.  86:811-823. 

Doyle,  M.  J.,  W.  W.  Morse,  and  A.  W.  Kendall  Jr. 

1993.  A comparison  of  larval  fish  assemblages  in  the 
temperate  zone  of  the  northeast  Pacific  and  northwest 
Atlantic  oceans.  Bull.  Mar.  Sci.  53:588-644. 

ENTRIX. 

2006.  Final  environmental  impact  statement  for 
the  Compass  Port  LLC  Deepwater  Port  License 
Application.  United  States  Coast  Guard,  Docket  number 
USCG-2004-17659. 

Farooqi,  T.  W.,  J.  G.  Ditty,  and  R.  F.  Shaw. 

2006a.  Engraulidae:  anchovies.  In  early  stages  of  Atlan- 
tic fishes:  an  identification  guide  for  the  western  Central 
North  Atlantic  (W.  J.  Richards,  ed.),  p.  101-127.  Taylor 
and  Francis  Group,  Boca  Raton,  FL. 

Farooqi,  T.  W.,  R.  F.  Shaw,  J.  G.  Ditty,  and  J.  Lyczkowski- 
Shultz. 

2006b.  Cynoglossidae:  tonguefishes.  In  Early  stages  of 
Atlantic  fishes:  an  identification  guide  for  the  western 
Central  North  Atlantic  (W.  J.  Richards,  ed.),  p.  2367- 
2379.  Taylor  and  Francis  Group,  Boca  Raton,  FL. 

Fodrie,  F.  J.,  K.  L.  Heck,  S.  P.  Powers,  W.  M.  Graham,  and  K.  L. 
Robinson. 

2009.  Climate-related,  decadal-scale  assemblage  changes 
of  seagrass-associated  fishes  in  the  northern  Gulf  of 
Mexico.  Glob.  Change  Biol.  16:48-59. 

Franks,  J.  S.,  K.  E.  VanderKooy,  and  N.  M.  Garber. 

2003.  Diet  of  tripletail,  Lobotes  surinamensis,  from  Mis- 
sissippi coastal  waters.  Gulf  Caribb.  Res.  15:27-32. 

Genner,  M.  J.,  N.  C.  Halliday,  S.  D.  Simpson,  A.  J.  Southward, 
S.  J.  Hawkins,  and  D.  W.  Sims. 

2009.  Temperature-driven  phenological  changes  within 
a marine  larval  fish  assemblage.  J.  Plankton  Res. 
doi:10.1093/plankt/fbp082. 

Govoni  J.  J.,  D.  E.  Hoss,  and  D.  R.  Colby. 

1989.  The  spatial  distribution  of  larval  fishes  about  the 
Mississippi  River  plume.  Limnol.  Oceanogr.  34:178- 
187. 


Gray,  C.  A.,  and  A.  G.  Miskiewicz. 

2000.  Larval  fish  assemblages  in  south-east  Australian 
coastal  waters:  seasonal  and  spatial  structure.  Estuar. 
Coast.  Shelf  Sci.  50:549-570. 

Hare,  J.  A.,  and  J.  J.  Govoni. 

2005.  Comparison  of  average  larval  fish  vertical  distri- 
butions among  species  exhibiting  different  transport 
pathways  on  the  southeast  United  States  continental 
shelf.  Fish.  Bull.  103:728-736. 

Hernandez,  F.  J.,  Jr.,  R.  F.  Shaw,  J.  S.  Cope,  J.  G.  Ditty, 
T.  W.  Farooqi,  and  M.  C.  Benfield. 

2003.  The  across-shelf  larval,  postlarval  and  juvenile 
fish  community  associated  with  offshore  oil  and  gas 
platforms  west  of  the  Mississippi  River  Delta.  In  Fish- 
eries, reefs  and  offshore  development  (D.  Stanley,  and 
A.  Scarborough-Bull,  eds.),  p.  39-72.  Am.  Fish.  Soc. 
Symp.  36,  Bethesda,  MD. 

Hernandez-Miranda,  E.,  A.  T.  Palma,  and  F.  P.  Ojeda. 

2003.  Larval  fish  assemblages  in  nearshore  coastal  waters 
off  central  Chile:  temporal  and  spatial  patterns.  Estuar. 
Coast.  Shelf  Sci.  56:1075-1092. 

Houde,  E.  D.,  and  J.  A.  Lovdal. 

1984.  Seasonality  of  occurrence,  foods  and  food  prefer- 
ences of  ichthyoplankton  in  Biscayne  Bay,  Florida.  Es- 
tuar. Coast.  Shelf  Sci.  18:403-419. 

Johnson,  D.  R.,  H.  M.  Perry,  J.  Lyczkowski-Shultz,  and 
D.  Hanisko. 

2009.  Red  snapper  larval  transport  in  the  northern  Gulf 
of  Mexico.  Trans.  Am.  Fish.  Soc.  138:458-470. 

Jost,  L. 

2006.  Entropy  and  diversity.  Oikos  113:363-375. 

Leslie,  J.  K.,  and  C.  A.  Timmins. 

1989.  Double  nets  for  mesh  aperture  selection  and  sam- 
pling in  ichthyoplankton  studies.  Fish.  Res.  7:225-232. 

Livingston,  R.  J. 

1985.  Organization  of  fishes  in  coastal  seagrass  system: 
the  response  to  stress.  In  Fish  community  ecology  in 
estuaries  and  coastal  lagoons:  towards  an  ecosystem  inte- 
gration (A.  Yanez-Arancibia,  ed.),  p.  367-382.  UNAM 
Press,  Mexico  City,  Mexico. 

Lyczkowski-Shultz,  J.,  and  P.  J.  Bond. 

2006.  Paralichthyidae:  sand  flounders.  In  Early  stages 
of  Atlantic  fishes:  an  identification  guide  for  the  western 
Central  North  Atlantic  (W.J.  Richards,  ed.),  p.  2291- 
2325.  Taylor  and  Francis  Group,  Boca  Raton,  FL. 

Lyczkowski-Shultz,  J.,  and  D.  S.  Hanisko. 

2007.  A time  series  of  observations  on  red  snapper  larvae 
from  SEAMAP  surveys,  1982-  2003:  seasonal  occur- 
rence, distribution,  abundance  and  size.  Am.  Fish. 
Soc.  Symp.  60:3—23. 

Lyczkowski-Shultz,  J.,  and  G.  W.  Ingram  Jr. 

2003.  Preliminary  guide  to  the  identification  of  the 
early  life  stages  of  balistid  fishes  of  the  western  Central 
North  Atlantic.  NOAA  Tech.  Memo.  NMFS-SEFC- 
493,  13  p. 

Marancik,  K.  E.,  L.  M.  Clough,  and  J.  A.  Hare. 

2005.  Cross-shelf  and  seasonal  variation  in  larval  fish 
assemblages  on  the  southeast  United  States  continen- 
tal shelf  off  the  coast  of  Georgia.  Fish.  Bull.  103: 
108-129. 

Muhling,  B.  A.,  L.  E.  Beckley,  J.  A.  Koslow,  and  A.  F.  Pearce. 

2008.  Larval  fish  assemblages  and  water  mass  struc- 
ture off  the  oligotrophic  south-  western  Australian 
coast.  Fish.  Oceanogr.  17:16-31. 

Naughton,  S.  P,  and  C.  H.  Saloman. 

1981.  Stomach  contents  of  juveniles  of  king  mackerel 


Hernandez  et  al.:  Variability  in  ichthyoplankton  abundance  and  composition  in  the  northern  Gulf  of  Mexico 


207 


(Scomberomorus  cavalla)  and  Spanish  mackerel  ( S . 
maculatus).  Northeast  Gulf  Sci.  5:71-74. 

O’Connor,  M.  I.,  J.  F.  Bruno,  S.  D.  Gaines,  B.  S.  Halpern, 
S.  E.  Lester,  B.  P.  Kinlan  and  J.  M.  Weiss. 

2007.  Temperature  control  of  larval  dispersal  and 
the  implications  for  marine  ecology,  evolution,  and 
conservation.  Proc.  Natl.  Acad.  Sci.  USA  104:1266— 
1271. 

Orton,  J.  H. 

1920.  Sea-temperature,  breeding  and  distribution  in 
marine  animals.  J.  Mar.  Biol.  Assoc.  U.K.  12:339- 
366. 

Overstreet,  R.  M.,  and  R.  W.  Heard. 

1982.  Food  contents  of  six  commercial  fishes  from  Mis- 
sissippi Sound.  Gulf  Res.  Rep.  7:137-149. 

Park,  K.,  Kim,  C.-K.,  and  W.  W.  Schroeder. 

2007.  Temporal  variability  in  summertime  bottom  hypoxia 
in  shallow  areas  of  Mobile  Bay,  Alabama.  Estuaries 
Coasts  30:54-65. 

Rakocinski,  C.  F.,  J.  Lyczkowski- Shultz,  and  S.  L.  Richardson. 

1996.  Ichthyoplankton  assemblage  structure  in  Missis- 
sippi Sound  as  revealed  by  canonical  correspondence 
analysis.  Estuar.  Coast.  Shelf  Sci.  43:237-257. 

Raynie,  R.  C.,  and  R.  F.  Shaw. 

1994.  Ichthyoplankton  abundance  along  a recruitment 
corridor  from  offshore  spawning  to  estuarine  nursery 
ground.  Estuar.  Coast.  Shelf  Sci.  39:421-450. 

Rester,  J.  K.,  D.  Hanisko,  N.  Sanders,  Jr.,  and  B.  Pellegrin. 

2000.  SEAMAP  environmental  and  biological  atlas  of 
the  Gulf  of  Mexico,  1998.  Gulf  States  Marine  Fisher- 
ies Commission,  Ocean  Springs,  MS. 

Richards,  W.  J.,  M.  F.  McGowan,  T.  Leming,  J.  T.  Lamkin,  and 
S.  Kelley. 

1993.  Larval  fish  assemblages  at  the  loop  current  bound- 
ary in  the  Gulf  of  Mexico.  Bull.  Mar.  Sci.  53:475- 
537. 

Scott,  G.  P,  S.  C.  Turner,  C.  B.  Grimes,  W.  J.  Richards,  and 
E.  B.  Brothers. 

1993.  Indices  of  larval  bluefin  tuna,  Tliunnus  thynnus, 
abundance  in  the  Gulf  of  Mexico:  modeling  variability 
in  growth,  mortality,  and  gear  selectivity.  Bull.  Mar. 
Sci.  53:912-929. 


Sheaves,  M. 

2006.  Is  the  timing  of  spawning  in  sparid  fishes  a 
response  to  sea  temperature  regimes?  Coral  Reefs 
25:655-669. 

Sheridan,  P.  F.,  D.  L.  Trimm,  and  B.  M.  Baker. 

1984.  Reproduction  and  food  habits  of  seven  species  of 
northern  Gulf  of  Mexico  fishes.  Contrib.  Mar.  Sci. 
27:175-204. 

Sherman,  K.  W.,  W.  Smith,  W.  Morse,  M.  Berman,  J.  Green,  and 
L.  Ejsymont. 

1984.  Spawning  strategies  of  fishes  in  relation  to  cir- 
culation, phytoplankton  production,  and  pulses  in  zoo- 
plankton off  the  northeastern  United  States.  Mar. 
Ecol.  Prog.  Ser.  18:1-19. 

Shipp,  R.  L. 

1992.  Introduction:  biogeography  of  Alabama’s  marine 
fishes.  In  Catalog  of  freshwater  and  marine  fishes  of 
Alabama  (H.  T.  Boschung,  ed.),  p.  7-9.  Alabama  Mus. 
Nat.  Hist.  14,  Univ.  Alabama,  Tuscaloosa,  AL. 

Smith,  P.  E.,  R.  C.  Counts,  and  R.  I.  Clutter. 

1968.  Changes  in  filtering  efficiency  of  plankton  nets 
due  to  clogging  under  tow.  J.  Cons.  Int.  Explor.  Mer 
32:232-248. 

Sogard,  S.  M.,  D.  E.  Hoss,  and  J.  J.  Govoni. 

1987.  Density  and  depth  distribution  of  larval  gulf  men- 
haden, Brevoortia  patronus,  Atlantic  croaker,  Micropogo- 
nias  undulatus , and  spot,  Leiostomus  xanthurus,  in  the 
northern  Gulf  of  Mexico.  Fish.  Bull.  85:601-609. 

Tolan,  J.  M.,  S.  A.  Holt,  and  C.  P.  Onuf. 

1997.  Distribution  and  community  structure  of  ichthyo- 
plankton in  Laguna  Madre  Seagrass  meadows:  potential 
impact  of  seagrass  species  change.  Estuaries  Coasts 
20:450-464. 

Tranter,  D.  J.  and  P.  E.  Smith. 

1968.  Filtration  performance.  In  Reviews  on  zooplank- 
ton sampling  methods,  part  I (D.  J.  Tranter,  ed.),  p. 
27-56.  Monogr.  Oceanogr.  Methodol.  2,  UNESCO 
Press,  Paris. 

Williams,  L.  W. 

1983.  Larval  fish  assemblages  of  lower  Mobile  Bay. 
M.S.  thesis,  55  p.  Dep.  Biology,  Univ.  South  Alabama, 
Mobile,  AL. 


208 


Observer-reported  skate  bycatch 

in  the  commercial  groundfish  fisheries  of  Alaska 


Kristy  A.  Lewis 

Email  address  for  contact  author:  duane.stevenson@noaa.gov 

NMFS,  Alaska  Fisheries  Science  Center 

Fisheries  Monitoring  and  Analysis  Division 

7600  Sand  Point  Way  NE 

Seattle,  Washington  98115 


Abstract — We  analyzed  skate  catch 
data  collected  by  observers  in  the 
North  Pacific  Groundfish  Observer 
Program  (NPGOP)  from  1998  through 
2008  to  document  recent  changes  in 
the  identification  of  skates  by  observ- 
ers and  to  examine  the  species  com- 
position of  observed  skate  catch  in 
Alaska’s  groundfish  fisheries  as  well 
as  recent  trends  in  skate  retention 
by  commercial  fishermen.  Histori- 
cally, almost  all  skate  bycatch  has 
been  reported  by  NPGOP  observers  as 
“skate  unidentified.”  However,  since 
2004  observers  have  been  trained  to 
identify  skates  to  the  genus  and  spe- 
cies level.  In  2008  over  95%  of  all 
skates  were  identified  at  least  to  the 
genus  level,  and  over  50%  were  iden- 
tified to  species.  The  most  common 
species  of  skates  identified  by  observ- 
ers in  groundfish  fisheries  are  Bathy- 
raja parmifera  (Alaska  skate),  Raja 
binoculata  (big  skate),  and  Bathyraja 
aleutica  (Aleutian  skate).  Species  com- 
position of  reported  skate  catch  gen- 
erally reflects  recent  survey-derived 
biomass  estimates,  with  B.  parmifera 
dominating  the  catches  in  the  Bering 
Sea  and,  to  a lesser  extent,  in  the 
Aleutian  Islands  region,  and  species 
of  the  genus  Raja  dominating  catches 
in  the  Gulf  of  Alaska.  A relatively 
high  percentage  of  the  skate  catch 
on  longline  vessels  is  still  reported 
at  the  family  or  genus  level  because 
of  difficulties  in  the  identification 
of  skates  not  brought  onboard  the 
vessel.  For  the  larger  skate  species, 
the  proportion  retained  for  processing 
has  increased  in  recent  years  as  the 
market  price  for  skate  product  has 
increased.  Although  observed  skate 
catch  does  not  give  a complete  account 
of  skate  bycatch  in  the  fisheries  of  the 
region,  observer  data  provide  criti- 
cal information  for  the  appropriate 
management  of  skate  populations  in 
Alaska. 


Manuscript  submitted  28  September  2009. 
Manuscript  accepted  19  January  2010. 
Fish.  Bull.  108:208-217  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


Duane  E.  Stevenson  (contact  author) 


Skates  are  large,  long  lived  fishes  with 
relatively  slow  growth  rates  and  low 
reproductive  potential  (Ebert,  2005; 
Ebert  et  al.,  2008).  These  aspects  of 
their  life  history,  combined  with  their 
relatively  low  mobility  and  benthic 
habitat,  make  skates  particularly 
vulnerable  to  fishing  pressure  and 
slow  their  recovery  from  population 
declines;  yet  few  countries  have  man- 
agement plans  for  skates  or  other 
chondrichthyan  species  (Stevens  et  ah, 
2000).  In  cases  where  skates  are  tar- 
geted by  fisheries,  population  declines 
can  be  rapid  (Agnew  et  ah,  2000). 
Moreover,  because  discard  mortality 
can  be  high  (Stobutzki  et  al.,  2002; 
Laptikhovsky,  2004),  skate  popula- 
tions may  be  dramatically  affected 
by  fishing  activity  even  if  they  are 
not  targeted  directly  (Brander,  1981; 
Casey  and  Myers,  1998;  Dulvy  et  ah, 
2000;  Stevens  et  al.,  2000).  In  addition 
to  population  declines,  fishing  pres- 
sure may  lead  to  significant  shifts  in 
community  structure  because  declines 
in  some  species  of  a skate  assemblage 
may  be  masked  by  increases  in  other, 
more  resilient  species  (Agnew  et  al., 
2000;  Dulvy  et  al.,  2000;  Stevens  et 
al.,  2000).  Therefore,  effective  man- 
agement of  skate  populations  requires 
species-specific  data  on  abundance 
trends  and  exploitation  rates. 

Skates  are  regularly  caught  in 
nearly  all  of  the  commercial  ground- 
fish fisheries  currently  prosecuted  in 
Alaska  waters,  including  fisheries 
targeting  Pacific  cod  ( Gadus  mac- 
rocephalus),  walleye  pollock  ( Ther - 
agra  chalcogramma),  yellowfin  sole 
(Limanda  aspera),  and  other  species 
(Ormseth  et  al.,  2009).  In  addition  to 
their  ubiquitous  presence  as  bycatch 


species,  skates  have  been  targeted  in 
Alaska  waters  on  a short-term  region- 
al basis.  An  unregulated  fishery  tar- 
geting Raja  binoculata  (big  skate),  R. 
rhina  (longnose  skate),  and  assorted 
species  of  Bathyraja  (including  Alaska 
skates)  developed  in  the  central  Gulf 
of  Alaska  (GOA)  in  February  2003. 
Shifting  economic  conditions  and  fish- 
ing seasons  soon  made  other  target 
species  more  valuable,  but  this  short- 
lived fishery  revealed  that  skates  can 
quickly  become  an  attractive  alterna- 
tive target  when  other  fisheries  are 
closed  (Matta,  2006).  More  recently, 
the  Alaska  Department  of  Fish  and 
Game  (ADF&G)  approved  a pilot  fish- 
ery for  big  and  longnose  skates  in 
the  state-managed  waters  of  Prince 
William  Sound  (ADF&G  Emergency 
Order  #2-G-E-04-09)  in  2009.  Else- 
where in  Alaska  skates  are  still  man- 
aged as  part  of  a large  nontarget  spe- 
cies complex,  although  beginning  in 
2011  skates  in  the  Bering  Sea  and 
Aleutian  Islands  will  be  managed  as 
a separate  unit. 

Recent  advances  in  the  taxonomy 
of  the  skates  of  the  North  Pacific  and 
Bering  Sea  (Ishiyama  and  Ishihara, 
1977;  Ishihara  and  Ishiyama,  1985, 
1986;  Stevenson  et  al.,  2004,  2007, 
2008)  have  facilitated  increasingly 
detailed  identification  of  skates  by 
observers  in  the  commercial  fisher- 
ies of  Alaska.  The  resulting  wealth 
of  detailed  catch  data  now  permits 
an  examination  of  skate  bycatch  on 
a level  that  was  not  previously  pos- 
sible. The  objectives  of  this  study  are 
1)  to  document  recent  changes  for  the 
identification  of  skates  in  the  NPGOP, 
and  2)  to  provide  an  overview  of  po- 
tential management  concerns  by  ex- 


Stevenson  and  Lewis:  Skate  bycatch  in  the  commercial  groundfish  fisheries  of  Alaska 


209 


amining  the  species  composition  of  observed  skate  catch 
(OSC)  in  Alaska’s  groundfish  fisheries  and  recent  trends 
in  skate  retention  by  commercial  fishermen. 

Materials  and  methods 

All  data  used  for  this  study  were  extracted  from  the 
North  Pacific  Groundfish  Observer  Program  (NPGOP) 
database  maintained  by  the  Fisheries  Monitoring  and 
Analysis  (FMA)  Division  of  the  National  Marine  Fisher- 
ies Service’s  (NMFS)  Alaska  Fisheries  Science  Center. 
This  database  houses  all  biological  data  collected  by 
groundfish  observers  onboard  commercial  fishing  vessels 
operating  in  the  waters  of  Alaska’s  federal  Exclusive  Eco- 
nomic Zone  (EEZ).  For  an  overview  of  the  database,  see 
the  FMA  Division  website  (National  Marine  Fisheries 
Service,  http://www.afsc.noaa.gov/FMA/fma_database. 
htm,  accessed  November  2009). 

Federal  law  requires  observers  to  be  present  at  all 
times  on  commercial  fishing  vessels  of  125  ft  (38.1  m) 
or  more  in  length  overall  (LOA)  operating  in  the  fed- 
eral EEZ.  For  vessels  from  60  to  124  ft  (18.3  to  37.8 
m)  LOA,  observer  coverage  is  required  for  only  30%  of 
fishing  days  and  for  vessels  less  than  60  ft  (18.3  m) 
LOA  no  observer  coverage  is  required.  The  catch  data 
used  for  this  study  were  taken  from  trawl  hauls  and 
longline  sets  during  which  an  observer  was  present  and 
was  sampling,  so  that  the  catch  statistics  presented 
here  do  not  represent  the  total  catch  of  the  fisheries  in 
this  region,  nor  do  they  represent  biomass  estimates. 
For  some  commercial  fisheries  in  the  area,  pot  gear  is 
used,  but  observers  rarely  encounter  skates  in  these 
fisheries,  and  therefore  such  data  are  not  included  in 
this  study. 

The  process  used  by  observers  to  determine  the  spe- 
cies composition  and  catch  weights  of  sampled  hauls 
depends  on  gear  type.  Observers  on  trawlers  may  de- 
termine the  species  composition  of  a haul  by  identifying 
and  weighing  the  entire  catch,  which  is  usually  not 
possible,  or  by  choosing  a random  sample  (generally  300 
kg  or  more)  of  the  catch  and  identifying  and  weighing 
all  taxa  within  the  sample.  The  proportion  by  weight 
of  each  taxon  in  the  sample  is  then  extrapolated  to  the 
total  catch  weight,  which  may  be  determined  by  a num- 
ber of  methods,  including  flow  scale  readings,  codend 
measurements,  or  bin  volume  estimates.  On  longline 
vessels,  observers  randomly  select  a “tally  period”  as 
the  gear  is  being  retrieved.  During  this  tally  period, 
the  observer  identifies  and  counts  specimens,  including 
specimens  that  drop  off  the  line  or  are  intentionally  dis- 
carded. A subset  of  the  specimens  identified  during  the 
tally  period  (15  or  more  per  species,  when  possible)  is 
retained  onboard  the  vessel  and  weighed  to  determine 
an  average  weight  for  each  taxon.  That  average  weight 
is  then  applied  to  all  specimens  identified  during  the 
tally  period,  and  the  resulting  proportional  species  com- 
position is  extrapolated  to  the  total  gear  set  to  obtain 
a total  catch  weight  for  each  species  for  each  set.  The 
basic  data  unit  used  for  this  study  is  the  extrapolated 


catch  weight  for  each  taxon  from  each  observed  haul 
or  set  (hereafter  trawl  hauls  and  longline  sets  will  be 
collectively  referred  to  as  “hauls”).  The  total  observed 
skate  catch  (OSC)  was  calculated  by  summing  extrapo- 
lated catch  weights  for  all  skate  taxa  (including  the 
following  unidentified  skate  groups:  “skate  unidenti- 
fied,” “Bathyraja  sp.,”  and  “ Raja  sp.”)  across  all  hauls 
in  which  skates  were  identified.  Scientific  and  common 
names  for  skate  taxa  follow  Stevenson  et  al.  (2007). 

From  the  inception  of  the  NPGOP  through  the  sam- 
pling year  2002,  observers  were  not  trained  to  identify 
skates  and  were  therefore  not  required  to  identify  them 
beyond  the  family  level.  During  2002  and  2003,  a field 
identification  key  was  developed  (Stevenson,  2004)  and 
experienced  observers  began  receiving  training  in  skate 
identification  during  annual  briefings.  Feedback  from 
experienced  observers  was  used  to  refine  the  identifica- 
tion materials  and  classroom  training,  and  beginning 
with  the  2004  sampling  year,  all  new  and  returning 
observers  were  provided  with  skate  identification  train- 
ing and  materials  for  identification  of  skate  in  the  field. 
Since  1 January  2004  all  observers  have  been  required 
to  identify  skates  to  the  species  level  when  possible. 
Because  of  these  changes  in  observer  identification 
training  and  policies,  two  separate  but  overlapping 
time  frames  were  used  in  this  study.  To  investigate  the 
trends  in  observed  skate  catch  and  the  history  of  skate 
identification  by  observers  an  11-year  time  frame  was 
chosen  and  queries  were  restricted  to  data  collected 
from  1 January  1998  through  31  December  2008.  For 
investigations  of  species-level  trends  in  observer  data, 
queries  were  restricted  to  data  collected  from  1 Janu- 
ary 2004  through  31  December  2008 — a period  that 
corresponds  with  the  time  period  in  which  all  new  and 
returning  observers  have  been  trained  to  identify  skates 
to  the  species  level.  Regions  were  defined  on  the  basis 
of  NMFS  management  areas:  Bering  Sea  comprises  the 
Bering  Sea  NMFS  management  areas  509-524;  the 
Aleutian  Islands  region  comprises  NMFS  management 
areas  541-543;  and  the  Gulf  of  Alaska  comprises  NMFS 
management  areas  610-650  (Fig.  1).  All  catch  propor- 
tions are  presented  as  a percentage  of  total  observer 
reported  catch  weight. 

The  targeted  resource  was  not  directly  recorded  in  ob- 
server catch  data,  so  that  for  the  purposes  of  this  study, 
the  term  “target  species”  is  defined  as  the  predominant 
species  in  the  catch.  “Predominant  species”  was  defined 
as  the  species  accounting  for  the  highest  percentage 
of  the  extrapolated  weight  in  the  species  composition 
sample  and  was  determined  on  a haul-by-haul  basis. 
Percent  retained  data  are  subjective  estimates  made 
by  observers  using  visual  approximations,  along  with 
information  provided  by  the  vessel’s  captain  or  factory 
manager.  Mean  retention  rates  used  here  are  weighted 
averages  calculated  annually  for  each  species  with  the 
following  equation: 

xaa 
Xa  ' 


210 


Fishery  Bulletin  108(2) 


150°E  160°E  170°E  180°  170°W  160°W  150°W  140°W  130°W  120°W 


0 250  500  1,000  km 

1  l L l I 


, 1 1 1 1 

180°  170°W  160°W  150°W  140°W 

Figure  1 

Map  showing  NMFS  management  areas  in  which  observed  skate  catch  was  examined  from  1998  through  2008. 
Stippled  areas  = Bering  Sea,  shaded  areas=Aleutian  Islands,  diagonal  hatching=Gulf  of  Alaska. 


where  i?;/  = the  observer  reported  retention  rate  of  spe- 
cies i in  haul  j\  and 

Ci;  = the  extrapolated  catch  weight  of  species  i 
in  haul  j. 

Historical  skate  price  information  was  derived  from 
Alaska  state  fish-ticket  data,  and  was  compiled  for  the 
study  period  by  Terry  Hiatt  (unpubl.  data1).  An  annual 
mean  price  was  determined  for  each  taxon  by  1)  calcu- 
lating the  exvessel  price  paid  per  pound  round  weight 
at  each  delivery  to  all  processors  where  the  purchase 
of  raw  skates  from  Alaska  waters  was  recorded,  and 
then  2)  calculating  the  simple  average  of  those  delivery 
price  points  over  the  calendar  year.  Round  weight  refers 
to  intact  whole  specimens.  For  deliveries  consisting  of 
nonwhole  specimens,  round  weight  (in  pounds)  was 
calculated  from  net  delivery  weight  by  using  a product 


1 Hiatt,  T.  2009.  NMFS  Alaska  Fisheries  Science  Center, 
Seattle,  WA  98115. 


recovery  rate  (PRR)  of  0.32  for  “wings”  and  0.9  for 
gutted  animals  (National  Marine  Fisheries  Service, 
http://www.fakr.noaa.gov/rr/tables/tabl3.pdf,  accessed 
November  2009).  Each  annual  mean  represents  at  least 
334  (range:  334-2247)  data  points. 

Results 

Skate  species  composition  reported  by  observers  over 
the  past  decade  has  changed  considerably.  Up  to  and 
including  2002,  over  98%  of  OSC  was  reported  as  “skate 
unidentified”  (Table  1).  In  2003,  less  than  90%  of  OSC 
was  unidentified,  and  the  proportion  of  unidentified 
skates  has  continued  to  drop  through  2008,  a year  in 
which  only  2%  of  OSC  was  unidentified.  Because  the 
proportion  of  unidentified  skates  has  dropped,  the  pro- 
portions of  skates  identified  to  the  genus  level  ( Bathy - 
raja)  and  to  the  species  level  (Bathyraja  parmifera.  Raja 
binoculata,  etc.)  have  continued  to  rise.  In  2008,  46%  of 
OSC  was  identified  to  the  genus  level  and  approximately 


Stevenson  and  Lewis:  Skate  bycatch  in  the  commercial  groundfish  fisheries  of  Alaska 


211 


Table  1 

Species  composition  (%  by  weight)  of  observed  skate  catch  by  year  reported  in  Alaska’s  groundfish  fisheries  for  1998- 
less  than  0.1%. 

-2008.  * = 

Taxon 

1998 

1999 

2000 

2001 

2002 

2003 

2004 

2005 

2006 

2007 

2008  2004-2008 

Skate  unidentified 

99.7 

99.6 

99.5 

98.6 

98.6 

88.7 

61.3 

25.2 

21.4 

7.1 

2.4 

23.2 

Bathyraja  sp. 

* 

* 

0.1 

1.1 

0.2 

0.4 

0.5 

39.3 

34.2 

42.6 

47.4 

33.1 

Bathyraja  parmifera 
(Alaska  skate) 

* 

* 

0.1 

0.2 

0.7 

7.9 

30.2 

27.2 

36.6 

40.0 

40.1 

34.8 

Bathyraja  aleutica 
(Aleutian  skate) 

* 

* 

* 

* 

0.1 

0.7 

2.2 

2.6 

1.9 

2.5 

2.7 

2.4 

Bathyraja  interrupta 
(Bering  skate) 

* 

* 

* 

* 

* 

0.3 

1.6 

1.5 

1.1 

1.3 

2.7 

1.7 

Bathyraja  maculata 
(whiteblotched  skate) 

* 

* 

* 

* 

* 

0.1 

1.1 

0.4 

0.7 

0.5 

1.0 

0.7 

Bathyraja  lindbergi 
(Commander  skate) 

* 

* 

* 

* 

* 

* 

0.1 

0.2 

0.1 

0.2 

0.2 

0.2 

Bathyraja  taranetzi 
(mud  skate) 

* 

* 

* 

* 

* 

* 

0.2 

0.1 

* 

0.1 

0.3 

0.2 

Bathyraja  trachura 
(roughtail  skate) 

* 

* 

* 

* 

* 

* 

0.1 

* 

0.1 

0.1 

* 

0.1 

Bathryaja  minispinosa 
(whitebrow  skate) 

* 

* 

* 

* 

* 

* 

* 

* 

* 

* 

* 

* 

Raja  sp. 

* 

* 

* 

* 

* 

* 

* 

0.1 

0.5 

* 

0.1 

0.1 

Raja  binoculata 
(big  skate) 

0.3 

0.4 

0.3 

* 

0.2 

1.7 

2.3 

2.3 

2.4 

3.7 

2.1 

2.5 

Raja  rhina 

(longnose  skate) 

* 

* 

* 

* 

* 

0.2 

0.5 

1.0 

1.0 

1.9 

1.0 

1.1 

52%  was  identified  to  species  (i.e.,  Bathyraja  parmifera 
and  other  species). 

The  portion  of  the  OSC  that  was  identified  to  the  spe- 
cies level  was  dominated  by  Bathyraja  parmifera,  Raja 
binoculata,  and  Bathyraja  aleutica  (Aleutian  skate), 
which  accounted  for  40.1%,  2.1%,  and  2.7%,  respectively, 
of  OSC  in  2008  (Table  1).  These  proportions  have  re- 
mained relatively  stable  since  observers  began  identi- 
fying skates  in  2004,  with  B.  parmifera,  R.  binoculata, 
and  B.  aleutica  averaging  34.8%,  2.5%,  and  2.4%,  re- 
spectively, of  the  annual  OSC  from  2004  through  2008. 
Seven  other  species  of  skates,  including  R.  rhina  and 
six  species  of  Bathyraja  ( B . interrupta,  B.  maculata,  B. 
lindbergi,  B.  taranetzi,  B.  trachura,  B.  minispinosa ), 
have  been  regularly  reported  in  smaller  proportions 
by  observers  since  2004.  Although  unidentified  skates 
now  constitute  less  than  5%  of  OSC,  a large  propor- 
tion of  skates  are  still  identified  only  to  the  genus  level 
(“ Bathyraja  sp.”  and  “ Raja  sp.”). 

The  species  composition  of  OSC  varied  by  region  and 
by  gear  type  within  each  region.  During  the  2004-08 
period,  Bathyraja  parmifera  was  the  most  commonly 
observed  species  in  both  the  Bering  Sea  and  Aleutian 
Islands  region  (Table  2).  In  the  Bering  Sea,  no  other 
single  species  made  up  more  than  1.7%  of  OSC,  and  a 
large  percentage  of  skates  were  identified  only  to  the 
genus  level.  Species  composition  profiles  were  similar 


for  both  types  of  trawl,  but  for  fisheries  using  longline 
gear  a much  higher  percentage  of  skates  were  not  iden- 
tified to  the  species  level. 

In  the  Aleutian  Islands,  B.  parmifera  again  accounted 
for  a higher  proportion  of  OSC  than  any  other  species 
(Table  2).  However,  notable  proportions  of  B.  maculata 
and  B.  aleutica  were  reported  in  this  region  as  well. 
As  in  the  Bering  Sea,  a large  proportion  of  the  skates 
were  not  identified  to  the  species  level,  and  most  of 
the  unidentified  skates  and  skates  identified  to  genus 
were  encountered  in  fisheries  using  longline  gear.  The 
species  composition  profile  for  pelagic  trawl  gear  in  the 
Aleutian  Islands,  with  only  two  species  reported  and  B. 
interrupta  accounting  for  over  80%  of  OSC,  was  mark- 
edly different  from  any  of  the  other  region-gear  combi- 
nations reported  in  our  study.  However,  that  profile  was 
based  on  only  two  species  composition  samples  in  which 
skates  were  reported. 

The  species  composition  of  OSC  was  quite  different 
in  the  Gulf  of  Alaska,  where  the  two  species  of  Raja  (R. 
binoculata  and  R.  rhina)  are  more  common,  accounting 
for  over  half  of  OSC  in  the  region  (Table  2).  Among 
species  of  Bathyraja,  B.  aleutica  accounted  for  the  high- 
est proportion  in  the  Gulf  of  Alaska.  The  proportion  of 
skates  not  identified  to  the  species  level  was  consider- 
ably lower  in  the  Gulf  of  Alaska  than  in  either  the  Ber- 
ing Sea  or  Aleutian  Islands,  and  the  species  composition 


212 


Fishery  Bulletin  108(2) 


Table  2 

Species  composition  (%  by  weight)  of  observed  skate  catch  by  region  and  by  gear  type  within  each  region  of  Alaska  for  2004-2008. 
Regions:  BS=Bering  Sea,  AI=Aleutian  Islands,  GOA=Gulf  of  Alaska.  Gear  types:  l=Nonpelagic  trawl,  2=Pelagic  trawl,  3=Long- 
line.  * = less  than  0.1%. 

Taxon 

BS 

AI 

GOA 

1 

2 

3 

All 

1 

2 

3 

All 

1 

2 

3 

All 

Skate  unidentified 

0.6 

1.4 

33.5 

24.2 

3.9 

* 

25.1 

17.6 

2.4 

3.5 

16.7 

9.6 

Bathyraja  sp. 

1.9 

1.5 

47.5 

34.5 

4.7 

* 

40.7 

28.0 

2.2 

3.3 

16.5 

9.4 

Bathyraja  parmifera 

90.0 

92.8 

14.9 

36.6 

50.9 

* 

18.0 

29.6 

3.6 

9.4 

2.8 

3.2 

Bathyraja  aleutica 

3.0 

2.4 

1.2 

1.7 

16.1 

* 

3.7 

8.0 

9.5 

9.8 

13.7 

11.6 

Bathyraja  interrupta 

1.3 

1.3 

1.7 

1.6 

1.1 

80.6 

0.4 

0.7 

5.9 

6.3 

3.0 

4.4 

Bathyraja  minispinosa 

0.1 

* 

* 

* 

0.1 

* 

0.1 

0.1 

* 

* 

* 

* 

Bathyraja  maculata 

0.4 

* 

0.2 

0.2 

17.2 

* 

9.3 

12.1 

* 

* 

0.2 

0.1 

Bathyraja  lindbergi 

* 

* 

0.1 

0.1 

0.2 

* 

1.7 

1.1 

* 

* 

0.2 

0.1 

Bathyraja  taranetzi 

0.2 

0.1 

* 

0.1 

5.2 

19.4 

0.6 

2.2 

* 

* 

* 

* 

Bathyraja  trachura 

* 

* 

* 

* 

0.1 

* 

0.2 

0.2 

* 

* 

1.4 

0.7 

Raja  sp. 

* 

* 

* 

* 

* 

* 

* 

* 

0.6 

0.3 

4.6 

2.6 

Raja  binoculata 

2.4 

0.4 

0.7 

1.0 

0.4 

* 

0.2 

0.3 

52.3 

24.8 

19.6 

35.7 

Raja  rhina 

0.1 

* 

* 

* 

* 

* 

* 

* 

23.4 

42.6 

21.3 

22.4 

Total 

100 

100 

100 

100 

100 

100 

100 

100 

100 

100 

100 

100 

profiles  varied  more  by  gear  type  than  in  the  other  two 
regions.  All  three  gear  types  were  dominated  by  species 
of  Raja,  but  R.  binoculata  accounted  for  over  50%  of 
OSC  from  nonpelagic  trawl  gear,  whereas  R.  rhina  was 
the  dominant  species  in  pelagic  trawl  and  longline  gear. 
As  in  the  other  two  regions,  proportions  of  unidentified 
skates  were  much  higher  on  longliners  than  on  vessels 
with  other  gear  types,  although  a much  higher  percent- 
age of  skates  were  identified  to  the  species  level  even 
with  longline  gear  in  the  Gulf  of  Alaska. 

Significant  amounts  of  skate  bycatch  were  encoun- 
tered by  observers  in  fisheries  targeting  a variety  of 
commercial  groundfish  species,  including  Pacific  cod, 
walleye  pollock,  Atka  mackerel,  shallow-water  flatfishes 
(primarily  yellowfin  and  rock  soles),  and  others.  During 
the  1998-2008  study  period,  nearly  72%  of  OSC  was  re- 
ported in  longline  fisheries,  and  over  65%  was  reported 
in  longline  hauls  targeting  Pacific  cod  (Table  3).  Non- 
pelagic trawl  fisheries  accounted  for  only  22%  of  OSC, 
most  of  which  was  reported  in  hauls  targeting  miscel- 
laneous flatfishes.  Pelagic  trawl  fisheries,  essentially  all 
of  which  target  walleye  pollock,  accounted  for  very  little 
of  OSC  (6%).  These  results  reflect  the  percentages  for 
the  Bering  Sea,  a region  in  which  over  90%  of  OSC  was 
reported.  In  the  Aleutian  Islands  significant  numbers 
of  skates  were  also  encountered  on  trawlers  targeting 
Atka  mackerel,  and  in  the  Gulf  of  Alaska  on  trawlers 
targeting  deepwater  flatfishes  (arrowtooth  flounder  and 
Greenland  turbot). 

The  percentage  of  OSC  retained  by  commercial  fish- 
ermen has  increased  over  the  past  decade  (Fig.  2).  In 
1998,  overall  mean  skate  retention  was  just  over  12%, 


and  that  figure  steadily  increased  to  a peak  of  nearly 
40%  in  2003.  For  the  most  recent  4 years  (2005-08) 
overall  skate  retention  has  remained  relatively  consis- 
tent at  around  30-35%.  Species-level  retention  data 
were  erratic  from  1998  through  2003.  They  have  be- 
come more  stable  since  2004  when  observers  began 
consistently  identifying  skates  to  the  species  level,  but 
the  annual  mean  retention  for  some  of  the  species, 
particularly  the  genus  Raja,  still  appears  relatively 
inconsistent  from  year  to  year.  Since  2004,  the  largest 
species  of  skates  ( Raja  binoculata,  R.  rhina,  Bathyraja 
parmifera,  B.  aleutica,  and  B.  maculata)  have  generally 
been  retained  at  30%  of  OSC  or  above,  and  smaller  spe- 
cies, such  as  B.  interrupta,  B.  lindbergi,  B.  taranetzi, 
and  B.  minispinosa,  have  been  retained  at  lower  levels 
(5-15%). 

Discussion 

From  the  inception  of  the  NPGOP  through  2003,  field 
identification  tools  for  the  skates  of  Alaska  were  limited, 
and  skate  bycatch  data  were  collected  at  a very  basic 
level.  Almost  all  skates  were  reported  by  observers  as 
“skate  unidentified.”  However,  from  2004  through  2008 
this  situation  changed  rapidly.  With  the  development 
and  deployment  of  a field  guide  and  the  implementation 
of  an  observer  training  protocol  (Stevenson,  2004),  the 
proportion  of  skates  identified  to  the  species  level  has 
increased  dramatically.  For  the  last  year  included  in  this 
study,  over  95%  of  OSC  was  identified  at  least  to  genus, 
and  that  proportion  may  continue  to  rise  in  future  years 


Stevenson  and  Lewis:  Skate  bycatch  in  the  commercial  groundfish  fisheries  of  Alaska 


213 


Table  3 

Observed  skate  catch  (in  tons)  by  region,  gear  type,  and  target  species  reported  in  Alaska’s  groundfish  fisheries  for  1998-2008. 
Target  species  is  defined  as  the  predominant  species  (by  % weight)  in  the  catch.  * = less  than  100  tons. 


Region 
Gear  type 


Target  species 


>*/ 


\-y 

cjT  „ 

A/  / 

& 


$ 


.21 


sfi 


<?'>" 


A 


0° 


^v 


<t%o" 

V 


:> 


o & cy 

■ %•*  %y' 


Other 

Total 

1367 

26,675 

* 

8967 

2714 

94,766 

* 

2067 

* 

* 

793 

5597 

448 

3445 

* 

* 

286 

2566 

1823 

32,187 

* 

9061 

3794 

102,929 

5630 

144,177 

Bering  Sea 


Nonpelagic  trawl 

2606 

18,400 

2476 

1677 

* 

Pelagic  trawl 

* 

* 

8912 

* 

* 

Longline 

90,314 

* 

206 

967 

461 

Aleutian  Islands 

Nonpelagic  trawl 

491 

* 

* 

* 

* 

Pelagic  trawl 

* 

* 

* 

* 

* 

Longline 

3850 

* 

* 

256 

497 

Gulf  of  Alaska 

Nonpelagic  trawl 

437 

756 

* 

1393 

* 

Pelagic  trawl 

* 

* 

* 

* 

* 

Longline 

1486 

* 

* 

* 

443 

All  areas 

Nonpelagic  trawl 

3534 

19,177 

2563 

3123 

101 

Pelagic  trawl 

* 

* 

8967 

* 

* 

Longline 

95,650 

* 

208 

1229 

1401 

Total 

99,222 

19,253 

11,738 

4366 

1504 

1021 

* 


1072 


1077 


443 


262 


744 

* 

* 

843 


* 

140 


319 


493 

544 


as  training  methods  and  identification  tools 
are  further  refined. 

Patterns  of  species  composition  in  OSC 
generally  parallel  recent  biomass  estimates 
for  regional  skate  populations  derived  from 
bottom  trawl  surveys.  Bathyraja  parmifera 
accounts  for  the  large  majority  of  OSC, 
which  is  not  surprising  given  that  B.  par- 
mifera is  the  most  abundant  species  of 
skate  encountered  in  bottom  trawl  surveys 
conducted  in  Alaska  waters  (Stevenson  et 
al.,  2008).  In  fact,  B.  parmifera  is  particu- 
larly common  on  the  Bering  Sea  continental 
shelf,  where  its  populations  make  up  about 
95%  of  the  total  skate  biomass  (Acuna  and 
Lauth,  2008;  Lauth  and  Acuna,  2009)  and 
where  commercial  fishing  effort  for  wall- 
eye pollock,  Pacific  cod,  and  flatfishes  is 
concentrated.  Many  of  the  other  species 
encountered  by  observers  in  the  Bering  Sea 
are  recorded  from  fishing  activity  on  the 
upper  continental  slope,  where  B.  aleutica, 
B.  maculata,  and  B.  interrupts  populations 


100 


All  skates 
q R.  binoculata 
o-  B.  rhina 
• — B.  parmifera 
« — B.  aleutica 
+ — B.  maculata 


1998  1999  2000  2001  2002  2003  2004  2005  2006  2007  2008 

Year 

Figure  2 

Overall  mean  percent  retention  of  skate  catch  in  commercial  fisheries 
for  each  year  from  1998  through  2008  (gray  bars),  as  well  as  mean 
percent  retention  for  Raja  binoculata , R.  rhina , Bathyraja  parmifera, 
B.  aleutica,  and  B.  maculata. 


214 


Fishery  Bulletin  108(2) 


are  concentrated  (Hoff  and  Britt,  2003,  2005,  2009; 
Stevenson  et  al.,  2008). 

In  the  Aleutian  Islands,  over  50%  of  OSC  consists 
of  B.  parmifera,  B.  maculata,  and  B.  aleutica  (Table 
2),  which  are  the  top  three  species  in  terms  of  recent 
biomass  estimates  for  the  region  (Zenger,  2004;  Rooper, 
2008;  Rooper  and  Wilkins,  2008).  However,  the  pro- 
portion of  B.  parmifera  is  higher  (29.6%  of  observed 
skate  catch)  and  that  of  B.  maculata  considerably  lower 
(12.1%)  in  commercial  catches  in  the  Aleutian  Islands 
than  their  biomass  estimates  in  the  region  (20-25% 
and  48%  of  total  skate  biomass,  respectively)  would 
indicate.  The  reasons  for  these  differences  in  relative 
catch  weight  are  unclear,  but  may  be  due  to  geographi- 
cally and  bathymetrically  concentrated  commercial  fish- 
ing effort.  Skate  populations  in  Alaska  are  primarily 
segregated  by  depth,  and  B.  maculata  tends  to  be  found 
in  deeper  waters  than  those  inhabited  by  B.  parmifera 
(Rooper,  2008;  Stevenson  et  al.,  2008).  Therefore,  shal- 
low-water fisheries  are  more  likely  to  catch  B.  par- 
mifera, and  although  observers  reported  skates  in  the 
Aleutian  Islands  from  depths  to  2000  m,  the  majority  of 
OSC  came  from  200  m or  less.  Thus,  Aleutian  popula- 
tions of  B.  parmifera  may  be  disproportionately  affected 
by  fishing  activity  because  of  the  shallow  depth  distri- 
bution of  this  species. 

The  two  species  of  Raja  (and  unidentified  Raja — “ Raja 
sp.”)  account  for  over  60%  of  OSC  in  the  Gulf  of  Alaska. 
These  results  are  also  consistent  with  fishery-indepen- 
dent  survey  data,  which  indicate  that  Raja  binoculata 
and  R.  rhina  are  the  most  abundant  species  in  the 
Gulf  of  Alaska,  making  up  about  37%  and  33%,  respec- 
tively, of  the  skate  biomass  in  the  region  (Stevenson  et 
al.,  2008;  von  Szalay  et  al.,  2009).  Among  species  of 
Bathyraja  in  the  Gulf  of  Alaska,  survey-derived  biomass 
estimates  indicate  that  B.  aleutica  is  the  most  common, 
and  indeed  B.  aleutica  accounts  for  a greater  proportion 
of  OSC  in  this  region  than  all  other  species  of  Bathyraja 
combined. 

Deepwater  skate  species,  such  as  B.  lindbergi,  B. 
minispinosa,  and  B.  trachura,  are  rarely  reported  by 
observers  in  any  of  the  three  regions,  probably  due  to 
the  relatively  small  amount  of  fishing  effort  targeting 
deepwater  species.  Other  species  known  to  be  rare  in 
Alaska  waters,  such  as  B.  abyssicola,  B.  mariposa,  and 
Amblyraja  badia,  have  been  only  rarely  reported  by 
observers,  and  only  B.  mariposa  has  been  confirmed  by 
photographs  and  collected  specimens. 

Although  the  percentage  of  unidentified  skates  in 
observer  species  composition  data  has  declined  to  a very 
low  level,  a large  percentage  of  OSC  is  still  identified 
only  to  genus.  These  less  specific  skate  identifications 
are  largely  the  result  of  uncertainty  with  identification 
in  the  field.  Because  observers  encounter  a relatively 
high  diversity  of  skates,  particularly  of  the  genus  Bathy- 
raja, and  must  often  interpret  subtle  characteristics  to 
identify  skates  to  the  species  level,  they  are  encour- 
aged to  identify  a skate  only  to  the  genus  level  if  the 
specimen  is  not  brought  to  hand  for  inspection  or  if 
the  identification  of  the  specimen  is  questionable.  As  a 


result,  species  composition  of  OSC  is  clearly  affected  by 
fish-handling  practices  and  observer  sampling  methods 
on  vessels  with  different  gear  types.  Observers  in  trawl 
fisheries  select  their  species  composition  samples  at 
random  from  the  catch  after  it  is  onboard  the  vessel. 
Therefore,  the  entire  composition  sample  is  weighed, 
and  all  specimens  in  the  composition  sample  are  iden- 
tified in  hand.  In  contrast,  on  longline  vessels  species 
composition  data  are  collected  as  the  gear  is  being  re- 
trieved, and  not  all  of  the  specimens  in  the  composi- 
tion sample  are  brought  on  board  and  weighed.  Some 
specimens  counted  during  the  tally  period,  particularly 
larger  species  such  as  many  of  the  skates  common  in 
Alaska  waters,  become  “drop-offs.”  These  specimens  are 
retrieved  to  the  surface  on  the  line  but  either  fall  off 
before  they  can  be  brought  onboard  or  are  intentionally 
released  to  save  strain  on  the  gear,  the  personnel,  and 
the  fishes.  Therefore,  many  of  the  skates  in  the  compo- 
sition sample  from  longline  vessels  are  not  brought  to 
hand  for  identification,  and  are  recorded  at  the  genus 
level.  Thus,  the  way  the  catch  is  handled  and  sampled 
in  longline  fisheries  largely  explain  the  influence  of 
gear  type  on  the  species  composition  profiles  reported 
here  (Table  2). 

The  influence  of  longline  data  is  significant  because 
the  majority  of  OSC  in  Alaska  waters  comes  from  long- 
liners.  In  fact,  the  data  presented  here  (Table  3)  indi- 
cate that  the  longline  fishery  for  Pacific  cod  in  the  Ber- 
ing Sea  accounts  for  more  skate  bycatch  than  all  other 
federally  managed  groundfish  fisheries  combined.  This 
result  must  be  interpreted  with  some  caution  because 
differences  in  observer  coverage  for  different  fisheries 
and  regions  may  have  influenced  these  figures,  and 
a predominant  species  is  not  a precise  indicator  of  a 
target  fishery.  But  it  is  clear  that  longliners  targeting 
Pacific  cod  catch  a lot  of  skates.  Moreover,  longline  gear 
is  often  fished  deeper  than  trawl  gear,  and  therefore 
may  affect  a greater  diversity  of  skate  species  than  gear 
fished  in  more  shallow  water  because  skate  diversity  in 
Alaska  waters  tends  to  be  highest  on  the  continental 
slope  (Stevenson  et  al.,  2008).  Therefore,  as  long  as  a 
high  proportion  of  skates  encountered  on  longliners  are 
identified  only  to  genus,  a potentially  important  seg- 
ment of  species-specific  catch  data  is  still  not  available 
for  analysis. 

The  presence  of  skates  in  the  catch  of  pelagic  trawls 
may  seem  counterintuitive  because  skates  are  generally 
benthic,  substrate-oriented  fishes  unlikely  to  be  found  in 
the  path  of  midwater  nets.  Indeed,  the  amount  of  skate 
catch  reported  in  pelagic  trawls  (about  6%  of  OSC)  is 
much  lower  than  in  the  other  two  gear  types.  There 
are  two  general  explanations  for  the  skates  that  are 
collected  in  pelagic  nets:  either  the  skates  were  swim- 
ming up  in  the  water  column  or  the  net  contacted  the 
seafloor.  The  target  of  most  pelagic  trawling  in  Alaska 
is  walleye  pollock,  a species  that  is  often  found  very 
close  to  the  bottom,  and  catch  data  from  pelagic  trawl- 
ers often  include  a variety  of  benthic  species,  such  as 
flatfishes  and  sculpins,  in  addition  to  skates.  Therefore, 
it  is  likely  that  at  least  a large  proportion  of  the  skate 


Stevenson  and  Lewis:  Skate  bycatch  in  the  commercial  groundfish  fisheries  of  Alaska 


215 


Figure  3 

Annual  mean  exvessel  price  paid  by  processors  in  Alaska  for  big  skate 
( Raja  binoculata),  longnose  skate  ( Raja  rhina),  and  miscellaneous  skates 
from  1998  through  2007. 


catch  in  pelagic  trawls  is  the  result  of  the  net  contact- 
ing, or  at  least  coming  very  close  to,  the  seafloor. 

Historically,  skates  have  not  been  considered  valuable 
by  Alaska’s  commercial  fishermen.  Even  though  skates 
are  large  fishes  that  represent  a significant  potential 
source  of  protein,  retention  of  skates  in  the  commercial 
fisheries  of  Alaska  has  been  low.  However,  groundfish 
observer  data,  coupled  with  exvessel  pricing  informa- 
tion, may  indicate  that  this  situation  is  beginning  to 
change.  Overall  mean  retention  was  less  than  15%  in 
the  late  1990s,  and  presumably  before  that  time  as  well; 
however,  it  has  increased  to  30-35%  in  recent  years. 
Species-level  catch  data  collected  since  2004  indicate 
that  the  large  species  (such  as  both  species  of  Raja, 
Bathyraja  parmifera , and  B.  aleutica)  are  retained  at 
a higher  rate  than  smaller  species,  and  that  retention 
rates  for  the  large  species  are  not  necessarily  consistent 
from  year  to  year.  The  general  increase  in  retention 
rates  may  reflect  changes  in  the  market  value  for  skate 
products.  Although  the  mean  exvessel  price  for  general 
skate  catch  has  remained  fairly  stable  over  the  past 
decade  (Fig.  3),  the  price  paid  to  Alaskan  fishermen 
for  big  skates  and  longnose  skates  has  risen  sharply. 
Since  2004,  when  processors  began  reporting  landings 
data  by  species  owing  to  changes  in  the  Fishery  Man- 
agement Plan  for  groundfish  of  the  Gulf  of  Alaska,  the 
mean  annual  price  paid  for  big  and  longnose  skates  has 
nearly  tripled. 

Although  the  data  presented  here  signify  a dramatic 
improvement  in  the  information  available  to  fishery 
managers,  some  noteworthy  gaps  persist.  The  data  pre- 
sented here  represent  only  sampled  hauls  on  vessels 
requiring  observer  coverage  in  federally  managed  fish- 
eries, and  therefore  other  sources  of  skate  bycatch  are 
not  represented.  Commercial  fishing  activity  in  the  Ber- 
ing Sea  and  Aleutian  Islands  is  conducted  primarily  on 
large  vessels,  which  are  required  to  have  100%  observer 
coverage,  and  therefore  observer  data  should  provide  a 


good  representation  of  skate  bycatch  in  those  regions. 
In  contrast,  many  of  the  commercial  vessels  operating 
in  the  Gulf  of  Alaska  are  small  enough  that  observer 
coverage  is  only  required  on  30%  of  fishing  days  or  is 
not  required  at  all.  Therefore,  observer  data  for  this 
region  may  provide  much  less  reliable  estimates  of  skate 
bycatch.  Because  the  two  species  of  the  genus  Raja  are 
common  in  the  Gulf  of  Alaska,  and  are  among  the  larg- 
est skate  species  in  the  region,  the  unobserved  catch  of 
those  species  is  of  particular  concern.  Disproportionate 
retention  of  larger  skates  is  prevalent  in  many  fisheries 
worldwide,  and  as  larger,  more  vulnerable  species  are 
removed,  smaller  species  may  become  more  abundant 
(Russ,  1991;  Agnew  et  al.,  2000;  Cedrola  et  al.,  2005; 
Swain  et  al.,  2005).  In  the  North  Atlantic,  severe  reduc- 
tion in  biomass  for  some  larger,  less  resilient  skate  spe- 
cies has  been  accompanied  by  an  increased  biomass  for 
smaller,  more  resilient  species  (Casey  and  Myers,  1998; 
Walker  and  Hislop,  1998;  Dulvy  et  al.,  2000).  Species- 
specific  observer  data  on  skate  bycatch  can  document 
this  phenomenon,  but  only  if  the  data  are  representa- 
tive of  total  fishing  effort.  Therefore,  undocumented 
sources  of  skate  bycatch,  as  well  as  nonspecific  data 
from  observed  longline  fisheries  (see  above  comments  on 
longline  species  composition  data),  present  significant 
remaining  challenges  to  fishery  managers. 

Observer  data  on  skate  bycatch  in  the  groundfish  fish- 
eries of  Alaska  represent  a rich  source  of  information 
for  managers  charged  with  protecting  skate  populations 
from  future  overexploitation.  The  species-level  catch 
data  now  being  collected  by  observers  have  facilitated 
the  development  of  an  age-structured  stock  assessment 
model  for  B.  parmifera  (B.  Matta,  personal  commun.2), 
which  is  a critical  aid  in  setting  appropriate  catch  lim- 
its for  the  species,  and  similar  models  for  other  species 


2 Matta,  Beth.  2009.  NMFS  Alaska  Fisheries  Science  Center, 
Seattle,  WA  98115. 


216 


Fishery  Bulletin  108(2) 


are  on  the  horizon.  These  fishery-dependent  data  can 
now  be  compared  directly  with  fishery-independent  sur- 
vey data,  creating  two  independent  lines  of  evidence  for 
management  strategies.  Specific  catch  data  may  also 
be  used  to  identify  areas  in  which  the  most  vulnerable 
species  may  be  most  heavily  impacted  and  thus  can 
help  identify  areas  in  which  restrictions  or  closures 
are  necessary.  Although  observer  data  do  not  give  a 
complete  account  of  skate  bycatch  in  the  fisheries  of 
Alaska,  the  information  currently  provided  allows  this 
diverse  assemblage  of  species  to  be  managed  in  a more 
biologically  appropriate  way  than  was  possible  in  the 
past.  As  fishing  pressure  on  Alaska’s  skate  populations 
increases,  the  consequences  of  data  deficiencies  will  be 
magnified,  and  observer  data  will  play  an  increasingly 
important  role  in  protecting  skates  from  the  declines  in 
biomass  and  shifts  in  community  structure  that  have 
befallen  these  fishes  in  other  parts  of  the  world. 

Acknowledgments 

We  thank  the  multitude  of  staff  and  observers  of  the 
North  Pacific  Groundfish  Observer  Program  that  have 
helped  to  collect  the  data  used  here.  We  also  thank  S. 
Gaichas,  O.  Ormseth,  and  B.  Matta  for  discussions  about 
skate  stock  assessments,  R.  Narita  for  assistance  with 
data  retrieval,  and  T.  Hiatt  for  providing  skate  price 
information.  For  comments  on  an  early  draft  of  the 
manuscript,  we  thank  M.  Loefflad,  B.  Mason,  B.  Matta, 
P.  Nelson,  O.  Ormseth,  and  J.  Orr. 

Literature  cited 

Acuna,  E.,  and  R.  R.  Lauth. 

2008.  Results  of  the  2007  eastern  Bering  Sea  conti- 
nental shelf  bottom  trawl  survey  of  groundfish  and 
invertebrate  resources.  NOAA  Tech.  Memo.  NMFS- 
AFSC-181,  195  p. 

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. 

Brander,  K. 

1981.  Disappearance  of  common  skate  Raia  batis  from 
Irish  Sea.  Nature  290:48-49. 

Casey,  J.  M.,  and  R.  A.  Myers. 

1998.  Near  extinction  of  a large,  widely  distributed 
fish.  Science  281:690-691. 

Cedrola,  P.  V.,  A.  M.  Gonzalez,  and  A.  D.  Pettovello. 

2005.  Bycatch  of  skates  (Elasmobranchii:  Arhyncho- 
batidae,  Rajidae)  in  the  Patagonian  red  shrimp 
fishery.  Fish.  Res.  75:141-150. 

Dulvy,  N.  K.,  J.  D.  Metcalfe,  J.  Glanville,  M.  G.  Pawson,  and 
J.  D.  Reynolds. 

2000.  Fishery  stability,  local  extinctions,  and  shifts  in  com- 
munity structure  in  skates.  Conserv.  Biol.  14:283—293. 
Ebert,  D.  A. 

2005.  Reproductive  biology  of  skates,  Bathyraja  (Ishi- 
yama),  along  the  eastern  Bering  Sea  continental 
slope.  J.  Fish  Biol.  66:618-649. 


Ebert,  D.  A.,  W.  D.  Smith,  and  G.  M.  Cailliet. 

2008.  Reproductive  biology  of  two  commercially  ex- 
ploited skates,  Raja  binoculata  and  Raja  binoculata,  in 
the  western  Gulf  of  Alaska.  Fish.  Res.  94:48-57. 

Hoff,  G.  R.,  and  L.  L.  Britt. 

2003.  The  2002  eastern  Bering  Sea  upper  continental  slope 
survey  of  groundfish  and  invertebrate  resources.  NOAA 
Tech.  Memo.  NMFS-AFSC-141,  261  p. 

2005.  Results  of  the  2004  eastern  Bering  Sea  upper 
continental  slope  survey  of  groundfish  and  inverte- 
brate resources.  NOAA  Tech.  Memo.  NMFS-AFSC-156, 
276  p. 

2009.  Results  of  the  2008  eastern  Bering  Sea  upper 
continental  slope  survey  of  groundfish  and  invertebrate 
resources.  NOAA  Tech.  Memo.  NMFS-AFSC-197, 
294  p. 

Ishihara,  H.,  and  R.  Ishiyama. 

1985.  Two  new  North  Pacific  skates  (Rajidae)  and  a 
revised  key  to  Bathyraja  in  the  area.  Jpn.  J.  Ichthyol. 
32:143-179. 

1986.  Systematics  and  distribution  of  the  skates  of  the 
North  Pacific  (Chondrichthyes,  Rajoidei).  In  Indo-Pacific 
fish  biology:  proceedings  of  the  second  international 
conference  on  Indo-Pacific  fishes  (T.  Uyeno,  R.  Aria,  T. 
Taniuchi,  and  K.  Matsuura,  eds.),  p.  269-280.  Ichthyol. 
Soc.  Japan,  Tokyo. 

Ishiyama,  R.,  and  H.  Ishihara. 

1977.  Five  new  species  of  skates  in  the  genus  Bathy- 
raja from  the  western  North  Pacific,  with  reference 
to  their  interspecific  relationships.  Jpn.  J.  Ichthyol. 
24:71-90. 

Laptikhovsky,  V.  V. 

2004.  Survival  rates  for  rays  discarded  by  the  bottom 
trawl  squid  fishery  off  the  Falkland  Islands.  Fish. 
Bull.  102:757-759. 

Lauth,  R.  R.,  and  E.  Acuna. 

2009.  Results  of  the  2008  eastern  Bering  Sea  conti- 
nental shelf  bottom  trawl  survey  of  groundfish  and 
invertebrate  resources.  NOAA  Tech.  Memo.  NMFS- 
AFSC-195,  229  p. 

Matta,  M.  E. 

2006.  Aspects  of  the  life  history  of  the  Alaska  skate, 
Bathyraja  parmifera,  in  the  eastern  Bering  Sea.  M.S. 
thesis,  92  p.  Univ.  Washington,  Seattle,  WA. 

Ormseth,  O.,  B.  Matta,  and  J.  Hoff. 

2009.  Bering  Sea  and  Aleutian  Islands  skates.  In  Stock 
assessment  and  fishery  evaluation  report  for  the  ground- 
fish resources  of  the  Bering  Sea/Aleutian  Islands  region, 
chapter  18a,  p.  1087-1178.  [Available  from  North  Pacific 
Fishery  Management  Council,  605  West  4th  Ave,  Suite 
306,  Anchorage,  AK  99501.] 

Rooper,  C.  N. 

2008.  Data  report:  2006  Aleutian  Islands  bottom  trawl 
survey.  NOAA  Tech.  Memo.  NMFS-AFSC-179,  237  p. 

Rooper,  C.  N.,  and  M.  E.  Wilkins. 

2008.  Data  report:  2004  Aleutian  Islands  bottom 
trawl  survey.  NOAA  Tech.  Memo.  NMFS-AFSC-185, 
207  p. 

Russ,  G.  R. 

1991.  Coral  reef  fisheries:  effects  and  yields.  In  The 
ecology  of  fishes  on  coral  reefs,  p.  601-635.  Academic 
Press,  San  Diego,  California. 

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  implications  for  marine 
ecosystems.  ICES  J.  Mar.  Sci.  57:476-494. 


Stevenson  and  Lewis:  Skate  bycatch  in  the  commercial  groundfish  fisheries  of  Alaska 


217 


Stevenson,  D.  E. 

2004.  Identification  of  skates,  sculpins,  and  smelts  by 
observers  in  North  Pacific  groundfish  fisheries  (2002- 
2003).  NOAA  Tech.  Memo.  NMFS-AFSC-142,  67  p. 

Stevenson,  D.  E.,  J.  W.  Orr,  G.  R.  Hoff,  and  J.  D.  McEachran. 

2004.  Bathyraja  mariposa : a new  species  of  skate  (Rajidae: 
Arhynchobatinae)  from  the  Aleutian  Islands.  Copeia 
2004:305-314. 

2007.  Field  guide  to  sharks,  skates,  and  ratfish  of  Alaska, 
77  p.  Alaska  Sea  Grant  College  Program,  Fairbanks, 
AK. 

2008.  Emerging  patterns  of  species  richness,  diversity, 
population  density,  and  distribution  in  the  skates  (Raji- 
dae) of  Alaska.  Fish.  Bull.  106:24-39. 

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. 


Swain,  D.  P.,  T.  Hurlburt,  and  H.  P.  Benoit. 

2005.  Changes  in  the  abundance  and  size  of  skates  in  the 
Southern  Gulf  of  St.  Lawrence,  1971-2002.  J.  Northw. 
Atl.  Fish.  Sci.  36:19-30. 

von  Szalay,  P.  G.,  M.  E.  Wilkins,  and  M.  M.  Martin. 

2009.  Data  report:  2007  Gulf  of  Alaska  bottom  trawl 
survey.  NOAA  Tech.  Memo.  NMFS-AFSC-189, 
247  p. 

Walker,  P.  A.,  and  J.  R.  G.  Hislop. 

1998.  Sensitive  skates  or  resilient  rays?  Spatial  and 
temporal  shifts  in  ray  species  composition  in  the  central 
and  north-western  North  Sea  between  1930  and  the 
present  day.  ICES  J.  Mar.  Sci.  55:392-402. 

Zenger,  H.  H.,  Jr. 

2004.  Data  report:  2002  Aleutian  Islands  bottom 
trawl  survey.  NOAA  Tech.  Memo.  NMFS-AFSC-143, 
247  p. 


218 


Effects  of  starvation  on  energy  density 
of  juvenile  chum  salmon  ( Oncorhynchus  keta ) 
captured  in  marine  waters 
of  Southeastern  Alaska 

Emily  A.  Fergusson  (contact  author) 

Molly  V.  Sturdevant 
Joseph  A.  Orsi 

E-mail  address  for  contact  author:  emily.fergusson@noaa.gov 

Auke  Bay  Laboratories 
Alaska  Fisheries  Science  Center 
National  Marine  Fisheries  Service 
17109  Point  Lena  Loop  Road 
Juneau,  Alaska  99801 


Abstract — We  conducted  laboratory 
starvation  experiments  on  juvenile 
chum  salmon  ( Oncorhynchus  keta) 
captured  in  the  neritic  marine  waters 
of  northern  Southeast  Alaska  in  June 
and  July  2003.  Temporal  changes  in 
fish  energy  density  (whole  body  energy 
content  [WBEC],  cal/g  dry  weight), 
percent  moisture  content,  wet  weight 
(g),  length  (mm),  and  size-related  con- 
dition residuals  were  measured  in  the 
laboratory  and  were  then  compared 
to  long-term  field  data.  Laboratory 
water  temperatures  and  salinities 
averaged  9°C  and  32  psu  in  both 
months.  Trends  in  response  variables 
were  similar  for  both  experimental 
groups,  although  sampling  intervals 
were  limited  in  July  because  fewer 
fish  were  available  (n  = 54)  than  in 
June  («  = 101).  Overall,  for  June  (45- 
d experimental  period,  9 intervals), 
WBEC,  wet  weight,  and  condition 
residuals  decreased  and  percent 
moisture  content  increased,  whereas 
fork  length  did  not  change.  For  July 
(20-d  experimental  period,  5 inter- 
vals), WBEC  and  condition  residuals 
decreased,  percent  moisture  content 
and  fork  length  increased,  and  wet 
weight  did  not  change.  WBEC,  per- 
cent moisture  content,  and  condition 
residuals  fell  outside  the  norm  of  long- 
term data  ranges  within  10-15  days 
of  starvation,  and  may  be  more  useful 
than  fork  length  and  wet  weight  for 
detecting  fish  condition  responses  to 
suboptimal  environments. 


Manuscript  submitted  19  May  2009. 
Manuscript  accepted  20  January  2010. 
Fish.  Bull.  218-225(3020). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


Energy  density  is  an  important  mea- 
sure of  fish  nutritional  condition  and 
is  used  to  assess  growth,  construct 
energy  budgets,  and  measure  energy 
flow  in  ecosystems  (Brett  et  al.,  1969; 
Jobling,  1994;  Ban  et  al.,  1996;  Edsall 
et  al.,  1999).  Energy  density  is  also 
a critical  parameter  for  bioenergetic 
models  (Orsi  et  al.,  2004;  Trudel  et  al., 
2005;  Wuenschel  et  al.,  2006;  Breck, 
2008).  Along  with  other  measures  of 
fish  condition,  such  as  body  composi- 
tion, growth,  and  length-weight  condi- 
tion indices,  energy  density  integrates 
and  reflects  the  history  of  fish  feed- 
ing environments  before  the  time  of 
sampling  (LeBrasseur,  1969;  Edsall  et 
al.,  1999;  Breck,  2008).  During  good 
feeding  periods,  fish  condition  will  be 
high,  whereas  the  reverse  is  expected 
during  poor  feeding  periods  as  energy 
reserves  are  depleted  to  maintain 
standard  metabolic  needs  (Jobling, 

1994) .  However,  an  examination  of 
how  quickly  energy  density  responds 
during  periods  of  poor  feeding  that  are 
usually  associated  with  low  growth 
has  been  limited  to  a few  studies.  In 
general,  a balanced  energy  budget  is 
expressed  as  the  equation:  ingestion 
= metabolism  + growth  + excretion, 
which  outlines  how  an  energy  source 
is  used  by  an  organism  and  what  pro- 
portion is  allocated  to  each  component 
of  the  equation  (Jobling,  1994;  Brett, 

1995) .  These  allocations  depend  on  the 
initial  amount  of  energy,  as  well  as  the 
environmental  conditions  that  affect 


physiological  rates,  such  as  tempera- 
ture and  salinity  (Brett  et  al.,  1969; 
Hoar,  1988;  Jobling,  1994).  When  fish 
are  starved,  growth  typically  ceases 
and  energy  density  declines;  when 
energy  stores  are  used,  the  percent- 
ages of  fat  and  protein  in  the  fish 
decrease  as  the  relative  water  content 
increases  (Brett,  1995;  Breck,  2008). 
Changes  in  fish  energy  density  may  be 
more  detectable  on  small  scales  than 
other  fish  parameters,  such  as  growth, 
during  periods  of  poor  feeding  condi- 
tions in  marginal  habitats. 

Juvenile  Pacific  salmon  (Oncorhyn- 
chus spp.)  use  transitional  habitats 
along  their  seaward  migration  from 
near  shore  to  the  open  ocean  and 
can  experience  rapid  environmen- 
tal changes  that  may  affect  growth 
and  energy  allocation  (Orsi  et  al., 
2000;  Cross  et  al.,  2008).  Fish  tran- 
sit these  demanding  environments  at 
the  same  time  that  they  are  experi- 
encing increasing  energy  demands 
while  undergoing  ontogenetic  changes 
in  metabolic  rate  related  to  salinity 
and  smoltification  (Hoar,  1998).  These 
transitional  habitats  are  presumed  to 
be  critical  feeding  areas  because  prey 
fields  also  change  dramatically,  and 
juvenile  salmon  are  often  found  in 
the  presence  of  planktivorous  forage 
fish  species  that  potentially  impact 
carrying  capacity  (Purcell  and  Stur- 
devant, 2001;  Park  et  al.,  2004;  Orsi 
et  al.,  2004).  Therefore,  understand- 
ing how  changes  in  juvenile  salmon 


Fergusson  et  al.:  Effects  of  starvation  on  energy  density  of  Oncorhynchus  keta 


219 


energy  density  reflect  habitat  quality  may  give  insight 
into  factors  that  affect  their  growth  and  survival,  par- 
ticularly if  food  resources  may  be  limited  during  this 
critical  time  in  their  life  history  (Paul  and  Willette, 
1997;  Boldt  and  Haldorson,  2004;  Cross  et  ah,  2008). 

We  initiated  a study  to  measure  changes  in  condition 
of  juvenile  chum  salmon  ( O . keta)  captured  at  sea  and 
later  denied  food  resources  in  the  laboratory.  In  previ- 
ous studies  on  fish  starvation,  juvenile  chum  salmon 
were  reared  entirely  in  the  laboratory  (LeBrasseur, 
1969;  Akiyama  and  Nose,  1980;  Murai  et  al.,  1983; 
Ban  et  al.,  1996);  however,  in  our  study  they  experi- 
enced variable  environmental  conditions  at  sea  before 
being  captured  and  transported  back  to  the  labora- 
tory. Thus,  these  salmon  from  field  collections  represent 
natural  variation  of  fish  in  marine  waters  better  than 
fish  reared  in  controlled  laboratory  environments.  Our 
primary  objective  was  to  measure  changes  in  energy 
density,  moisture  content,  weight,  length,  and  a size- 
related  condition  residual  index  for  field-caught  juvenile 
chum  salmon  in  response  to  starvation  in  the  labora- 
tory over  time.  We  also  compared  the  condition  of  these 
experimentally  starved  fish  to  that  determined  from  a 
long-term  data  series  on  field-caught  fish  1)  to  assess 
the  range  of  normally  occurring  condition  values  and  2) 
to  identify  the  length  of  time  before  experimental  values 
fell  outside  the  observed  range. 

Methods 

Juvenile  chum  salmon  for  the  experiments  were  cap- 
tured in  the  vicinity  of  Icy  Strait  (58°N  latitude,  135°W 
longitude)  about  50  km  west  of  Juneau,  Alaska,  in  June 
and  July  2003.  Fish  were  obtained  during  the  South- 
east Alaska  Coastal  Monitoring  (SECM)  Project  long- 
term annual  survey  of  juvenile  salmon  by  the  Auke  Bay 
Laboratories  (ABL)  aboard  the  NOAA  ship  John  N.  Cobb 
(Orsi  et  al.,  2004).  Juvenile  chum  salmon  were  collected 
from  the  neritic  waters  of  Icy  Strait  and  Upper  Chatham 
Strait,  along  the  primary  seaward  migration  corridor 
in  the  northern  region  of  Southeast  Alaska  (Orsi  et  al., 
2000,  2004).  Preliminary  observations  along  this  corridor 
showed  that  juvenile  chum  salmon  exhibit  approximately 
a five-fold  increase  in  body  length,  100-fold  increase  in 
weight,  25%  increase  in  energy  density,  and  more  than 
6%  decline  in  body  moisture  content  between  May  and 
September.  We  used  fish  from  this  locality  in  June  and 
July,  the  periods  of  highest  abundance  and  greatest 
interaction  with  other  juvenile  salmon  species.  In  June, 
fish  were  captured  with  a Kodiak  pair-trawl  fished  at  1 
m/sec  for  10  min  (Mortensen  et  al.,  2000).  In  July,  fish 
were  captured  with  a Nordic  264  rope  trawl  fished  at 
1.5  m/sec  for  20  min  (Orsi  et  al.,  2000).  All  fish  caught 
were  immediately  transferred  from  the  trawl  codend  to 
static  live  tanks  containing  sea  water.  Juvenile  chum 
salmon  were  then  identified  and  sorted  into  flow-through 
“live”  tanks.  The  sea  water  for  the  tanks  was  pumped 
from  a depth  of  3 m and  then  filtered  to  prevent  feeding 
on  zooplankton  prey.  Before  transfer  to  the  laboratory, 


the  juvenile  chum  salmon  were  held  onboard  for  one 
day  in  June  and  four  days  in  July  while  the  surveys 
were  completed.  To  establish  a baseline  for  the  start 
of  the  starvation  experiments,  on  the  day  of  capture  a 
subsample  of  fish  were  measured  (fork  length,  FL,  mm) 
and  frozen  (-5°C)  for  later  laboratory  analysis.  Daily 
temperature  and  salinity  measurements  were  recorded 
and  averaged  11.4°C  and  26.1  psu  in  June  and  12.7°C 
and  23.2  psu  in  July. 

In  the  laboratory,  the  juvenile  chum  salmon  were 
placed  in  two  living-stream  tanks  (Frigid  Units,  Inc., 
Toledo,  OH)  (200x50x48  cm)  with  screened  baffles  sepa- 
rating the  inflow  and  outflow  pipes.  One  unit  was  allo- 
cated the  salmon  captured  in  June;  the  other  unit — the 
salmon  captured  in  July.  Ambient  sea  water  from  a 
25-m  depth  in  Auke  Bay  was  supplied  to  the  tanks  at 
a rate  of  3 L/min.  Daily  temperature  and  salinity  mea- 
surements were  recorded  in  the  laboratory  tanks  and 
averaged  8.6°C  and  31.7  psu  for  June  and  8.6°C  and 
32.1  psu  for  July.  Sea  water  was  filtered  to  prevent  feed- 
ing on  zooplankton  prey.  The  fish  were  not  subjected  to 
any  strong  currents  that  would  increase  activity  costs. 
To  best  mimic  the  photoperiod  in  the  natural  environ- 
ment at  the  time  of  capture,  light  conditions  in  the  labo- 
ratory were  set  at  a standard  eight  hours  of  darkness, 
one  hour  of  dusk,  one  hour  of  dawn,  and  14  hours  of 
daylight.  Subsamples  of  10-15  fish  were  removed  from 
the  tank  at  predetermined  intervals  and  sacrificed  with 
an  overdose  of  tricaine  methanesulfonate  (MS-222), 
then  frozen  (-5°C)  individually  for  later  size  and  calo- 
rimetric analyses.  Fish  that  had  died  between  sacrifice 
intervals  were  not  included  in  the  experiments. 

Frozen  juvenile  chum  salmon  were  processed  for  data, 
including  energy  density  in  terms  of  whole  body  energy 
content  (WBEC,  cal/g  wet  weight  [WW]),  dry  weight 
(DW,  mg),  percent  moisture  content  (%<MC),  FL,  and  wet 
weight  (mg).  After  excising  each  stomach  and  removing 
and  weighing  its  contents,  we  dried  the  fish  to  obtain 
DW  (full  gut  minus  empty  gut,  nearest  mg)  so  that  un- 
digested prey  would  not  bias  the  final  values.  Stomachs 
examined  from  fish  sacrificed  after  the  first  time  inter- 
val were  devoid  of  prey  and  therefore  stomachs  were  not 
excised  in  subsequent  time  intervals.  All  viscera  were 
replaced  in  the  body  cavity  before  the  fish  were  dried  to 
a stable  weight  (<5  mg  change),  requiring  a minimum 
of  48  hours  at  55°C.  The  DW  was  recorded  and  %MC 
of  each  fish  was  calculated  as  ([1  -DW/WW]  x 100). 
Each  fish  was  homogenized  with  a Waring  pulverizer, 
then  finely  ground  with  a mortar  and  pestle  to  yield 
a uniform  powder.  Susamples  of  15  mg  were  formed 
into  pellets  with  a pellet  press  and  stored  in  a desic- 
cator to  prevent  rehydration.  A 1425  Parr  micro-bomb 
calorimeter  was  used  to  obtain  cal/g  DW  for  each  fish; 
this  measure  was  converted  to  WBEC  by  multiplying  by 
DW/WW.  Estimates  of  WBEC  from  replicate  subsamples 
were  consistent  (<2%  coefficient  of  variation).  To  ac- 
count for  potential  effects  of  size  variation  on  WBEC 
and  %MC,  size-related  condition  residuals  (CR)  were 
calculated  by  using  the  ln-transformed  experimental 
FL  and  WW  measures  for  each  fish.  We  first  derived 


220 


Fishery  Bulletin  108(2) 


a regression  equation  from  all  paired  ln-weights  and 
ln-lengths  (rc  = 8475;  -700  per  year)  of  field-caught  ju- 
venile chum  salmon  collected  during  June-August  for 
the  SECM  project  from  1997  to  2008.  We  then  used 
this  regression  equation  to  predict  ln(WW)  for  each 
experimental  ln(FL).  Finally,  we  obtained  the  CRs  by 
subtracting  the  predicted  ln(WW)  from  the  observed 
ln(WW)  (Jakob  et  al.,  1996;  Brodeur  et  al.,  2004). 

To  account  for  potential  stock-related  differences 
in  condition  of  the  experimental  chum  salmon  (of  un- 
known stocks),  WBEC  was  determined  for  additional 
field-caught  fish  of  known  stocks.  Historically,  between 
70%  and  90%  of  fish  caught  in  June  originated  from 
Macaulay  Hatchery  (MH),  whereas  mixed  hatchery 
stocks  were  present  during  July  (Orsi  et  al.,  2004). 
Otoliths  were  not  retained  from  the  fish  used  in  the 
experimental  groups;  however,  stock  of  origin  was  de- 
termined from  thermal  marks  present  on  the  otoliths 
of  juvenile  chum  salmon  captured  in  the  study  area  in 
July  and  these  marks  indicated  that  the  fish  were  from 
unmarked  stocks  (UM,  presumably  wild)  and  MH  and 


Hidden  Falls  Hatchery  (HF)  stocks.  Both  hatcheries 
mark  100%  of  chum  salmon  released.  Energy  densities 
were  determined  (as  described  above)  for  these  three 
stock  groups. 

One-way  analyses  of  variance  (ANOVA)  were  used  for 
initial  statistical  analyses  to  compare  WBEC,  %MC, 
FL,  and  WW  of  fish  across  sampling  intervals  for  each 
experimental  group  and  for  July  stock  groups.  If  sig- 
nificant differences  were  detected,  Tukey’s  paired  com- 
parison tests  were  performed  to  identify  the  interval 
in  which  they  were  found.  We  used  graphical  analyses 
to  compare  the  WBEC  and  %MC  for  each  experimen- 
tal group  to  the  norms  (one  standard  deviation  about 
the  mean)  derived  from  the  entire  SECM  field  data 
set  (1997-2008)  from  June  and  July  (n  = 1257;  WBEC: 
993.4  ±72.3  and  %MC:  79.4  ±1.2).  The  temporal  data 
from  the  experiments  were  compared  to  these  norms 
to  identify  the  duration  of  starvation  before  the  ex- 
perimental measures  fell  outside  the  long-term  range 
of  field  values. 


Days  starved 

Figure  t 

Average  whole-body  energy  content  (WBEC,  cal/g  wet  weight)  and  one 
standard  error  about  the  mean  for  juvenile  chum  salmon  ( Oncorhyn - 
chus  keta)  starved  over  time  in  the  laboratory  after  capture  in 
the  marine  waters  of  Icy  Strait  and  Upper  Chatham  Strait  in  the 
northern  region  of  southeastern  Alaska,  June  and  July  2003.  The 
grey  band  indicates  one  standard  deviation  about  the  mean  for 
all  field-caught  juvenile  chum  salmon  examined  for  WBEC  during 
the  Southeast  Coastal  Monitoring  project,  June-July  (n  = 1257), 
1997-2008.  Significant  differences  (Tukey’s  paired  comparisons; 
P<0.05)  and  percent  change  between  sample  intervals  are  shown 
in  inset  boxes. 


Results 

The  numbers  of  juvenile  chum  salmon 
obtained  for  the  two  starvation  trials  included 
101  fish  for  June  and  54  fish  for  July.  The 
higher  number  of  juvenile  chum  salmon 
available  in  June  allowed  nine  experimental 
time  intervals  to  be  tested,  spanning  45  days 
(mean  of  five  days  per  interval,  range  of  1-16 
days  between  intervals).  The  smaller  number 
of  juvenile  chum  salmon  available  in  July 
allowed  only  five  experimental  time  intervals 
to  be  tested,  spanning  20  days  (mean  of  five 
days  per  interval,  range  of  1-10  days  between 
intervals).  Both  experimental  groups  had 
common  intervals  at  about  10  and  20  days. 
Mortality  between  sampling  intervals  was 
minimal  in  both  groups:  13  fish  died  in  June 
(70%  during  the  first  10  days  of  the  experi- 
ment) and  two  died  in  July  (both  during  the 
first  2 days). 

The  energy  content  of  juvenile  chum  salm- 
on declined  over  time  in  both  experimental 
groups  (Fig.  1).  Initial  WBEC  was  significant- 
ly higher  in  June  than  in  July  (1081.2  cal/g 
WW  compared  to  960.5  cal/g  WW;  P<0.001). 
For  the  June  sample  group,  WBEC  decreased 
significantly  (P<0.001)  by  19%  between  days 
zero  and  19  and  by  40%  between  days  zero 
and  45;  see  table  insets  in  figures  for  signifi- 
cant differences  (Tukey’s  paired  comparisons) 
between  intervals.  For  the  July  sample  group, 
WBEC  decreased  significantly  (P<0.001)  by 
11%  between  days  zero  and  20.  Overall,  the 
relative  loss  of  energy  content  was  almost 
twice  as  great  in  June  as  in  July  at  day  20. 

In  contrast  to  WBEC,  %MC  of  juvenile 
chum  salmon  increased  over  time  in  both 


Fergusson  et  al.:  Effects  of  starvation  on  energy  density  of  Oncorhynchus  keta 


221 


Table  1 

Average  fork  length  (FL,  mm),  wet  weight  (WW,  g),  percent  moisture  content  (%MC,  [(1  -dry  weight/WW)x  100]),  and  whole 
body  energy  content  (WBEC,  cal/g  WW),  for  unmarked  (presumably  wild)  and  hatchery  stock  groups  of  juvenile  chum  salmon 
( Oncorhynchus  keta ) captured  in  the  marine  waters  of  Icy  Strait  and  Upper  Chatham  Strait  in  the  northern  region  of  southeast- 
ern Alaska,  July  2003.  Standard  errors  are  given  in  parentheses. 


Stock  group 

n 

FL 

WW 

%MC 

WBEC 

Unmarked 

13 

120  (1.7) 

17.5  (0.8) 

80.4  (0.1) 

954.0(5.7) 

Macaulay  Hatchery 

10 

137  (3.0) 

29.0(1.5) 

80.3  (0.2) 

957.5  (14.3) 

Hidden  Falls  Hatchery 

10 

127  (2.9) 

22.1  (1.6) 

80.4  (0.1) 

959.5  (9.6) 

experimental  groups  (Fig.  2).  Initial  %MC  was 
significantly  lower  (P<0.001)  in  June  than  in 
July  (77.8%  compared  to  80.1%).  For  the  June 
sample  group,  %MC  increased  significantly 
(P<0.001)  by  4%  between  days  zero  and  19  and 
by  9%  between  days  zero  and  45.  For  the  July 
sample  group,  %MC  increased  significantly 
(P<0.001)  by  1%  between  days  zero  and  20. 
Overall,  the  increase  in  %MC  was  four  times 
as  great  in  June  as  in  July  at  day  20. 

Changes  in  the  WW  and  FL  of  juvenile  chum 
salmon  over  time  were  not  consistent  between 
the  experimental  groups  (Fig.  3).  For  WW, 
initial  values  did  not  differ  (P>0.05)  between 
June  and  July  (14.2  compared  to  13.6  g).  For 
the  June  sample  group,  WW  decreased  signifi- 
cantly (PcO.Ol)  by  39%  between  days  zero  and 
45.  For  the  July  sample  group,  no  significant 
(P>0.05)  differences  in  WW  were  observed. 
Similarly,  initial  FL  values  did  not  differ 
(P>0.05)  between  June  and  July  (112  com- 
pared to  110  mm).  For  the  June  sample  group, 
FL  did  not  change  significantly  (P>0.05)  be- 
tween days  zero  and  45.  For  the  July  sample 
group,  FL  increased  significantly  (PcO.001)  by 
19%  between  days  zero  and  20. 

The  CR  of  juvenile  chum  salmon  became 
increasingly  negative  over  time  in  both  ex- 
perimental groups  (Fig.  4).  Initial  CRs  were 
positive  in  both  months,  but  June  CRs  were 
lower  than  those  for  July.  For  the  June  sample 
group,  CR  declined  significantly  (P<0.001)  be- 
tween days  zero  and  19  and  between  days 
zero  and  45.  For  the  July  sample  group,  CR 
declined  significantly  (P<0.001)  between  days 
zero  and  20.  Mean  CRs  shifted  from  positive 
to  negative  after  approximately  10  days  of 


Days  starved 


Figure  2 

Average  percent  moisture  content  (%MC,  [(1-dry  weight / wet 
weight)  x 100])  and  one  standard  error  about  the  mean  for  juve- 
nile chum  salmon  ( Oncorhynchus  keta)  starved  over  time  in  the 
laboratory  after  capture  in  the  marine  waters  of  Icy  Strait  and 
Upper  Chatham  Strait  in  the  northern  region  of  southeastern 
Alaska,  June  and  July  2003.  The  grey  band  indicates  one  stan- 
dard deviation  about  the  mean  for  all  field-caught  juvenile  chum 
salmon  examined  for  %MC  during  the  Southeast  Coastal  Monitoring 
project,  June-July  (n  = 1257),  1997-2008.  Significant  differences 
(Tukey’s  paired  comparisons;  P<0.05)  and  percent  change  between 
sample  intervals  are  shown  in  inset  boxes. 


starvation  in  each  sample  group  and  continued  to  de- 
cline, indicating  increasingly  poor  condition  for  a given 
size  fish. 

Hatchery  stock  group  did  not  affect  the  WBEC  or 
%MC  of  the  July-caught  juvenile  chum  salmon.  A total 
of  33  fish  were  examined:  UM  (n- 13),  MH  (n=10),  and 
HF  (n  = 10)  (Table  1).  Stock  had  no  effect  on  WBEC  or 
%MC  (P>0.05).  However,  WW  and  FL  did  differ  signifi- 


cantly (P<0.001)  among  stocks  and  were  highest  for  the 
MH  stock  and  lowest  for  the  UM  stock  (Table  1). 


Discussion 

To  our  knowledge,  this  is  the  first  published  study  of 
the  change  in  energy  density  and  %MC  of  field-captured 


222 


Fishery  Bulletin  108(2) 


Days  starved 


Days  starved 


Figure  3 

Average  fork  length  (mm,  top  panels)  and  wet  weight  (g,  bottom  panels)  for  juvenile  chum 
salmon  ( Oncorhynclius  keta)  starved  over  time  in  the  laboratory  after  capture  in  the 
marine  waters  of  Icy  Strait  and  Upper  Chatham  Strait  in  the  northern  region  of  south- 
eastern Alaska,  during  June  (left  panels)  and  July  (right  panels)  2003.  Error  bars  are 
one  standard  error  about  the  mean.  Significant  differences  (Tukey’s  paired  comparisons; 
P<0.05)  and  percent  change  between  sample  intervals  are  shown  in  inset  boxes. 


juvenile  chum  salmon  during  starvation.  Limited  infor- 
mation has  been  published  on  the  changes  in  condition  of 
laboratory-reared  chum  salmon  due  to  starvation.  Such 
studies  typically  show  depletion  of  stored  nutrients  and 
declines  in  condition  and  size  over  time,  despite  differ- 
ences in  methods  (LeBrasseur,  1969;  Akiyama  and  Nose, 
1980;  Murai  et  ah,  1983;  Ban  et  al.,  1996).  For  nutrient 
responses,  lipid  and  serum  protein  levels  of  laboratory- 
reared  juvenile  chum  salmon  were  lowest  after  10  and 
20  days  of  starvation,  respectively  (Ban  et  ah,  1996); 
unfortunately,  however,  energy  content  was  not  deter- 
mined. We  did  not  directly  measure  lipid  and  protein, 
but  the  decline  in  WBEC  that  we  observed  between  days 
zero  and  10  and  between  days  20  and  45  in  June  could 
reflect  similar  declines  in  these  nutrient  measures.  For 
condition  responses,  two  studies  showed  that  %MC  of 
small  starved  juvenile  chum  salmon  increased  by  4.3% 
(41  mm  and  0.45  g initial  size;  42-d  starvation;  LeBras- 
seur, 1969)  to  5.4%  (0.26  g initial  size;  28-d  starvation; 
Murai  et  ah,  1983)  at  ~15°C;  another  study  showed  that 
% MC  of  larger  starved  juvenile  chum  salmon  increased 
by  12%  (94.5  mm  and  7.9  g initial  size;  91-d  starvation; 
Akiyama  and  Nose,  1980)  at  17°C.  Trends  in  %MC  of 
our  juvenile  chum  salmon  were  comparable  despite  the 
differences  in  fish  size,  duration  of  starvation,  and  water 


temperature.  For  size  responses,  weight  decreased  for 
five  size-groups  of  juvenile  chum  salmon  (0.46-7.95  g 
initial  size;  5-13  wk  starvation);  however,  the  percentage 
weight  loss  decreased  as  fish  size  increased  (Akiyama 
and  Nose,  1980).  These  differences  in  weight  loss  among 
fish  sizes  indicate  that  physiological  responses  to  starva- 
tion may  vary  with  ontogeny. 

Our  results  are  also  comparable  to  information  avail- 
able for  other  salmonid  species  and  stages.  For  starved 
juvenile  sockeye  salmon  (O.  nerka),  energy  density  de- 
clined more  rapidly  and  %MC  increased  more  rapidly 
with  increasing  temperatures  (Brett  et  ah,  1969).  In  our 
study,  chum  salmon  in  June  exhibited  a 40%  decline  in 
WBEC  and  a 9%  increase  in  %MC  after  45  days  of  star- 
vation at  an  average  temperature  of  ~9°C.  By  compari- 
son, at  similar  temperatures  (10°C),  laboratory-reared 
juvenile  sockeye  salmon  lost  37%  of  initial  WBEC  and 
gained  9%  MC  during  99  days  of  starvation  (Table  3 in 
Brett  et  al.,  1969).  Such  inverse  relationships  between 
fraction  water  and  fraction  lipid  or  energy  content  are 
often  reported  during  starvation  (Miglavs  and  Jobling, 
1989;  Simpkins  et  al.,  2004;  Breck,  2008).  In  a few 
studies,  size  changes  similar  to  those  that  we  observed 
have  also  been  reported  among  other  starved  salmo- 
nids.  Weight  decreased  for  starved  juvenile  Arctic  charr 


Fergusson  et  al.:  Effects  of  starvation  on  energy  density  of  Oncorhynchus  keta 


223 


DC 

O 


0.6-1 

0.4- 

°'2 1 
0.0 

-0.2- 

-0.4- 


June 


$ 

! i 


1 i 


ra  -0.6J 


o 

O 


Figure  4 

Condition  residuals  (CR)  for  individual  juvenile  chum  salmon  (Oncorhyn- 
chus  keta)  starved  over  time  in  the  laboratory  after  capture  in  the  marine 
waters  of  Icy  Strait  and  Upper  Chatham  Strait  in  the  northern  region 
of  southeastern  Alaska,  June  and  July  2003.  The  CRs  were  calculated 
by  using  the  In-transformed  experimental  fork  length  and  wet  weight 
measures  for  each  fish  in  a regression  equation  derived  from  all  paired 
ln-weights  and  ln-lengths  of  field-caught  juvenile  chum  salmon  col- 
lected during  the  Southeast  Coastal  Monitoring  project,  June-August 
(n  = 8476)  from  1997  to  2008.  The  0.0-line  represents  the  expected  CR 
of  an  average  fish;  therefore,  positive  values  indicate  above  average 
condition  and  negative  values  indicate  below  average  condition. 


( Salvelinus  alpinus;  Miglavs  and  Jobling, 

1989),  rainbow  trout  ( O . mykiss;  Simp- 
kins et  al.,  2004),  and  Atlantic  salmon 
(Salmo  salar\  Stefansson  et  al.,  2009)  for 
starvation  periods  of  4-6  weeks.  Length 
and  weight  of  small  (30.1-mm  and  0.14-g) 
sockeye  salmon  decreased  significantly 
after  14-49  days  of  starvation  in  colder 
water  (7.9°C;  Bilton  and  Robins,  1973) 
than  that  used  in  our  experiment.  Like 
the  salmonids  in  the  above  studies, 
weight  of  our  juvenile  chum  salmon  de- 
creased for  the  June  experimental  group, 
but  similar  conclusions  about  the  July 
fish  could  not  be  reported  because  of  the 
shorter  experimental  period. 

The  chum  salmon  caught  in  June  ini- 
tially had  approximately  11%  higher 
WBEC  and  approximately  3%  lower  %MC 
than  fish  caught  in  July — differences  that 
could  be  accounted  for  by  both  environ- 
mental and  biological  variables.  In  both 
the  June  and  July  experimental  groups, 
a measurable  increase  in  WBEC  and  de- 
crease in  %MC  occurred  between  days 
zero  and  one.  These  changes  may  have 
been  attributed  to  a physiological  stress 
response  that  caused  the  fish  to  lose  wa- 
ter and  therefore  increased  the  relative 
WBEC  and  decreased  the  %MC  (Breck, 

2008).  Temperature  and  salinity  both  af- 
fect fish  physiological  rates  and  influence 
ingestion,  metabolism,  and  growth  (Brett 
et  al.,  1969;  Mason,  1974;  Sheridan  et  al., 

1983;  Jobling,  1994;  Weatherley  and  Gill, 

1995).  In  our  study,  field  temperature 
was  cooler  and  salinity  was  higher  in 
June  (11°C;  26  psu)  than  in  July  (13°C; 

23  psu),  but  fish  captured  in  both  months 
were  transferred  into  identical,  colder 
(9°C)  and  more  saline  (32  psu)  environments  in  the 
laboratory.  Monthly  differences  in  temperature  and 
salinity  were  therefore  eliminated  as  variables  in  the 
experiments.  However,  the  fish  captured  in  June  had 
probably  smolted  more  recently  (Zaporozhec  and  Za- 
porozhec,  1993;  Hoar,  1998)  and  spent  less  time  in  the 
marine  environment,  and  probably  had  lower  growth 
rates  (Orsi  et  al.,  2000)  and  energy  requirements  than 
fish  captured  in  July,  when  it  was  warmer. 

We  accounted  for  potential  size-related  effects  on 
WBEC  and  %MC  by  using  length-weight  regression 
analysis,  which  corrected  for  natural  variation  in  fish 
size;  however,  the  results  may  still  be  misleading  be- 
cause this  regression  did  not  account  for  differences  in 
actual  nutritional  status  or  body  composition,  such  as 
protein,  lipid,  and  water  content  (Miglavs  and  Jobling, 
1989;  Edsall  et  al.,  1999;  Kotiaho,  1999;  Trudel  et  al., 
2005;  Congleton  and  Wagner,  2006).  Length-weight  re- 
gression analysis  is  useful  for  initially  identifying  con- 
dition in  relation  to  a long-term  index  and  to  anticipate 


trends  in  energy  density,  but  to  account  for  changes  in 
nutritional  status  or  body  composition  WBEC,  %MC, 
or  proximate  composition,  should  be  used  to  verify  the 
CR  results. 

In  our  study,  stocks  of  juvenile  chum  salmon  sampled 
from  the  same  habitat  did  not  differ  in  WBEC  or  %MC, 
but  size  did  differ  significantly.  By  comparison,  for  ju- 
venile pink  salmon  (O.  gorbuscha ) captured  together 
in  marine  habitats  of  Prince  William  Sound,  Alaska, 
differences  in  length  and  WBEC  between  stock  groups 
have  not  been  consistent  (Paul  and  Willette,  1997;  Boldt 
and  Haldorson,  2004;  Cross  et  al.,  2008).  For  fish  ~80 
mm  in  length,  the  occurrence  of  length  differences  be- 
tween juvenile  pink  salmon  stocks  depended  on  the  size 
of  hatchery  fish  at  time  of  release  (Cross  et  al.,  2008). 
In  a concurrent  study,  juvenile  pink  salmon  length  dif- 
fered between  stock  groups,  but  WBEC  did  not  (Boldt 
and  Haldorson,  2004).  Conversely,  energy  content  (so- 
matic) of  smaller  juvenile  pink  salmon  (~35  mm)  did 
differ  between  stock  groups  (Paul  and  Willette,  1997). 


224 


Fishery  Bulletin  108(2) 


These  studies,  along  with  ours,  support  the  idea  that 
different  stock  groups  of  juvenile  salmon  may  have 
similar  WBEC  in  common  habitats  despite  stock-specific 
size  differences,  and  thus  emphasize  the  importance  of 
habitat  quality  on  fish  condition.  These  different  results 
could  also  be  related  to  ontogenetic  changes  in  physiol- 
ogy (Hoar,  1998;  Wuenschel  et  al.,  2006). 

Because  so  little  mortality  occurred  within  each  ex- 
perimental group,  we  conclude  that  juvenile  salmon 
can  survive  for  prolonged  periods  without  food  during 
the  summer  months,  as  has  also  been  reported  by  Ste- 
fansson  et  al.  (2009).  Most  of  the  mortalities  occurred 
within  the  first  eight  days  of  the  June  experiment.  As 
discussed  previously,  the  June  fish  were  younger  and 
less  robust  (lower  CR)  and  could  have  been  more  sus- 
ceptible to  environmental  stresses  because  of  scale  loss 
(Bouck  and  Smith,  1979)  from  net  abrasion  during  cap- 
ture, for  example.  However,  even  though  juvenile  chum 
salmon  were  still  alive  after  45  days  of  starvation,  many 
salmonids  cannot  recover  physiologically  after  extended 
periods  of  starvation  because  of  compromised  seawater 
tolerance  or  impaired  compensatory  growth  (Bilton  and 
Robins,  1973;  Ban  et  al.,  1996;  Stefansson  et  al.,  2009); 
such  recovery  capabilities  in  juvenile  chum  salmon  re- 
main unclear. 

The  experimental  WBEC,  %MC,  and  CR  differed  from 
the  long-term  average  of  the  SECM  data  sets  during 
both  months.  After  about  10  days  of  starvation,  WBEC 
was  below  the  normal  range,  %MC  was  above  the  nor- 
mal range,  and  CR  shifted  from  positive  to  negative, 
in  both  months.  More  specifically,  by  day  20,  the  June 
fish  had  lost  twice  their  WBEC  and  CR,  and  had  gained 
four  times  %MC  as  the  July  fish.  The  WBEC  of  the 
June  fish  required  only  3-7  days  of  starvation  before 
dropping  to  the  lower  initial  level  of  the  July  fish. 

Our  study  on  the  effects  of  starvation  on  field-caught 
juvenile  chum  salmon  indicates  that  WBEC,  %MC,  and 
CR  are  more  responsive  measures  than  WW  and  FL  to 
prolonged  food  deprivation  in  a controlled  laboratory 
environment.  Although  starvation  is  an  extreme  case 
of  limited  food  resources,  clearly  juvenile  chum  salmon 
can  survive  these  conditions  for  extended  periods,  but 
may  consequently  be  less  tolerant  of  variable  environ- 
mental conditions  and  more  susceptible  to  other  sources 
of  mortality,  such  as  predation.  Future  studies  will 
focus  on  monitoring  the  seasonal  response  of  juvenile 
salmon  condition  measures,  such  as  WBEC,  %MC,  and 
CR,  in  different  habitats  at  sea. 

Acknowledgments 

We  thank  the  command  and  crew  of  the  NOAA  ship 
John  N.  Cobb  for  help  in  collecting  samples.  We  thank  D. 
Tersteeg  and  the  staff  at  the  Macaulay  Hatchery  otolith 
laboratory  for  decoding  all  of  the  otoliths  used  in  this 
study.  This  manuscript  was  improved  with  suggestions 
from  three  anonymous  reviewers.  Finally,  we  thank  A. 
Wertheimer  and  A.  Moles  for  statistical  and  editorial 
help  with  this  manuscript. 


Literature  cited 

Akiyama,  T.,  and  T.  Nose. 

1980.  Changes  in  body  weight,  condition  factor  and  body 
composition  of  fingerling  chum  salmon  with  various 
sizes  during  starvation.  Bull.  Natl.  Res.  Inst.  Aquae. 
1:71-78. 

Ban,  M.,  H.  Hasegawa,  and  M.  Ezure. 

1996.  Effects  of  starvation  and  refeeding  on  physiologi- 
cal condition  of  juvenile  chum  salmon,  Oncorhynchus 
keta.  Sci.  Rep.  Hokkaido  Salmon  Hatch.  50:117-123. 

Bilton,  H.  T.,  and  G.  L.  Robins. 

1973.  The  effects  of  starvation  and  subsequent  feed- 
ing on  survival  and  growth  of  Fulton  channel  sockeye 
salmon  fry  (Oncorhynchus  nerka).  J.  Fish.  Res.  Board 
Can.  30:1-5. 

Boldt,  J.  L.,  and  L.  J.  Haldorson. 

2004.  Size  and  condition  of  wild  and  hatchery  pink  salmon 
juveniles  in  Prince  William  Sound,  Alaska.  Trans. 
Am.  Fish.  Soc.  133:173-184. 

Bouck,  G.  R.,  and  S.  D.  Smith. 

1979.  Mortality  of  experimentally  descaled  smolts  of 
coho  salmon  (Oncorhynchus  kisutch)  in  fresh  and  salt 
water.  Trans.  Am.  Fish.  Soc.  108:67-69. 

Breck,  J.  E. 

2008.  Enhancing  bioenergetics  models  to  account  for 
dynamic  changes  in  fish  body  composition  and  energy 
density.  Trans.  Am.  Fish.  Soc.  137:340-356. 

Brett,  J.  R. 

1995.  Energetics.  In  Physiological  ecology  of  Pacific 
salmon  (C.  Groot,  L.  Margolis,  and  W.  C.  Clarke,  eds.), 
p.  3-68.  UBC  Press,  Vancouver,  B.C.,  Canada. 

Brett,  J.  R.,  J.  E.  Shelbourn,  and  C.  T.  Shoop. 

1969.  Growth  rate  and  body  composition  of  fingerling 
sockeye  salmon,  Oncorhynchus  nerka,  in  relation  to  tem- 
perature and  ration  size.  J.  Fish.  Res.  Board  Can. 
26:2363-2394. 

Brodeur,  R.  D.,  J.  P.  Fisher,  D.  J.  Teel,  R.  L.  Emmett,  E.  Casillas, 
and  T.  W.  Miller. 

2004.  Juvenile  salmonid  distribution,  growth,  condition, 
origin,  and  environmental  and  species  associations  in  the 
Northern  California  Current.  Fish.  Bull.  102:25-46. 

Congleton,  J.  L.,  and  T.  Wagner. 

2006.  Blood-chemistry  indicators  of  nutritional  status  in 
juvenile  salmonids.  J.  Fish  Biol.  69:473-490. 

Cross,  A.  D.,  D.  A.  Beauchamp,  K.  W.  Myers,  and  J.  H.  Moss. 

2008.  Early  marine  growth  of  pink  salmon  in  Prince 
William  Sound  and  the  Coastal  Gulf  of  Alaska  during 
years  of  low  and  high  survival.  Trans.  Am.  Fish.  Soc. 
137:927-939. 

Edsall,  T.  A.,  A.  M.  Frank,  D.  V.  Rottiers,  and  J.  V.  Adams. 

1999.  The  effect  of  temperature  and  ration  size  on  the 
growth,  body  composition,  and  energy  content  of  juvenile 
coho  salmon.  J.  Great  Lakes  Res.  25:355-362. 

Hoar,  W.  S. 

1988.  The  physiology  of  smolting  salmonids.  In  Fish 
physiology,  vol  11  (Hoar,  W.  S.,  and  D.  J.  Randall,  eds.), 
p.  375-343.  Academic  Press,  New  York,  NY. 

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

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

Jobling,  M. 

1994.  Fish  bioenergetics,  309  p.  Chapman  and  Hill, 
London,  U.K. 


Fergusson  et  al.:  Effects  of  starvation  on  energy  density  of  Oncorhynchus  keta 


225 


Kotiaho,  J.  S. 

1999.  Estimating  fitness:  comparison  of  body  condition 
indices  revisited.  Oikos  87:399-400. 

LeBrasseur,  R.  J. 

1969.  Growth  of  juvenile  chum  salmon  (Oncorhynchus 
keta)  under  different  feeding  regimes.  J.  Fish.  Res. 
Board  Can.  26:1631-1645. 

Mason,  J.  C. 

1974.  Behavioral  ecology  of  chum  salmon  fry  ( Oncorhyn- 
chus keta)  in  a small  estuary.  J.  Fish.  Res.  Board  Can. 
31:83-92. 

Miglavs,  I.,  and  M.  Jobling 

1989.  The  effects  of  feeding  regime  on  proximate  body 
composition  and  patterns  of  energy  deposition  in  juve- 
nile Arctic  charr,  Salvelinus  alpinus.  J.  Fish  Biol. 
35:1-11. 

Mortensen,  D.,  A.  Wertheimer,  S.  Taylor,  and  J.  Landingham. 

2000.  The  relation  between  early  marine  growth  of  pink 
salmon,  Oncorhynchus  gorbuscha,  and  marine  water 
temperature,  secondary  production,  and  survival  to 
adulthood.  Fish.  Bull.  98:319-335. 

Murai,  T.,  Y.  Hirasawa,  T.  Akiyama,  and  T.  Nose. 

1983.  Effects  of  previous  dietary  history  on  the  mortality 
and  changes  in  body  compositions  of  chum  salmon  fry 
during  starvation  in  seawater.  Bull.  Natl.  Res.  Inst. 
Aquae.  4:79-86. 

Orsi,  J.  A.,  M.  V.  Sturdevant,  J.  M.  Murphy,  D.  G.  Mortensen, 
and  B.  L.  Wing. 

2000.  Seasonal  habitat  use  and  early  marine  ecology 
of  juvenile  Pacific  salmon  in  southeastern  Alaska.  N. 
Pac.  Anadr.  Fish  Comm.  Bull.  No.  2:111-122. 

Orsi,  J.  A.,  A.  C.  Wertheimer,  M.  V.  Sturdevant,  E.  A.  Fergusson, 
D.  G.  Mortensen,  and  B.  L.  Wing. 

2004.  Juvenile  chum  salmon  consumption  of  zooplank- 
ton in  marine  waters  of  southeastern  Alaska:  a bio- 
energetics approach  to  implications  of  hatchery  stock 
interactions.  Rev.  Fish  Biol.  Fish.  14:335-359. 

Park,  W.,  M.  Sturdevant,  J.  Orsi,  A.  Wertheimer,  E.  Fergusson, 
W.  Heard,  and  T.  Shirley. 

2004.  Interannual  abundance  patterns  of  copepods 
during  an  ENSO  event  in  Icy  Strait,  southeastern 
Alaska.  ICES  J.  Mar.  Sci.  61:464—477. 


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

1997.  Geographical  variation  in  somatic  energy  content  of 
migrating  pink  salmon  fry  from  Prince  William  Sound: 
A tool  to  measure  nutritional  status.  In  Forage  fishes 
in  marine  ecosystems:  proceedings  of  the  international 
symposium  on  the  role  of  forage  fishes  in  marine  eco- 
systems (C.  W.  Mecklenburg,  ed.),  p.  707-720.  Alaska 
Sea  Grant  College  Program  Report  97-01,  Univ.  Alaska, 
Fairbanks,  AK. 

Purcell,  J.  E.,  and  M.  V.  Sturdevant. 

2001.  Prey  selection  and  dietary  overlap  among  zooplank- 
tivorous  jellyfish  and  juvenile  fishes  in  Prince  William 
Sound,  Alaska.  Mar.  Ecol.  Prog.  Ser.  210:67-83. 

Sheridan,  M.  A.,  W.  V.  Allen,  and  T.  H.  Kerstetter. 

1983.  Seasonal  variations  in  the  lipid  composition  of  the 
steelhead  trout,  Sal/no  gairdneri  Richardson,  associ- 
ated with  the  parr-smolt  transformation.  J.  Fish  Biol. 
23:125-134. 

Simpkins,  D.  G„  W.  A.  Hubert,  C.  M.  Del  Rio,  and  D.  C.  Rule. 

2004.  Constraints  of  body  size  and  swimming  velocity  on 
the  ability  of  juvenile  rainbow  trout  to  endure  periods 
without  food.  J.  Fish  Biol.  65:530-544. 

Stefansson,  S.  O.,  A.  K.  Imsland,  and  S.  O.  Handeland. 

2009.  Food-deprivation,  compensatory  growth  and  hydro- 
mineral balance  in  Atlantic  salmon  ( Salmo  salar)  post- 
smolts  in  sea  water.  Aquaculture  290:243-249. 

Trudel,  M.,  S.  Tucker,  J.  F.  T.  Morris,  D.  A.  Higgs,  and  D.  W. 

Welch. 

2005.  Indicators  of  energetic  status  in  juvenile  coho 
salmon  and  Chinook  salmon.  N.  Am.  J.  Fish.  Manag. 
25:374-390. 

Weatherley,  A.  H„  and  H.  S.  Gill. 

1995.  Growth.  In  Physiological  ecology  of  Pacific  salmon 
(C.  Groot,  L.  Margolis,  and  W.  C.  Clarke,  eds.),  p.  103- 
158.  UBC  Press,  Vancouver,  B.C.,  Canada. 

Wuenschel,  M.  J.,  A.  R.  Jugovich,  and  J.  A.  Hare. 

2006.  Estimating  the  energy  density  of  fish:  the  impor- 
tance of  ontogeny.  Trans.  Am.  Fish.  Soc.  135:379-385. 

Zaporozhec,  O.  M.,  and  G.  V.  Zaporozhec. 

1993.  Preparation  of  hatchery-reared  chum  fry  for  life 
at  sea:  osmoregulation  dynamics.  Fish.  Oceanog. 
2:91-96. 


226 


Accuracy  of  sex  determination 

for  northeastern  Pacific  Ocean  thornyheads 

(Sebastolobus  altivelis  and  S.  alascanus) 

Erica  L.  Fruh  (contact  author)1 
Aimee  Keller2 
Jessica  Trantham3 
Victor  Simon2 

Email  address  for  contact  author:  Erica.Fruh@noaa.gov 

' National  Oceanographic  and  Atmospheric  Administration 
National  Marine  Fisheries  Service 
Northwest  Fisheries  Science  Center 
Fishery  Resource  Analysis  and  Monitoring  Division 
2032  SE  OSU  Drive 
Newport,  Oregon  97365 

2 National  Oceanographic  and  Atmospheric  Administration 
National  Marine  Fisheries  Service 

Northwest  Fisheries  Science  Center 
Fishery  Resource  Analysis  and  Monitoring  Division 
2725  Montlake  Blvd.  East 
Seattle,  Washington  98112 

3 Husbandry  Department 
Underwater  World 

1245  Pate  San  Vitores  RD  Ste  400 
Turnon,  Guam  96913 


Abstract — Determining  the  sex  of 
thornyheads  ( Sebastolobus  alasca- 
nus and  S.  altivelis ) can  be  difficult 
under  field  conditions.  We  assessed 
our  ability  to  correctly  assign  sex  in 
the  field  by  comparing  results  from 
field  observations  to  results  obtained 
in  the  laboratory  through  both  mac- 
roscopic and  microscopic  examination 
of  gonads.  Sex  of  longspine  thorny- 
heads was  more  difficult  to  determine 
than  that  of  shortspine  thornyheads 
and  correct  determination  of  sex 
was  significantly  related  to  size. 
By  restricting  the  minimum  size  of 
thornyheads  to  18  cm  for  macroscopic 
determination  of  sex  we  reduced  the 
number  of  fish  with  misidentified  sex 
by  approximately  65%. 


Manuscript  submitted  25  June  2009. 
Manuscript  accepted  11  February  2010. 
Fish.  Bull.  108:226-232  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National 
Marine  Fisheries  Service,  NOAA. 


Accurate  sex-specific  data  are  essen- 
tial for  fitting  age-structured  popula- 
tion dynamic  models  and  estimating 
spawning  biomass  (Methot,  2000). 
Assessing  sex  ratio  is  of  added 
importance  if  sex-based  selectiv- 
ity occurs  within  a fishery;  because 
separate  management  measures  may 
be  required  for  male  and  female  fish 
(Cochrane,  2009). 

Thornyheads  are  a common  conti- 
nental slope  species  and  support  a 
large  commercial  fishery  (Gunder- 
son, 1997).  Longspine  thornyheads 
( Sebastolobus  altivelis)  are  found 
from  the  Gulf  of  Alaska  to  southern 
Baja  California,  whereas  shortspine 
thornyheads  (Sebastolobus  alasca- 
nus) are  distributed  from  the  Ber- 
ing Sea  to  northern  Baja  (Orr  et  al., 
2000).  Longspine  thornyheads  gener- 
ally inhabit  depths  greater  than  400 
m,  have  a distribution  range  to  about 
1400  m depth  (Jacobson  and  Vetter, 
1996),  and  a peak  in  abundance  and 
spawning  biomass  at  about  1000  m 
depth  (Wakefield,  1990;  Jacobson 
and  Vetter,  1996).  Shortspine  thorny- 
heads are  found  from  20  m to  over 


1500  m in  depth,  are  most  abundant 
in  the  range  of  180  to  450  m,  and 
the  majority  of  the  spawning  bio- 
mass occurs  between  600  and  1400 
m,  where  longspine  thornyheads 
are  most  abundant  (Jacobson  and 
Vetter,  1996).  The  maximum  size  of 
shortspine  thornyheads  (>70  cm)  is 
larger  than  that  of  longspine  thorny- 
heads (—38  cm).  Shortspine  thorny- 
heads migrate  to  deeper  water  as 
their  body  size  increases,  whereas 
longspine  thornyheads  do  not  mi- 
grate to  deeper  water  with  increas- 
ing size. 

Identifying  the  sex  of  mature  long- 
spine thornyheads  and  shortspine 
thornyheads  by  gross  visual  exami- 
nation is  difficult  when  gonads  re- 
gress to  a resting  state  (Pearson  and 
Gunderson,  2003)  because  male  and 
female  gonads  are  small,  not  fully 
developed,  and  are  morphologically 
similar.  Determining  the  sex  of  in- 
dividual thornyheads  collected  dur- 
ing the  annual  Northwest  Fisheries 
Science  Center  (NWFSC)  West  Coast 
Groundfish  Bottom  Trawl  Survey  is 
difficult  because  the  survey  occurs 


Fruh  et  al.:  Accuracy  of  sex  determination  for  Sebastolobus  altivelis  and  S.  aloscanus 


227 


from  May  to  October  when  thornyheads  are  not  re- 
productively  active  and  gonads  are  in  a resting  state 
(Moser,  1974;  Wakefield,  1990). 

The  addition  of  sex  identification  for  both  thornyhead 
species  to  survey  sampling  protocols  will  improve  the 
information  available  for  management  of  the  resource. 
To  address  concerns  about  the  ability  of  field  person- 
nel to  correctly  determine  sex  of  thornyheads  while  at 
sea,  we  examined  the  sex  of  longspine  and  shortspine 
thornyheads  in  the  laboratory  using  macroscopic  ex- 
amination of  gonads  (as  a correlate  for  field  work)  in 
contrast  to  microscopic  techniques  (for  confirmation  of 
results).  An  additional  goal  was  to  determine  a mini- 
mum size  below  which  the  error  rate  for  classification 
of  sex  of  thornyheads  in  the  field  was  judged  to  be  too 
high  by  investigating  the  relationship  between  sex  mis- 
identification  and  length,  geographic  area,  and  month 
captured.  Because  assessment  scientists  are  interested 
in  the  actual  proportion  of  males  to  females,  we  also 
evaluated  absolute  percent  error  after  accounting  for 
the  portion  of  the  error  that  was  cancelled  out  by  bal- 
ancing the  number  of  misidentified  males  reported  as 
females  against  the  number  of  misidentified  females 
reported  as  males. 

Materials  and  methods 

The  2003  NWFSC  West  Coast  Groundfish  Bottom  Trawl 
Survey  was  conducted  between  24  June  and  23  October, 
from  the  area  off  Cape  Flattery,  Washington  (48°10'N 
lat.)  to  the  U.S. -Mexico  border  (32°30'N  lat.)  at  water 
depths  of  55-1280  m.  The  survey  area  was  covered  twice 
by  chartered  commercial  fishing  vessels  (20  to  28  m 
length).  The  first  sampling  period  was  from  24  June  to 
13  August  and  the  second  from  31  August  to  23  October. 
A stratified  random  sampling  design  was  used  and  the 
survey  area  was  subdivided  into  adjacent  cells  of  equal 
area  (1.5  nmi  long,  by  2.0  nmi  lat.,  Albers  equal  area 
projection).  A total  of  620  primary  sites  were  randomly 
selected  from  cells  stratified  by  geographic  location  and 
depth.  The  geographic  allocation  was  based  on  assign- 
ing 15-25%  of  the  cells  to  each  of  five  International 
North  Pacific  Fisheries  Commission  (INPFC)  statis- 
tical areas:  U.S. -Vancouver  (47°30'N  to  U.S. -Canada 
border),  Columbia  (43°00'  to  47°30'N),  Eureka  (40°30'  to 
43°00'N),  Monterey  (36°00'  to  40°30'N),  and  Conception 
(U.S. -Mexico  border  to  36°00'N).  The  survey  area  was 
further  stratified  into  depth  zones  with  45%  of  the  cells 
allocated  to  the  shallow  depth  zone  (55-183  m),  30%  to 
mid-depth  (184-549  m)  and  25%  to  the  deep  stratum 
(550-1280  m).  Each  of  four  chartered  fishing  vessels 
was  assigned  155  stations  to  sample. 

The  bottom  trawl  survey  is  a standardized  fishery  in- 
dependent survey  and  all  fishing  operations  are  conduct- 
ed in  strict  compliance  to  national  protocols  (Stauffer, 
2004).  Vessels  were  equipped  with  standard  Aberdeen- 
style  nets  with  small  mesh  (1.5-inch  stretched  measure) 
liner  in  the  codend.  All  thornyheads  randomly  selected 
for  biological  sampling  were  assigned  a unique  identi- 


fication number,  individually  weighed  (kg),  measured 
(fork  length,  cm),  and  frozen  while  at  sea.  All  frozen 
specimens  were  brought  back  to  the  laboratory  where 
fish  were  thawed,  dissected,  and  examined  macroscopi- 
cally  to  identify  sex.  For  macroscopic  examination  of 
gonads,  an  incision  was  made  with  a scalpel  on  the 
ventral  surface  of  each  thornyhead  from  the  vent  to  the 
base  of  the  pectoral  fin.  The  lateral  side  of  the  fish  was 
opened  to  expose  the  gonads,  and  a visual  identification 
of  sex  was  based  on  the  physical  structure  of  the  gonad- 
al tissue  as  described  by  Lagler  et  al.  (1962).  Sex  was 
recorded  as  male,  female,  or  unknown.  For  microscopic 
identification  of  sex,  a section  of  gonad  tissue  from  each 
fish  was  placed  on  a glass  microscope  slide,  stained  with 
acetocarmine  solution  and  compressed  with  a cover  slip. 
The  stain  acted  on  the  gonad  tissue  by  readily  staining 
oocytes  dark  pink  (Guerrero,  1974).  The  slides  were 
viewed  under  a lOx  power  microscope  (Leica  DM  LS2, 
Bannockburn,  IL),  and  females  were  distinguished  from 
males  by  the  presence  of  dark  pink  stained  oocytes. 

Accuracy  of  sex  determination  was  examined  in  rela- 
tion to  length  by  species,  geographic  region,  and  month 
of  capture  (June-October).  To  determine  a size  thresh- 
old below  which  sex  determination  should  not  be  at- 
tempted in  the  field,  we  examined  both  the  total  and 
absolute  percentage  of  incorrectly  sexed  thornyheads 
in  relation  to  length.  To  avoid  biasing  results,  we  did 
not  consider  our  ability  to  correctly  identify  female 
thornyheads  at  smaller  sizes,  as  opposed  to  our  ability 
to  correctly  identify  males  at  smaller  sizes.  Absolute  er- 
ror was  calculated  as  the  absolute  value  of  misidentified 
males  minus  misidentified  females  divided  by  the  total 
number  examined  at  each  1-cm  size  interval,  and  this 
value  was  then  expressed  as  a percentage.  Size  data 
were  transformed  (natural  logarithm)  to  reduce  hetero- 
geneity of  variance  before  statistical  analysis.  Data  were 
statistically  compared  by  analysis  of  variance  (ANOVA) 
by  using  SAS  for  Windows  (SAS  Institute,  Inc.,  Cary, 
NC).  Significant  ANOVAs  were  followed  by  a nonpara- 
metric  comparison  of  means  test  (Tukey’s  test).  Fish  in 
which  the  gonad  could  not  be  found,  stained,  or  micro- 
scopically identified  were  not  included  in  the  analyses. 

Results 

A total  of  574  successful  tows  were  completed.  Figure  1 
shows  the  distribution  and  relative  abundance  (kg/ha) 
of  thornyheads  from  the  2003  survey.  Both  species 
were  concentrated  in  the  mid-  and  deep  depth  strata 
(183-1280  m)  and  exhibited  higher  relative  abundance 
north  of  Pt.  Conception,  CA  (34°30'N  lat.).  Longspine 
thornyheads  were  collected  in  214  tows  at  depths  of 
328—1280  m (mean  depth  802  m)  and  shortspine  thorny- 
heads were  collected  in  311  tows  at  depths  of  88-1280  m 
(mean  depth  605  m).  A total  of  2325  thornyheads  were 
collected  for  later  processing  in  the  laboratory.  Sex  was 
determined  for  852  longspine  thornyheads  and  1148 
shortspine  thornyheads.  Sex  was  indeterminable  for 
189  longspine  and  136  shortspine  thornyheads  (average 


228 


Fishery  Bulletin  108(2) 


Figure  1 

Distribution  and  relative  abundance  (kg/ha)  of  (A)  longspine  thornyhead  ( Sebastolobus  altivelis)  and  (B)  shortspine 
thornyhead  ( Sebastolobus  alascanus)  determined  from  the  2003  Northwest  Fisheries  Science  Center  west  coast  ground- 
fish  trawl  survey.  SD  = standard  deviation. 


length  14.3  cm).  Longspine  thornyhead  sex  was  misiden- 
tified  by  visual  examination  in  23.1%  of  males  and  22.4% 
of  females,  and  for  shortspine  thornyheads,  in  9.4%  of 
males  and  9.3%  of  females. 

Average  lengths  of  longspine  and  shortspine  thorny- 
heads (females,  males,  and  total)  for  which  sex  was 
misidentified  were  significantly  lower  than  the  lengths 
for  fish  whose  sex  was  correctly  assigned  (Table  1). 
For  shortspine  thornyheads,  the  average  length  of  sex- 
misidentified  females  was  significantly  smaller  than 
that  of  males  (ANOVA:  df=6,  F- 5.5,  P= 0.02).  Similar 
tendencies  were  seen  for  longspine  thornyhead  lengths 
but  the  results  were  not  significant  (Table  1). 

Determining  sex  for  longspine  thornyheads  greater 
than  22  cm  would  eliminate  approximately  80%  of  the 
overall  error  rate,  but  would  also  eliminate  50%  of  the 
fish  whose  sex  was  correctly  determined.  By  proposing 
18  cm  as  the  minimum  size  for  examining  longspine 
thornyheads  in  the  field  we  eliminated  approximately 
65%  of  the  incorrectly  sexed  fish,  while  retaining  >70% 
of  those  correctly  sexed  (Fig.  2A).  On  average,  the  sex 
of  50.5%  of  longspine  thornyheads  ranging  in  size  from 
11  to  17  cm  was  incorrectly  determined.  This  average 
dropped  to  approximately  10%  for  longspine  thorny- 


heads at  lengths  from  18  to  34  cm.  A similar  result  was 
seen  for  shortspine  thornyheads  (Fig.  2A).  The  average 
percentage  of  shortspine  thornyheads  with  misidentified 
sex  was  53.7%  at  lengths  from  11  to  17  cm.  This  value 
decreased  to  5.9%  for  larger  fish  (18-71  cm)  (Fig.  2A). 

With  a single  exception,  more  males  were  misiden- 
tified as  females  in  every  size  category  for  both  spe- 
cies, and  the  absolute  percentage  of  sex-misidentified 
fish  decreased  at  fork  lengths  greater  than  17  cm  (Fig. 
2B).  For  longspine  thornyheads  the  average  decreased 
from  15.8%  for  fish  11-17  cm  to  2.2%  for  fish  18-34 
cm  length  and  the  average  percentage  for  shortspine 
thornyheads  dropped  from  24.5%  to  3.0%  in  the  larger 
size  category  (Fig.  2B). 

Sex  misidentification  in  longspine  thornyheads  did  not 
vary  significantly  by  month  from  June  through  October 
(ANOVA:  df=7,  F=1.74,  P=0.34;  Fig.  3A).  However,  sex 
misidentification  for  shortspine  thornyheads  was  signifi- 
cantly higher  in  August,  with  an  increasing  trend  from 
June  through  August  followed  by  a decline  (ANOVA: 
df=7,  F=15.5,  P=0.02;  Fig.  3A). 

The  accuracy  of  sex  determination  varied  by  geo- 
graphic area  for  both  species  (Fig.  3B).  The  sex  of  long- 
spine thornyheads  was  more  frequently  misidentified 


Fruh  et  al.:  Accuracy  of  sex  determination  for  Sebastolobus  altivelis  and  5.  alascanus 


229 


Table  1 

Number  ( n ),  mean  fork  length  (cm,  ±standard  error  [SE] ),  and  analyses  of  variance  (ANOVAs)  for  sizes  for  female,  male,  and 
total  longspine  ( Sebastolobus  altivelis)  and  for  female,  male,  and  total  shortspine  thornyheads  ( S . alascanus)  captured  during 
the  2003  Northwest  Fisheries  Science  Center  west  coast  groundfish  trawl  survey,  correctly  and  incorrectly  assigned  sex  based 
on  visual  examination. 


Correct 

Incorrect 

ANOVAs 

Species 

n 

Mean  length  ( ± SE ) 

n 

Mean  length  ( ± SE ) 

df 

F 

P 

Longspine  thornyhead 

female 

396 

21.6  (0.21) 

114 

18.5  (0.37) 

509 

48.9 

0.0001 

male 

259 

23.4  (0.20) 

83 

19.0  (0.47) 

341 

96.6 

0.0001 

total 

655 

22.3  (0.15) 

197 

18.7  (0.29) 

851 

52.7 

0.0001 

Shortspine  thornyhead 

female 

560 

35.5  (0.52) 

58 

23.6  (1.16) 

617 

50.8 

0.0001 

male 

481 

34.9  (0.45) 

49 

28.1  (1.52) 

529 

21.5 

0.0001 

total 

1041 

35.2  (0.35) 

107 

25.7  (0.96) 

1147 

36.2 

0.0001 

above  43°N  latitude,  and  the  U.S. -Vancouver  and  Co- 
lumbia areas  had  a significantly  higher  average  per- 
centage of  misidentification  than  the  Eureka,  Monterey, 
and  Conception  areas  (ANOVA:  df=4,  F=44.1,  P-0.007). 
The  sex  of  shortspine  thornyheads  became  more  diffi- 


cult to  correctly  identify  below  40°N  latitude,  and  both 
the  Monterey  and  Conception  areas  had  a significantly 
higher  average  percentage  of  misidentification  com- 
pared to  the  Eureka,  Columbia,  and  U.S. -Vancouver 
areas  (ANOVA:  df=4,  P=13.9,  P=0.03).  There  were  no 


230 


Fishery  Bulletin  108(2) 


Fork  length  (cm) 

Figure  2 

(A)  Total  percentage  of  sex-misidentified  longspine  ( Sebastolobus  altivelis) 
and  shortspine  thornyheads  (Sebastolobus  alascanus)  determined  by  compar- 
ing the  gross  morphological  features  of  gonads  to  a section  of  each  gonad 
subsequently  stained  and  viewed  microscopically,  by  size  (fork  length,  cm); 
and  (B)  the  absolute  percent  error  in  identifying  the  sex  of  thornyheads 
after  accounting  for  the  portion  of  the  total  error  that  is  cancelled  out  by 
balancing  the  number  of  sex-misidentified  males  against  the  number  of 
sex-misidentified  females. 


significant  differences  in  mean  fork  length  for  longspine 
thornyheads  between  the  different  areas  (ANOVA:  df= 
858,  F=0.3,  P=0.9),  but  for  shortspine  thornyheads, 
size  was  significantly  larger  in  the  Monterey  and  U.S.- 
Vancouver  areas  (ANOVA:  df=1140,  F=  4.7,  P=0.0009), 
and  large  fish  in  the  Monterey  area  had  a higher  rate 
of  individuals  for  which  sex  was  incorrectly  determined 
than  similar  size  shortspine  thornyheads  in  the  U.S.- 
Vancouver  area. 


Discussion 

This  study  provides  guidance  for  a minimum  size  limit 
below  which  sex  of  thornyheads  should  not  be  deter- 
mined at-sea  because  of  high  error  rates.  High  quality 
biological  information  is  important  for  management  and 
modeling  of  thornyhead  populations  along  the  U.S.  west 
coast  (Fay,  2005).  Fishery  scientists  need  estimates  of 


sex  ratio  for  fish  populations  because  shifts  in  these 
values  can  indicate  overfishing  on  one  sex  or  the  other 
due  to  selective  gear,  differential  growth  rates,  segrega- 
tion by  sex  or  any  combination  of  these  (Cochrane,  2009). 

In  previous  studies  of  the  reproductive  biology  of 
thornyheads,  the  longspine  thornyhead  spawning  was 
determined  to  begin  in  January,  peak  in  February  and 
March,  and  continue  at  least  through  April  (Wakefield, 
1990;  Pearson  and  Gunderson,  2003;  Cooper  et  al., 
2005).  Shortspine  thornyheads  spawn  between  Decem- 
ber and  May  along  the  U.S.  west  coast.  The  onset  of 
sexual  maturity  occurs  at  17-19  cm  total  length  (10% 
mature  females)  in  both  species  and  90%  are  mature 
at  25-27  cm  (Pearson  and  Gunderson,  2003).  Sex  of 
smaller  thornyheads  is  difficult  to  determine,  particu- 
larly during  the  summer,  because  of  the  small  size  of 
the  gonads — size  being  a function  of  the  annual  spawn- 
ing cycle.  Pearson  and  Gunderson  (2003)  noted  that 
of  36  longspine  thornyheads  designated  as  immature 


Fruh  et  at:  Accuracy  of  sex  determination  for  Sebastolobus  altivelis  and  S.  alascanus 


231 


females  in  the  field  on  the  basis  of  gross  morphological 
features,  nine  were  actually  males. 

Correct  visual  identification  of  sex  for  both  shortspine 
and  longspine  thornyheads  increased  in  fish  longer  than 
17  cm.  Overall  accuracy  is  greater  for  shortspine  than 
for  longspine  thornyheads,  and  greater  for  females  than 
for  males,  and  this  accuracy  is  related  to  size  in  both 
instances.  For  both  species,  18  cm  was  selected  as  the 
lower  limit  for  determining  the  sex  of  thornyheads  in 
the  field  because  the  majority  of  sex-misidentified  fish 
fell  below  this  value.  In  2003,  66%  of  the  longspine 
thornyheads  and  90%  of  the  shortspine  thornyheads 
measured  in  the  field  throughout  the  survey  period  were 
greater  than  17  cm.  The  selected  size  falls  within  the 
range  of  lengths  noted  for  the  onset  of  sexual  maturity 
in  both  species. 

Because  the  survey  is  conducted  after  the  completion 
of  the  spawning  season  for  longspine  thornyheads  (Janu- 
ary-April),  the  samples  are  collected  exclusively  during 
the  reproductive  resting  stage.  Sex  misidentification  was 
relatively  constant  for  longspine  thornyheads  through- 
out the  sample  period  and  there  were  no  significant 
differences  among  months.  Sex  misidentification  was 
greater  for  longspine  than  for  shortspine  thornyheads 
for  each  time  period.  The  lower  rate  of  sex  misidentifica- 
tion for  shortspine  thornyheads  may  be  related  to  their 
longer  spawning  season  (December— May).  Differences 
in  the  reproductive  cycles  of  the  two  species  resulted 
in  the  cessation  of  spawning  coinciding  with  the  start 
of  the  survey  sampling  for  shortspine  thornyheads  and 
may  partially  explain  the  observed  overall  lower  rate  of 
sex  misidentification  for  this  species.  The  middle  of  the 
reproductive  resting-stage  period  correlated  with  high 
levels  of  sex  misidentification  for  both  species,  although 
only  for  shortspine  thornyheads  was  the  difference  sig- 
nificant (in  August). 

The  differences  in  sex  misidentification  among  geo- 
graphic areas  are  more  difficult  to  explain.  Sex  of  long- 
spine thornyhead  was  more  frequently  misidentified  in 
the  U.S. -Vancouver  and  Columbia  areas.  Samples  in 
these  areas  were  collected  primarily  in  June  and  Sep- 
tember, the  periods  with  the  highest  rates  of  sex  mis- 
identification. The  lack  of  any  significant  differences  in 
mean  length  for  longspine  thornyheads  between  INPFC 
areas  indicates  that  the  higher  rates  of  misidentification 
of  sex  farther  north  were  not  a function  of  size,  but  were 
related  to  the  timing  of  the  annual  spawning  cycle  at 
differing  latitudes. 

Shortspine  thornyhead  samples  collected  in  the  Eure- 
ka, Columbia,  and  U.S. -Vancouver  areas  (i.e.,  those  with 
significantly  lower  rates  of  sex  misidentification)  were 
primarily  taken  in  June,  July,  and  September  when  the 
rate  of  sex  misidentification  for  shortspine  thornyheads 
was  lowest.  Additionally,  there  were  significant  differ- 
ences in  the  lengths  of  shortspine  thornyheads  among 
areas,  indicating  that  the  lower  rates  of  sex  misidentifi- 
cation in  the  U.S. -Vancouver  area  may  also  be  partially 
related  to  size  (although  similar  size  differences  were 
not  observed  in  the  Eureka  and  Columbia  areas).  Be- 
cause differences  in  geographic  area  were  related  to  size 


0) 

O 

q3  35 

Q_ 

30 
25 
20 
15 
10 
5 
0 

Conception  Monterey  Eureka  Columbia  US-Vancouver 

Figure  3 

Percentage  of  sex-misidentified  longspine  ( Sebastolo- 
bus altivelis)  and  shortspine  (Sebastolobus  alascanus ) 
thornyheads  determined  by  comparing  the  gross  mor- 
phological features  of  gonads  to  a section  of  each  gonad 
subsequently  stained  and  viewed  microscopically  (A) 
by  month,  and  (B)  by  geographic  area  as  defined  by 
the  International  North  Pacific  Fisheries  Commission 
regions:  Conception  (U.S. -Mexico  border  to  36°00'N  lat.), 
Monterey  (36°00'  to  40°30'N  lat.),  Eureka  (40°30'  to 
43°00'N  lat.),  Columbia  (43°00'  to  47°30'N  lat.),  and 
U.S. -Vancouver  (47°30'N  lat.  to  U.S. -Canada  border). 


for  at  least  one  thornyhead  species  and  the  differences 
in  seasonal  determination  of  sex  were  variable,  we  rec- 
ommend that  sex  determination  of  thornyheads  <18  cm 
not  be  attempted  in  the  field.  This  is  likely  a conserva- 
tive estimate  because  identifying  sex  in  fresh  specimens 
at  sea  is  somewhat  more  reliable  than  examining  frozen 
and  thawed  specimens  in  the  laboratory.  The  approach 
described  here  establishes  a protocol  for  determining 
a minimum  size  for  at-sea  sex  identification  of  thorny- 
heads, but  may  be  applicable  for  use  with  any  species 
where  ambiguity  may  exist  in  correctly  identifying  the 
sex  of  fish  at  smaller  sizes,  within  different  regions,  or 
across  spawning  or  other  seasonal  cycles. 


Acknowledgments 

We  thank  the  captains  and  crew  of  the  fishing  vessels 
Ms.  Julie , Excalibur,  Captain  Jack,  and  Blue  Horizon  for 
their  effort  during  the  2003  NWFSC  West  Coast  Ground- 
fish  Bottom  Trawl  Survey.  We  also  thank  the  biologists 
who  participated  in  this  study,  including  K.  Bosley,  J. 
Buchanan,  D.  Kamikawa,  and  V.  Tuttle. 


232 


Fishery  Bulletin  108(2) 


Literature  cited 

Cochrane,  K. 

2009.  Current  paradigms  and  forms  of  advice.  In  Fish 
reproductive  biology  implications  for  assessment  and 
management  (Jakobsen,  T.,  M.  J.  Fogerty,  B.  A.  Megrey, 
and  E.  Moksness,  eds.),  p.  335-354.  Wiley-Blackwell, 
Ames,  IA. 

Cooper,  D.  W.,  K.  E.  Pearson,  and  D.  R.  Gunderson. 

2005.  Fecundity  of  shortspine  thornyhead  ( Sebastolo - 
bus  alascanus)  and  longspine  thornyhead  (S.  altive- 
lis)  ( Scorpaenidae)  from  the  northeastern  Pacific 
Ocean,  determined  by  stereological  and  gravimetric 
techniques.  Fish.  Bull.  103:15-22. 

Fay,  G. 

2005.  Stock  assessment  and  status  of  longspine  thorny- 
head (Sebastolobus  altivelis)  off  California,  Oregon,  and 
Washington  in  2005,  90  p.  Pacific  Fishery  Management 
Council,  Portland,  OR. 

Guerrero  III,  R.  D. 

1974.  An  aceto-carmine  squash  method  for  sexing  juve- 
nile fishes.  Prog.  Fish  Cult.  36:56. 

Gunderson,  D.  R. 

1997.  Spatial  patterns  in  the  dynamics  of  slope  rockfish 
stocks  and  their  implications  for  management.  Fish. 
Bull.  95:219-230. 

Jacobson,  L.  D.,  and  R.  D.  Vetter. 

1996.  Bathymetric  demography  and  niche  separa- 
tion of  thornyhead  rockfish:  Sebastolobus  alascanus 
and  Sebastolobus  altivelis.  Can.  J.  Fish.  Aquat.  Sci. 
53:600-609. 


Lagler,  K.  F.,  J.  E.  Bardach  and  R.  R.  Miller. 

1962.  Ichthyology,  545  p.  John  Wiley  and  Sons,  Inc., 
New  York. 

Methot,  R. 

2000.  Technical  description  of  the  stock  synthesis  assess- 
ment program.  U.S.  Dep.  Commer.,  NOAA.  Tech.  Memo. 
NMFS-NWFSC-43,  46  p. 

Moser,  H.  G. 

1974.  Development  and  distribution  of  larvae  and  juveniles 
of  Sebastolobus  (Pisces:  Family  Scorpaenidae).  Fish. 
Bull.  72:865-884. 

Orr,  J.  W.,  M.  A.  Brown,  and  D.  C.  Baker. 

2000.  Guide  to  rockfishes  (Scorpaenidae)  of  the  genera 
Sebastes , Sebastolobus,  and  Adelosebastes  of  the  North- 
east Pacific  Ocean,  2nd  ed.  NOAA  Tech.  Memo.  NMFS- 
AFSC-117,  48  p. 

Pearson,  K.  E.,  and  D.  R.  Gunderson. 

2003.  Reproductive  biology  and  ecology  of  shortspine 
thornyhead  rockfish,  Sebastolobus  alascanus  and  long- 
spine thornyhead  rockfish,  S.  altivelis,  from  the  north- 
easter Pacific  Ocean.  Environ.  Biol.  Fish.  62:117-136. 

Stauffer,  G. 

2004.  NOAA  protocols  for  groundfish  bottom  trawl  sur- 
veys of  the  nation’s  fishery  resources.  U.S.  Dep.  Com- 
merce, NOAA  Tech.  Memo.  NMFS-F/SPO-65,  205  p. 

Wakefield,  W.  W. 

1990.  Patterns  in  the  distribution  of  demersal  fishes 
on  the  upper  continental  shelf  off  central  California 
with  studies  of  ontogenetic  vertical  migration  in  par- 
ticle flux.  Ph.D.  thesis,  281  p.  Scripps  Institution  of 
Oceanography,  Univ.  California,  San  Diego,  CA. 


233 


Abstract — Body-size  measurement 
errors  are  usually  ignored  in  stock 
assessments,  but  may  be  important 
when  body-size  data  (e.g.,  from  visual 
surveys)  are  imprecise.  We  used 
experiments  and  models  to  quantify 
measurement  errors  and  their  effects 
on  assessment  models  for  sea  scallops 
( Placopecten  magellanieus).  Errors  in 
size  data  obscured  modes  from  strong 
year  classes  and  increased  frequency 
and  size  of  the  largest  and  smallest 
sizes,  potentially  biasing  growth,  mor- 
tality, and  biomass  estimates.  Model- 
ing techniques  for  errors  in  age  data 
proved  useful  for  errors  in  size  data. 
In  terms  of  a goodness  of  model  fit 
to  the  assessment  data,  it  was  more 
important  to  accommodate  variance 
than  bias.  Models  that  accommodated 
size  errors  fitted  size  data  substan- 
tially better.  We  recommend  experi- 
mental quantification  of  errors  along 
with  a modeling  approach  that  accom- 
modates measurement  errors  because 
a direct  algebraic  approach  was  not 
robust  and  because  error  parameters 
were  difficult  to  estimate  in  our 
assessment  model.  The  importance 
of  measurement  errors  depends  on 
many  factors  and  should  be  evaluated 
on  a case  by  case  basis. 


Manuscript  submitted  22  June  2009. 
Manuscript  accepted  25  January  2010. 
Fish.  Bull.  108:233-247  (2010). 

The  views  and  opinions  expressed  or 
implied  in  this  article  are  those  of  the 
author  (or  authors)  and  do  not  necessarily 
reflect  the  position  of  the  National  Marine 
Fisheries  Service,  NOAA. 


Measurement  errors  in  body  size 

of  sea  scallops  ( Placopecten  magellanieus ) 

and  their  effect  on  stock  assessment  models 


Larry  D.  Jacobson  (contact  author)1  Deborah  Hart1 


Kevin  D.  E.  Stokesbury2 
Melissa  A.  Allard2 
Antonie  Chute1 
Bradley  P.  Harris2 


Tom  Jaffarian2 
Michael  C.  Marino  II2 
Jacob  I.  Nogueira2 
Paul  Rago1 


Email  address  for  contact  author:  Larry.Jacobson@noaa.gov 

1 NOAA  Fisheries 

Northeast  Fisheries  Science  Center 
166  Water  Street 

Woods  Hole,  Massachusetts  02543-1026 

2 Department  of  Fisheries  Oceanography 
School  for  Marine  Science  and  Technology 
University  of  Massachusetts  School  of  Marine  Sciences 
838  South  Rodney  French  Boulevard 

New  Bedford,  Massachusetts  02744-1221 


Two  fishery-independent  surveys  are 
important  for  monitoring  Atlantic 
sea  scallop  (Placopecten  magellani- 
cus)  abundance  and  biomass  levels 
off  the  northeastern  coast  of  the 
United  States  because  they  provide 
abundance,  body  size,1  meat  weight 
(weight  of  marketable  adductor  mus- 
cles), and  other  data  (NEFSC2’3).  The 
National  Marine  Fisheries  Service, 
Northeast  Fisheries  Science  Center 
(NEFSC)  sea  scallop  dredge  survey 
has  been  conducted  annually  since 
1977  (Serchuk  et  ah,  1979;  Serchuk 
and  Wigley,  1986).  In  addition,  an 
underwater  video  survey  for  sea  scal- 
lops and  other  benthic  organisms  has 
been  conducted  annually  since  2003 
(Stokesbury,  2002;  Stokesbury  et  al., 
2004)  by  the  University  of  Massachu- 
setts Dartmouth,  School  for  Marine 
Science  and  Technology  (SMAST).  The 
dredge  and  video  surveys  are  carried 
out  across  the  range  of  sea  scallops  in 
U.S.  waters. 

In  this  analysis,  we  used  sea  scal- 
lops to  draw  attention  to  errors  in 
body-size  data  when  the  data  are 
used  in  a length-structured  stock 
assessment  model.  The  topic  of  mea- 
surement errors  in  body-size  data 
has  received  relatively  little  atten- 
tion, although  Heery  and  Berkson 


(2009)  evaluated  effects  of  systematic 
errors  (biased  sampling)  in  fishery 
size-composition  data  used  in  an  age- 
structured  model.  Our  work  was  mo- 
tivated by  questions  that  arose  from 
examining  video  survey  shell-height 
data  in  sea  scallop  stock  assessments 
(NEFSC2’3).  Our  experimental  and 
analytical  results  may  be  important 
and  useful  in  other  situations  where 
body-size  data  are  imprecise.  Body- 


1 Shell  height  (SH,  the  distance  in  mm 
between  the  umbo  and  shell  margin) 
is  the  body  size  measurement  for  sea 
scallops. 

2 NEFSC  (Northeast  Fisheries  Science 
Center).  2004.  Stock  assessment  for 
Atlantic  sea  scallops.  In  39th  north- 
east regional  stock  assessment  workshop 
(39th  SAW)  assessment  summary  report 
and  assessment  report.  Northeast  Fish- 
eries Science  Center,  National  Marine 
Fisheries  Service,  Woods  Hole  Labora- 
tory, 166  Water  St.,  Woods  Hole,  MA 
02543.  Ref.  Doc.  04-10,  p.  87-211. 

3 NEFSC  (Northeast  Fisheries  Science 
Center).  2007.  Stock  assessment  for 
Atlantic  sea  scallops.  In  45th  northeast 
regional  stock  assessment  workshop  (45th 
SAW)  assessment  summary  report  and 
assessment  report.  Northeast  Fisheries 
Science  Center,  National  Marine  Fisher- 
ies Service,  Woods  Hole  Laboratory,  166 
Water  St.,  Woods  Hole,  MA  02543.  Ref. 
Doc.  07-16,  p.  139-370. 


234 


Fishery  Bulletin  108(2) 


size  data  may  be  imprecise,  for  example,  when  col- 
lected by  scuba  (St.  John  et  al.,  1990;  Edgar  et  al., 
2004),  remotely  operated  underwater  vehicles  (ROV; 
Butler  et  al.,  2006),  camera  sleds  (Rosenkranz  and 
Byersdorfer,  2004),  or  in  other  optical  surveys  where 
body-size  measurements  are  obtained  without  handling 
individual  specimens. 

In  fishery  stock  assessment  modeling,  body-size  mea- 
surements are  almost  always  assumed  to  be  without 
error.  In  contrast,  statistical  sampling  errors  that  arise 
from  too  few  are  often  considered  in  modeling  (Fournier 
and  Archibald,  1982;  Pennington  et  al.,  2001).  Measure- 
ment errors  in  fishery  age  data  have  received  substan- 
tial attention  and  are  often  addressed  in  stock  assess- 
ment modeling  (Methot,  1989,  1990).  Approaches  to 
dealing  with  measurement  error  in  body-size  data  have 
not  been  explored. 

Shell-height  composition  data  for  sea  scallops  are 
of  two  types:  1)  distributions  of  shell-height  measure- 
ments, which  include  measurement  errors  and  true 
variability  among  individuals  in  size;  and  2)  distri- 
butions of  shell-height  measurements,  which  include 
measurement  errors  only.  It  is  important  to  distinguish 
between  these  two  types  of  data.  In  particular,  shell- 
height  compositions  are  sample  specific  and  depend  on 
the  underlying  distribution  of  true  sizes.  In  our  study 
measurement  errors  are  the  difference  between  the 
video  or  board  measurements  and  the  true  shell  height 
of  individual  specimens  (i.e.,  after  removing  differences 
in  true  shell  height  among  individuals).  Shell-height 
composition  data  are  important  because  they  are  in- 
terpreted in  stock  assessments  to  estimate  year-class 
strength,  mortality,  and  other  biological  characteris- 
tics. In  our  study  measurement  errors  are  important 
because  they  can  be  used  to  quantify  the  accuracy  of 
the  measurement  process  itself  and  because  they  affect 
shell-height  data  from  all  samples. 

Two  types  of  measurement  errors  are  considered  in 
this  study.  The  first  type  is  bias  that  causes  individual 
shell-height  measurements  and  estimated  sample  means 
to  differ,  on  average,  from  their  true  values  (Cochran, 
1977).  The  second  type  is  random  errors,  which  cause 
variability  in  shell-height  measurements  and  affect  the 
precision  of  measurements  and  estimated  mean  values 
(Cochran,  1977). 

Figure  1 shows  how  hypothetical  errors  in  sea  scal- 
lop shell-height  measurements  tend  to  smooth  the  true 
underlying  distribution  of  the  data.  Measurement  errors 
tend  to  smooth  modes  in  the  data  (which  usually  cor- 
respond to  recruitment  events)  by  moving  individuals 
from  size  bins  with  relatively  high  numbers  into  adja- 
cent bins  with  lower  numbers.  Random  measurement 
errors  also  tend  to  expand  the  range  of  observed  sizes 
by  decreasing  the  smallest  observed  size  and  increas- 
ing the  largest  (Fig.  1).  Bias  degrades  body-size  data 
by  making  measurements  consistently  larger  or  smaller 
than  the  true  value.  Methot  (1989,  1990)  highlighted 
these  issues  in  the  context  of  age  data  from  survey  and 
fishery  samples.  We  use  Methot’s  modeling  methods  in 
our  analysis  for  shell-height  data. 


In  principal,  body-size  measurement  errors  can  cause 
errors  in  a wide  range  of  important  fishery  estimates 
but  biomass  estimates  are  of  particular  importance.  In 
the  absence  of  bias,  imprecise  body-size  data  tend  to 
cause  positive  bias  in  mean  weight  and  biomass  esti- 
mates because  of  the  nonlinear  relationship  between 
size  and  biomass  and  Jensen’s  inequality  (Feller,  1966). 
For  example,  according  to  Jensen’s  inequality,  if  body 
weight  is  a cubic  function  of  body  size,  then  a -10% 
error  in  body  size  will  cause  a 0.93-l  = -27%  error  in 
estimated  body  weight  for  one  individual.  In  contrast, 
a +10%  error  in  body  size  will  cause  a 1.13-1  = +33% 
error  in  body  weight.  The  combined  effect  of  the  two 
errors  for  two  scallops  of  the  same  size  would  be  a posi- 
tive bias  of  +6%. 

The  length-based  Beverton-Holt  mortality  estimator 
involves  equilibrium  and  other  assumptions  that  may 
make  it  inappropriate  to  use  in  some  cases  (Gedamke 
and  Hoenig,  2006),  but  it  clearly  demonstrates  the  po- 
tential effects  of  errors  in  body-size  measurements  on 
stock  assessment  model  mortality  estimates: 


L-L 


(1) 


where  Z = 
= 

K = 
L = 

Lc  = 


the  instantaneous  rate  of  mortality  from  all 
sources; 

asymptotic  length; 

rate  parameter  from  the  von  Bertalanffy 
growth  equation; 

average  length  of  individuals  in  a sample 
from  the  fishery;  and 

the  “critical”  length  at  which  individuals 
are  fully  vulnerable  to  the  fishery  (Quinn 
and  Deriso,  1999). 


With  all  other  factors  held  constant,  a positive  bias  in 
L will  make  the  numerator  in  Equation  1 too  small,  the 
denominator  too  large,  and  the  mortality  estimate  will 
be  biased  low.  Conversely,  a negative  bias  in  L will  bias 
the  mortality  estimate  high. 

In  this  article,  we  characterize  measurement 
errors  in  shell-height  data  for  sea  scallops  in  two 
types  of  surveys,  using  experimental  data.  The 
experimental  results  are  used  to  evaluate  effects 
on  mean  body  weight  and  swept-area  biomass  es- 
timates, and  on  biomass  and  mortality  estimates 
from  a modern  size-structured  stock  assessment 
model.  The  assessment  model  demonstrates  a 
promising  approach  (used  originally  for  age  data) 
for  accommodating  measurement  errors  in  body- 
size  data.  In  the  appendices,  we  use  numerical  and 
bootstrap  techniques  to  evaluate  robustness  of  the 
assessment  model  approach  in  comparison  to  an 
algebraic  one.  Our  purpose  is  not  to  evaluate  the 
merits  of  any  particular  survey,  rather,  we  use  sea 
scallops  as  an  example  for  dealing  with  general 
problems  arising  from  body-size  measurement  er- 
rors in  survey  and  fishery-dependent  data,  and  for 
suggesting  possible  approaches  to  using  such  data. 


Jacobson  et  al.:  Measurement  errors  in  body  size  of  Placopecten  magellanicus 


235 


Simulated  shell  heights  Simulated  shell  heights 

with  and  without  errors  with  out  errors  and  residuals 

Bias  only 


Figure  1 

Rootograms  (Tukey,  1977)  showing  hypothetical  distributions  of  Atlantic  sea  scallop  ( Placopecten 
magellanicus)  shell-height  (SH)  measurements  with  and  without  simulated  measurement  errors.  The 
black  line  in  each  panel  shows  the  distribution  of  measurements  with  no  errors  (5-mm  size  bins). 
In  the  left  column,  bars  show  distributions  of  shell  heights  with  measurement  errors.  In  the  right 
column,  bars  show  residuals  (measurement  with  no  errors  minus  measurements  with  errors).  For 
the  “bias  only”  scenario  (A  and  B),  precise  measurement  errors  were  assumed  with  a bias  of -4.1 
mm.  For  the  “imprecision  only”  scenario  (C  and  D)  unbiased  measurement  errors  were  assumed 
with  a standard  deviation  of  6.1  mm.  For  the  “imprecision  and  bias”  scenario  (E  and  F),  measure- 
ment errors  were  assumed  with  a bias  of -4.1  mm  and  standard  deviation  of  6.1  mm. 


Materials  and  methods 

The  SMAST  sea  scallop  survey  is  conducted  with  video 
cameras  mounted  on  a steel  pyramid  frame  to  provide 
a 3.24-m2  view  of  the  sea  floor  and  associated  macro- 
benthos (Stokesbury,  2002;  Stokesbury  et  al.,  2004). 
Video  images  are  recorded  at  sea  on  high-resolution 


S-VHS  videotape  and  then  replayed  in  the  laboratory 
where  digitized  images  are  created.  All  sea  scallops 
are  counted,  and  all  clearly  visible  sea  scallops  (with 
the  hinge  and  opposite  edge  visible)  within  the  digi- 
tized images  are  measured  to  the  nearest  mm  by  using 
Image  Pro  Plus®  software  (Media  Cybernetics,  Inc., 
Bethesda,  MD). 


236 


Fishery  Bulletin  108(2) 


In  previous  analyses,  correction  factors  were  applied 
to  the  raw  video  shell-height  measurements  to  account 
for  distance  from  the  origin  (DFO),  which  is  the  dis- 
tance of  a specimen  from  the  “origin”  (center)  of  the 
sampling  frame  (Stokesbury  et  ah,  2004).  Subsequent 
work  during  routine  stock  assessments  (unpublished) 
indicated  that  adjustments  were  unnecessary  because 
the  distributions  of  measurement  errors  were  simpler 
and  easier  to  describe  statistically,  and  data  were  easier 
to  model  without  adjustments.  Moreover,  adjusted  data 
were  sometimes  less  accurate  than  the  unadjusted  data. 
Additional  research  may  result  in  more  accurate  adjust- 
ments or  transformations  of  body-size  data.  However, 
unadjusted  video  data  from  the  “large”  camera  on  the 
sampling  frame  are  used  in  current  stock  assessments 
and  in  this  analysis. 

NEFSC  sea  scallop  surveys  are  conducted  with  a 
2.44-m  New  Bedford  sea  scallop  dredge  with  a 38-mm 
liner.  The  catch  is  sorted,  counted,  and  measured  on  the 
deck  of  the  research  vessel.  In  most  cases,  the  entire 
catch  is  counted  and  measured,  but  a few  large  catches 
were  subsampled.  During  the  early  1980s  through  2003, 
sea  scallops  in  the  catch  were  measured  to  the  nearest 
5-mm  shell-height  interval  with  a standard  NEFSC  sea 
scallop  measuring  board. 

Experiments 

Two  experiments  were  conducted  during  20  and  23  Feb- 
ruary 2003  when  the  SMAST  video  pyramid  was  placed 
in  a 341,000-L  tank  filled  with  seawater  in  the  SMAST 
laboratory.  NEFSC  sea  scallop  measuring  boards  and 
SMAST  video  equipment  in  the  experiments  were  con- 
figured and  used  in  a realistic  manner  that  was  similar 
to  use  during  actual  surveys  at  sea.  Accurate  measure- 
ments used  as  true  shell  heights  in  this  analysis  were 
made  to  the  nearest  mm  by  using  scientific  calipers 
under  laboratory  conditions  with  adequate  lighting. 

We  used  the  experimental  data  to  evaluate  statisti- 
cal characteristics  of  shell-height  composition  data  and 
shell-height  measurement  errors. 

Accuracy,  bias,  and  precision  of  measurements  were 
quantified  by  comparing  data  obtained  from  the  mea- 
suring board  and  video  camera  with  data  from  the 
caliper.  Accuracy  is  the  closeness  to  the  true  underlying 
value  and  is  measured  by  mean  square  error  (MSE).  For 
shell-height  composition  data, 

MSE  = (h-H)2,  (2) 

where  h = the  mean  of  the  measurements;  and 

H = the  mean  of  the  true  values  for  the  sample 
(Cochran,  1977). 

For  measurement  errors  in  our  analysis, 


n 


MSE  = — — , (3) 

n 


where  e-  - h—Hj  = the  error  for  the  jth  observation  (where 
hj  is  the  measurement  and  is  the 
true  value). 

Bias  and  variance  both  contribute  to  MSE.  In  fact, 
MSE  = s2  + b2,  where  s2  is  the  variance  and  b is  bias 
(Cochran,  1977).  In  our  study,  b-h-H  where  h is  the 
mean  of  shell-height  measurements  and  H is  the  mean 
of  the  true  shell  heights  in  the  sample.  Bias  is  the  same 
for  shell-height  composition  data  and  measurement  er- 
rors as  shown  below: 

n 

^ ](hj-Hj)/n  = h-H . (4) 

7=1 

Variance  (s2)  was  computed  from  shell-height  composi- 
tion data  or  measurement  errors  by  using  the  standard 
formula.  Variance  of  shell-height  composition  data  and 
measurement  errors  will  generally  be  different  because 
true  shell  heights  usually  differ  among  specimens  in  a 
sample. 

It  is  convenient  to  express  accuracy,  bias,  and  preci- 
sion in  terms  of  the  square  root  of  the  MSE  (RMSE), 
bias  (b),  and  standard  deviation  (s)  because  all  three 
are  absolute  measures  with  the  same  units  (mm  for  sea 
scallop  shell-height  data).  Percent  RMSE  (RMSE /htrue), 
percent  bias  ( blhtrue ),  and  the  CV  (s/h)  are  useful  for 
making  comparisons  on  a relative  basis. 

The  third  and  fourth  moment  statistics,  gl  and  g2 
were  used  to  measure  skewness  (asymmetry)  and  kur- 
tosis  (peakedness)  of  shell-height  composition  data  and 
measurement  errors,  in  relation  to  what  would  be  ex- 
pected from  a normal  distribution  (Sokal  and  Rohlf, 
1995).  Skewness  and  kurtosis  statistics  for  shell-height 
composition  data  and  measurement  errors  from  the 
same  sample  differ  if  there  is  variability  in  size  among 
specimens.  For  normally  distributed  random  variables 
with  no  skewness,  gr  - 0.  Negative  g1  values  indicate 
skewness  to  the  left  (a  distribution  with  a long  left 
tail  and  more  small  values  than  expected  in  a normal 
distribution).  Positive  g1  values  indicate  skewness  to 
the  right  (long  right  tail  with  more  large  values  than 
expected).  Similarly,  positive  g2  values  indicate  dis- 
tributions more  peaked  than  expected  for  a normal 
distribution,  and  negative  g2  values  indicate  distribu- 
tions that  are  less  peaked  (flatter)  than  expected.  The 
two  statistics  convey  information  about  the  shape  of 
any  distribution  in  relation  to  a normal  distribution, 
but  care  is  required  in  interpreting  g 1 and  g2,  particu- 
larly for  data  that  are  far  from  normally  distributed. 
The  skewness  and  kurtosis  statistics  were  easier  to 
interpret  for  measurement  errors  than  for  shell-height 
measurements  because  the  latter  were  not  normally 
distributed. 

We  used  a test  for  normally  distributed  statistics 
(Sokal  and  Rohlf,  1995)  to  evaluate  the  statistical  sig- 
nificance of  skewness  and  kurtosis  for  distributions  of 
measurement  errors  that  might  be  otherwise  assumed 
normally  distributed.  Statistical  tests  were  carried  out 


Jacobson  et  at:  Measurement  errors  in  body  size  of  Placopecten  magellanicus 


237 


for  distributions  of  measurement  errors  because 
they  were  closer  to  normally  distributed. 

Multiple  shell  height-measurements  were  usu- 
ally made  from  single  specimens  in  our  experi- 
ments. We  made  allowance  for  repeated  sampling 
when  testing  skewness  and  kurtosis  by  using  the 
number  of  unique  specimens  in  the  experiment  as 
the  degrees  of  freedom  instead  of  the  number  of 
measurements  (i.e.,  if  n measurements  were  made 
on  each  of  k specimens,  we  used  k as  the  degrees 
of  freedom  in  statistical  tests).  The  effect  of  this 
adjustment  was  to  make  the  statistical  tests  more 
conservative  (less  likely  to  reject  the  null  hypoth- 
esis of  no  difference).  The  number  of  specimens 
is  a reasonable  lower  bound  estimate  of  the  true 
effective  sample  size. 

Body  weights  for  sea  scallops  and  other  marine 
organisms  are  often  computed  from  body  size.  For 
sea  scallops  in  this  analysis, 

w = ea+pUh)^  (5) 

where  W = sea  scallop  meat  weight  ( g , the  weight 
of  the  marketable  adductor  muscle); 
h = shell  height  (mm);  and  the  parameter 
values  a=-12.01  and  3.22. 

Bland-Altman  plots  (1986,  1995)  were  used  to 
characterize  shell-height  measurement  errors.  In 
the  case  of  measuring  boards,  for  example,  the  dif- 
ference between  the  measuring  board  and  caliper 
shell-height  measurements  for  each  sea  scallop  was 
plotted  on  the  y-axis  against  the  average  of  the 
two  measures  for  the  same  individual  on  the  x- 
axis.  Bland-Altman  plots  are  typically  presented 
as  scatter  plots  with  a point  for  each  difference 
(pair  of  measurements);  however,  boxplots  may  be 
more  useful  in  some  circumstances  (see  below). 
Bland-Altman  plots  are  useful  because  they  elimi- 
nate spurious  correlations  when  the  difference  of 
y-x  is  plotted  against  the  more  precise  measure  (x) 
and  because  patterns  are  easier  to  discern  along  a 
horizontal  line  (the  x-axis)  than  along  a diagonal 
line.  Spurious  correlations  occur  because  the  mea- 
surement error  in  x affects  the  variables  plotted  on 
both  the  x-  and  y-axes. 

Experiment  1 was  designed  to  measure  the  accuracy 
of  video  measurements  for  objects  of  known  size  (square 
ceramic  tiles)  as  a function  of  position  in  the  video 
frame  as  measured  by  DFO  (Fig.  2).  Scuba  divers  in 
experiment  1 placed  black  and  white  ceramic  floor  tiles 
(all  were  48.5x48.5  mm)  in  a closely  packed  square  grid 
on  the  bottom  of  the  tank,  starting  at  the  center  of  the 
video  pyramid  and  covering  the  entire  range  of  view 
in  actual  surveys  (Fig.  2).  The  width  and  height  of  91 
tiles  across  the  field  of  view  and  at  various  distances 
and  positions  from  the  center  of  the  sampling  frame 
(Fig.  2)  were  estimated  from  video  images  by  using  the 
standard  video  survey  procedures  described  above.  Data 
were  recorded  in  such  a way  that  the  length  and  height 


measurements  from  the  same  tile  could  be  associated 
with  each  other  and  with  the  particular  position  of  the 
tile  in  the  video  image.  The  tiles  used  in  experiment  1 
(48.5x48.5  mm)  corresponded  roughly  with  the  size  of 
the  smallest  scallops  fully  recruited  to  the  dredge  and 
video  surveys  (about  40  mm  SH)  and  included  in  stock 
assessment  analyses.  Sea  scallops,  according  to  actual 
survey  data,  cover  a much  wider  range  of  shell  heights 
(to  about  190  mm  SH  in  experiment  2,  see  Discussion 
section). 

Experiment  2 was  designed  to  measure  the  accu- 
racy of  video  shell-height  measurements  for  sea  scal- 
lop shells  of  varying  sizes  (39  to  192  mm  SH)  placed 
randomly  on  a sand-granule-pebble  substrate,  similar 


238 


Fishery  Bulletin  108(2) 


to  the  random  aggregations  observed  on  Georges  Bank. 
All  shell-height  measurements  could  be  linked  with 
each  individual  sea  scallop  in  experiment  2 because 
the  right  valve  of  172  individual  sea  scallop  shells  was 
numbered  uniquely.  The  identification  numbers  were 
large  and  written  under  the  valve  with  dark  indelible 
ink  and  clearly  visible  with  video  equipment  when  the 
sea  scallops  were  turned  over  so  that  the  labels  faced 
the  camera.  The  numbered  sea  scallops  were  assigned 
randomly  to  fifteen  groups.  All  members  of  the  same 
group  were  stored  together  in  a bag  with  a unique  label 
for  group  identification. 

In  each  experimental  replicate,  a group  of  shell 
valves  was  placed  randomly  on  the  bottom  of  the  tank. 
Two  video  images  were  made  for  each  group.  The  first 
image  (with  the  valve  turned  towards  the  sediment 
and  identification  numbers  hidden)  was  used  by  four 
technicians  to  independently  measure  shell  heights. 
The  second  image  was  taken  with  identification  num- 
bers visible  after  divers  turned  the  shells  over  and 
replaced  them  in  their  original  positions.  After  video 
images  were  recorded,  the  shell  valves  were  measured 
with  measuring  boards  by  two  technicians  who  could 
not  see  the  identification  numbers  and  once  by  a third 
technician  with  calipers. 

A stock  assessment  model  that  incorporates 
errors  from  shell-height  measurements 

Following  NEFSC2’3  procedures,  we  used  results  from 
experiment  2 and  a modified  version  of  the  CASA  (catch- 
at-size-analysis,  Sullivan  et  al.,  1990)  stock  assessment 
model  (Appendix  1)  to  investigate  potential  effects  of 
shell-height  measurement  errors  on  model-based  bio- 
mass and  fishing  mortality  estimates  for  two  sea  scallop 
stocks.  Assessment  model  results  in  this  article  should 
not  be  used  by  managers  because  model  runs  were 
tailored  to  investigate  potential  effects  of  shell-height 
measurement  errors  and  because  some  types  of  data 
were  omitted. 

As  described  in  Appendix  1,  the  CASA  model  that 
is  routinely  used  for  sea  scallop  stock  assessments  ac- 
commodates both  bias  and  imprecision  in  shell-height 
measurements.  CASA  models  were  run  for  sea  scal- 
lops in  the  Mid-Atlantic  Bight  during  1982-2006.  In 
contrast  to  NEFSC2,  measurement  error  parameters 
were  obtained  from  experiments  and  not  estimated  in 
the  CASA  model  itself.  The  data  used  in  modeling  in- 
cluded commercial  landings  in  metric  tons  (t),  survey 
trend  data  (numbers  per  unit  of  sampling  effort)  from 
the  camera  video  and  dredge  surveys,  and  shell-height 
composition  data  from  the  commercial  fishery,  video, 
and  dredge  surveys.  Survey  selectivity  patterns  were 
not  estimated  because  the  video  and  dredge  surveys 
have  flat  selectivity  patterns  (catch  sea  scallops  equally 
well)  at  shell  height  >40  mm,  and  goodness-of-fit  calcu- 
lations were  restricted  to  this  size  range  (Appendices 
B7-B8  in  NEFSC3).  Measurement  errors  in  commercial 
shell-height  data  were  assumed  to  be  the  same  as  those 
in  the  dredge  survey  for  lack  of  better  information  and 


because  procedures  for  measuring  sea  scallops  on  land 
in  port  samples  and  at-sea  in  fishery  observer  samples 
are  similar  to  procedures  followed  in  surveys. 

As  described  in  Appendix  1,  bias  and  precision  of 
shell-height  measurements  are  represented  in  the  CASA 
model  by  an  error  matrix  ( E ) that  gives  the  probability 
that  a sea  scallop  in  each  true  shell-height  bin  is  as- 
signed to  a range  of  observed  shell-height  bins  (a  range 
that  accommodates  measurement  errors).  As  described 
by  Methot  (1989,  1990)  for  age  data,  the  error  matrix 
E can  be  set  up  to  deal  with  a wide  range  of  situations 
for  bias  and  variance  (e.g.,  both  can  vary  among  shell- 
height  bins  or  over  time). 

For  the  calculation  of  E for  sea  scallops  in  this  analy- 
sis, shell-height  measurement  error  distributions  were 
assumed  to  be  normally  distributed  with  means  and 
standard  deviations  from  experiment  2.  The  normal 
distributions  for  measurement  errors  were  truncated 
three  standard  deviations  above  and  below  the  mean. 
In  calculating  distributions  of  measurement  errors,  true 
shell  heights  were  assumed  with  or  without  bias  to  be 
uniformly  distributed  within  each  true  5-mm  SH  bin 
so  that,  for  example,  the  frequency  of  sea  scallops  with 
true  shell  heights  of  70,  71,  72,  73,  and  74  mm  (in  the 
70-74.9  mm  SH  bin  with  midpoint  72.5)  was  the  same. 
Distributions  for  measurement  errors  were  normalized 
to  sum  to  one  before  use  in  the  CASA  model. 


Results 

Height  and  width  measurements  from  the  same  tiles 
in  experiment  1 were  not  significantly  different  by  a 
paired  t-test  (£=-0.23,  P=0.30,  91  df).  Therefore,  height 
and  width  measurements  from  91  tiles  in  experiment  1 
were  combined  to  form  a single  set  of  video  data  (a  total 
of  182  measurements)  (Table  1). 

The  RMSE  statistic  for  video  tile-size  composition 
and  measurement  errors  in  experiment  1 (Table  1)  was 
3.5  mm  (%RMSE  = 7%,  Table  1).  Bias  (-2.2  mm)  and 
imprecision  (standard  deviation  2.7  mm)  of  video  tile 
measurements  were  similar.  In  comparison  to  the  true 
size  of  the  tiles  (48.5  mm),  the  smallest  measurement 
was  38  mm,  and  the  largest  measurement  was  50  mm. 
The  video  size-composition  data  and  measurement  er- 
rors were  left  skewed  (g1=-0.28)  and  flatter  (g2=-0.53) 
than  expected  for  a normal  distribution.  There  were 
gaps  in  the  distribution  of  the  video  tile  measurements 
(Fig.  3)  due  to  the  resolution  of  the  video  images  used 
in  digitizing  (each  pixef=3x3  mm). 

Measurement  error  increased  with  DFO  for  the  video 
tile  measurements  (Fig.  3).  Bias  was  positive  for  DFO 
<400  mm  and  negative  at  larger  DFO  levels. 

RMSE  for  shell-height  composition  data  in  experi- 
ment 2 was  33  mm  (%RMSE  30%)  for  video  and  34  mm 
(%RMSE  = 31%)  for  measuring  board  data  (Table  2). 
Mean  shell  height  was  106  mm  for  video  and  109  mm 
for  measuring  boards,  compared  to  110  mm  for  calipers. 
Minimum  shell  height  was  34  mm  for  video,  38  mm  for 
measuring  boards,  and  39  mm  for  calipers.  Maximum 


Jacobson  et  al.:  Measurement  errors  in  body  size  of  Placopecten  mage/lanicus 


239 


shell  height  was  201  mm  for  video,  193  mm  for  measur- 
ing boards,  and  192  mm  for  calipers. 

Bland-Altman  plots  for  experiment  2 show  that  mea- 
suring board  shell  heights  were  more  accurate  than 
video  measurements,  and  that  bias  in  video  and  mea- 
suring board  data  was  relatively  constant  across  the 
range  of  shell  heights  in  experiment  2 (Fig.  4).  However, 
relatively  large  outliers  sometimes  occurred  in  video 
measurements  at  80-130  mm  SH  (Fig.  4). 

Video  and  measuring-board  shell-height  compositions 
in  experiment  2 were  similar  in  terms  of  skewness  with 
gj=-0.41  for  video  measurements  and  -0.47  for  measur- 
ing boards  compared  to  -0.46  for  calipers  (Table  2).  The 
video  shell-height  distribution  was  more  peaked  with 
g2~- 0.65  compared  to  g2=- 0.85  for  measuring  boards, 
and  g2=-0.84  for  calipers  (Table  2).  Video  measurement 
errors  were  skewed  to  the  left  (g1=-0.60)  compared  to 
measuring-board  errors  which  were  nearly  symmetrical 
(^1=-0.05).  The  distribution  of  errors  for  measuring 
boards  was  flatter  (g2~- 0.85)  and  video  measurement 
errors  were  more  peaked  (g2- 1-84)  than  would  be  ex- 
pected for  normal  distribution.  The  error  distribution 
for  measuring  boards  had  a nearly  flat  mode  about 
5-mm  wide  because  shell  heights  are  automatically 
truncated  by  measuring  boards  to  the  next  lowest  5-mm 
shell-height  bin. 

On  a proportional  basis,  meat  weights  calculated  from 
shell  heights  in  experiment  2 were  much  less  accu- 
rate than  the  original  shell-height  measurements.  In 


Table  1 

Summary  of  size-composition  data  and  measurement 
errors  for  182  tile  measurements  (height  and  width  from 
91  tiles,  each  48.5x48.5  mm)  by  video  equipment  in  exper- 
iment 1. 

Statistic 

Video 

Measurements  and  measurement  errors 

Bias 

-2.2 

Standard  deviation 

2.7 

Square  root  of  the  mean  squared  error 

3.5 

Skewness  (gq) 

-0.28 

Kurtosis  (g2) 

-0.53 

Measurements 

Minimum 

38.3 

5%  quantile 

41.2 

95%  quantile 

50.1 

Maximum 

50.1 

Average 

46.3 

Percent  bias 

-5% 

Coefficient  of  variation 

6% 

Percent  square  root  of  the  mean  squared 

error  7% 

particular,  %RMSE  values  for  meat  weights  were  71% 
and  74%  for  video  and  measuring  boards,  respectively 
(Table  3),  compared  to  30%  and  31%  for  the  original 


Table  2 

Summary  statistics  for  shell-height  composition  data  and  measurement  errors  (in  mm)  from  172  uniquely  identified  Atlantic  sea 
scallop  ( Placopecten  magellanicus)  shell  valves  in  experiment  2.  “NA”  means  that  a statistic  is  not  applicable. 

Statistic 

True  shell  height  (calipers) 

Video 

Measuring  boards 

Shell  heights  and  measurement  errors 

n measurements  used 

172 

670 

344 

n omitted 

0 

18 

0 

Bias 

NA 

-4.5 

-0.6 

Shell  heights 

Minimum 

38.5 

34.3 

37.5 

5%  quantile 

54.8 

48.8 

52.5 

95%  quantile 

149.6 

147.3 

147.5 

Maximum 

192.0 

200.6 

192.5 

Average 

109.9 

106.5 

109.3 

Percent  bias 

NA 

-4% 

-1% 

Standard  deviation 

33.5 

33.1 

33.6 

Coefficient  of  variation 

30% 

31% 

31% 

Square  root  of  the  mean  squared  error 

NA 

33.4 

33.6 

Percent  square  root  of  the  mean  squared  error 

NA 

30% 

31% 

Skewness  (gx) 

-0.46 

-0.41 

-0.47 

Kurtosis  (g2) 

-0.84 

-0.65 

-0.85 

Measurement  errors 

Standard  deviation 

NA 

6.1 

1.7 

Square  root  of  the  mean  squared  error 

NA 

7.6 

1.8 

Skewness  (gj) 

NA 

-0.60 

-0.044 

Kurtosis  (g2) 

NA 

1.84 

-0.85 

240 


Fishery  Bulletin  108(2) 


shell  heights  (Table  2).  The  nonlinear  shell-height  to 
meat-weight  relationship  showed  exaggerated  extremes 
of  the  distributions  so  that  the  ratio  of  maximum  to 
mean  meat  weight  was  158/27=5.9  for  video  data  and 
138/29  = 4.8  for  measuring  boards  (Table  3)  compared  to 
201/106=1.9  and  193/109=1.8  for  shell  heights  (Table  2). 
Variance  in  meat-weight  measurements  increases  as 
true  meat-weight  increases  for  video  data  and,  to  a 
lesser  extent,  for  measuring  boards  (Fig.  5). 

The  meat-weight  composition  data  were  more  right 
skewed  (gj=1.53)  and  flatter  (g2= 6.22)  than  the  meat- 
weight  composition  data  from  measuring  boards 
(^1=0.92  and  ^2=2.61)  or  calipers  (^=0.99  and  g2=3.00). 
Errors  in  meat-weight  data  were  left  skewed  and  not  as 
peaked  for  video  (^1=-0.80  and  g2=2.48)  than  measur- 
ing board  data  (^1=-1.06  andg\2  = 4.68). 


801  A 


Measurement  (mm) 


i 1 1 1 1 1 


0 200  400  600  800  1000 

Distance  from  origin  (mm) 

Figure  3 

(A)  Video  measurements  for  tiles  in  experiment  1.  The  verti- 
cal line  shows  the  true  value  at  48.5  mm.  ( B)  Measurement 
errors  (video  measurement  minus  caliper  measurement) 
for  tiles  in  experiment  1 as  a function  of  distance  from  the 
origin  (DFO).  The  nonlinear  LOESS  regression  line  shows 
the  overall  trend  in  measurement  errors  as  a function  of 
DFO. 


Results  from  the  assessment  models 

Based  on  results  from  experiment  2 (Table  2)  and 
assumptions  listed  above,  video  shell-height  measure- 
ments for  sea  scallops  with  true  sizes  evenly  distributed 
over  100-104.99  mm  SH  (i.e.,  the  100-mm  bin  with 
midpoint  102.5  mm)  would  fall  into  nine  observed  shell- 
height  bins  with  midpoints  from  77.5  to  117.5  mm  (Table 
4).  Measuring  board  shell-height  measurements  would 
fall  into  four  observed  shell-height  bins  with  midpoints 
ranging  from  92.5  to  107.5  mm  (Table  4). 

Four  model  configurations  were  used.  The  “no  mea- 
surement error”  model  configuration  was  fitted  by  as- 
suming no  errors  in  shell-height  data.  The  “bias  only” 
model  was  fitted  by  assuming  that  shell-height  data 
were  biased  (to  the  extent  measured  in  experiment  2), 
but  precise  (with  zero  variance).  The  “imprecision 
only”  model  was  fitted  by  assuming  that  shell-height 
measurements  were  imprecise  (standard  deviations 
from  experiment  2),  but  not  biased.  The  “impreci- 
sion and  bias”  model  was  fitted  by  assuming  both 
types  of  shell-height  measurement  errors. 

Models  which  accommodated  measurement  errors 
fitted  better,  with  substantially  lower  negative  log 
likelihoods  for  both  stocks,  than  models  that  ig- 
nored measurement  errors.  Differences  in  negative 
log  likelihood  were  mostly  for  shell-height  compo- 
sition data.  Mean  2004-06  biomass  and  fishing 
mortality  rates  and  coefficients  of  variation  (CV) 
for  biomass  and  fishing  mortality  estimates  were 
similar  for  all  model  configurations  (Table  5). 

Discussion 

The  importance  of  body-size  measurement  errors 
and  the  need  to  accommodate  them  in  modeling 
probably  depends  on  the  situation.  Biological  factors 
(growth  rate,  recruitment  variability),  assessment 
model  type,  quality  and  quantity  of  fishery  and  fish- 
ery-independent data  may  be  important.  Sea  scallops 
may  be  an  atypical  case  because  they  are  a data-rich 
species.  We  suggest  that  the  potential  importance  of 
body  size  measurement  errors  should  be  evaluated 
on  a case  by  case  basis,  particularly  if  body-size  data 
may  be  imprecise  or  biased.  Simulation  studies  may 
be  useful  in  determining  the  importance  of  experi- 
mentally derived  body-size  measurement  errors  on 
stock  assessment  results. 

In  the  sea  scallop  case,  models  that  accommodat- 
ed measurement  errors  fitted  substantially  better, 
but  there  was  little  effect  on  point  estimates  and 
variances  for  recent  biomass  and  fishing  mortality. 
We  hypothesize  that  effects  on  biomass  and  mortal- 
ity estimates  would  be  larger  in  cases  with  positive 
biases  in  body-size  measurements.  For  both  video 
and  measuring  boards,  the  positive  bias  in  meat 
weights  due  to  the  nonlinear  relationship  between 
body  size  and  meat  weight  was  mitigated  to  some 
extent  by  the  negative  bias  in  shell-height  mea- 


Jacobson  et  al.:  Measurement  errors  in  body  size  of  Placopecten  magellanicus 


241 


Measurement 


Figure  4 

Modified  Bland-Altman  plots  for  Atlantic  sea  scallop  ( Placopecten 
magellanicus)  shell-height  (SH)  measurements  in  experiment  2.  The 
y-axis  shows  the  difference  between  the  experimental  measurement 
(measuring  boards  in  A or  video  in  B)  and  the  caliper  measure- 
ment. The  jc-axis  shows  the  average  of  the  experimental  and  caliper 
measurement.  Boxplots  and  30-mm  shell-height  bins  were  used 
instead  of  traditional  scatter  plots  for  shell  height  measurements 
in  experiment  2 because  the  large  number  of  samples  between  120 
and  150  mm  SH  gave  the  impression  that  variance  was  higher  for 
those  sizes.  Boxplots  show  the  interquartile  range  (a  robust  vari- 
ance measure)  and  are  not  sensitive  to  sample  size.  The  width  of 
the  boxplots  is  proportional  to  the  number  of  observations  for  the 
shell-height  bin. 


surements.  In  contrast,  Heery  and  Berkson 
(2009)  used  simulations  to  evaluate  effects  of 
systematic  sampling  errors  (too  many  small 
or  too  many  large  individuals)  in  size-com- 
position data  from  commercial  catches  and 
three  simulated  stocks.  The  simulated  data 
were  used  in  a forward-projecting  age-struc- 
tured stock  assessment  and  in  projection 
models  to  estimate  stock  size  and  fishing 
mortality  in  relation  to  threshold  values,  and 
rebuilding  trajectories.  Body-size  data  with 
too  many  large  individuals  biased  stock  size 
high  and  fishing  mortality  low  and  tended  to 
support  management  measures  that  did  not 
meet  management  goals,  particularly  for  lon- 
ger lived  and  depleted  stocks.  Body-size  data 
with  too  many  small  individuals  were  less 
problematic,  but  tended  to  support  overly 
restrictive  management  actions  in  extreme 
cases.  Heery  and  Berkson’s  (2009)  results 
indicate  that  systematic  errors  in  sampling 
may  be  more  important  than  errors  in  indi- 
vidual measurements  of  body  size. 

Variance  in  calculated  meat  weights  in- 
creased rapidly  with  shell  height  with  both 
video  and  measuring  board  techniques,  in 
contrast  to  the  variance  in  shell  heights 
(Figs.  4 and  5).  This  additional  source  of 
variability  likely  increases  variance  in  bio- 
mass estimates,  particularly  for  relatively 
large  fishable  sea  scallops. 

In  our  analysis,  assessment  models  that 
accommodated  shell-height  measurement  er- 
rors fitted  better,  even  though  no  additional 
parameters  were  estimated.  The  Mid-Atlan- 
tic Bight  model  that  accommodated  impre- 
cise (but  not  biased)  shell-height  measure- 
ment errors  had  a negative  log  likelihood 
that  was  15  units  smaller  than  the  negative 
log  likelihood  for  the  no  measurement  er- 
ror model  (Table  5).  Results  for  the  Georges 
Bank  stock  (not  shown  to  conserve  space)  were  similar. 
In  contrast  and  based  on  likelihood  theory,  a difference 
in  negative  log  likelihoods  of  just  1.92  units  is  sufficient 
to  justify  an  additional  parameter  in  a statistical  model 
at  the  P= 0.05  level  (Venzon  and  Moolgavkar,  1988). 
Comparing  results  of  the  “bias  only”  scenario  to  results 
from  the  “imprecision  only”  and  “imprecision  and  bias” 
scenarios,  we  found  that  improvements  in  goodness  of 
fit  were  mostly  due  to  accommodating  imprecision;  bias 
was  less  important  (Table  5). 

Experiments 

Our  results  highlight  the  value  and  information  that 
may  be  gained  from  evaluating  body  size  measurement 
errors  experimentally.  Body-size  measurement  error 
experiments  should  be  conducted  when  survey  equip- 
ment is  changed,  particularly  if  body-size  measurements 
are  imprecise.  In  some  cases,  frequent  “mini-experi- 


ments” may  be  required  if  the  accuracy  of  the  equip- 
ment tends  to  drift  over  time  or  change  in  response  to 
environmental  conditions. 

Our  results  indicate  the  importance  of  designing  mea- 
surement error  experiments  so  that  individual  speci- 
mens can  be  identified  and  associated  with  individual 
measurements;  otherwise  measurement  errors  can  not 
be  estimated  individually  and  evaluated  directly.  Data 
from  experiment  2 were  most  useful  because  individual 
sea  scallops  were  numbered  and  replicate  measure- 
ments of  different  types  could  be  linked  and  analyzed 
in  detail.  In  addition,  the  full  range  of  variability  for  all 
important  factors  (i.e.,  distance  from  the  origin  (DFO), 
shell  height,  and  identity  of  individual  technicians) 
should  be  included  in  the  experimental  design. 

We  ignored  skewness  and  kurtosis  in  measurement 
errors  in  calculating  measurement  error  matrices  for 
use  in  the  CASA  stock  assessment  model.  In  future 
modeling,  it  may  be  better  to  use  the  experimental  dis- 


242 


Fishery  Bulletin  108(2) 


Table  3 

Summary  statistics  of  meat  weights  and  meat  weight  measurement  errors  (g)  for  Atlantic  sea  scallop  (Placopecten  magellanicus) 
shell-height  measurements  in  experiment  2 (sample  sizes  are  the  same  as  those  for  shell-height  measurements  in  Table  2).  The 
original  shell  heights  were  obtained  with  calipers,  video  camera,  and  measure  boards.  “NA”  means  that  a statistic  is  not  applicable. 

Statistic 

True  (calipers) 

Video 

Measuring  boards 

Meat  weights  and  measurement  errors 

Bias 

NA 

-3.2 

-0.4 

Meat  weights 

Minimum 

0.8 

0.5 

0.7 

5%  quantile 

2.4 

1.7 

2.1 

95%  quantile 

61.3 

58.3 

58.6 

Maximum 

136.9 

157.7 

138.0 

Average 

29.8 

27.3 

29.4 

Percent  bias 

NA 

-10% 

-1% 

Standard  deviation 

22.2 

21.4 

21.8 

Coefficient  of  deviation 

74% 

78% 

74% 

Square  root  of  the  mean  squared  error 

NA 

21.6 

21.8 

Percent  square  root  of  the  mean  squared  error 

NA 

71% 

74% 

Skewness  (gx) 

0.99 

1.53 

0.92 

Kurtosis  (g2) 

3.00 

6.22 

2.61 

Measurement  errors 

Standard  deviation 

NA 

5.1 

1.5 

Square  root  of  the  mean  squared  error 

NA 

6.0 

1.6 

Skewness  (g^) 

NA 

-0.80 

-1.06 

Kurtosis  ( g2 ) 

NA 

2.48 

4.68 

Table  4 

Estimated  probability  distributions  for  Atlantic  sea  scallop  ( Placopecten  magellanicus)  shell-height  (SH)  measurements  based  on 
bias  and  standard  deviations  from  experiment  2.  Condition  factors  for  error  matrices  used  in  the  catch-at-size-analysis  (CASA) 
stock  assessment  model  scenarios  are  given  also.  The  shell-height  bins  are  5-mm  wide  and  identified  by  their  midpoint.  For 
example,  sea  scallops  80-84.9  mm  SH  fall  into  a bin  whose  midpoint  is  82.5  mm. 

Video  scenario 

Measuring  board  scenario 

Calipers 

Imprecision 

Imprecision 

Imprecision 

Imprecision 

Statistic  (true  shell  height) 

Bias  only 

only 

and  bias 

Bias  only 

only 

and  bias 

Condition  factor  (k) 

NA 

3xl015 

5457 

2638 

1.6 

2.1 

2.3 

Bias  (mm) 

0 

-4.5 

0 

-4.5 

-0.6 

0 

-0.6 

Standard  deviation  (mm) 

0 

0 

6.1 

6.1 

0 

1.7 

1.7 

Shell  height  bin  (mm) 

Probability  of  observed  bins 

72.5 

77.5 

0.0009 

82.5 

0.0014 

0.0167 

87.5 

0.0203 

0.0820 

92.5 

0.0929 

0.2158 

0.0001 

97.5 

0.8000 

0.2300 

0.3101 

0.2000 

0.1325 

0.2008 

102.5 

1.0000 

0.2000 

0.3110 

0.2436 

0.8000 

0.7349 

0.7181 

107.5 

0.2300 

0.1045 

0.1325 

0.0810 

112.5 

0.0929 

0.0243 

117.5 

0.0203 

0.0020 

122.5 

0.0014 

127.5 

Jacobson  et  al.:  Measurement  errors  in  body  size  of  Placopecten  magellanicus 


243 


50  100 

Mean  shell  height  (mm) 

Figure  5 

Bland-Altman  plots  for  Atlantic  sea  scallop  ( Placopecten  mag- 
ellanicus) meat  weights  calculated  from  experimental  shell- 
height  measurements  in  experiment  2 (measuring  boards  in 
panel  A and  video  in  panel  B).  The  y-axis  shows  the  difference 
between  the  meat  weights  calculated  from  the  experimental 
(video  or  measuring  board)  shell  height  measurements  and  the 
meat  weights  calculated  from  caliper  measurements.  The  x-axis 
shows  the  average  of  the  experimental  and  caliper-derived 
measurements. 


tributions  of  measurement  errors  directly  in  er- 
ror matrices,  particularly  if  experimental  sample 
sizes  are  large. 

Drouineau  et  al.  (2008)  used  simulation  analy- 
sis to  show  the  importance  of  alternative  as- 
sumptions about  the  distribution  of  individuals 
within  size  groups  and  the  statistical  distribu- 
tion of  growth  increments  in  length-structured 
models  like  the  CASA  (catch-at-size-analysis) 
model.  Our  experience  indicates  that  the  same 
types  of  assumptions  are  important  in  calcu- 
lating body-size  measurement-error  matrices. 

In  particular,  it  was  important  to  assume  that 
individuals  were  uniformly  distributed  within 
size  groups,  to  make  realistic  assumptions  about 
the  distributions  of  measurement  errors,  and  to 
be  careful  in  programming  to  ensure  consistent 
calculations  at  the  boundaries  of  length  bins  for 
calculating  error  matrices  and  for  the  stock  as- 
sessment model. 

Statistical  methods  for  repeated  measurements 
or  random  effects  may  be  suitable  for  analysis  of 
our  experimental  data.  We  made  allowances  for 
repeated  measures  in  bootstrap  calculations  (Ap- 
pendix 2)  and  in  calculating  P-values  for  skew- 
ness and  kurtosis  tests,  but  not  in  calculating 
other  statistics  (Tables  1-3). 

Our  experiments  were  conducted  under  ideal 
conditions  with  tiles  and  shell  valves,  rather 
than  live  sea  scallops.  Our  results  may  under- 
estimate the  magnitude  of  errors  under  more 
realistic  field  conditions. 

Model  results  may  depend  on  shell-height  bin 
width  such  that  larger  shell  height  bins  would 
cause  measurement  errors  to  have  a greater  im- 
pact on  biomass  and  mortality  estimates.  We 
used  5-mm  SH  bins  for  sea  scallops  because  5- 
mm  is  the  resolution  and  approximate  accuracy 
for  the  survey  shell-height  data.  In  general,  it 
may  be  important  to  consider  the  magnitude  of 
measurement  errors  in  making  decisions  about  size  bins 
used  in  stock  assessment  modeling. 

Body-size  measurement  errors 

Random  measurement  errors  are  unavoidable.  One  may 
conclude  that  it  is  incumbent  on  the  researcher  to  search 
out  and  correct  sources  of  bias,  whatever  the  source.  We 
suggest  that  it  may  be  more  cost  effective  to  quantify 
measurement  errors  experimentally  and  to  accommodate 
them  in  modeling.  Time  series  with  consistent  body-size 
measurement  errors  are  probably  easiest  to  interpret. 
Models  may  become  overly  complex  if  multiple  sets  of 
assumptions  about  measurement  errors  are  required 
to  interpret  one  survey  time  series.  Resources  required 
to  quantify  measurement  errors  after  each  adjustment 
to  survey  procedures  or  equipment  may  be  better  spent 
on  more  accurately  characterizing  the  measurement 
errors  for  survey  gear  that  remains  the  same  for  longer 
periods  of  time. 


Bootstrap  results  also  showed  that  an  algebraic  ap- 
proach to  removing  errors  from  the  data  by  using  the 
inverse  error  matrix  E _1  gave  negative  proportions  for 
both  video  and  measuring  board  data  in  at  least  some 
size  groups  (Appendix  2).  The  sampling  distribution  for 
algebraically  adjusted  shell-height  data  may  be  difficult 
to  characterize.  These  results  indicate  that  it  may  be 
difficult  to  remove  measurement  errors  directly  from 
body-size  data  and  we  hypothesize  that  approaches  like 
the  one  used  in  the  CASA  model  will  generally  perform 
better.  Bootstrap  results  showed  that  estimates  of  pre- 
dicted shell-height  composition  data  with  measurement 
errors  as  carried  out  in  the  CASA  model  were  robust 
to  uncertainties  in  the  measurement-error  matrix  E 
(Appendix  2).  Models  can  be  designed  to  be  robust  to 
measurement  errors.  For  example,  the  last  size  bin  in 
the  CASA  model  is  a plus-group  that  absorbs  data  for 
large  scallops  that  may  have  been  strongly  affected 
by  measurement  errors.  Other  data  in  the  model  may 
have  also  contributed  to  the  robustness  of  biomass  and 


244 


Fishery  Bulletin  108(2) 


Table  5 

Results  from  the  catch-at-size-analysis  (CASA)  model  for  Mid-Atlantic  Bight  sea  scallops  (Placopecten  magellanicus)  and  four 
model  configurations.  The  “no  measurement  error”  model  configuration  does  not  accommodate  shell-height  measurement  errors. 
Other  model  configurations  accommodate  bias  and  imprecise  measurement  errors  in  various  combinations  as  shown  in  the  table. 
Lower  negative  log  likelihood  (NLL)  values  indicate  better  model  fit.  Coefficients  of  variation  (CV)  shown  in  parenthesis  are 
asymptotic  variances  calculated  by  the  delta  method.  For  ease  of  comparison,  the  “no  measurement  error”  configuration  NLL 
values  were  subtracted  from  corresponding  NLL  statistics  for  all  three  configurations.  The  lowest  NLL,  biomass  or  fishing  mor- 
tality estimates  in  each  row  are  printed  in  boldface. 

Variable  or  estimate 

No 

measurement 

error 

Bias  only 

Imprecision 

only 

Imprecision 

and 

bias 

Bias  and  precision  (mm)  assumed  in  modeling 

Standard  deviation — video  survey 

0.0 

0.0 

6.1 

6.1 

Bias — video  survey 

0.0 

-4.5 

0.0 

-4.5 

Standard  deviation — dredge  survey 

0.0 

0.0 

1.7 

1.7 

Bias — survey 

0.0 

-0.6 

0.0 

-0.6 

Negative  log  likelihood  (NLL) 

Total 

0.00 

20.92 

-14.62 

-1.16 

Commercial  fishery  shell-height  data 

0.00 

4.99 

-0.34 

2.06 

Dredge  survey  shell-height  data 

0.00 

-4.14 

-10.66 

-6.97 

Video  survey  shell-height  data 

0.00 

19.45 

-3.00 

4.59 

Mean  biomass  and  fishing  mortality  during  2004-06 

Fishing  mortality  (y-1) 

0.45 

0.41 

0.46 

0.42 

(8%) 

(7%) 

(8%) 

(8%) 

Biomass  (t  meats) 

81,211 

84,650 

80,844 

83,602 

(5%) 

(5%) 

(5%) 

(5%) 

fishing  mortality  estimates  to  assumptions  about  shell- 
height  measurement  errors. 

In  principal,  measurement-error  parameters  could  be 
estimated  directly  in  stock  assessment  models  without 
resorting  to  experiments.  Measurement-error  param- 
eters in  the  CASA  model  were  estimated  in  the  NEFSC 
study,2  but  the  estimates  proved  to  be  unstable  (NEF- 
SC3). Without  at  least  one  source  of  accurate  body-size 
data,  there  may  be  too  little  information  about  mea- 
surement errors  to  estimate  parameters.  In  addition, 
there  may  be  strong  correlations  between  estimated 
measurement  errors  and  estimates  of  other  factors  that 
affect  interpretation  of  body-size  data,  such  as  survey 
and  fishery  selectivity,  natural  mortality,  and  recruit- 
ment variability. 

Acknowledgements 

We  thank  F.  Serchuk  (Northeast  Fisheries  Science 
Center,  Woods  Hole,  MA),  S.  Correia  (Massachusetts 
Division  of  Marine  Fisheries,  New  Bedford,  MA),  C. 
O’Keefe  and  C.  Adams  (SMAST,  New  Bedford,  MA), 
and  five  anonymous  reviewers  for  useful  technical  and 
editorial  suggestions.  We  are  grateful  for  support  from 
the  School  of  Marine  Science  and  Technology,  the  Mas- 
sachusetts Division  of  Marine  Fisheries,  and  NOAA 
awards:  NA04NMF4720332,  NA05NMF4721131,  and 


NA06NMF4720097.  We  are  grateful  to  the  crews  and 
scientific  staff  who  collected  and  measured  sea  scallops 
in  NEFSC  and  SMAST  surveys.  Live  sea  scallops  used  in 
the  experiments  were  provided  by  commercial  sea  scallop 
vessels  from  New  Bedford  and  Fairhaven,  MA. 

Literature  cited 

Bland,  J.  M.,  and  D.  G.  Altman. 

1986.  Statistical  methods  for  assessing  agreement 
between  two  methods  of  clinical  measurement.  Lancet 
1:307-310. 

1995.  Comparing  methods  of  measurement:  why  plotting 
difference  against  standard  method  is  misleading.  Lan- 
cet 346:1085-1087. 

Butler,  J.,  M.  Neuman,  D.  Pinkard,  R.  Kvitek,  and  G.  Cochrane. 

2006.  The  use  of  sonar  mapping  techniques  to  refine 
population  estimates  of  the  endangered  white  abalone 
( Haliotis  sorenseni).  Fish.  Bull.  104:521-532. 
Cochran,  W.  G. 

1977.  Sampling  techniques,  428  p.  John  Wiley  and  Sons, 
Inc.,  New  York. 

Drouineau,  H.,  S.  Mahevas,  M.  Bertignac,  and  A.  Fertin. 

2008.  Assessing  the  impact  of  discretisation  assumptions 
in  a length-structured  population  growth  model.  Fish. 
Res.  91:160-167. 

Edgar,  G.  J.,  N.  S.  Barrett,  and  A.  J.  Morton. 

2004.  Bias  associated  with  the  use  of  underwater 
visual  census  techniques  to  quantify  the  density  and 


Jacobson  et  al.:  Measurement  errors  in  body  size  of  Placopecten  magellanicus 


245 


size-structure  of  fish  populations.  J.  Exp.  Biol.  Ecol. 
308:269-290. 

Feller.  W. 

1966.  An  introduction  to  probability  theory  and  its  appli- 
cations. Volume  II.  626  p.  John  Wiley  and  Sons,  Inc., 
New  York. 

Fournier,  D.,  and  C .P.  Archibald. 

1982.  General  theory  for  analyzing  catch  at  age 
data.  Can.  J.  Fish.  Aquat.  Sci.  39:1195-1207. 

Gedamke,  T.,  and  J.  M.  Hoenig. 

2006.  Estimating  mortality  from  mean  length  data 
in  nonequilibrium  situations,  with  application  to 
the  assessment  of  goosefish.  Trans.  Am.  Fish.  Soc. 
135:476-487. 

Heery,  E.  C.,  and  J.  Berkson. 

2009.  Systematic  errors  in  length  frequency  data  and  their 
effect  on  age-structured  stock  assessment  models  and 
management.  Trans.  Am.  Fish.  Soc.  138:218-232. 

Horn,  R.  A.,  and  C.  R.  Johnson. 

1985.  Matrix  analysis,  561  p.  Cambridge  Univ.  Press, 
New  York. 

Methot,  R.  D. 

1989.  Synthetic  estimates  of  historical  abundance  and 
mortality  for  northern  anchovy.  Am.  Fish.  Soc.  Symp. 
6:66-82. 

1990.  Synthesis  model:  an  adaptive  framework  for  analy- 
sis of  diverse  stock  assessment  data.  Int.  N.  Pacific 
Fish.  Comm.  Bull.  50:259-277. 

Pennington,  M.,  L.-M.  Burmeister,  and  V.  Hjellvik. 

2001.  Assessing  the  precision  of  frequency  distribu- 
tions estimated  from  trawl-survey  samples.  Fish. 
Bull.  100:74-80. 

Quinn,  T.,  and  R.  B.  Deriso. 

1999.  Quantitative  fish  dynamics,  570  p.  Oxford  Univ. 
Press,  Oxford,  U K. 

Ricker,  W.  E. 

1975.  Comparison  and  interpretation  of  biological  sta- 
tistics of  fish  populations.  Fish.  Res.  Board  Can.  Bull. 
191:1-382. 

Rosenkranz,  G.  E.,  and  S.  C.  Byersdorfer. 

2004.  Video  scallop  survey  in  the  eastern  Gulf  of  Alaska, 
USA.  Fish.  Res.  69:131-140. 

Serchuk,  F.  M,  and  S.  E.  Wigley. 

1986.  Evaluation  of  USA  and  Canadian  research  vessel 
surveys  for  sea  scallops,  (Placopecten  magellanicus)  on 
Georges  Bank.  J.  Northw.  Atl.  Fish.  Sci.  7:1-13. 

Serchuk,  F.  M.,  P.  W.  Wood  Jr.,  J.  A.  Posgay,  and  B.  E.  Brown. 

1979.  Assessment  and  status  of  sea  scallop  (Placopec- 
ten magellanicus),  populations  off  the  Northeast  coast 
of  the  United  States.  Proc.  Natl.  Shellfish.  Assoc. 
69:161-191. 

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

1995.  Biometry,  859  p.  Freeman,  New  York. 

St.  John,  J.,  G.  R.  Russ,  and  W.  Gladstone. 

1990.  Accuracy  and  bias  of  visual  estimates  of  numbers, 
size  structure,  and  biomass  of  a coral  reef  fish.  Mar. 
Ecol.  Prog.  Ser.  64:253-262. 

Stokesbury,  K.  D.  E. 

2002.  Estimation  of  sea  scallop  abundance  in  closed 
areas  of  Georges  Bank,  USA.  Trans.  Am.  Fish.  Soc. 
131:1081-1092. 

Stokesbury,  K.  D.  E.,  B.  P.  Harris,  M.  C.  Marino  II,  and  J.  I. 

Nogueira. 

2004.  Estimation  of  sea  scallop  abundance  using  a 
video  survey  in  off-shore  US  waters.  J.  Shellfish  Res. 
23:33-44. 


Sullivan,  P.  J.,  H.-L.  Lai,  and  V.  F.  Gallucci. 

1990.  A catch-at-length  analysis  that  incorporates  a 
stochastic  model  of  growth.  Can.  J.  Fish.  Aquat.  Sci. 
47:184-198. 

Venzon,  D.  J.,  and  S.  H.  Moolgavkar. 

1988.  A method  for  computing  profile-likelihood  based 
confidence  intervals.  Appl.  Stat.  37:87-94. 

Tukey,  J.  W. 

1977.  Exploratory  data  analysis,  688  p.  Addison-Wesley 
Pub.  Co.,  Reading,  MA. 

Appendix  1 

Following  the  approach  of  the  Northeast  Fisheries  Sci- 
ence Center  (NEFSC, 2>3)  we  used  a likelihood  approach 
to  fitting  the  CASA  model  to  sea  scallop  stock  assessment 
data.  The  best  estimates  from  the  model  minimized  the 
combined  negative  log  likelihood  of  all  the  data.  Relevant 
details  are  described  below.  Appendix  B10  in  the  NEFSC 
report  (NEFSC3)  is  a complete  technical  description  of 
the  CASA  model  for  sea  scallops.  Appendix  B12  in  that 
same  report  (NEFSC3)  describes  CASA  model  perfor- 
mance with  simulated  stock  assessment  data. 

Estimates  of  population  abundance  and  survey  size 
selectivity  are  available  for  each  shell  height  and  year 
as  the  CASA  model  is  fitted.  In  a single  year,  for  ex- 
ample, we  calculated  the  number  of  sea  scallops  in  the 
population  that  were  available  or  selected  by  the  video 
gear  with  the  following  equation: 

nh=QhNh > (A1> 

where  Nh  = the  predicted  number  of  sea  scallops  in  the 
population  for  shell  height  bin  h; 
qh  = the  size-specific  probability  of  detection 
(selectivity)  in  the  video  survey  (on  a scale 
of  0 to  1 and  relative  to  the  bin  with  maxi- 
mum probability  of  detection);  and 
nh  = the  estimated  number  of  sea  scallops  in  the 
population  that  are  available  to  the  video 
survey  gear. 

In  the  absence  of  measurement  error,  the  predicted  shell- 
height  composition  jih  for  the  survey  is 


1=1 


where  L = the  number  of  shell-height  bins  in  the  model. 

If  if  is  a row  vector  of  length  L containing  the  predicted 
proportions  (before  measurement  errors)  for  each  length 
group  in  the  survey,  then 

p = nE,  (A3) 

where  p the  row  vector  of  predicted  proportions  (includ- 
ing measurement  errors). 


246 


Fishery  Bulletin  108(2) 


In  Equation  A3,  E is  a square  measurement  error  ma- 
trix with  L rows  and  columns  that  distributes  numbers 
at  true  shell  height  into  observed  shell  heights  bins  that 
are  larger  and  smaller  than  the  true  shell  height.  For 
example,  the  first  row  of  E sums  to  one  and  gives  the 
probability  of  observed  shell  heights  for  sea  scallops  in 
the  first  true  shell  height  bin.  The  last  row  of  E sums 
to  one  and  gives  the  probabilities  that  sea  scallops  in 
each  shell  height  bin  would  be  assigned  to  the  “plus 
group”  because  of  measurement  error.  As  described  in 
the  text,  we  estimated  E for  sea  scallops  using  results 
from  experiment  2. 

Appendix  2 

Equation  A3  in  Appendix  1 indicates  the  possibility  of 
correcting  shell-height  data  measurement  algebraically, 


without  resorting  to  an  approach  like  the  CASA  model. 
In  particular,  if  the  matrix  E is  invertible,  then  it  may 
be  possible  to  estimate  the  true  sample  proportions  n 
by  multiplying  both  sides  of  Equation  A3  by  the  inverse 
matrix  E~u. 

n - pE~x.  (A4) 

However,  the  inverse  calculation  in  Equation  A4  will  be 
unreliable  if  the  estimated  error  matrix  E is  poorly  con- 
ditioned. If  the  error  matrix  is  poorly  conditioned,  then 
small  inaccuracies  in  the  estimate  of  E will  propagate 
into  larger  errors  in  the  inverse  E~l  and  the  predicted 
proportions  k. 

As  described  by  Horn  and  Johnson  (1985),  the  condi- 
tion factor  for  an  invertible  matrix  E is 

k=\e  ||||e'1||,  (A5) 


where  ||Z?||  = the  matrix  norm  of  E. 

The  condition  factor  k is  always  at  least  one  and 
is  an  upper  bound  measure  of  the  extent  to  which 
errors  in  the  original  error  matrix  E (ignoring 
errors  in  p ) will  propagate  to  its  inverse.  If  k is 
slightly  larger  than  one,  then  uncertainty  in  E 1 
and  n from  Equation  A4  will  be  at  most  slightly 
greater  then  uncertainty  in  E.  If  k is  large,  then 
uncertainty  in  E -1  and  n may  be  much  larger 
than  uncertainty  in  E. 

The  measurement-error  matrices  that  included 
both  bias  and  imprecision  are  the  most  realistic 
according  to  results  from  experiment  2.  The  con- 
dition factors  for  these  error  matrices  were  2638 
for  video  and  2.3  for  measuring  boards  (Table  4). 
These  condition  factors  indicate  that  uncertainty 
in  E -1  and  “corrected”  shell-height  composition 
data  could  be  much  higher  than  uncertainty  in 
the  original  error  matrix  E for  video  and  at  most 
2.3  times  higher  for  measuring  boards. 

Bootstrap  analyses  show  the  practical  signifi- 
cance of  condition  factors  for  video  and  measur- 
ing board  data  in  our  study.  For  example,  for  the 
video  shell-height  measurements  in  experiment 
2,  the  first  step  was  to  resample  n data  records 
(including  one  video  measurement  and  the  corre- 
sponding caliper  measurement)  with  replacement 
from  the  data  in  experiment  2. 

Sample  sizes  (77  = 670  for  video  and  77  = 344  for 
measuring  boards)  were  the  same  as  the  number 
of  experimental  measurements  and  constituted 
an  upper  bound  on  the  true  effective  sample  size 
because  they  ignore  repeated  measurements  on 
the  same  specimens  (Table  2).  The  effect  of  us- 
ing an  upper  bound  estimate  for  effective  sample 
size  was  to  understate  effects  of  uncertainty  in 
error  matrices.  Our  interest  was,  however,  in  a 
“best  case”  scenario  with  relatively  large  sample 
sizes.  Next,  the  measurement  errors  (e.g.  video 
or  measuring  board  minus  caliper  measure- 


0.10- 


0.15 


0.00 


Measurement 

boards 

\ 


S- 


' VA 


Video 


100 


50  - 


-50  - 


-100 


B— — — 

& 

1 1 1 1 1 1 
30  50 

1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 
70  90  110  130  150 

1 1 1 1 
185 

B 

• . 

o § 8 

° 

§ 1 ? 
i i T : i 
I — : R_l|— 

if s 8 r i i * ° 

Rath  ■ ^ „ o a . - „ » „ .« 

1 - : V 

8 tig 
“ 8 § 

Borers  . _ t ; * 8 

lift  s ° T T T o 

? § s 

o § 

° ° o 

• 

i i i i i i i 

30  50 


70 


150  185 


90  110  130 

Shell  height  (mm) 

Appendix  Figure  1 

Boxplots  showing  bootstrap  distributions  (1000  iterations)  of 
estimated  true  shell-height  (SH)  composition  for  Atlantic  sea 
scallops  ( Placopecten  magellanicus ) in  experiment  2,  based  on 
measurement  boards  (A)  and  video  (B)  shell-height  data.  True 
shell-height  compositions  were  estimated  by  using  bootstrap 
estimates  of  the  inverse  of  the  measurement  error  matrix  E 
and  Equation  A4.  The  solid  line  in  (A)  shows  the  actual  caliper- 
derived  shell-height  data  in  the  experiment.  The  solid  line  is 
not  visible  in  (B)  because  of  the  scale  of  the  y-axis. 


Jacobson  et  al.:  Measurement  errors  in  body  size  of  Placopecten  magellanicus 


247 


Shell  height  (mm) 

Appendix  Figure  2 

Bootstrap  distributions  (1000  iterations)  for  Atlantic  sea  scal- 
lop ( Placopecten  magellanicus ) shell-height  data  obtained  from 
measurement  boards  (A)  and  video  (B),  with  measurement 
errors.  The  solid  line  shows  the  actual  caliper-derived  shell- 
height  data  in  experiment  2. 


ments),  their  mean  (bias),  and  variance  were 
used  to  calculate  the  bootstrap  measurement  er- 
ror matrix  and  its  inverse.  Finally,  the  original 
video  shell-height  composition  data  used  in  ex- 
periment 2 (expressed  as  proportions)  were  then 
multiplied  by  the  bootstrap  inverse  matrix  (Eq. 

A4)  to  remove  measurement  errors  and  obtain  a 
bootstrap  estimate  of  the  true  shell-height  com- 
position. There  were  1000  bootstrap  iterations 
for  both  the  video  and  measurement  board  data. 

The  variability  among  bootstrap  estimates  of  the 
true  shell-height  composition  was  due  entirely  to 
errors  in  the  measurement  error  matrix  E and 
its  inverse  . 

As  expected,  based  on  condition  factors  (see 
above)  and  measurement  error  statistics  (Table 
2),  bootstrap  estimates  of  true  caliper  shell- 
height  composition  data  from  video  data  were 
highly  variable  and  predicted  proportions  ranged 
from  -188  to  195  (i.e.,  outside  the  feasible  range 
for  proportions).  Bootstrap  estimates  from  mea- 
surement board  data  resembled  the  correspond- 
ing true  caliper  measurements.  However,  the 
estimated  proportions  for  both  measurement 
methods  were  often  negative  and  infeasible  (Ap- 
pdx.  Fig.  1). 

We  used  a similar  bootstrap  procedure  to  eval- 
uate effects  of  uncertainty  in  predicted  length 
compositions  with  measurement  errors  (Eq.  A3 
in  Appdx.  1),  which  is  the  approach  used  in  the 
CASA  model.  In  this  bootstrap  analysis,  the 
caliper  shell  height  composition  data  from  ex- 
periment 2 were  assumed  to  be  true  and  error 
matrices  were  generated  by  bootstrapping  the 
experimental  and  video  and  measuring  board 
data  as  described  above.  The  sample  size  was 
n=172  for  both  video  and  measuring  boards  and 
the  same  as  the  number  of  individual  specimens 
in  experiment  2.  This  lower  bound  estimate  of 
the  effective  sample  size  was  used  in  order  to 
overstate  effects  of  uncertainty  in  error  matrices.  Re- 
sults indicated  that  the  calculations  used  in  the  CASA 
model  for  measurement  errors  were  robust  to  uncer- 


tainty about  the  error  matrices  and  the  magnitude  of 
the  errors  because  variability  in  predicted  shell  height 
compositions  was  relatively  minor  (Appdx.  Fig.  2). 


248 


Fishery  Bulletin 

Guidelines  for  authors 


Manuscript  Preparation 

Contributions  published  in  Fishery  Bulletin  describe 
original  research  in  marine  fishery  science,  fishery 
engineering  and  economics,  as  well  as  the  areas  of 
marine  environmental  and  ecological  sciences  (including 
modeling).  Preference  will  be  given  to  manuscripts  that 
examine  processes  and  underlying  patterns.  Descriptive 
reports,  surveys,  and  observational  papers  may  occa- 
sionally be  published  but  should  appeal  to  an  audience 
outside  the  locale  in  which  the  study  was  conducted. 
Although  all  contributions  are  subject  to  peer  review, 
responsibility  for  the  contents  of  papers  rests  upon  the 
authors  and  not  on  the  editor  or  publisher.  Submission 
of  an  article  implies  that  the  article  is  original  and  is  not 
being  considered  for  publication  elsewhere.  Articles  may 
range  from  relatively  short  contributions  (10-15  typed, 
double-spaced  pages,  tables  and  figures  not  included) 
to  extensive  contributions  (20-30  typed  pages).  Manu- 
scripts must  be  written  in  English;  authors  whose  native 
language  is  not  English  are  strongly  advised  to  have 
their  manuscripts  checked  by  English-speaking  col- 
leagues before  submission. 

Title  page  should  include  authors’  full  names  and 
mailing  addresses  and  the  senior  author’s  telephone, 
fax  number,  and  e-mail  address,  and  a list  of  key  words 
to  describe  the  contents  of  the  manuscript.  Abstract 
should  be  limited  to  200  words  (one-half  typed  page), 
state  the  main  scope  of  the  research,  and  emphasize 
the  author’s  conclusions  and  relevant  findings.  Do  not 
review  the  methods  of  the  study  or  list  the  contents  of 
the  paper.  Because  abstracts  are  circulated  by  abstract- 
ing agencies,  it  is  important  that  they  represent  the 
research  clearly  and  concisely.  Text  must  be  typed  in 
12  point  Times  New  Roman  font  throughout.  A brief 
introduction  should  convey  the  broad  significance  of 
the  paper;  the  remainder  of  the  paper  should  be  divided 
into  the  following  sections:  Materials  and  methods, 
Results,  Discussion  (or  Conclusions),  and  Acknowl- 
edgments. Headings  within  each  section  must  be  short, 
reflect  a logical  sequence,  and  follow  the  rules  of  multi- 
ple subdivision  (i.e.,  there  can  be  no  subdivision  without 
at  least  two  items).  The  entire  text  should  be  intelligible 
to  interdisciplinary  readers;  therefore,  all  acronyms, 
abbreviations,  and  technical  terms  should  be  written 
out  in  full  the  first  time  they  are  mentioned.  Include 
FAO  common  names  for  species  in  the  list  of  keywords 
and  in  the  introduction.  Regional  common  names  may 
be  used  throughout  the  rest  of  the  text  if  they  are  dif- 
ferent from  FAO  common  names  which  can  be  found  at 
http://www.fishbase.org/search.html.  Follow  the  U.S. 
Government  Printing  Office  Style  Manual  (2000  ed.) 
and  Scientific  Style  and  Format:  the  CSE  Manual  for 
Authors,  Editors,  and  Publishers  (7th  ed.)  for  editorial 


style;  for  fish  nomenclature  follow  the  most  current  issue 
of  the  American  Fisheries  Society’s  Common  and  Scien- 
tific Names  of  Fishes  from  the  United  States,  Canada, 
and  Mexico,  6th  ed.  Dates  should  be  written  as  follows: 
11  November  2000.  Measurements  should  be  expressed 
in  metric  units,  e.g.,  58  metric  tons  (t);  if  other  units  of 
measurement  are  used,  please  make  this  fact  explicit 
to  the  reader.  Write  out  the  numbers  zero  through 
nine  unless  they  form  part  of  measurement  units  (e.g., 
nine  fish  but  9 mm).  Refrain  from  using  the  shorthand 
slash  (/),  an  ambiguous  symbol,  in  the  general  text 
Literature  cited  comprises  published  works  and 
those  accepted  for  publication  in  peer-reviewed  literature 
(in  press).  Follow  the  name  and  year  system  for  citation 
format  in  the  “Literature  cited”  section  (that  is  say, 
citations  should  be  listed  alphabetically  by  the  authors’ 
last  names,  and  then  by  year  if  there  is  more  than  one 
citation  with  the  same  authorship).  If  there  is  a sequence 
of  citations  in  the  text,  list  chronologically:  (Smith, 
1932;  Green,  1947;  Smith  and  Jones,  1985).  Abbrevia- 
tions of  serials  should  conform  to  abbreviations  given  in 
the  Serial  Sources  for  the  BIOSIS  Previews  Database. 
Authors  are  responsible  for  the  accuracy  and  complete- 
ness of  all  citations.  Literature  citation  format:  Author 
(last  name,  followed  by  first-name  initials).  Year.  Title 
of  report  or  manuscript.  Abbreviated  title  of  the  series 
to  which  it  belongs.  Always  include  number  of  pages. 
Cite  all  software  and  special  equipment  or  chemical 
solutions  used  in  the  study,  not  in  a footnote  but  within 
parentheses  in  the  text  (e.g.,  SAS,  vers.  6.03,  SAS  Inst., 
Inc.,  Cary,  NC). 

Tables  are  often  overused  in  scientific  papers;  it 
is  seldom  necessary  or  even  desirable  to  present  all 
the  data  associated  with  a study.  Tables  should  not 
be  excessive  in  size  and  must  be  cited  in  numerical 
order  in  the  text.  Headings  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 
with  italic  numeral  footnote  markers.  Use  asterisks  to 
indicate  probability  in  statistical  data.  Do  not  type  table 
legends  on  a separate  page;  place  them  above  the  table 
data.  Do  not  submit  tables  in  photo  mode. 

• Zeros  should  precede  all  decimal  points  for  values 
less  than  one. 

• Sample  size,  n,  should  be  italicized. 

• Capitalize  the  first  letter  of  the  first  word  in  all  labels 
within  figures. 

• Do  not  use  overly  large  font  sizes  in  maps  and  for  units 
of  measurements  along  axes  in  figures. 

• Do  not  use  bold  fonts  or  bold  lines  in  figures. 

• Do  not  place  outline  rules  around  graphs. 

• Do  not  use  horizontal  lines  through  graphs  to  indicate 
measurement  units. 

• Use  a comma  in  numbers  of  five  digits  or  more  (e.g. 
13,000  but  3000). 

• Maps  require  a North  arrow  and  degrees  latitude- 
longitude  (e.g.,  170°E). 


Fishery  Bulletin  108(2) 


249 


Figures  include  line  illustrations,  photographs  (or 
slides),  and  computer-generated  graphs  and  must  be 
cited  in  numerical  order  in  the  text.  Graphics  should 
aid  in  the  comprehension  of  the  text,  but  they  should 
be  limited  to  presenting  patterns  rather  than  raw  data. 
Figures  should  not  exceed  one  figure  for  every  four 
pages  of  text.  Figures  must  be  labeled  with  author’s 
name  and  number  of  the  figure.  Avoid  placing  labels 
vertically  (except  of  y axis).  Figure  legends  should 
explain  all  symbols  and  abbreviations  and  should  be 
double-spaced  on  a separate  page  at  the  end  of  the 
manuscript.  Color  is  allowed  in  figures  to  show  morpho- 
logical differences  among  species  (for  species  identifica- 
tion), to  show  stain  reactions,  and  to  show  gradations  in 
temperature  contours  within  maps.  Color  is  discouraged 
in  graphs,  and  for  the  few  instances  where  color  may 
be  allowed,  the  use  of  color  will  be  determined  by  the 
Managing  Editor. 

Failure  to  follow  these  guidelines 
and  failure  to  correspond  with  editors 
in  a timely  manner  will  delay 
publication  of  a manuscript. 

Copyright  law  does  not  apply  to  Fishery  Bulletin, 
which  falls  within  the  public  domain.  However,  if  an 
author  reproduces  any  part  of  an  article  from  Fishery 
Bulletin  in  his  or  her  work,  reference  to  source  is  consid- 
ered correct  form  (e.g..  Source:  Fish.  Bull  97:105). 

Submission 

The  Scientific  Editorial  Office  encourages  authors 
to  submit  their  manuscripts  as  a single  PDF  (pre- 
ferred) or  Word  (zipped)  document  by  e-mail  to 


Fishery.Bulletin@noaa.gov.  Please  use  the  subject 
heading,  “Fishery  Bulletin  manuscript  submission”. 
Do  not  send  encrypted  files.  Please  provide  names 
and  contact  information  for  3-4  suggested  reviewers. 
Commerce  Department  personnel  should  submit  papers 
under  a completed  NOAA  Form  25-700.  Or  you  may 
send  your  manuscript  on  a compact  disc  in  one  of  the 
above  formats.  For  further  details  on  electronic  sub- 
mission, please  contact  the  Scientific  Editorial  Office 
directly  (see  address  below). 

Richard  D.  Brodeur,  Ph.D. 

Scientific  Editor,  Fishery  Bulletin 
Northwest  Fisheries  Science  Center 

2030  S.  Marine  Science  Dr. 

Newport,  Oregon  97365-5296 

Once  the  manuscript  has  been  accepted  for  publication, 
you  will  be  asked  to  submit  a final  electronic  copy  of 
your  manuscript.  When  requested,  the  text  and  tables 
should  be  submitted  in  Word  or  Word  Rich  Text  Format. 
Figures  should  be  sent  as  PDF  files,  Windows  metafiles, 
tiff  files,  or  EPS  files.  Send  a copy  of  figures  in  the  origi- 
nal software  if  conversion  to  any  of  these  formats  yields 
a degraded  version. 

Questions?  If  you  have  questions  regarding  these 
guidelines,  please  contact  the  Managing  Editor,  Sharyn 
Matriotti,  at 

Sharyn.Matriotti@noaa.gov 

Questions  regarding  manuscripts  under  review  should 
be  addressed  to  Richard  Brodeur,  Scientific  Editor,  at 
Rick.Brodeur@noaa.gov. 


I. 


Fishery  Bulletin 

Subscription  form 


Superintendent  of  Documents  Publications  Order  Form 
*5178 

I I YES,  please  send  me  the  following  publications: 

Subscriptions  to  Fishery  Bulletin 

for  $36.00  per  year  ($50.40  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.) 


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  your  orders 

| | | | | p|  | | | | | | | | | | | | | | (202)  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! 


Also  available  online  at  http://bookstore.gpo.gov/collections/fishery-bulletin