Skip to main content

Full text of "The Wilson journal of ornithology"

See other formats


harvard  university 


Ernst  Mayr  Library 
of  the  Museum  of 
Comparative  Zoology 


MCZ 

LIBRARY 

FEB  15  2011 

HARVARD 

UNIVERSITY 


U U- 


^he  Wilson  Journal 

of  Ornithology 

Volume  118,  Number  1,  March  2006 


Published  by  the 
Wilson  Ornithological  Society 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  Ornithologist. 

President — Doris  J.  Watt,  Dept,  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  IN  46556,  USA;  e-mail: 
dwatt@saintmarys.edu 

First  Vice-President — James  D.  Rising,  Dept,  of  Zoology,  Univ.  of  Toronto,  Toronto,  ON  M5S  3G5, 
Canada;  e-mail:  rising@zoo.utoronto.ca 

Second  Vice-President — E.  Dale  Kennedy,  Biology  Dept.,  Albion  College,  Albion,  MI  49224,  USA; 
e-mail:  dkennedy@albion.edu 

Editor — James  A.  Sedgwick,  U.S.  Geological  Survey,  Fort  Collins  Science  Center,  2150  Centre  Ave., 
Bldg.  C,  Fort  Collins,  CO  80526,  USA;  e-mail:  wjo@usgs.gov 

Secretary — Sara  R.  Morris,  Dept,  of  Biology,  Canisius  College,  Buffalo,  NY  14208,  USA;  e-mail: 
morriss@canisius.edu 

Treasurer — Melinda  M.  Clark,  52684  Highland  Dr.,  South  Bend,  IN  46635,  USA;  e-mail:  MClark@tcservices.biz 

Elected  Council  Members — Robert  C.  Beason,  Mary  Gustafson,  and  Timothy  O’Connell  (terms  expire 
2006);  Mary  Bomberger  Brown,  Robert  L.  Curry,  and  James  R.  Hill,  III  (terms  expire  2007);  Kathy  G. 
Beal,  Daniel  Klem,  Jr.,  and  Douglas  W.  White  (terms  expire  2008). 

Membership  dues  per  calendar  year  are:  Active,  $21.00;  Student,  $15.00;  Family,  $25.00;  Sustaining, 
$30.00;  Life  memberships  $500  (payable  in  four  installments). 

The  Wilson  Journal  of  Ornithology  is  sent  to  all  members  not  in  arrears  for  dues. 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 
(formerly  The  Wilson  Bulletin ) 

THE  WILSON  JOURNAL  OF  ORNITHOLOGY  (ISSN  1559-4491)  is  published  quarterly  in  March,  June, 
September,  and  December  by  the  Wilson  Ornithological  Society,  810  East  10th  St.,  Lawrence,  KS  66044-8897.  The 
subscription  price,  both  in  the  United  States  and  elsewhere,  is  $40.00  per  year.  Periodicals  postage  paid  at  Lawrence,  KS. 
POSTMASTER:  Send  address  changes  to  OSNA,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

All  articles  and  communications  for  publications  should  be  addressed  to  the  Editor.  Exchanges  should  be  addressed 
to  The  Josselyn  Van  Tyne  Memorial  Library,  Museum  of  Zoology,  Ann  Arbor,  Michigan  48109. 

Subscriptions,  changes  of  address,  and  claims  for  undelivered  copies  should  be  sent  to  OSNA,  5400  Bosque  Blvd., 
Ste.  680,  Waco,  TX  76710.  Phone:  (254)  399-9636;  e-mail:  business@osnabirds.org.  Back  issues  or  single  copies  are 
available  for  $12.00  each.  Most  back  issues  of  the  journal  are  available  and  may  be  ordered  from  OSNA.  Special  prices 
will  be  quoted  for  quantity  orders.  All  issues  of  the  journal  published  before  2000  are  accessible  on  a free  Web  site  at  the 
Univ.  of  New  Mexico  library  (http://elibrary.unm. edu/sora/).  The  site  is  fully  searchable,  and  full-text  reproductions  of  all 
papers  (including  illustrations)  are  available  as  either  PDF  or  DjVu  files. 


© Copyright  2006  by  the  Wilson  Ornithological  Society 
Printed  by  Allen  Press,  Inc.,  Lawrence,  Kansas  66044,  U.S. A. 


COVER:  Wilson’s  Snipe  ( Gallinago  delicata).  Illustration  by  Scott  Rashid. 


© This  paper  meets  the  requirements  of  ANSI/NISO  Z39.48-1992  (Permanence  of  Paper). 


FRONTISPIECE.  An  adult  White-masked  Antbird  ( Pithys  castaneus ) above  and  a juvenile  below.  Previously 
known  from  only  the  type  specimen,  the  species  was  rediscovered  in  2001  in  northwestern  Department  Loreto, 
Peru  (see  p.  13).  Original  painting  (watercolor  and  gouache)  by  Daniel  F.  Lane. 


VOL.  118,  NO.  1 


ne  Wilson  Journal 
of  Ornithology 

Published  by  the  Wilson  Ornithological  Society 
March  2006  PAGES  1—130 


The  Wilson  Journal  of  Ornithology  118(1):  1-2,  2006 


MESSAGE  FROM  THE  EDITOR:  THE  NEW 
WILSON  JOURNAL  OF  ORNITHOLOGY 


This  issue  of  your  journal — 118(1),  March 
2006 — is  the  debut  issue  of  The  Wilson  Journal 
of  Ornithology.  As  indicated  in  the  insert  letter 
that  came  with  your  December  2003  issue,  the 
Wilson  Council,  Wilson  Society  officers,  and  I 
spent  considerable  time  over  the  last  year  de- 
bating— and  eventually  agreeing  on — the  need 
to  update  the  journal’s  name  and  appearance. 
We  believe  that  the  new  name  maintains  the 
tradition  of  honoring  Alexander  Wilson,  more 
clearly  reflects  the  journal’s  theme  and  content, 
and  is  more  contemporary.  In  addition  to  the 
new  journal  name,  the  front  and  back  covers 
have  been  redesigned,  the  title  page  is  new, 
and  we  have  added  a new  feature  to  The  Wil- 
son Journal  of  Ornithology. 

The  front  cover  of  each  issue  will  portray  a 
different  illustration  of  one  of  the  species 
named  after  Alexander  Wilson.  Pen  and  ink  or 
halftone  artwork  was  solicited  from  over  two 
dozen  artists,  and  we  selected  those  illustra- 
tions that  we  believe  demonstrate  both  orni- 
thological and  artistic  merit.  The  Wilson’s  Snipe 
on  the  March  cover  is  a halftone  by  artist  Scott 
Rashid.  Pen  and  ink  illustrations  of  the  Wilson’s 
Phalarope,  Wilson’s  Plover,  and  Wilson’s 
Storm-Petrel  will  appear  on  the  covers  of  the 
June,  September,  and  December  issues,  re- 
spectively. The  fifth  species  named  after  Al- 


exander Wilson,  Wilson’s  Warbler,  will  appear 
on  the  cover  of  each  issue  in  a logo  designed 
by  George  Miksch  Sutton,  and  the  Wilson’s 
Warblers  that  appeared  on  the  cover  from  1962 
to  2005 — also  by  G.  M.  Sutton — will  now  ap- 
pear on  the  title  page  of  the  first  article  in  each 
issue. 

The  back  cover  (Contents)  has  also  been  re- 
designed, to  make  it  more  aesthetically  pleas- 
ing and  easier  to  read.  A new  feature,  “Once 
Upon  a Time  in  American  Ornithology,”  de- 
buts, as  well.  This  feature  will  put  forward  the 
observations  and  reflections  of  naturalists  from 
times  past — to  afford  retrospection  and  to  re- 
mind us  all  of  the  exhilaration  that  comes  from 
being  afield  and  how  it  once  was  in  American 
ornithology.  I encourage  Wilson  Ornithologi- 
cal Society  members  and  other  readers  of  the 
journal  to  submit  favorite  historical  field  ac- 
counts (including  a brief  introductory  state- 
ment) for  consideration  of  publication  in  a fu- 
ture issue. 

I realize  that  such  cosmetic  modifications 
will  have  little  long-term  effect  on  subscrip- 
tions, membership,  or  the  ornithological  sci- 
ence offered  in  The  Wilson  Journal  of  Orni- 
thology. Combined  with  a renewed  commit- 
ment and  more  substantive  changes  behind  the 
scenes,  however,  I believe  that  the  publication 


1 


2 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


of  this  issue  does  mark  a new  beginning  for 
the  Wilson  Ornithological  Society  and  its  jour- 
nal: (1)  the  journal  has  been  published  on  time 
beginning  with  the  June  2005  issue,  (2)  most 
authors  are  receiving  an  initial  decision  on 
their  work  within  3-4  months,  (3)  the  time 
from  manuscript  submission  to  publication 
now  averages  only  about  12  months,  and  (4) 
manuscript  submissions  are  up  >20%  from 
2004.  I sincerely  hope  that  you,  the  readers 
and  authors,  welcome  the  new  look  and  the 


improvements  we  continue  to  make  to  The 
Wilson  Journal  of  Ornithology.  I thank  Wilson 
Council  and  officers;  Keith  Parsons,  Karen 
Ridgway,  and  the  graphics  department  at  Allen 
Press;  Teri  Kman;  and  The  Wilson  Journal  of 
Ornithology  Editorial  Office  staff — Beth  Dillon, 
Alison  Goffredi,  and  Cynthia  Melcher.  All  were 
instrumental  in  the  execution  and  realization  of 
the  new  design  changes  and  in  helping  to 
bring  The  Wilson  Journal  of  Ornithology  back 
on  schedule. — James  A.  Sedgwick,  Editor. 


The  Wilson  Journal  of  Ornithology  1 18(1):3-12,  2006 


VARIATION  IN  MASS  OF  FEMALE  PROTHONOTARY  WARBLERS 

DURING  NESTING 

CHARLES  R.  B LEM 123  AND  LEANN  B.  BLEM1  2 


ABSTRACT. — Over  an  18-year  period  (1987-2004),  we  examined  variation  in  body  mass  of  female  Protho- 
notary  Warblers  ( Protonotaria  citrea)  captured  throughout  their  nesting  cycle.  As  is  typical  for  many  small 
passerine  birds,  body  mass  was  greatest  during  egg  laying  and  decreased  throughout  incubation  and  feeding  of 
young.  Mass  decreased  significantly  between  the  onset  of  incubation  and  fledging  of  both  first  and  second  broods. 
Mass  loss  was  gradual  during  incubation,  noteworthy  during  the  first  2 days  of  feeding  nestlings,  but  did  not 
continue  to  decrease  throughout  the  feeding  period.  Mass  lost  while  raising  the  first  brood  was  regained  before 
initiating  the  second  brood.  Mass  of  female  warblers,  adjusted  for  effects  of  nest  attempt,  year,  clutch  size,  and 
day  and  stage  of  nesting,  increased  slightly  with  age.  Body  mass  of  nesting  female  warblers  varied  significantly 
with  day  of  the  nest  cycle  during  incubation  but  not  during  egg  laying  or  feeding  of  young.  Mass  was  associated 
with  clutch  size  during  incubation  in  both  first  and  second  broods,  but  was  not  associated  significantly  with 
brood  size  when  females  were  feeding  nestlings.  Frequency  of  food  delivery  to  nestlings  was  associated  nega- 
tively with  female  body  mass.  Females  typically  made  more  feeding  trips  per  day  than  males.  Feeding  rates 
were  correlated  among  pairs;  that  is,  females  with  higher  rates  of  delivery  were  mated  to  males  that  made  a 
higher  number  of  trips.  Received  18  February  2005,  accepted  21  October  2005. 


Mass  loss  is  often  used  as  an  index  of  re- 
productive costs  in  birds  (see  review  in  Mer- 
kle  and  Barclay  1996),  largely  because  it  is  a 
consistent  factor  in  patterns  of  avian  life  his- 
tory. During  the  breeding  season,  female  pas- 
serine birds  typically  gain  mass  in  the  period 
before  egg  laying,  maintain  or  gradually  lose 
a small  amount  during  incubation,  and  then 
lose  a significant  amount  of  mass  during 
brooding  (e.g.,  Ricklefs  1974;  Freed  1981; 
Moreno  1989a,  1989b).  A similar  pattern  of 
change  during  breeding  has  been  documented 
in  several  passerine  birds  (e.g..  Freed  1981, 
Ricklefs  and  Hussell  1984,  Hillstrom  1995, 
Merila  and  Wiggins  1997).  Researchers  have 
hypothesized  that  mass  loss  may  be  a proxi- 
mate response  to  energetic  demands  (e.g., 
Nice  1937,  Hussell  1972,  Askenmo  1977). 
Specifically,  mass  loss  should  be  greatest  dur- 
ing periods  when  energy  demands  are  great- 
est, particularly  near  fledging  when  nestlings 
have  acquired  the  ability  to  thermoregulate, 
and  are  relatively  large.  According  to  this  hy- 
pothesis, mass  loss  should  be  a function  of 
brood  size.  A second  hypothesis  suggests  that 
decreased  mass  reduces  the  energy  required 


1 Dept,  of  Biology,  Virginia  Commonwealth  Univ., 
1000  W.  Cary  St.,  Richmond,  VA  23284-2012,  USA. 

2 Current  address:  Flathead  Lake  Biological  Station, 
311  Bio  Station  Lane,  Poison,  MT  59860,  USA. 

3 Corresponding  author;  e-mail: 
cblem@saturn.vcu.edu 


for  flight  when  food  demands  of  nestlings  are 
greatest,  thus  reducing  energy  requirements  of 
females  and  increasing  the  efficiency  of  feed- 
ing the  young  (e.g..  Freed  1981,  Norberg 
1981,  Hinsley  2000).  In  this  instance,  body 
mass  should  decrease  shortly  after  eggs  hatch 
and  should  be  independent  of  brood  size.  A 
final  hypothesis  is  that  mass  loss  results  from 
degeneration  of  female  reproductive  tissues 
during  the  nesting  cycle  (Ricklefs  1974,  Rick- 
lefs and  Hussell  1984),  and  should  not  pro- 
gressively occur  during  incubation  or  feeding 
of  young.  Some  studies  have  eliminated  the 
tissue  degeneration  hypothesis  because  gonad- 
al atrophy  is  over  before  the  period  when 
mass  loss  is  greatest  (Moreno  1989a,  1989b; 
Merkle  and  Barclay  1996).  It  is  difficult  to 
isolate  these  three  hypotheses,  however,  and 
some  researchers  have  not  found  them  to  be 
mutually  exclusive  (e.g.,  Hillstrom  1995,  Mer- 
ila and  Wiggins  1997). 

The  question  that  usually  has  been  ad- 
dressed is:  “Is  mass  loss  evidence  of  energy 
demand  and/or  does  it  reduce  costs  of  flight 
and  enhance  parental  fitness?”  It  has  been 
shown  that  energy  expenditure  is  related  sig- 
nificantly to  rates  of  nest  visitation,  but  not 
always  in  a linear  manner  (Bryant  1988).  Fur- 
thermore, decreased  body  mass  of  adults  rear- 
ing young  may  enhance  their  fitness  through 
reduction  of  energy  demand  during  the  period 
of  feeding  nestlings.  Our  study  examined 


3 


4 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


measurements  of  body  mass  of  female  Pro- 
thonotary  Warblers  ( Protonotaria  citrea)  ob- 
tained over  an  1 8-year  period.  With  these  data, 
we  attempted  to  answer  three  questions:  (1) 
How  does  female  body  mass  in  this  species 
vary  over  the  breeding  season?  (2)  Does  body 
mass  vary  significantly  among  stages  of  nest- 
ing and  among  years?  (3)  What  are  the  roles 
of  brood  size,  stage  of  reproduction,  and  nest 
attempt  in  determining  body  mass  in  this  spe- 
cies? 

METHODS 

Study  area  and  measurement  of  mass. — Be- 
ginning in  March  1987,  we  placed  wooden 
nest  boxes  along  tidal  creeks  in  swamp  forest 
on  and  near  Presquile  National  Wildlife  Ref- 
uge (37°  20'  N,  77°  15'  W)  near  Hopewell, 
Virginia  (Blem  and  Blem  1991,  1992,  1994). 
The  dominant  tree  species  were  black  gum 
(Nyssa  sylvatica),  red  maple  ( Acer  rubrum), 
and  ash  ( Fraxinus  sp.).  Tidal  amplitude  in  the 
swamp  during  spring  tides  was  >1  m.  Nest 
boxes  were  placed  on  metal  poles  at  approx- 
imately 100-m  intervals  along  creek  banks. 
Box  dimensions  were  28LX9WX6D  cm 
and  the  entrance  hole  was  3.8  cm  in  diameter 
(see  Blem  and  Blem  1991).  We  determined 
optimal  nest-box  sites  during  the  first  2 years 
of  the  study  (Blem  and  Blem  1991)  and  boxes 
were  adjusted  accordingly  to  maximize  their 
usage  by  warblers.  The  number  of  nest  boxes 
used  in  the  study  was  gradually  increased 
from  141  in  1987  to  320  in  2004. 

The  contents  of  boxes  were  documented  6— 
20  times  during  the  breeding  season  each  year, 
depending  upon  the  demands  of  other  inves- 
tigations of  reproductive  output.  Females  were 
captured  as  they  exited  nest  boxes,  weighed 
to  the  nearest  0.1  g on  a portable  electronic 
balance,  and  banded  with  federal  bands.  No 
warbler  in  these  analyses  was  weighed  twice 
per  stage,  and  usually  not  more  than  once  dur- 
ing the  same  nest  attempt.  Midday  (10:00- 
14:00  EST)  masses  (g)  did  not  vary  signifi- 
cantly with  time  of  day  (mass  = -0.04  hr  + 
16.3,  P = 0.49,  R 2 = 0.008,  n = 2,124).  Only 
midday  masses  were  used  in  the  following 
analyses.  We  recorded  dates  of  first  eggs  and 
clutch  sizes  for  those  nests  visited  often 
enough  that  we  could  be  certain  of  the  timing. 
Clutch  size  throughout  the  study  was  consid- 
ered to  be  the  number  of  eggs  present  at  the 


onset  of  incubation.  We  converted  first  egg 
(nest  start)  dates  into  Julian  days  for  analysis. 
Prothonotary  Warblers  generally  produce  two 
clutches  each  season  (Petit  1989),  and  second 
clutches  typically  include  fewer  eggs  (Blem  et 
al.  1999).  We  therefore  divided  nests  with 
eggs  in  two  groups — “first  nests,”  in  which 
first  eggs  were  laid  from  25  April  through  20 
May,  and  “second  nests,”  in  which  first  eggs 
were  laid  after  20  May  (see  Petit  1989).  Some 
of  the  second  nests  may  have  been  replace- 
ment clutches  for  first  nests  that  had  been  dep- 
redated, but  we  are  certain  that  many  of  them 
were  produced  by  females  that  had  success- 
fully fledged  young  (Podlesak  and  Blem  2001, 
2002).  We  used  20  May  as  the  separation  date 
because  it  represents  a major  hiatus  in  laying 
and  is  the  date  after  which  few  first  clutches 
have  been  laid  at  our  study  site.  It  also  was 
used  because  of  the  length  of  time  necessary 
for  Prothonotary  Warblers  to  complete  one 
nesting  cycle  (approximately  27  days)  after  a 
mean  potential  starting  date  of  24  April  (Blem 
and  Blem  1992).  We  divided  nesting  into  three 
phases:  laying  (and  egg  formation),  incuba- 
tion, and  feeding  young.  The  first  phase  ended 
with  the  first  day  of  incubation  and  included 
birds  that  were  building  nests  as  well  as  laying 
eggs.  The  second  phase  began  with  the  first 
egg  and  ended  with  hatching  (Fig.  1). 

Feeding  visits. — In  2002,  we  recorded  feed- 
ing visits  by  warblers  at  individual  boxes  dur- 
ing first  broods  by  means  of  battery-powered 
remote  video  cameras  with  programmable, 
portable  videocassette  recorders.  We  obtained 
>500  hr  of  nest-activity  records  at  eight  nests 
(four  broods  of  three  young  and  four  broods 
of  five  young)  on  days  7 through  10.  Video 
cameras  were  small  and  camouflaged  and  did 
not  noticeably  alter  behavior  of  the  warblers. 
Individual  visits  (see  Figs.  2-3)  were  tran- 
scribed from  replays  of  the  recordings  in  the 
lab.  We  totaled  all  feeding  visits  made  by  both 
parents  from  dawn-to-dark  for  all  4 days.  We 
could  not  accurately  assess  prey  size  from  the 
recordings,  but  we  did  count  the  number  of 
items — mostly  caterpillars — that  were  dis- 
tinctly larger  than  2 cm  (“large  prey”),  as 
judged  by  the  entry  hole  in  the  nest  box.  Fe- 
male warblers  were  weighed  2 days  before 
nestlings  fledged. 

Analyses. — Over  the  18-year  period,  we  ob- 
tained 2,124  measurements  of  body  mass  from 


20 

18 

16 

14 

12 

20 

18 

16 

14 

12 

}dy  r 


and  Blent  • MASS  OF  FEMALE  PROTHONOTARY  WARBLERS 


5 


First  broods 


n=  145 


i • *•.}(  .i.  ..... 

t ••  • *! • ir 


n = 1 ,343 


n = 237 


Second  broods 


0 -5 

Egg  laying 


5 10 

Incubation 


Days 


15  20  25 

Feeding  nestlings 


(g)  of  female  Prothonotary  Warblers  during  nesting  in  eastern  Virginia,  1987-2004  (day 
tion).  Numerous  circles  are  hidden  under  duplicate  values  ( n = 2,124). 


6 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


FIG.  2.  Feeding  visits/nest/day  made  by  female  Prothonotary  Warblers  during  days  7-10  of  feeding  nestlings 
versus  female  body  mass  at  the  end  of  incubation,  eastern  Virginia,  2002.  Open  circles  represent  broods  of  three 
nestlings;  solid  circles  represent  broods  of  five  nestlings.  Nest  visitation  was  a function  of  female  body  mass, 
regardless  of  brood  size. 


977  different  adult  female  warblers.  For  anal- 
ysis, we  partitioned  these  measurements 
among  nesting  attempts  (first  and  second 
nests,  n — 1,344  and  780,  respectively)  and 
stages  of  nesting  (egg  formation  and  laying, 
incubating,  and  feeding).  The  number  of  mea- 
surements in  each  stage-year  combination  var- 
ied from  24  during  laying  in  second  nests  to 
1,344  during  incubation  in  first  nests.  Clutch 
size  varied  from  two  to  six  eggs  and  ages  of 
females  ranged  from  1 to  8 years. 

To  examine  differences  in  mass  between 
nests  and  among  stages  of  nesting  and  brood 
sizes  (adjusted  for  day  of  nesting),  we  used 
univariate  ANCOVA  with  multiple  indepen- 
dent variables  in  PROC  GLM  (SAS  Institute, 
Inc.  2000).  Brood  size,  nest  attempt,  age, 
stage  of  nesting,  and  their  interactions  were 
considered  fixed  (categorical)  effects  in  vari- 
ous models.  Day  of  nesting  (range  = —9  to 
24;  0 = day  of  onset  of  incubation)  was  a 
continuous  variable.  Analysis  of  covariance 


was  done  using  the  PROC  GLM  procedure 
because  the  data  set  was  unbalanced  among 
effects  (Zar  1999).  Type  III  sums  of  squares 
were  used,  adjusting  significance  of  each  fac- 
tor for  the  effects  of  all  other  variables.  Single 
comparisons  of  means  were  done  by  means  of 
appropriate  Mests  based  on  tests  of  equality 
of  variances  (SAS  Institute,  Inc.  2000).  Few 
females  were  measured  more  than  once  during 
the  same  stage  of  nesting  in  a given  nest  in 
the  same  year;  therefore,  we  did  not  use  re- 
peated measures  analyses.  Because  some  of 
the  associated  variables  were  not  measured 
with  each  measurement  of  body  mass,  sample 
sizes  vary  among  analyses.  All  r-tests  were 
two-tailed.  Means  are  presented  ± SD.  Statis- 
tical significance  was  set  at  P < 0.05. 

RESULTS 

Body  mass. — In  the  following  analyses  and 
comparisons,  we  assumed  that  patterns  found 
between  specific  points  along  a regression 


Blem  and  Blem  • MASS  OF  FEMALE  PROTHONOTARY  WARBLERS 


7 


Female  visits/nest/day 

FIG.  3.  Feeding  visits/nest/day  by  mated  pairs  of  Prothonotary  Warblers  during  days  7-10  of  feeding  nest- 
lings, eastern  Virginia,  2002.  Open  circles  represent  broods  of  three  nestlings;  solid  circles  represent  broods  of 
five  nestlings.  Males  brought  food  less  often  than  females,  but  the  frequency  of  male  visits/nest/day  was  a 
function  of  that  of  females. 


were  representative  of  patterns  deduced  from 
single  measurements  of  numerous  females. 
This  was  confirmed  in  our  observations  of 
multiple  measurements  of  a few  single  fe- 
males (CRB  unpubl.  data). 

Body  mass  of  female  Prothonotary  War- 
blers varied  over  the  breeding  season  in  the 
typical  passerine  pattern.  That  is,  variation 
was  greatest  during  egg  laying,  mass  de- 
creased gradually  during  incubation,  and  then 
there  was  a noteworthy  decrease  in  mass  im- 
mediately after  the  eggs  hatched  (Fig.  1).  Af- 
ter the  decline  immediately  after  hatching, 
adult  female  mass  did  not  change  over  time 
throughout  the  period  of  feeding  nestlings. 
Mean  body  masses  did  not  differ  between  nest 
attempts  during  egg  formation  and  laying 
(first  nests:  16.9  ± 1.2,  n = 143;  second  nests: 
16.8  ± 1.9,  n = 93,  Fh235  = 0.20,  P = 0.65), 
but  did  differ  between  nests  during  incubation 
(first  nests:  16.2  ± 0.9,  n = 1,225;  second 
nests:  15.6  ± 0.9,  n = 304,  FU526  = 6.7,  P = 
0.011)  and  during  the  feeding  phase  (first 


nests:  15.2  ± 1.0,  n = 238;  second  nests:  14.9 
± 0.8,  n = 121;  F1>358  = 6.7,  P = 0.012).  Mass 
did  not  vary  with  day  of  nesting  in  the  laying 
or  feeding  stages  of  either  nesting  attempt,  but 
it  did  decline  significantly  with  day  of  incu- 
bation (first  nests:  Ful3 42  = 18.0,  P < 0.001; 
second  nests:  F1303  = 33.5,  P < 0.001). 

As  judged  by  the  collective  scatter  of  in- 
dividual masses  over  time,  females  collective- 
ly lost  10.1%  of  their  body  mass  between  the 
onset  of  incubation  and  fledging  of  first 
broods  and  1 1 .3%  in  second  broods.  Much  of 
this  loss  appeared  to  occur  during  the  first  2 
days  of  feeding  nestlings  (5.4  and  7.7%,  re- 
spectively). Mass  lost  during  first  broods  was 
regained  before  the  initiation  of  second 
broods.  Body  mass  extremes  were  11.9  g for 
an  incubating  bird  and  21.0  g for  a female 
during  the  early  days  of  egg  laying. 

When  the  data  set  including  all  variables 
was  considered  (n  = 1,814;  Fig.  1),  mass  var- 
ied significantly  with  nest  attempt,  stage  of 
nesting,  clutch  size  (2-6),  female  age  (1-8 


8 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


TABLE  1.  Analysis  of  covariance  of  body  mass  of  female  Prothonotary  Warblers  in  eastern  Virginia,  1987- 
2004  (n  = 1,814).  All  two-way  and  three-way  interactions  were  statistically  insignificant  except  for  nesting 
attempt  X stage  of  nesting.  Clutch  sizes  were  2-6  and  ages  were  1-8  years.  Days  of  nesting  ranged  from  -9 
through  24. 


Source 

df 

F 

P > F 

Nesting  attempt 

1 

7.6 

0.006 

Stage  of  nesting 

2 

27.0 

<0.001 

Clutch  size 

4 

10.4 

<0.001 

Age 

5 

6.8 

<0.001 

Day  of  nest  cycle 

1 

35.7 

<0.001 

Year 

17 

2.6 

0.015 

Nesting  attempt  X stage  of  nesting 

1 

2.8 

0.050 

years),  day  of  the  nest  cycle,  and  year  (Table 
1).  There  was  a significant  interaction  between 
nesting  attempt  (first/second  nest)  and  stage  of 
nesting,  but  no  other  two-way  and  three-way 
interactions  were  statistically  significant. 
When  stages  of  nesting  were  analyzed  indi- 
vidually, body  mass  during  the  laying  and 
feeding  stages  did  not  differ  among  clutches/ 
broods  of  different  sizes  and  mass  did  not 
vary  significantly  with  day  of  nesting  in  these 
stages. 

Body  mass  adjusted  for  effects  of  nest  at- 
tempt, year,  clutch  size,  and  day  and  stage  of 
nesting  varied  significantly  with  female  age 
(Fu 213  = 15.0,  P < 0.001;  Table  2).  Unad- 
justed masses  indicated  that  much  of  this 
change  occurred  between  birds  in  their  first 
year  (SY  birds)  and  all  older  age  classes 
(ASY).  Measurements  of  mass  were  obtained 
from  a large  range  of  ages,  including  64  mea- 
surements that  exceeded  the  published  maxi- 
mum age  (5  years  1 1 months)  for  the  species 
(Kennard  1975). 

During  incubation,  mass  was  significantly 


TABLE  2.  Least-squares  means  of  body  mass 
among  incubating  female  Prothonotary  Warblers  dur- 
ing mid-incubation  (days  3-8)  as  a function  of  age 
(years)  in  eastern  Virginia,  1987-2004  (/?  = 1,540). 
All  means  were  adjusted  for  the  effects  of  nest  attempt, 
clutch  size,  and  day  and  stage  of  nesting. 


Age 

Mean  mass  (g) 

n 

1 

16.0 

275 

2 

16.3 

565 

3 

16.4 

420 

4 

16.4 

147 

5 

16.1 

80 

>6 

16.1 

48 

associated  with  day  of  nesting  and  clutch  size 
(Table  3).  Mass  tended  to  decrease  gradually 
throughout  incubation.  Birds  with  larger 
clutches  during  first  nesting  attempts  tended 
to  have  greater  body  mass;  birds  with  small 
clutches  in  second  nests  had  the  lowest  body 
mass. 

Feeding  visits. — Total  nest  visits  per  day 
made  by  females  during  days  7-10  of  feeding 
nestlings  was  a function  of  female  body  mass, 
regardless  of  brood  size  (three  young;  Fl3  = 
13.8,  P = 0.023,  R2  = 0.80;  five  young:  F13 
= 15.5,  P = 0.034,  R2  = 0.85;  Fig.  2).  Males 
brought  food  less  often  than  females  (three 
young:  x2  = 38.2,  df  = 1,  P < 0.052;  five 


TABLE  3.  Analysis  of  covariance  of  body  mass 
among  female  Prothonotary  Warblers  in  eastern  Vir- 
ginia, 1987-2004  by  stage  of  nesting  ( n = 2,124  in  all 
analyses).  Clutch  and  brood  sizes  were  2-6  and  ages 
were  1-6  years;  days  of  nesting  ranged  from  —9 
through  24  (day  0 = first  day  of  incubation). 

Source 

df 

F 

P > F 

Egg  formation  and  laying  (/? 

= 169) 

Nesting  attempt 

1 

0.9 

0.34 

Clutch  size 

4 

2.2 

0.092 

Day  of  nesting 

1 

0.2 

0.70 

Age 

5 

1.7 

0.13 

Incubation  {n  = 1,647) 
Nesting  attempt 

1 

52.3 

<0.001 

Clutch  size 

4 

9.3 

<0.001 

Day  of  nesting 

1 

40.4 

<0.001 

Age 

5 

6.3 

<0.001 

Feeding  nestlings  ( n = 
Nesting  attempt 

308) 

1 

4.3 

0.039 

Brood  size 

4 

1.0 

0.45 

Day  of  nesting 

1 

0.3 

0.58 

Age 

5 

1.3 

0.26 

Blem  and  Blem  • MASS  OF  FEMALE  PROTHONOTARY  WARBLERS 


9 


TABLE  4.  Mean  visitation  rates  (no./day 
total)  for  days  7-10  of  nestling  development  in 

± SD)  of  male 
eastern  Virginia 

and  female  Prothonotary  Warblers  (percent  of 
l,  2002. 

Female  visits 

Male  visits 

Brood  size 

Per  nest 

Per  nestling 

Per  nest 

Per  nestling 

3 {n  = 4) 
5 (n  = 4) 

306  ± 95  (63.8) 
396  ± 148  (56.5) 

102.0 

79.2 

171  ± 40  (36.2) 
295  ± 108  (43.5) 

57.0 

59.0 

young:  x2  = 12.1,  df  = 1,  P < 0.054;  Table 
4),  but  frequency  of  male  visits  per  day  was 
a function  of  that  of  females  (female  visits  = 
1.0  ± 1.06  X male  visits;  R2  = 0.75,  Fl3  = 
17.7,  p = 0.006;  Fig.  3).  Female  feeding  trips 
per  nestling  decreased  with  brood  size  (x2  = 
9.3,  df  = 1,  P < 0.05;  Table  4),  but  male  trips 
per  nestling  did  not  decrease  (x2  = 0.034,  df 
= 1,  P > 0.05).  The  percentage  of  total  pa- 
rental visits  made  by  males  declined  from  a 
high  of  44.0%  on  day  7 to  a low  of  34.8%  on 
day  10.  Males  brought  significantly  more 
“large  prey  items”  to  the  nest  than  did  fe- 
males (males:  330,  females:  210;  x2  = 26.7, 
df  = 1,  p < 0.05).  These  prey  items  were 
mostly  Hexagenia  sp.  mayflies  and  lepidop- 
teran  caterpillars.  There  was  no  significant  dif- 
ference in  the  number  of  larger  prey  delivered 
by  males  to  different  brood  sizes  (175  in 
broods  of  three,  155  in  broods  of  five;  x2  = 
1.2,  df  = 1,  P > 0.05). 

DISCUSSION 

Body  mass  clearly  is  associated  with  stage 
of  breeding  activity  in  small  passerines  (Freed 
1981,  Ricklefs  and  Hussell  1984,  Cichon 
2001),  and  each  stage — egg  formation  and 
laying,  incubation,  and  feeding  of  nestlings 
is  characterized  by  a different  pattern  of  mass 
change  (e.g.,  Fig.  1).  Mass  change  of  female 
Prothonotary  Warblers  in  our  study  was  sim- 
ilar to  that  reported  in  several  other  studies  of 
passerine  species  (e.g..  Freed  1981,  Ricklefs 
and  Hussell  1984,  Johnson  et  al.  1990,  Hills- 
trom  1995).  During  egg  laying,  body  mass 
varied  greatly  with  follicle  formation  and  re- 
lease of  eggs,  then  declined  progressively 
throughout  incubation  (Fig.  1),  and  dropped 
sharply  at  hatching.  Female  mass  then  re- 
mained relatively  constant  throughout  the  pe- 
riod of  feeding  nestlings.  Mass  changes  in 
Prothonotary  Warblers  during  egg  laying  and 
incubation  were  similar  to  those  of  all  small 
passerines  and  require  little  explanation.  Mass 


loss  at  hatching  is  more  complex  and  differs 
among  species.  Because  the  significance  of 
this  loss  is  uncertain,  the  behavior  and  com- 
positional dynamics  of  females  requires  closer 
scrutiny. 

Two  potential  hypotheses  have  been  pro- 
posed to  explain  mass  loss  of  female  birds 
during  feeding  of  nestlings:  (1)  energy  de- 
mand (cost  of  reproduction  hypothesis  = re- 
serve mobilization  hypothesis;  Cavitt  and 
Thompson  1997),  and  (2)  long-term  benefits 
from  reduction  of  power  demands  for  flight 
during  feeding  (mass  adjustment  hypothesis  = 
flight  efficiency  hypothesis).  Forming  and  lay- 
ing eggs,  incubating,  and  feeding  nestlings  re- 
quires additional  collection  and  expenditure  of 
energy,  whereas  adjusting  mass  to  save  energy 
expended  in  flight  during  the  numerous  trips 
made  while  feeding  young  is  an  adaptive  loss. 

It  has  become  obvious  that  body  mass  can 
vary  as  a result  of  energy  demand  during  ex- 
treme years  (Merila  and  Wiggins  1997)  or 
with  larger  broods  (Nur  1984).  It  appears  to 
be  axiomatic  that  reserves  should  be  depleted 
during  times  of  high-energy  demand  and  it  is 
well  known  that  body  mass  and  energy  re- 
serves are  closely  related  (Blem  1990).  Part 
of  the  variation  in  mass  within  stages  of  the 
nest  cycle  may  result  from  differences  in  an- 
nual factors,  such  as  temperature  extremes,  in- 
clement weather  (Merila  and  Wiggins  1997), 
or  brood  number  (De  Laet  and  Dhondt  1989). 
Because  of  our  large  sample  size,  we  were 
able  to  detect  annual  variation  within  the  in- 
cubation period  of  first  nests,  largely  by  elim- 
inating much  of  the  variation  associated  with 
several  other  variables.  Others  (e.g.,  Johnson 
et  al.  1990)  have  likewise  found  significant 
annual  variations  in  mass  of  breeding  birds, 
and  extreme  environmental  conditions  in  ex- 
ceptional years  have  important  influences  on 
body  mass  (Merila  and  Wiggins  1997). 

Not  all  studies,  however,  have  shown  that 
energy  demand  is  an  important  factor  in  body 


10 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


mass.  For  example,  larger  broods  are  not  al- 
ways associated  with  greater  mass  loss  of  fe- 
males (Pinkowski  1978,  this  study),  even 
though  energy  expenditure  by  females  in- 
creases with  brood  size  (Sanz  et  al.  1998). 
Furthermore,  food-supplementation  studies 
have  provided  mixed  results.  Food  supple- 
ments did  not  affect  female  mass,  brood  mass, 
or  length  of  the  nestling  period  among  House 
Wrens  ( Troglodytes  aedon\  Cavitt  and 
Thompson  1997)  or  Northern  Wheatears 
( Oenanthe  oenanthe;  Moreno  1989a).  How- 
ever, food-supplemented  female  Mountain 
Bluebirds  ( Sialia  currucoides;  Garcia  et  al. 
1993)  maintained  greater  body  mass  and 
fledged  larger  young  than  females  receiving 
no  food  supplementation.  Some  studies  have 
found  that  female  mass  is  a negative  function 
of  brood  size  (Nur  1984,  Merila  and  Wiggins 
1997),  and  that  energy  demand  during  first 
broods  may  influence  the  probability  of  hav- 
ing a second  brood  in  some  species  (De  Laet 
and  Dhondt  1989).  In  Prothonotary  Warblers, 
it  appears  that  many  females  totally  recover 
lost  mass  fairly  quickly  between  nest  attempts. 
It  has  been  suggested  that  species  breeding  in 
different  environments  may  respond  different- 
ly to  stress  associated  with  increased  energy 
requirements  and  there  may  not  be  selection 
for  adaptive  mass  loss  (Cavitt  and  Thompson 
1997). 

The  pattern  of  mass  change  in  female  Pro- 
thonotary Warblers  in  our  study  does  not  sup- 
port the  cost  of  reproduction  hypothesis,  but 
it  does  support  the  mass  adjustment  hypothe- 
sis. Important  supporting  observations  includ- 
ed ( 1 ) the  regular  loss  of  mass  after  hatching 
in  both  nesting  attempts,  (2)  the  lack  of  influ- 
ence of  brood  size  on  female  mass,  (3)  no 
increasing  loss  in  female  mass  as  young  grew 
and  when  feeding  activity  levels  were  great- 
est, (4)  more  feeding  trips  made  by  females 
that  weighed  less,  and  (5)  little  evidence  that 
males  adjusted  their  feeding  efforts  to  offset 
demands  on  females.  Trivers  (1972)  predicted 
that,  within  breeding  pairs,  females  would 
provide  the  largest  proportion  of  nestling  care 
because  they  had  a larger  share  of  investment 
of  energy  than  males.  In  our  study,  female 
Prothonotary  Warblers  made  more  feeding 
trips  than  males  (both  broods).  Male  Protho- 
notary Warblers,  however,  brought  a greater 
proportion  of  large  prey,  which  may  have  sig- 


nificantly offset  female  effort  during  later 
stages  in  the  nesting  cycle  even  though  males 
made  fewer  trips  as  nestlings  neared  fledging. 

The  mass  adjustment  hypothesis  suggests 
that  birds  benefit  from  mass  loss  due  to  de- 
creased wing  loading  (e.g..  Freed  1981,  Nor- 
berg  1981,  Ricklefs  and  Hussell  1984,  Cavitt 
and  Thompson  1997).  Energy  saved  by  mass 
reduction  may  enable  parent  birds  to  raise 
more  young  faster  or  produce  fledglings  with 
greater  mass.  Observations  supporting  the 
mass  adjustment  hypothesis  include  (1)  great- 
er loss  of  mass  before  the  period  of  maximum 
energy  requirement  (e.g.,  Freed  1981,  Ricklefs 
and  Hussell  1984,  Merkle  and  Barclay  1996, 
this  study),  (2)  loss  of  mass  independent  of 
brood  size  (e.g.,  Freed  1981,  this  study)  or 
length  of  incubation  (Sanz  and  Moreno  1995, 
this  study),  and  (3)  no  increase  in  body  mass 
among  food-supplemented  females  feeding 
nestlings  (Cavitt  and  Thompson  1997).  In  our 
study,  mass  loss  of  females  during  incubation 
was  correlated  with  clutch  size,  but  mass  of 
females  feeding  nestlings  was  not  affected  by 
brood  size,  nor  did  female  mass  decrease 
throughout  nestling  development.  If  increased 
energy  demand  is  important,  then  female  mass 
should  decline  significantly  as  nestlings  grow, 
although  it  is  possible  that  males  may  “pick 
up  the  slack.”  That  is,  male  warblers  might 
feed  young  more  frequently  or  with  higher- 
quality  food  in  large  broods  than  small,  thus 
reducing  energy  demands  on  females  and  al- 
lowing them  to  maintain  their  mass  and  fit- 
ness. Our  observations  weakly  support  these 
ideas.  Males  did  bring  more  large  prey  items 
than  females,  but  this  did  not  vary  with  brood 
size  or  with  nestling  age.  Furthermore,  males 
made  fewer  visits  late  in  the  nesting  cycle 
than  females.  This  pattern  is  nearly  identical 
with  that  documented  for  Willow  Tits  ( Poecile 
montanus ; Rytkonen  et  al.  1996).  Similar 
studies  have  shown  that  nest  visitation  rates 
may  be  greater  in  males  of  some  species 
(Grundel  1987),  greater  in  females  of  others 
(Pinkowski  1978,  Conrad  and  Robertson 
1993),  or  may  not  differ  between  the  sexes 
(Best  1977,  Knapton  1984,  Omland  and  Sher- 
ry 1994).  The  significance  of  the  age:body 
mass  relationship  during  the  reproductive  pe- 
riod is  not  clear.  We  are  aware  of  few  studies 
that  have  demonstrated  an  age  effect  on  mass 
(see  De  Laet  and  Dhondt  1989,  Merila  and 


Blem  and  Blem  • MASS  OF  FEMALE  PROTHONOTARY  WARBLERS 


1 1 


Wiggins  1997).  In  our  study,  however,  female 
age  had  a significant  effect  on  body  mass, 
even  after  mass  was  adjusted  for  the  effects 
of  many  other  variables. 

Mass  variation  of  female  birds  during  nest- 
ing obviously  is  a complex  phenomenon. 
Deeper  insight  into  mass  variations  will  be  ob- 
tained only  with  studies  that  combine  mea- 
sures of  body  composition,  condition  of  re- 
production tracts,  and  measures  of  hormone 
levels  with  stage  of  nesting.  While  time-con- 
suming, collecting  large  data  sets  over  nu- 
merous years  is  well  worth  the  trouble,  but 
would  be  even  more  valuable  if  simultaneous 
studies  could  be  carried  out  at  several  sites 
over  the  range  of  the  species. 

ACKNOWLEDGMENTS 

We  thank  the  officials  of  the  Eastern  Virginia  Rivers 
National  Wildlife  Refuge  (NWR)  Complex  for  per- 
mission to  conduct  this  study  at  Presquile  NWR.  The 
North  American  Bluebird  Society  financially  support- 
ed box  construction.  More  than  100  students,  friends, 
and  faculty  colleagues  assisted  in  this  project  and  we 
thank  them  all,  especially  A.  S.  and  K.  C.  Seidenberg 
and  J.  R.  and  R.  J.  Reilly  for  their  continuous  help 
over  many  years.  We  thank  L.  B.  Williams,  K.  R.  Guis- 
inger,  and  D.  S.  Stevens  for  transcribing  visits  of  war- 
blers from  long,  boring  tape  recordings.  The  Virginia 
Society  of  Ornithology  and  several  of  its  chapters 
helped  fund  our  efforts.  This  is  Rice  Center  for  Envi- 
ronmental Life  Sciences  Research  Contribution  No. 
001. 

LITERATURE  CITED 

Askenmo,  C.  1977.  Effects  of  addition  and  removal  of 
nestlings  on  nestling  weight,  nestling  survival, 
and  female  weight  loss  in  the  Pied  Flycatcher  Fi- 
cedula  hypoleuca  (Pallas).  Ornis  Scandinavica  8: 
1-8. 

Best,  L.  B.  1977.  Patterns  of  feeding  Field  Sparrow 
young.  Wilson  Bulletin  89:625-627. 

Blem,  C.  R.  1990.  Avian  energy  storage.  Current  Or- 
nithology 7:59-1 14. 

Blem,  C.  R.  and  L.  B.  Blem.  1991.  Nest  box  selection 
by  Prothonotary  Warblers.  Journal  of  Field  Orni- 
thology 62:299-307. 

Blem,  C.  R.  and  L.  B.  Blem.  1992.  Prothonotary  War- 
blers nesting  in  nest  boxes:  clutch  size  and  timing 
in  Virginia.  Raven  63:15-20. 

Blem,  C.  R.  and  L.  B.  Blem.  1994.  Composition  and 
microclimate  of  Prothonotary  Warbler  nests.  Auk 
111:197-200. 

Blem,  C.  R.,  L.  B.  Blem,  and  C.  I.  Barrientos.  1999. 
Relationships  of  clutch  size  and  hatching  success 
to  age  of  female  Prothonotary  Warblers.  Wilson 
Bulletin  111:577-581. 

Bryant,  D.  M.  1988.  Energy  expenditure  and  body 


mass  changes  as  measures  of  reproductive  costs 
in  birds.  Functional  Ecology  2:23-34. 

Cavitt,  J.  F.  and  C.  F.  Thompson.  1997.  Mass  loss  in 
breeding  House  Wrens:  effects  of  food  supple- 
ments. Ecology  78:2512-2523. 

Cichon,  M.  2001.  Body-mass  changes  in  female  Col- 
lared Flycatchers:  state-dependent  strategy.  Auk 
1 18:550-552. 

Conrad,  K.  F.  and  R.  J.  Robertson.  1993.  Patterns  of 
parental  provisioning  by  Eastern  Phoebes.  Condor 
95:57-62. 

De  Laet,  J.  F.  and  A.  A.  Dhondt.  1989.  Weight  loss 
of  the  female  during  the  first  brood  as  a factor 
influencing  second  brood  initiation  in  Great  Tits 
Parus  major  and  Blue  Tits  P.  caeruleus.  Ibis  131: 
281-289. 

Freed,  L.  A.  1981.  Loss  of  mass  in  breeding  wrens: 
stress  or  adaptation?  Ecology  62:1 179-1 186. 

Garcia,  P.  F.  J.,  M.  S.  Merkle,  and  R.  M.  R.  Barclay. 
1993.  Energy  allocation  to  reproduction  and  main- 
tenance in  Mountain  Bluebirds  ( Sialia  currocoi- 
des ):  a food  supplementation  experiment.  Cana- 
dian Journal  of  Zoology  71:2352-2357. 

Grundel,  R.  1987.  Determinants  of  nestling  feeding 
rates  and  parental  investment  in  the  Mountain 
Chickadee.  Condor  89:319-328. 

Hillstrom,  L.  1995.  Body  mass  reduction  during  re- 
production in  the  Pied  Flycatcher  Ficedula  hypo- 
leuca: physiological  stress  or  adaptation  for  low- 
ered costs  of  locomotion?  Functional  Ecology  9: 
807-817. 

Hinsley,  S.  A.  2000.  The  costs  of  multiple  patch  use 
by  birds.  Landscape  Ecology  15:765-775. 

Hussell,  D.  J.  T.  1972.  Factors  affecting  clutch  size  in 
Arctic  passerines.  Ecological  Monographs  42: 
317-364. 

Johnson,  R.  K.,  R.  R.  Roth,  and  J.  T.  Paul,  Jr.  1990. 
Mass  variation  in  breeding  Wood  Thrushes.  Con- 
dor 92:89-96. 

Kennard,  J.  H.  1975.  Longevity  records  of  North 
American  birds.  Bird-Banding  46:55-73. 

Knapton,  R.  W.  1984.  Parental  feeding  of  nestling 
Nashville  Warblers:  the  effects  of  food  type, 
brood-size,  nestling  age,  and  time  of  day.  Wilson 
Bulletin  96:594-602. 

Merila,  J.  and  D.  A.  Wiggins.  1997.  Mass  loss  in 
breeding  Blue  Tits:  the  role  of  energetic  stress. 
Journal  of  Animal  Ecology  66:452-460. 

Merkle,  M.  S.  and  R.  M.  R.  Barclay.  1996.  Body 
mass  variation  in  breeding  Mountain  Bluebirds 
Sialia  currucoides : evidence  of  stress  or  adapta- 
tion for  flight?  Journal  of  Animal  Ecology  65: 
401-413. 

Moreno,  J.  1989a.  Body-mass  variation  in  breeding 
Northern  Wheatears:  a field  experiment  with  sup- 
plementary food.  Condor  91:178-186. 

Moreno,  J.  1989b.  Strategies  of  mass  change  in  breed- 
ing birds.  Biological  Journal  of  the  Linnean  So- 
ciety 37:297-310. 

Nice,  M.  M.  1937.  Studies  in  the  life  history  of  the 


12 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


Song  Sparrow.  I.  Transactions  of  the  Linnean  So- 
ciety of  New  York  4:1-247. 

Norberg,  R.  A.  1981.  Temporary  weight  decrease  in 
breeding  birds  may  result  in  more  fledged  young. 
American  Naturalist  118:838-850. 

Nur,  N.  1984.  The  consequences  of  brood  size  for 
breeding  Blue  Tits.  I.  Adult  survival,  weight 
change  and  the  cost  of  reproduction.  Journal  of 
Animal  Ecology  53:479-496. 

Omland,  K.  E.  and  T.  W.  Sherry.  1994.  Parental  care 
at  nests  of  two  age  classes  of  male  American  Red- 
start: implications  for  female  mate  choice.  Condor 
96:606-613. 

Petit,  L.  1989.  Breeding  biology  of  Prothonotary  War- 
blers in  riverine  habitat  in  Tennessee.  Wilson  Bul- 
letin 101:51-61. 

Pinkowski,  B.  C.  1978.  Feeding  of  nestling  and  fledg- 
ling Eastern  Bluebirds.  Wilson  Bulletin  90:84-98. 

Podlesak,  D.  W.  and  C.  R.  Blem.  2001.  Factors  af- 
fecting growth  of  Prothonotary  Warblers.  Wilson 
Bulletin  113:263-272. 

Podlesak,  D.  W.  and  C.  R.  Blem.  2002.  Determina- 
tion of  age  of  nestling  Prothonotary  Warblers. 
Journal  of  Field  Ornithology  73:33-37. 

Ricklefs,  R.  E.  1974.  Energetics  of  reproduction  in 
birds.  Pages  152-291  in  Avian  energetics  (R.  A. 


Paynter,  Jr.,  Ed.).  Publications  of  the  Nuttall  Or- 
nithological Club,  no.  15.  Cambridge,  Massachu- 
setts. 

Ricklefs,  R.  E.  and  D.  J.  T.  Hussell.  1984.  Changes 
in  adult  mass  associated  with  the  nesting  cycle  in 
the  European  Starling.  Ornis  Scandinavica  15: 
155-161. 

Rytkonen,  S.,  K.  Koivula,  and  M.  Orell.  1996.  Pat- 
terns of  per-brood  and  per-offspring  provisioning 
efforts  in  the  Willow  Tit  Parus  montanus.  Journal 
of  Avian  Biology  27:21-30. 

Sanz,  J.  J.  and  J.  Moreno.  1995.  Mass  loss  in  brood- 
ing female  Pied  Flycatchers  Ficedula  hypoleuca : 
no  evidence  for  reproductive  stress.  Journal  of 
Avian  Biology  26:313-320. 

Sanz,  J.  J.,  J.  M.  Tinbergen,  M.  Orell,  and  S.  Ryt- 
konen. 1998.  Daily  energy  expenditure  during 
brood  rearing  of  Great  Tits  Parus  major  in  north- 
ern Finland.  Ardea  86:101-107. 

SAS  Institute,  Inc.  2000.  SAS/STAT  user’s  guide, 
ver.  8.2.  SAS  Institute,  Inc.,  Cary,  North  Carolina. 

Trivers,  R.  L.  1972.  Parental  investment  and  sexual 
selection.  Pages  136-179  in  Sexual  selection  and 
the  descent  of  man  (B.  Campbell,  Ed.).  Aldine 
Press,  Chicago,  Illinois. 

Zar,  J.  J.  1999.  Biostatistical  analysis.  Prentice-Hall, 
Upper  Saddle  River,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  1 18(1):  13— 22,  2006 


THE  REDISCOVERY  AND  NATURAL  HISTORY  OF  THE 
WHITE-MASKED  ANTBIRD  ( PITHYS  CASTANEUS) 

DANIEL  F.  LANE,1  6 THOMAS  VALQUI  H.,1 2 JOSE  ALVAREZ  A.,1 2 3 
JESSICA  ARMENTA,25  AND  KAREN  ECKHARDT4 5 6 


ABSTRACT. — In  July  2001,  a Louisiana  State  University  Museum  of  Natural  Science  expedition  rediscovered 
the  White-masked  Antbird  ( Pithys  castaneus ) at  a site  along  the  Rfo  Morona  in  northwestern  Departmento 
Loreto,  Peru.  Prior  to  this  rediscovery,  the  species  was  known  only  from  the  type  specimen,  taken  in  1937,  and 
nothing  was  recorded  concerning  its  natural  history.  The  lack  of  additional  specimens  led  to  speculation  that  P. 
castaneus  was  a hybrid.  Here,  we  present  data  demonstrating  that  the  White-masked  Antbird  is  a valid  species, 
and  we  report  the  first  observations  of  its  behavior,  habitat,  morphology,  and  voice.  Received  14  January  2005, 
accepted  1 1 October  2005. 


In  1938,  Berlioz  (1938)  described  a distinc- 
tive new  species  of  antbird  in  the  genus  Pith- 
ys— until  then  considered  monotypic — from  a 
single  specimen  collected  by  Ramon  Olalla  on 
16  September  1937  at  “Andoas,  lower  [Rio] 
Pastaza,  eastern  Ecuador.”  This  new  species, 
the  White-masked  Antbird  ( Pithys  castaneus ), 
has  remained  one  of  the  most  intriguing  mys- 
teries of  Neotropical  ornithology  for  over  60 
years  (see  David  and  Gosselin  2002  for  gen- 
der of  scientific  name).  Besides  the  collector, 
no  biologist  had  ever  seen  the  bird  alive,  and 
there  is  no  information  on  the  species’  natural 
history  or  preferred  habitat.  The  type  locality, 
“Andoas,”  is  particularly  intriguing  in  that  at 
least  three  sites  in  the  Pastaza  area  bear  this 
name  (Stevens  and  Traylor  1983,  Paynter 
1993),  and  according  to  T.  Mark  {in  lift.),  we 
may  never  really  know  the  true  location  of  the 
type  locality. 

The  type  specimen,  a male  (contra  Ridgely 
and  Tudor  1994),  is  housed  at  the  Paris  Mu- 
seum in  France.  According  to  Berlioz  (1938, 
1948),  it  was  part  of  a collection  that  included 
three  specimens  of  White-plumed  Antbird  (P. 
albifrons  peruvianas ) and  therefore  appeared 


1 Louisiana  State  Univ.  Museum  of  Natural  Science, 
1 19  Foster  Hall,  Baton  Rouge,  LA  70803,  USA. 

2 Dept,  of  Biological  Sciences,  Louisiana  State 
Univ.,  Baton  Rouge,  LA  70803,  USA. 

3 Inst,  de  Investigaciones  de  la  Amazonia  Peruana 
(IIAP),  Av.  Quinones  Km.  2.5,  Iquitos,  Peru. 

4 Museo  de  Historia  Natural  de  la  Univ.  Nacional 
Mayor  de  San  Marcos,  Apartado  14-0434,  Lima,  Peru. 

5 Current  address:  Dept,  of  Biological  Sciences,  P.O. 
Box  413,  Lapham  Hall,  Univ.  of  Wisconsin,  Milwau- 
kee, WI  53201,  USA. 

6 Corresponding  author;  e-mail:  dlane@lsu.edu 


to  be  a sympatric  congener.  It  differed  from 
P.  albifrons  in  its  larger  size,  its  lack  of  any 
gray  on  the  body,  and  its  lack  of  elongated 
plumes  on  the  face  or  throat. 

Decades  passed  without  any  additional  re- 
cords of  P.  castaneus.  Subsequent  authors 
doubted  the  validity  of  the  species,  and  many 
suggested  that  it  represented  nothing  more 
than  a hybrid  of  P.  albifrons  and  another  ant- 
bird species  (Sibley  and  Monroe  1990,  Schu- 
lenberg  and  Stotz  1991,  Collar  et  al.  1992, 
Stattersfield  and  Capper  2000,  Ridgely  and 
Greenfield  2001b).  Willis  (1984)  and  person- 
nel at  the  Philadelphia  Academy  of  Natural 
Sciences  (ANSP;  Collar  et  al.  1992,  Ridgely 
and  Tudor  1994)  searched  without  success  for 
P.  castaneus  along  the  upper  Rio  Pastaza  in 
Peru  and  Ecuador,  respectively. 

Thus,  when  our  Louisiana  State  University 
Museum  of  Natural  Science  (LSUMZ)  orni- 
thological field  team  visited  several  sites  in 
northwestern  Departamento  Loreto,  Peru, 
from  May  through  July  2001 , it  was  with  great 
surprise  that  we  found  P.  castaneus  to  be  fair- 
ly common  at  one  of  our  field  sites.  The  main 
goal  of  our  fieldwork  was  to  inventory  the 
avifauna  of  two  isolated  patches  of  varillal 
(white  sand)  forest  (see  Whitney  and  Alvarez 
1998;  Alvarez  and  Whitney  2001,  2003).  One 
of  these  forest  patches  was  in  the  interfluvium 
between  the  Morona  and  Santiago  rivers  in 
northern  Peru,  north  of  the  Rio  Maranon,  only 
about  60  km  west  of  the  Rfo  Pastaza,  and  it 
was  there  that  we  found  P.  castaneus. 

Remarkably,  while  reviewing  specimen  ma- 
terial at  the  Museo  de  Historia  Natural  de  la 
Universidad  Mayor  San  Marcos  (MUSM), 


13 


14 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


FIG.  1 . Known  localities  for  Pithy s castaneus  in  northwestern  Departmento  Loreto,  Peru.  The  star  represents 
suspected  location  of  “Andoas,”  the  type  locality,  on  the  Rio  Pastaza  (Berlioz  1938).  The  square  represents  the 
location  of  the  species’  rediscovery  in  July  2001  on  the  west  bank  of  the  Rio  Morona  (04°  17'  S,  77°  14'  W). 
The  Cordillera  Campanquis  lies  between  the  ribs  Morona  and  Santiago,  immediately  to  the  west  of  our  field 
site. 


Lima,  in  November  2002,  we  discovered  two 
additional  specimens  of  P.  castaneus  (one 
adult  and  one  juvenile).  These  specimens  were 
reportedly  taken  somewhere  in  the  Cordillera 
Campanquis  region  on  the  border  of  Depart- 
mentos  Amazonas  and  Loreto  between  the 
Morona  and  Santiago  rivers  (see  Fig.  1),  in 
the  mid-  to  late  1990s  by  a Peruvian  anthro- 
pologist, Andres  Treneman  (I.  Franke  J.  pers. 
comm.).  Unfortunately,  no  additional  speci- 
men data  are  available,  and  the  collector  could 
not  be  contacted  for  additional  information. 

METHODS 

Locality. — We  established  a campsite  on  the 
west  bank  of  the  Rio  Morona  about  54  km 
north-northwest  of  its  mouth  (04°  17'  S,  77° 
14' W;  Fig.  1),  Departmento  Loreto.  The 


study  site  was  on  the  south  side  of  the  mouth 
of  Quebrada  Cashacano,  a right-bank  tributary 
of  the  Rfo  Morona,  about  2.3  km  north  of  the 
village  of  Tierra  Blanca.  We  observed  and 
made  a general  collection  of  birds  at  this  site 
between  2 and  21  July  2001.  Our  camp  was 
set  up  in  a clearing  of  a homestead  abandoned 
about  30  years  earlier  and  which,  reportedly, 
has  been  reinhabited  since  our  visit  (B.  Walker 
pers.  comm.).  A preexisting  trail,  used  for  the 
harvest  of  palm  fronds  for  thatched-roof  con- 
struction, led  directly  into  white-sand  forests 
for  about  2 km.  Another  trail,  cut  along  the 
bluff  above  the  Morona,  connected  the  camp 
with  the  village  of  Tierra  Blanca.  From  this 
trail,  at  least  another  three  trails  also  entered 
the  varillal  forest.  Additional  trails  were  cut 
near  camp  for  census  routes  and  net  lanes; 


Lane  et  al.  • REDISCOVERY  OF  WHITE-MASKED  ANTBIRD 


15 


most  trails  were  in  varillal,  but  three  also  en- 
tered the  adjacent  varzea  (seasonally  inundat- 
ed) forest.  We  also  found  two  patches  of  richer 
clay-soil  terra  firme  forest  north  and  south  of 
the  surveyed  varillal  forest  patch,  into  which 
we  cut  two  trails. 

Habitat. — Most  of  the  forest  where  P.  cas- 
taneus  was  observed — particularly  away  from 
major  water  bodies — grew  on  very  moist, 
white-sand  soils.  Numerous  areas  of  wet, 
swampy  conditions  indicated  a high  water  ta- 
ble. The  terrain  was  without  significant  relief, 
but  throughout  the  varillal  forest  were  many 
small  depressions  where  water  accumulated 
(particularly  after  rains),  presumably  pits  re- 
sulting from  tree-falls.  The  soil  consisted  of 
rather  coarse  sand  with  stones  of  up  to  5 cm 
in  diameter  (up  to  15  cm  in  the  small  creeks 
that  transected  the  forest  interior).  Using  a nat- 
ural cut  at  the  Rio  Morona  riverbank  for  ref- 
erence, the  sandy  soil  is  approximately  4 m 
deep  at  the  river’s  edge.  Typical  of  many  var- 
illal forests,  a thick  layer  of  dead  leaves  and 
humus  covered  the  forest  floor  (Ruokolainen 
and  Tuomisto  1993,  1998;  Richards  1996). 
The  forest  canopy  of  the  varillal  was  relative- 
ly even,  with  a height  of  about  20  to  30  m. 
The  relative  absence  of  buttressed  trees  is  typ- 
ical of  varillal  forests  (Richards  1996);  how- 
ever, many  such  trees  were  present  in  more 
humid  forest  areas  at  the  Morona  site.  As  has 
been  noted  in  other  varillal  forests  (Anderson 
1981,  Richards  1996),  there  were  few  lianas, 
and  epiphytic  growth  was  negligible. 

Data  collection. — We  collected  specimens 
using  mist  nets  and  shotguns.  Permits  for 
specimen  collection  were  issued  by  Peru’s  In- 
stitute Nacional  de  Recursos  Naturales  (IN- 
RENA).  Specimens  were  deposited  into  the 
collections  of  LSUMZ  and  MUSM.  Skull  os- 
sification, gonad  information,  and  presence  of 
fat  in  prepared  specimens  were  determined 
following  standard  LSUMZ  specimen  prepa- 
ration protocol.  Natural  history  information 
was  acquired  through  opportunistic  (not  sys- 
tematic) encounters  with  P.  castaneus.  Spec- 
trograms of  voice  recordings  were  prepared 
using  Canary  sound  analysis  software  (Charif 
et  al.  1995). 

Specimens  examined. — Pithy s castaneus : 
Peru:  Loreto;  west  bank  of  Rio  Morona,  —54 
km  NNW  of  the  mouth,  140  m elevation  (04° 
17'  S,  77°  14'  W)  (LSUMZ  172973,  172974, 


172975,  172976  [skeleton  and  partial  skin], 
172977,  172978,  172979  [skeleton  and  partial 
skin],  MUSM  23504,  23505,  23506,  23507; 
DFL  1646  [skeleton,  uncataloged],  TVH  399 
[alcohol,  uncataloged]). 

Pithys  albifrons : Ecuador:  Pastaza;  Coco- 
naco,  300  m elevation  (LSUMZ  83237);  Peru: 
Amazonas;  Huampami,  —215  m elevation 
(LSUMZ  84917),  Chiriaco,  -320  m elevation 
(LSUMZ  78514,  88018,  88019,  88022);  Lor- 
eto; Libertad,  S bank  of  Rio  Napo,  80  km  N 
of  Iquitos,  120  m elevation  (LSUMZ  1 10094, 
110096,  110097,  110098,  110099,  110100, 
1 10102,  1 10103,  1 10104,  1 10105);  157  km  by 
river  NNE  of  Iquitos,  N of  Rio  Napo,  110  m 
elevation  (LSUMZ  110106,  110109,  110112, 
110113). 

Gymnopithys  leucaspis:  Peru:  Loreto;  west 
bank  of  Rio  Morona,  —54  km  NNW  of  the 
mouth,  140  m elevation  (04°  17'  S,  IT  14'  W) 
(LSUMZ  172985);  Quebrada  Oran,  -5  km  N 
of  Rio  Amazonas,  85  km  NE  of  Iquitos,  1 10 
m elevation  (LSUMZ  119884,  1 19885, 
119886,  119887,  119890,  119891,  119892, 
119893). 

Phlegopsis  erythroptera:  Ecuador:  Sucum- 
bios;  Limoncocha,  300  m elevation  (00°  24' 
S,  76°  37'  W)  (LSUMZ  70916,  70917,  70919, 
83314).  Peru:  Loreto;  W bank  of  Rio  Morona, 
—54  km  NNW  of  the  mouth,  140  m elevation 
(04°  17'  S,  IT  14'  W)  (LSUMZ  173001);  1.5 
km  S of  Libertad,  S bank  of  Rio  Napo,  80  km 
N of  Iquitos,  120  m elevation  (LSUMZ 
110213,  110215,  110217);  1 km  N of  Rio 
Napo,  157  km  by  river  NNE  of  Iquitos,  110 
m elevation  (LSUMZ  110219);  lower  Rio 
Napo  region,  E bank  of  Rio  Yanayacu,  —90 
km  N of  Iquitos,  120  m elevation  (LSUMZ 
115573). 

Rhegmatorhina  melanosticta:  Peru:  Ama- 
zonas; headwaters  of  Rio  Kagka  (of  Rio  Ce- 
nepa),  —790  m elevation  (04°  16'  S,  78°  09' 
W)  (LSUMZ  88028,  88029);  San  Martin;  -15 
km  by  trail  NE  of  Jirillo  on  trail  to  Balsa- 
puerto,  1,350  m elevation  (LSUMZ  116947); 
Huanuco;  —35  km  NE  Tingo  of  Marfa,  Ha- 
cienda Santa  Elena,  —1,000  m elevation 
(LSUMZ);  Pasco;  Abra  Aguachini,  —30  km 
SW  of  Puerto  Bermudez,  1,020  m elevation 
(LSUMZ  130274);  Pasco;  Puellas,  km  41  on 
Villa  Rica-Puerto  Bermudez  highway,  950  m 
elevation  (LSUMZ  106073,  106074,  106078). 


16 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


RESULTS 

Specimen  data. — We  collected  13  speci- 
mens of  Pithys  castaneus  during  our  visit  to 
the  Rio  Morona  site.  We  prepared  nine  as 
study  skins  (from  which  several  trunk  skele- 
tons were  saved),  three  as  complete  skeletons 
(from  which  two  partial  skin  specimens  were 
saved),  and  one  was  preserved  whole  in  al- 
cohol. Mass  and  lengths  of  flat-wing,  tail,  tar- 
sus, and  culmen  (from  distal  edge  of  the  nares 
to  bill  tip)  of  all  specimens  are  presented  and 
compared  with  measurements  of  the  P.  cas- 
taneus holotype  and  other  Peruvian  ant 
swarm-following  antbirds  (Table  1). 

Three  of  the  12  specimens  in  “adult”  plum- 
age (LSUMZ  172973,  MUSM  23504,  MUSM 
23507)  still  possessed  a bursa  of  Fabricius  and 
one  had  an  incompletely  ossified  skull  (75% 
ossification),  suggesting  that  first  basic  plum- 
age is  acquired  quickly  and  is  nearly  indistin- 
guishable from  definitive  plumage  (but  see  be- 
low). One  specimen  (LSUMZ  172978)  was  a 
male  still  largely  in  juvenal  plumage  (skull  os- 
sification 50%,  bursa  8X6  mm).  Of  the  12 
specimens  dissected,  only  2 — both  with  im- 
mature characters — were  reported  to  have 
subcutaneous  fat  deposits:  “trace  fat”  in  one 
and  “light  fat”  in  the  other.  Six  of  12  speci- 
mens dissected  exhibited  trace  or  light  body 
molt.  Seven  individuals  had  asymmetrical 
wing  molt,  and  seven  had  asymmetrical  tail 
molt.  Stomach  contents  were  reported  as  “in- 
sect parts”  for  all  specimens  in  which  the 
stomachs  were  not  empty.  The  guts  of  two 
specimens  were  infested  with  nematodes. 

Variation  in  the  series. — Twelve  speci- 
mens— 5 males  and  7 females — exhibited 
similar  plumage,  with  no  sexual  dichroma- 
tism. All  these  adults  appeared  to  match  the 
description  of  P.  castaneus  and  the  photos  of 
the  holotype  very  closely.  Of  the  specimens 
in  “adult”  plumage,  two  that  appeared  to  be 
in  their  first  year  (see  above)  have  very  sparse, 
light-grayish  scaling  on  the  center  of  the 
throat  (unmarked  white  in  all  other  individu- 
als), suggesting  that  it  may  be  an  age-related 
character.  Otherwise,  plumage  characters  were 
uniform  among  all  the  “adult”  specimens. 
The  juvenal-plumaged  bird  differs  in  being 
washed  with  colder  brown  overall,  particular- 
ly on  the  breast,  vent,  and  center  of  the  back. 
Furthermore,  the  white  of  the  juvenile’s  face 


dddddbod 


r-  (N 
o 

+1  +1  +1  +|  +i  +|  +i  +i  +i  +i 
"qqa^qooqaoi;^ 
ri  n'  d O ri  ri  ri  ^ 


in  oo  io 
o o o o o 

ro  +|  +i  +i  +i  +i 
<N 

o i;  q in  o 
mi  oi  —i  no 

04  04  04  Ol  04 


On  On  ~ On 
04  — : — ' o 

+1  +1  +1  +1  +1 

oj  cq  ^ no  on 

oi  oi  b *n  cd 

in  m m 


O O-  O »/">  CO 


04  +|  +|  +|  +|  +| 
00 

oo  no  >n  >n  >n 


h n 1;  q h 

oi  — i oi  — oi 

+1  +1  +1  +1  +1 

q o in  o; 

On  d oo  oo  rd 


04  CO  < — 


04 


— no  m O O i/n 


a C 
p & 


a.  a 


£1 


5 c M 
S C « 2 
^ ^ c .5 


s c 

.a  £ 


"q  "q  5. 


£ £ £ £ £ 6' 


Gymnopithys  leucaspis  (females)  5 23.8  ± 2.3  71.9  ± 1.3  42.9  ± 2.1  25.8  ± 1.2 

Phlegopsis  erythroptera  (males)  5 58.4  ± 5.2  91.6  ± 1.7  63.3  ± 2.4  33.5  ±1.7 

Phlegopsis  erythroptera  (females)  5 58.2  ± 7.0  88.0  ± 1.3  59.0  ± 1.0  32.0  ± 1.4 

Rhegmatorhina  melanosticta  (males)  4 30.0  ±1.9  81.3  ± 5.2  53.0  ± 2.0  27.6  ±1.2 

Rhegmatorhina  melanosticta  (females)  5 33.0  ± 4.7  78.0  ± 2.8  49.8  ± 2.5  27.4  ± 0.8 


Lane  et  al.  • REDISCOVERY  OF  WHITE-MASKED  ANTBIRD 


17 


TABLE  2.  Number  of  individuals  per  species  attending  army  ant  swarms  ( Eciton  burchelli  and  Labidus 
praedator ) with  Pithys  castaneus , Departmento  Loreto,  Peru,  July  2001.  Columns  represent  individual  swarms. 
Only  swarms  observed  for  >15  min  were  included. 


Date  (ant  swarma) 


4 July 

6 July 

6 July 

8 July 

10  July 

11  July 

12  July 

14  July 

17  July 

(E) 

(E) 

(E) 

(E) 

(L.) 

(E) 

(E) 

(L) 

(E) 

Pithys  castaneus 

2 

4 

3 

3 

1 

1 

4 

4 

3 

Pithys  albifrons 

3 

5 

— 

— 

— 

— 

— 

— 

— 

Phlegopsis  erythroptera 

— 

2 

— 

— 

— 

— 

— 

— 

— 

Gymnopithys  leucaspis 

5 

4 

2 

2 

— 

3 

2 

4 

4 

Hylophylax  poecilinota 

— 

2 

2 

1 

1 

— 

— 

— 

— 

Percnostola  arenarum 

1 

— 

1 

— 

— 

— 

1 

1 

2 

Dendrocolaptes  certhia 

1 

3 

— 

— 

— 

— 

— 

— 

— 

Dendrocincla  merula 

— 

— 

— 

— 

— 

— 

— 

1 

1 

Xiphorhynchus  ocellatus 

2 

2 

— 

— 

— 

1 

— 

1 

1 

Deconychura  longicauda 

1 

— 

— 

— 

— 

— 

— 

— 

a E = Eciton  burchelli , L = Labidus  praedator. 


was  restricted  to  the  area  between  the  eye  and 
gape  and  a longitudinal  line  along  the  center 
of  the  throat.  This  specimen’s  dark  head  mark- 
ings were  more  extensive  than  those  on  defin- 
itive-plumaged  birds,  and  they  were  a duller, 
sooty,  dark  brown  (see  frontispiece). 

Soft-part  colors  were  relatively  uniform 
across  most  specimens.  The  irides  were  brown 
or  dark  brown  (all  soft-part  colors  taken  from 
tag  data  recorded  at  time  of  preparation)  in 
nine  specimens  with  adult  characters,  dark 
gray-brown  in  the  three  specimens  with  first- 
basic  characters,  and  dark  gray  in  the  juvenile. 
Thus,  iris  color  evidently  changes  from  gray 
to  dark  brown  as  an  individual  ages.  In  all 
specimens,  the  maxilla  was  blackish-slate 
with  a silvery-white  tomium,  the  latter  con- 
stricted at  mid-bill  in  some  individuals.  Man- 
dible coloration  varied  more.  Most  adults  had 
a mostly  silvery-white  tomium  with  blackish- 
slate  color  on  the  gonys  and  base  of  the  man- 
dible (except  the  tomium).  Approximately  the 
distal  half  of  the  juvenile’s  bill  was  silvery- 
white,  and  the  mouth  interior  was  dark  gray. 
The  tarsus  color  of  adult  individuals  was 
brownish-orange  or  ochre-orange;  the  juve- 
nile’s tarsi  were  dirty  yellow  with  a gray  tinge. 
The  toes  were  dirty  yellow,  pale  orange,  or 
dull  saffron  yellow;  the  claws  of  the  juvenile 
bird  were  gray. 

Behavioral  observations. — Our  initial  ob- 
servations of  P.  castaneus  were  made  by  TVH 
and  DFL  at  a swarm  of  Eciton  burchelli  army 
ants  on  4 July  2001,  when  the  first  specimens 
were  collected.  Based  on  our  observations,  we 


were  confident  in  labeling  P.  castaneus  a pro- 
fessional army  ant-follower  ( sensu  Willis 
1967).  We  never  saw  it  foraging  away  from 
army  ant  swarms  and  observed  it  attending 
swarms  of  two  army  ant  species:  Eciton  bur- 
chelli and  Labidus  praedator.  For  at  least  1 2— 
15  min  on  8 July,  JAA  observed  a single  in- 
dividual of  P.  castaneus  with  a female  Scale- 
backed  Antbird  ( Hylophylax  poecilinota ) fol- 
lowing a swarm  of  L.  praedator  ants  that  oc- 
cupied less  than  10  m2  of  the  forest  floor.  The 
bird’s  behavior  was  similar  to  that  of  others 
observed  following  swarms  of  E.  burchelli. 
Both  the  P.  castaneus  and  the  H.  poecilinota 
individuals  left  the  swarm  for  3-4  min,  only 
to  return  later.  Also  observed  attending 
swarms  of  L.  praedator  (although  independent 
of  the  above  observation)  were  Allpahuayo 
Antbirds  ( Percnostola  arenarum),  a species 
previously  unknown  as  an  ant-follower  (Isler 
et  al.  2001,  Zimmer  and  Isler  2003),  and  Bi- 
colored Antbirds  ( Gymnopithys  leucaspis).  On 
four  occasions  on  different  days,  we  observed 
a single  individual  of  P.  castaneus  quietly 
passing  through  the  forest  without  foraging, 
suggesting  movement  between  ant  swarms  or 
between  an  ant  swarm  and  a nest  (Willis 
1981).  In  Table  2,  we  present  the  attendance 
of  regular  ant-following  species  observed  at 
swarms  at  the  Morona  site. 

Most  often,  P.  castaneus  was  observed  at 
or  near  the  broad  front  of  a moving  ant  col- 
umn. Individuals  tended  to  perch  <0.5  m 
above  ground  and  frequently  dropped  to  the 
forest  floor  to  investigate  leaf  litter  or  capture 


18 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


arthropods.  Birds  often  were  observed  attend- 
ing a swarm  for  5 to  15  min  at  a time  and 
then  leaving  the  swarm  (at  least  once  while 
carrying  a food  item)  for  roughly  equal  peri- 
ods of  time.  On  at  least  one  such  occasion,  a 
pair  of  P.  castaneus  was  observed  joining  a 
family  group  of  G.  leucaspis  moving  between 
what  appeared  to  be  two  column  heads  (about 
30  m apart)  of  a single  E.  burchelli  ant  swarm. 
Willis  (1981)  reported  similar  behavior  for  P. 
albifrons.  On  another  occasion,  a single  indi- 
vidual was  seen  moving  around  a standing 
hollow  tree  in  which  a swarm  of  E.  burchelli 
had  bivouacked  the  previous  evening,  but  had 
not  yet  started  its  morning  activity. 

Most  of  the  professional  ant-following 
thamnophilids  at  the  Morona  site  regularly 
made  exaggerated  tail  “pounding”  or  “wag- 
ging” movements  (terms  following  Zimmer 
and  Isler  2003)  while  foraging  at  ant  swarms, 
especially  upon  returning  to  a perch  after 
pouncing  on  a prey  item,  or  when  agitated  by 
the  presence  of  an  observer.  P.  castaneus  was 
not  observed  regularly  using  any  such  tail 
movement.  Only  once  or  twice  did  we  notice 
an  individual  pound  its  tail,  usually  after  a 
pouncing  attack  on  prey;  the  tail  movement 
was  made  once  and  not  repeated.  By  contrast, 
DFL  noted  that  the  G.  leucaspis  almost  con- 
stantly wagged  their  tails  laterally,  although 
this  contrasts  with  the  published  observations 
of  others  (e.g.,  Zimmer  and  Isler  2003).  In  ad- 
dition, DFL  observed  both  P.  albifrons  and 
the  Reddish-winged  Bare-eye  ( Phlegopsis  er- 
ythroptera ) regularly  pounding  their  tails 
downward  (also  see  Willis  1981,  1984;  Zim- 
mer and  Isler  2003).  We  were  unable  to  de- 
termine whether  such  tail  movements  are  in- 
tended as  a form  of  inter-  or  intraspecific 
“body  language”  among  swarm  attendants,  as 
a sign  of  agitation,  or  as  a form  of  flushing 
insect  prey.  Nevertheless,  the  relative  lack  of 
such  tail-moving  behavior  in  P.  castaneus 
seems  noteworthy.  Willis  (1968)  reports  that 
the  monotypic  genus  Skutchia  also  lacks  ste- 
reotypic tail-moving  behavior,  but  other  ob- 
servers contest  this  (B.  M.  Whitney  pers. 
comm.). 

In  our  observations  of  ant-following  birds 
at  the  Morona  site  (Table  2),  we  noted  several 
occurrences  of  one  ant-following  species  sup- 
planting another  near  the  leading  edges  of  ant 
swarms  and  took  this  to  represent  a domi- 


nance hierarchy  among  the  attendant  species 
(see  Willis  1967,  1981).  From  our  observa- 
tions, we  concluded  that  the  dominance  hier- 
archy (from  most  to  least  dominant)  was  Phle- 
gopsis erythroptera , Pithys  castaneus , and  G. 
leucaspis.  Other  swarm-attending  antbirds,  in- 
cluding Pithys  albifrons,  noticeably  avoided 
the  leading  edge  of  the  swarm  when  any  of 
the  other  professional  ant-followers  were  pre- 
sent. Our  observations  of  the  last  species 
agree  with  those  of  Willis  (1981),  who  also 
termed  P.  albifrons  a subordinate  ant  swarm 
attendant.  Since  the  dominance  hierarchy  sug- 
gested above  has  a positive  correlation  to 
overall  body  size,  we  suggest  that  size  may  be 
the  ultimate  cause  (or,  alternatively,  a proxi- 
mate cause — i.e.,  a source  of  maintenance)  of 
the  hierarchy  (see  Table  1). 

Voice. — We  recorded  at  least  seven  distinct 
vocalizations  from  P.  castaneus  (Fig.  2),  in- 
cluding a mewed  whistle  that  rises  in  frequen- 
cy (Fig.  2A).  This  is  a single  note  often  given 
quietly,  although  occasionally  it  can  be  quite 
loud,  and  we  suspect  represents  the  species’ 
“loudsong”  (such  as  it  is).  To  our  knowledge, 
P.  albifrons  does  not  give  a true  loudsong 
(sensu  Willis  1967,  Isler  et  al.  1998,  Isler  and 
Whitney  2002,  Zimmer  and  Isler  2003)  as  do 
most  other  thamnophilids.  However,  the  spe- 
cies is  known  to  produce  a vocalization  sim- 
ilar to  that  described  above  for  P.  castaneus : 
a rarely  heard,  weak,  mewing  whistled  vocal- 
ization that  falls  in  frequency  and  is  suspected 
to  serve  as  a song  (Willis  1981,  Isler  and 
Whitney  2002;  Fig.  2B).  The  whistled  notes 
of  the  loudsong  of  P.  castaneus  appear  to  be 
punctuated  by  occasional  quiet,  chiming  notes 
(Fig.  2C),  perhaps  an  integral  part  of  the  loud- 
song. Song  intervals  can  be  as  short  as  2 sec 
but  often  are  longer. 

P.  castaneus  also  produced  two  vocaliza- 
tions when  alarmed  or  when  agitated  by  play- 
back of  what  we  believed  was  the  species’ 
song  (see  below).  These  notes  of  agitation 
were  interspersed  with  sharp  chattered  “chit!” 
calls  (Fig.  2D),  similar  to  the  “chip”  calls  de- 
scribed for  P.  albifrons  by  Willis  (1981).  An- 
other vocalization  given  by  agitated  birds  was 
a louder,  higher-pitched  “chirr,”  with  the  in- 
dividual notes  more  distinct  (Fig.  2E)  than  in 
the  undisturbed  chirr  call  (see  below).  Occa- 
sionally, the  agitated  chirr  commenced  with  a 
chit  note  (Fig.  2F).  While  giving  these  vocal- 


Lane  et  al.  • REDISCOVERY  OF  WHITE-MASKED  ANTBIRD 


19 


FIG.  2.  Sound  spectrograms  of  antbird  vocalizations.  Unless  otherwise  noted,  all  recordings  were  made  by 
D.  F.  Lane  at  our  Rfo  Morona  locality,  Departmento  Loreto,  Peru,  July  2001.  (A)  “Song”  of  Pithys  castaneus. 
(B)  “Song”  of  Pithys  albifrons  (T.  A.  Parker,  III,  and  G.  F.  Budney,  from  Isler  and  Whitney  2002).  (C)  “Chime” 
of  Pithys  castaneus.  (D)  “Chit”  of  Pithys  castaneus.  (E)  Agitated  “Chirr”  of  Pithys  castaneus.  (F)  “Chit-chirr” 
of  Pithys  castaneus.  (G)  “Mew”  of  Pithys  castaneus  (J.  Alvarez  A.).  (H)  “Chirr”  of  Pithys  castaneus.  (I) 
“Chirr”  of  Pithys  albifrons.  (J)  “Chirr”  of  Gymnopithys  leucaspis.  (K)  “Chirr”  of  Phlegopsis  erythroptera. 


izations  of  agitation,  one  male  (sex  confirmed 
by  collection),  was  observed  perched  on  a 
horizontal  branch  at  the  edge  of  a treefall  gap 
about  2 m above  the  ground.  This  was  the 
highest  we  ever  observed  the  species  to  perch, 
and  was  likely  an  agitation  response  to  play- 
back of  the  song.  On  one  occasion,  a distinct, 
quiet,  mewing  “eaaah”  call  was  given  by  two 
individuals  while  close  to  one  another;  we  in- 
terpret this  as  some  sort  of  contact  call  or 
“softsong”  within  the  pair  (Fig.  2G). 

The  most  common  vocalization  was  a call 
given  by  individuals  while  foraging  at  ant 
swarms.  This  was  a deep  chirr  call  (terms  fol- 
lowing Willis  1967,  Zimmer  and  Isler  2003; 
Fig.  2H),  similar  to  vocalizations  given  by 
most  professional  ant-following  thamnophil- 


ids  when  attending  ant  swarms,  and  suspected 
to  be  a means  of  maintaining  individual  for- 
aging space  at  the  swarm  (Willis  1967;  M.  L. 
Isler  in  litt.).  When  the  chirr  of  P.  castaneus 
was  heard  simultaneously  with  those  of  most 
of  the  other  species  of  professional  ant-fol- 
lowers  at  a swarm,  it  sounded  generally  loud- 
er, of  lower  overall  frequency,  and  descended 
less  obviously  (see  Fig.  2H-2K).  Only  the 
chirr  call  of  Phlegopsis  erythroptera  (Fig.  2K) 
reaches  a frequency  as  low  as  that  of  Pithys 
castaneus , but  the  former  can  be  distinguished 
easily  by  a higher,  more  metallic  introductory 
sound  and  a more  sharply  descending  com- 
ponent. The  chirr  call  of  Phlegopsis  erythrop- 
tera was  louder  than  that  of  Pithys  castaneus 
on  occasion,  but  this  appeared  to  be  influ- 


20 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


enced  by  emotional  state  and  was  not  always 
the  case. 

Playback  experiments  using  recordings  of 
the  suspected  song  elicited  varying  reactions 
from  individuals:  some  responded  immediate- 
ly, giving  agitated  calls  and  posing  on  exposed 
perches  that  were  higher  than  typical  perches 
(see  above),  while  others  approached  silently 
to  investigate.  On  two  occasions,  individuals 
approached  only  after  2-3  min  of  playback. 
Playback  of  chirr  calls  resulted  in  a quiet,  cu- 
rious approach  at  best. 

DISCUSSION 

Taxonomic  status  of  the  species. — Whereas 
the  generic  allocation  of  Pithys  castaneus  has 
been  considered  dubious,  we  believe  that  phe- 
notypic characters  such  as  the  species’  song- 
like vocalization,  its  bold  chestnut  plumage, 
black  hood  and  white  face,  and  its  saffron- 
yellow  legs  all  suggest  a close  relationship 
with  P.  albifrons.  Furthermore,  R.  T.  Brum- 
field and  J.  G.  Tello  (unpubl.  data)  have  been 
building  a molecular  phylogeny  of  the  Tham- 
nophilidae,  and  have  found  P.  castaneus  and 
P.  albifrons  to  be  sister  taxa. 

Potential  habitat  specialization. — Based  on 
our  observations,  we  suspect  that  P.  castaneus 
is  restricted  to  varillal  forests.  We  should 
note,  however,  that  we  observed  and  mist-net- 
ted P.  castaneus  individuals  that  had  followed 
ant  swarms  from  varillal  into  varzea  forest 
immediately  adjacent  to  our  campsite,  and 
twice  we  recorded  individuals  on  richer,  hilly 
terra  firme  forest  within  300  m of  typical  var- 
illal habitat.  We  never  encountered  Hairy- 
crested  Antbird  ( Rhegmatorhina  melanosticta ) 
at  the  Morona  site  and  wonder  whether  it  may 
be  replaced  by  the  similarly  sized  P.  casta- 
neus (see  Table  1)  in  the  region  or  (more  like- 
ly) habitat.  We  can  find  no  evidence  that  R. 
melanosticta  inhabits  the  region  between  the 
rios  Santiago  and  Pastaza,  but  it  is  quite  pos- 
sible that  this  is  due  to  poor  sampling  as  it  is 
to  true  absence.  If  R.  melanosticta  competi- 
tively excludes  P.  castaneus  outside  the  Mo- 
rona-Pastaza  varillal  forest,  this  may  explain 
the  restricted  distribution  of  the  latter  species. 
Furthermore,  if  varillal  forest  habitat  was  not 
included  in  the  searches  conducted  by  Willis 
and  the  ANSP  expedition  along  the  Pastaza, 
their  failure  to  encounter  the  species  may  be 


explained  by  the  possible  habitat  specializa- 
tion of  P.  castaneus. 

Potential  distribution  of  Pithys  castaneus. — 
Landsat  imagery,  complemented  with  infor- 
mation from  Instituto  de  Investigaciones  de  la 
Amazonia  Peruana  personnel  and  local  peo- 
ple, shows  what  we  interpret  to  be  fairly  large 
blocks  of  varillal  forest  embedded  within  a 
quadrangle  formed  by  the  Rio  Maranon  to  the 
south,  the  Rio  Morona  to  the  east,  the  Rio 
Mayuriaga  to  the  north,  and  the  Cordillera 
Campanquls  to  the  west.  Besides  this  area,  P. 
castaneus  populations  are  likely  to  occur  in 
similar  forest  along  the  Rio  Pastaza  in  Loreto 
and  probably  into  Ecuador.  At  present,  we 
have  no  information  about  the  existence  of 
varillal  forest  at  the  latter  sites.  However, 
some  indicator  species  of  varillal  forest  have 
been  found  along  the  upper  Rio  Pastaza  in  Ec- 
uador (e.g..  Pompadour  Cotinga,  Xipholena 
punicea,  and  Red-fan  Parrot,  Deroptyus  accip- 
itrinus\  Ridgely  and  Greenfield  2001a),  sug- 
gesting that  the  area  probably  supports  varillal 
forest  habitat.  We  suspect  that  once  such  for- 
ests along  the  upper  Rio  Pastaza  are  located 
and  surveyed,  the  mystery  of  the  true  position 
of  the  “Andoas”  collecting  locality  finally 
will  be  unraveled. 

Conservation. — The  west  bank  of  the  Rio 
Morona,  including  the  areas  of  varillal  forest 
where  our  work  was  conducted,  are  part  of  the 
recently  created  Zona  Reservada  Santiago  Co- 
maina,  created  in  1999.  According  to  Peruvian 
legislation,  its  new  status  is  temporary,  but 
supposedly,  it  will  be  ranked  as  a definitive 
conservation  unit  in  the  future  (National  Park, 
National  Reserve,  National  Sanctuary,  or 
Communal  Reserve).  However,  local  leaders 
of  the  Federacion  de  Comunidades  Indlgenas 
del  Rio  Morona  informed  us  that  they  strongly 
oppose  the  creation  of  a reserve  and  will  fight 
to  prevent  this  action. 

A branch  of  the  North-Peruvian  oil  pipeline 
that  transports  oil  from  the  upper  Rio  Pastaza 
passes  through  a large  portion  of  varillal  for- 
est as  it  crosses  the  Rio  Mayuriaga  on  its  way 
to  the  Rio  Maranon.  At  present,  this  has  meant 
the  destruction  of  only  a 50-m-wide  swath  of 
forest  along  the  pipeline.  However,  an  oil  spill 
could  have  drastic  consequences  for  this  rath- 
er delicate  habitat,  particularly  with  its  flat  ter- 
rain and  poor  drainage.  Furthermore,  the  pipe- 
line itself  could  represent  a potential  dispersal 


Lane  et  al.  • REDISCOVERY  OF  WHITE-MASKED  AN'I  BIRD 


21 


barrier  for  P.  castaneus.  It  is  also  worth  men- 
tioning that  there  are  several  plans  to  connect 
Ecuador’s  Amazonian  road  network  to  the  Rio 
Maranon.  Anecdotal  evidence  suggests  that 
many  bird  species  of  interior  forest  understory 
are  averse  to  crossing  large  openings  or  other 
similar  breaks,  such  as  rivers  or  roads  (Zim- 
mer and  Isler  2003).  Thus,  gaps  such  as  those 
associated  with  roads  and  pipelines  may  pose 
barriers  to  gene  flow  in  populations  of  these 
understory  species. 

Population  estimate. — During  our  stay  we 
surveyed  about  8 km2  of  white-sand  forests 
and  encountered  between  six  and  eight  differ- 
ent army  ant  swarms  of  E.  burchelli  and  two 
of  L.  praedator.  Based  on  our  extrapolations, 
we  estimate  the  number  of  P.  castaneus  to  be 
between  18  and  26  individuals  in  the  area  we 
surveyed.  If  we  consider  the  immediate  area 
(the  Morona-Santiago  interfluvium)  covered 
with  varillal,  then  the  population  estimate  of 
P.  castaneus  would  be  —1,300-2,500  individ- 
uals. Prior  to  our  rediscovery  of  P.  castaneus, 
the  species  was  considered  to  be  rare,  with  a 
very  restricted  global  distribution,  and  prob- 
ably threatened  (Bibby  1992,  Stattersfield  and 
Capper  2000).  Considering  the  population  es- 
timates and  the  potential  threats  presented 
here,  we  recommend  changing  the  species’ 
status  from  Data  Deficient  to  Vulnerable,  ac- 
cording to  the  ranking  criteria  presented  in 
Stattersfield  and  Capper  (2000).  If  a road  or 
any  other  invasive  construction  project  threat- 
ens the  white-sand  forests  between  the  rlos 
Morona  and  Santiago,  then  the  species’  status 
should  be  upgraded  to  a category  of  higher 
risk. 

Since  our  rediscovery  of  P.  castaneus  in 
July  2001,  and  our  discovery  of  the  two  Tre- 
neman  specimens  in  MUSM,  we  have  been 
informed  of  two  subsequent  observations  of 
P.  castaneus  by  colleagues  who  visited  our 
Morona  site.  Observers  visited  the  site  22-24 
June  2002  and  24  May  2003  (M.  Levy,  J. 
Nilsson,  M.  Sokol,  and  B.  Walker  pers. 
comm.).  Both  parties  saw  the  species,  but  the 
2002  observation  was  of  multiple  individuals 
and  the  observers  regarded  the  species  as 
“one  of  the  most  common  birds”  at  the  site. 
During  the  2003  visit,  however,  only  one  in- 
dividual was  observed,  possibly  because 
swarms  of  army  ants  were  not  easily  encoun- 
tered then  (an  artifact  of  the  season?). 


ACKNOWLEDGMENTS 

We  thank  J.  P.  O’Neill  as  the  initiator  and  organizer 
of  the  2001  expedition.  Funding  for  the  2001  expedi- 
tion was  received  from  the  Coypu  Foundation  and  a 
donation  from  the  late  R.  B.  Wallace.  Additional  fund- 
ing was  provided  to  JAA  by  a Fulbright  Scholarship. 
Permits  for  fieldwork  were  granted  by  INRENA,  and 
we  particularly  appreciate  the  efforts  of  M.  Prieto  C. 
and  R.  Acero  V.  of  that  institution.  D.  Huachaca,  M. 
Sanchez  S.,  A.  Urbay  T.,  M.  Pizango,  M.  Tenazoa,  H. 
Pizango,  F.  Salazar,  and  R.  Sandoval  all  provided  lo- 
gistical support  in  the  field.  M.  L.  and  P.  R.  Isler  gra- 
ciously allowed  us  access  to  their  data  and  recording 
collections,  and  freely  provided  the  fine  maps  and 
sonograms  for  our  figures.  M.  Levy,  J.  Nilsson,  M. 
Sokol,  and  B.  Walker  all  related  information  from  their 
visits  to  the  Morona  site.  T.  Mark  provided  us  with  his 
Andoas  manuscript,  and  he  and  J.  W.  Fitzpatrick  pro- 
vided information  and  photos  of  the  holotype  at  the 
Paris  Museum.  L.  B.  McQueen  produced  the  illustra- 
tion that  allowed  us  to  recognize  Pithys  castaneus 
while  in  the  field.  T.  S.  Schulenberg  provided  us  with 
rare  reprints  and  data  collected  on  Peruvian  antbirds. 
I.  Franke  J.  kindly  allowed  us  access  to  the  ornitho- 
logical collection  in  her  care  at  MUSM,  and  assisted 
us  in  trying  to  find  more  details  of  the  two  “mystery” 
Pithys  castaneus  specimens  there.  This  manuscript 
benefited  from  comments  by  M.  L.  and  P.  R.  Isler,  J. 
V.  Remsen,  Jr.,  T.  S.  Schulenberg,  B.  Walker,  B.  M. 
Whitney,  K.  J.  Zimmer,  and  two  anonymous  reviewers. 

LITERATURE  CITED 

Alvarez  A.,  J.  and  B.  M.  Whitney.  2001.  A new 
Zimmerius  tyrannulet  (Aves:  Tyrannidae)  from 
white  sand  forests  of  northern  Amazonian  Peru. 
Wilson  Bulletin  113:1-9. 

Alvarez  A.,  J.  and  B.  M.  Whitney.  2003.  New  dis- 
tributional records  of  birds  from  white-sand  for- 
ests of  the  northern  Peruvian  Amazon,  with  im- 
plications for  biogeography  of  northern  South 
America.  Condor  105:552-566. 

Anderson,  A.  B.  1981.  White-sand  vegetation  of  Bra- 
zilian Amazonia.  Biotropica  13:199-210. 

Berlioz,  J.  1938.  Pithys  castanea,  sp.  nov.  Bulletin  of 
the  British  Ornithologists’  Club  58:90-91. 
Berlioz,  J.  1948.  Note  critique  sur  le  genre  Pithys 
Vieillot  (Formicariides).  L’Oiseau  et  la  Revue 
Frangaise  d’Ornithologie  18:1-4.  [in  French] 
Bibby,  C.  J.  1992.  Putting  biodiversity  on  the  map: 
priority  areas  for  conservation.  International 
Council  for  Bird  Preservation,  Girton,  Cambridge, 
United  Kingdom. 

Charif,  R.  A.,  S.  Mitchell,  and  C.  W.  Clark.  1995. 
Canary  1.2.  Cornell  Laboratory  of  Ornithology, 
Ithaca,  New  York. 

Collar,  N.  J.,  L.  P.  Gonzaga,  N.  Krabbe,  A.  Mad- 
rono Nieto,  L.  J.  Naranjo,  T.  A.  Parker,  III,  and 
D.  C.  Wege.  1992.  Threatened  birds  of  the  Amer- 
icas: the  ICBP/IUCN  Red  Data  Book,  3rd  ed.,  part 


22 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  1,  March  2006 


2.  Smithsonian  Institution  Press,  Washington, 
D.C. 

David,  N.  and  M.  Gosselin.  2002.  The  grammatical 
gender  of  avian  genera.  Bulletin  of  the  British  Or- 
nithologists’ Club  122:257-282. 

Isler,  M.  L.,  J.  Alvarez  A.,  P.  R.  Isler,  and  B.  M. 
Whitney.  2001.  A new  species  of  Percnostola 
antbird  (Passeriformes:  Thamnophilidae)  from 
Amazonian  Peru,  and  an  analysis  of  species  limits 
within  Percnostola  rufifrons.  Wilson  Bulletin  1 13: 
164-176. 

Isler,  M.  L.,  P.  R.  Isler,  and  B.  M.  Whitney.  1998. 
Use  of  vocalizations  to  establish  species  limits  in 
antbirds  (Passeriformes:  Thamnophilidae).  Auk 
115:577-590. 

Isler,  P.  R.  and  B.  M.  Whitney.  2002.  Songs  of  the 
antbirds:  Thamnophilidae,  Formicariidae,  and 
Conopophagidae  (CD).  Library  of  Natural 
Sounds,  Cornell  University,  Ithaca,  New  York. 

Paynter,  R.  A.,  Jr.  1993.  Ornithological  gazetteer  of 
Ecuador.  Harvard  University  Press,  Cambridge, 
Massachusetts. 

Richards,  P.  W.  1996.  The  tropical  rainforests:  an  eco- 
logical study.  Cambridge  University  Press,  Cam- 
bridge, United  Kingdom. 

Ridgely,  R.  S.  and  P.  J.  Greenfield.  2001a.  The  birds 
of  Ecuador:  field  guide.  Cornell  University  Press, 
Ithaca,  New  York. 

Ridgely,  R.  S.  and  P.  J.  Greenfield.  2001b.  The  birds 
of  Ecuador:  status,  distribution,  and  taxonomy. 
Cornell  University  Press,  Ithaca,  New  York. 

Ridgely,  R.  S.  and  G.  Tudor.  1994.  The  birds  of 
South  America.  University  of  Texas  Press,  Austin. 

Ruokolainen,  K.  and  H.  Tuomisto.  1993.  La  vege- 
tacion  de  terrenos  no  indundables  (tierra  firme)  en 
la  selva  baja  de  la  Amazonia  peruana.  Pages  1 39 — 
1 53  in  Amazonia  Peruana:  vegetacion  humeda 
tropical  en  el  llano  subandino  (R.  Kalliola,  M.  Pu- 
hakka,  and  W.  Danjoy,  Eds.).  Amazon  Project  of 
the  University  of  Turku,  Jyvaskyla,  Finland. 

Ruokolainen,  K.  and  H.  Tuomisto.  1998.  La  vege- 


tacion natural  de  la  zona  de  Iquitos.  Pages  253- 
365  in  Geoecologia  y desarrollo  amazonico.  Es- 
tudio  integrado  de  la  zona  de  Iquitos,  Peru  (R. 
Kalliola  and  S.  Flores  P,  Eds.).  Annales  Univer- 
sitatis  Turkuensis  Series  A II,  vol.  1 14. 

Schulenberg,  T.  S.  and  D.  F.  Stotz.  1991.  The  taxo- 
nomic status  of  Myrmeciza  stictothorax  (Todd). 
Auk  108:731-733. 

Sibley,  C.  G.  and  B.  L.  Monroe,  Jr.  1990.  Distribu- 
tion and  taxonomy  of  birds  of  the  world.  Yale 
University  Press,  New  Haven,  Connecticut. 

Stattersfield,  A.  J.  and  D.  R.  Capper  (Eds.).  2000. 
Threatened  birds  of  the  world:  the  official  source 
for  birds  on  the  IUCN  red  list.  BirdLife  Interna- 
tional, Cambridge,  United  Kingdom,  and  Lynx 
Edicions,  Barcelona,  Spain. 

Stevens,  L.  and  M.  A.  Traylor,  Jr.  1983.  Ornitho- 
logical gazetteer  of  Peru.  Harvard  University 
Press,  Cambridge,  Massachusetts. 

Whitney,  B.  M.  and  J.  Alvarez  A.  1998.  A new 
Herpsilochmus  antwren  (Aves:  Thamnophilidae) 
from  northern  Peru  and  adjacent  Ecuador:  the  role 
of  edaphic  heterogeneity  of  terra  firme  forest.  Auk 
115:559-576. 

Willis,  E.  O.  1967.  The  behavior  of  Bicolored  Ant- 
birds. University  of  California  Publications  in  Zo- 
ology 79:1-132. 

Willis,  E.  O.  1968.  Taxonomy  and  behavior  of  Pale- 
faced  Antbirds.  Auk  85:253-264. 

Willis,  E.  O.  1981.  Diversity  in  adversity:  the  behav- 
iors of  two  subordinate  antbirds.  Arquivos  de 
Zoologia  30:159-234. 

Willis,  E.  O.  1984.  Phlegopsis  erythroptera  (Gould 
1855)  and  relatives  (Aves,  Formicariidae)  as  army 
ant  followers.  Re  vista  Brasiliera  Zoologia  2:165- 
170. 

Zimmer,  K.  J.  and  M.  L.  Isler.  2003.  Family  Tham- 
nophilidae (typical  antbirds).  Pages  448-681  in 
Handbook  of  birds  of  the  world,  vol.  8:  broadbills 
to  tapaculos  (J.  del  Hoyo,  A.  Elliott,  and  D.  Chris- 
tie, Eds.).  Lynx  Edicions,  Barcelona,  Spain. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):23 — 35,  2006 


NESTING  ECOLOGY  OF  LESSER  PRAIRIE-CHICKENS  IN  SAND 
SAGEBRUSH  PRAIRIE  OF  SOUTHWESTERN  KANSAS 

JAMES  C.  PITMAN,'-1 2 3 4-8  CHRISTIAN  A.  HAGEN,1-5  BRENT  E.  JAMISON,1-6 7 
ROBERT  J.  ROBEL,1  THOMAS  M.  LOUGHIN,2  AND  ROGER  D.  APPLEGATE57 


ABSTRACT. — Despite  the  fact  that  the  Lesser  Prairie-Chicken  ( Tympanuchus  pallidicinctus)  is  a species  of 
conservation  concern,  little  is  known  about  its  nesting  ecology,  particularly  in  sand  sagebrush  (. Artemisia filifolia) 
habitats.  To  find  and  monitor  nests,  we  captured  and  equipped  227  female  Lesser  Prairie-Chickens  with  trans- 
mitters (87  yearlings,  1 17  adults,  and  23  of  unknown  age)  from  1997  to  2002  in  southwestern  Kansas.  Apparent 
nest  success  was  similar  for  yearlings  (31%,  n = 74)  and  adults  (27%,  n = 97)  but  differed  marginally  ( P = 
0.090)  between  first  nests  (29%)  and  renests  (14%).  An  estimated  31%  of  females  that  were  unsuccessful  in 
their  first  nesting  attempt  initiated  a second  nest.  The  probability  that  a female  would  initiate  a second  nest  after 
failure  of  the  initial  attempt  was  negatively  influenced  by  the  day  of  incubation  on  which  the  initial  attempt 
failed.  Over  95%  of  all  nests  were  initiated  and  completed  between  5 May  and  2 July.  The  primary  cause  of 
nest  failure  was  predation  by  coyotes  ( Canis  latrans ) and  gopher  snakes  ( Pituophis  melanoleucus ).  Mean  clutch 
size,  egg  fertility,  hatching  success,  nesting  and  renesting  frequency,  and  incidence  of  interspecific  parasitism 
were  all  similar  across  years  and  between  yearlings  and  adults.  Distances  between  nest  sites  were  used  as  an 
index  to  nest-site  fidelity  between  first  nests  and  renests  and  for  across-year  nesting  attempts.  Mean  distances 
between  first  nests  and  renests  were  similar  for  yearlings  (1,071  m)  and  adults  (1,182  m).  Mean  distance  between 
nests  constructed  by  the  same  female  in  subsequent  years  (918  m)  did  not  differ  between  age  classes  or  success 
of  the  first  year’s  nest.  Most  females  (80%)  nested  closer  to  a lek  other  than  the  lek  where  they  were  captured. 
Received  24  January  2005,  accepted  21  September  2005. 


Range-wide,  Lesser  Prairie-Chickens  ( Tym- 
panuchus pallidicinctus ) have  declined  by  an 
estimated  97%  since  the  1800s  (Crawford 
1980,  Taylor  and  Guthery  1980).  In  Kansas, 
Lesser  Prairie-Chickens  are  most  abundant  in 
the  western  part  of  the  state — south  of  the  Ar- 
kansas River  in  mixed  and  shortgrass  prairie 
dominated  by  sand  sagebrush  {Artemisia  fili- 
folia). They  also  occur  in  mixed  grass  prairie 
north  of  the  Arkansas  River,  but  this  habitat 
is  generally  devoid  of  sand  sagebrush.  Lesser 


1 Div.  of  Biology,  Kansas  State  Univ.,  Manhattan, 
KS  66506,  USA. 

2 Dept,  of  Statistics,  Kansas  State  Univ.,  Manhattan, 
KS  66506,  USA. 

3 Survey  and  Research  Office,  Kansas  Dept,  of 
Wildlife  and  Parks,  P.O.  Box  1525,  Emporia,  KS 
66801,  USA. 

4 Current  address:  Survey  and  Research  Office,  Kan- 
sas Dept,  of  Wildlife  and  Parks,  P.O.  Box  1525,  Em- 
poria, KS  66801,  USA. 

5 Current  address:  Oregon  Dept,  of  Fish  and  Wild- 
life, 61374  Parrell  Rd.,  Bend,  OR  97702,  USA. 

6 Current  address:  Missouri  Dept,  of  Conservation, 
P.O.  Box  368,  Clinton,  MO  64735,  USA. 

7 Current  address:  Tennessee  Wildlife  Resources 
Agency,  Ellington  Agricultural  Center,  P.O.  Box 
40747,  Nashville,  TN  37204,  USA. 

8 Corresponding  author;  e-mail: 
jimp@wp.state.ks.us 


Prairie-Chickens  currently  occupy  31  of  39 
counties  believed  to  compose  their  historical 
distribution  in  Kansas,  but  counts  of  leks  and 
individual  birds  suggest  that  Lesser  Prairie- 
Chickens  have  experienced  significant  de- 
clines since  1964  (Jensen  et  al.  2000). 

The  mechanisms  responsible  for  Lesser 
Prairie-Chicken  population  declines  have  not 
been  identified;  however,  aspects  of  nesting 
ecology  could  be  influential  (Peterson  and  Sil- 
vy  1996,  Wisdom  and  Mills  1997).  Thus, 
identifying  age-specific  variation  in  nesting 
variables  is  important  to  understanding  a spe- 
cies’ demography  or  life-history  strategy  (Pat- 
ten et  al.  2005).  Most  research  on  Lesser  Prai- 
rie-Chicken nesting  ecology  has  been  con- 
ducted in  sand  shinnery  oak  {Quercus  havar- 
dii ) habitats  in  New  Mexico  and  Texas  (Davis 
et  al.  1979,  Haukos  and  Broda  1989,  Riley  et 
al.  1992).  The  objectives  of  our  study  were  to 
provide  baseline  information  on  age-specific 
variation  in  nesting  ecology,  record  fidelity  to 
previous  nest  sites  (within-year  renests  and 
across-year  attempts),  and  document  nest-site 
locations  relative  to  leks  of  Lesser  Prairie- 
Chickens  in  sand  sagebrush  prairie  of  south- 
western Kansas.  We  examined  annual  varia- 
tion and  the  effects  of  age  on  reproductive  pa- 
rameters and  nest-site  placement. 


23 


24 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


METHODS 

Study  area. — From  1997  to  2002,  we  stud- 
ied Lesser  Prairie-Chickens  inhabiting  sand 
sagebrush  habitat  south  of  the  Arkansas  River 
in  Finney  County,  Kansas  (37°  52'  N,  100° 
59'  W).  We  initiated  field  work  on  a 7,700-ha 
area  in  1997  and  on  a nearby  5,600-ha  area  in 
2000;  we  continued  work  on  both  areas 
through  summer  2002.  Vegetation  was  similar 
in  both  areas;  sand  sagebrush  was  the  most 
conspicuous  vegetation  present  and  was  inter- 
spersed with  grasses,  including  little  bluestem 
( Schizachyrium  scoparium),  needle-and- 
thread  ( Stipa  comata),  sand  lovegrass  (. Era - 
grostis  trichodes ),  sixweeks  fescue  ( Vulpia 
octoflora ),  blue  grama  ( Bouteloua  gracilis ), 
sand  dropseed  ( Sporobolus  cryptandrus ), 
sideoats  grama  ( B . curtipendula),  and  western 
wheatgrass  ( Agropyron  smithii).  The  most 
common  forb  species  were  Russian  thistle 
( Salsola  kali),  western  ragweed  ( Ambrosia 
psilostachya ),  sand  lily  ( Leucocrinum  montan- 
um),  and  common  sunflower  ( Helianthus  an- 
nuus).  Each  study  area  was  bounded  almost 
entirely  by  center-pivot  irrigated  cropland  and 
grazed  seasonally  by  livestock.  Annual  pre- 
cipitation averaged  50  cm  (U.S.  Department 
of  Commerce  2003)  and  ranged  from  42  cm 
(2000)  to  59  cm  (1997)  during  our  study. 

Locating  and  monitoring  nests. — Using 
walk-in  funnel  traps,  we  captured  female 
Lesser  Prairie-Chickens  on  leks  from  mid- 
March  through  mid-April  (Haukos  et  al. 
1990).  Except  in  1997  (when  age  was  not  de- 
termined), we  classified  captured  birds  as 
yearlings  (—10  months  of  age)  or  adults  (>21 
months  of  age)  by  examining  the  primaries 
(Copelin  1963).  We  equipped  birds  with  11-g 
necklace-style  transmitters  (life  expectancy  = 
6-12  months;  models  from  AVM  Instrument 
Company,  Colfax,  California;  Advanced  Te- 
lemetry Systems,  Isanti,  Minnesota;  and  Ho- 
lohil  Systems,  Carp,  Ontario)  and  released 
them  on-site  immediately  after  capture.  Each 
day,  we  determined  locations  of  transmitter- 
equipped  birds  by  triangulating  bearings  col- 
lected from  a truck-mounted,  null-peak  telem- 
etry system.  Bird  locations  were  determined 
until  transmitter  failure,  emigration  from  the 
primary  study  areas,  or  bird  death.  When  birds 
emigrated  from  our  study  area,  we  re-located 
them  by  extensive  ground  searches  or  from 


fixed-wing  aircraft.  We  monitored  females 
that  moved  off  our  study  area  two  to  three 
times  per  week  throughout  the  nesting  season. 

Using  a hand-held  antenna,  we  found  nests 
by  approaching  transmitter-equipped  females 
when  their  locations  had  remained  unchanged 
>3  consecutive  days.  If  the  female  was  in- 
cubating, she  was  flushed  so  the  eggs  could 
be  counted  and  the  clutch  examined  for  inter- 
specific parasitism  (Hagen  et  al.  2002).  We 
marked  nest  locations  with  flags  (1997)  or 
transmitters  (1998-1999)  at  a distance  of  5 m 
from  the  nest  bowl  (Jamison  2000),  or  we  re- 
corded locations  with  a global  positioning  sys- 
tem (2000-2002).  Nest  sites  were  not  visited 
again  until  the  female  departed  the  site  with  a 
brood  or  until  the  nest  was  depredated  or 
abandoned.  This  technique  allowed  us  to  es- 
timate apparent  nest  success  only.  Because  we 
did  not  determine  nest  status  throughout  in- 
cubation, we  did  not  estimate  daily  survival 
of  eggs  or  nests  according  to  the  Mayfield 
method  (Mayfield  1975). 

After  the  departure  of  each  nesting  female, 
we  classified  nest  fate  as  successful  (produced 
at  least  one  chick),  unsuccessful,  or  aban- 
doned. Beginning  in  2000,  we  opened  un- 
hatched eggs  to  determine  whether  embryos 
had  developed.  If  the  nest  was  depredated,  we 
systematically  searched  the  area  within  a 10- 
m radius  for  tracks,  scat,  or  eggshell  frag- 
ments to  help  determine  the  predator’s  identity 
(Sargeant  et  al.  1998). 

Statistical  analyses. — We  recorded  clutch 
size  and  estimated  the  start  of  incubation  for 
yearling  and  adult  nests.  We  defined  the  start 
of  incubation  as  the  first  day  on  which  we 
detected  no  changes  in  the  female’s  telemetry 
locations — typically,  3-5  days  before  a nest 
was  located.  We  estimated  the  initiation  date 
of  each  nest  by  backdating  from  the  start  of 
incubation  by  1 day  for  each  egg  in  the  clutch 
(Coats  1955).  We  also  calculated  apparent 
nest  success  (the  proportion  of  all  known  nests 
producing  at  least  one  chick  X 100),  hatching 
success,  egg  fertility,  percentage  of  females 
attempting  a nest,  percentage  of  females  re- 
nesting, and  the  incidence  of  interspecific  par- 
asitism— separately  for  yearlings  and  adults. 
We  defined  hatching  success  as  the  number  of 
eggs  hatched  divided  by  initial  clutch  size 
(Westemeier  et  al.  1998b).  We  defined  percent 
fertility  as  the  number  of  eggs  hatching  or 


Pitman  et  al.  • NESTING  ECOLOGY  OF  LESSER  PRAIRIE-CHICKENS 


25 


containing  a developed  embryo  divided  by  the 
total  number  of  eggs  in  the  nest  bowl  at  the 
time  of  hatching.  We  estimated  incubation 
length  as  the  time  (days)  between  the  start  of 
incubation  and  the  date  when  a female  left  the 
nest  with  a brood  (as  determined  from  telem- 
etry locations).  We  estimated  nesting  frequen- 
cy as  the  percentage  of  females  that  attempted 
a nest.  Females  that  did  not  attempt  a nest  and 
died  before  31  May  were  excluded  from  our 
estimate  of  nesting  frequency.  Because  we 
documented  some  first  nesting  attempts  after 
31  May,  it  was  uncertain  whether  birds  dying 
prior  to  this  date  would  have  subsequently  at- 
tempted a nest.  Interspecific  parasitism  was 
reported  as  the  percentage  of  nests  containing 
eggs  of  both  Lesser  Prairie-Chickens  and  oth- 
er bird  species.  Interspecific  nest  parasitism 
was  previously  described  for  the  1 997  to  1 999 
field  seasons  (Hagen  et  al.  2002);  here,  we 
summarize  all  records  of  parasitism  from 
1997  to  2002.  The  percentage  of  females  at- 
tempting to  renest  was  estimated  as  the  per- 
centage of  females  known  to  have  incubated 
and  lost  a first  clutch  and  that  subsequently 
incubated  a second.  Because  of  some  small 
expected  cell  counts,  we  used  a Fisher’s  exact 
test  for  all  comparisons  (Agresti  1996).  In  ad- 
dition, we  used  two-tailed  f-tests  for  unequal 
variances  (Zar  1999)  to  compare  clutch  size, 
incubation  date,  hatch  date,  and  incubation 
length  between  yearlings  and  adults. 

We  used  logistic  regression  to  assess  the  re- 
lationship between  the  likelihood  of  renesting 
and  (1)  age  class,  (2)  clutch  size  of  the  initial 
nest  attempt,  and  (3)  day  into  incubation  when 
the  initial  attempt  failed.  We  excluded  data 
from  1997  because  we  did  not  identify  age 
class  of  birds  that  year.  Initially,  we  fit  seven 
a priori  models  to  data  associated  with  59 
failed  first  nest  attempts  recorded  from  1998 
to  2002.  We  considered  all  four  additive  mod- 
els and  main  effect  models  for  each  of  the 
three  independent  terms.  We  used  the  mini- 
mization of  Akaike’s  Information  Criterion  for 
small  sample  sizes  (AICc.)  to  rank  the  models 
(Burnham  and  Anderson  1998).  All  models 
where  AAICc  < 2 were  considered  to  be  com- 
peting models  (Burnham  and  Anderson  1998). 
Because  age  class  was  not  included  in  any  of 
the  competing  models  (all  AAICc  > 2),  we 
excluded  this  variable  and  developed  models 
using  an  expanded  data  set  ( n = 69)  that  in- 


cluded failed  first  nest  attempts  recorded  from 
1997  to  2002.  We  used  the  same  model  pro- 
cedures previously  described  to  fit  three  of  our 
a priori  models  that  included  the  main  effects 
(1)  clutch  size  and  (2)  day  of  incubation  on 
which  the  initial  attempt  failed. 

We  calculated  distances  between  first  nests 
and  renests,  nesting  attempts  in  multiple 
years,  and  distances  from  nest  sites  to  the  lek 
of  capture  and  the  nearest  lek.  We  used  anal- 
ysis of  variance  (ANOVA)  to  determine 
whether  year  or  age  class  influenced  the  dis- 
tance between  an  initial  nest  site  and  the  re- 
nest location  and  the  affinity  of  nesting  fe- 
males to  lek  sites  (capture  lek  and  nearest  lek). 
We  also  used  ANOVA  to  determine  whether 
age  class  or  success  of  the  first-year  nest  af- 
fected distance  between  nest  sites  in  subse- 
quent years.  For  these  analyses,  we  excluded 
all  data  from  1997  because  we  did  not  identify 
age  class  that  year;  however,  we  included 
pooled  age-class  data  from  1997  in  the  data 
tables  to  provide  an  overview  of  nesting  pa- 
rameters for  the  duration  of  our  study.  We  in- 
terpreted simple  effects  with  two-sample  t- 
tests  when  significant  interactions  were  found 
(Zar  1999).  We  considered  all  differences  sig- 
nificant when  P < 0.05  and  marginally  sig- 
nificant when  0.05  < P < 0.10.  We  report 
parameter  estimates  and  means  as  ± SE  (or 
SD  as  noted). 

RESULTS 

Nesting  ecology. — We  captured  227  female 
Lesser  Prairie-Chickens  and  fitted  them  with 
transmitters  (87  yearlings,  117  adults,  and  23 
of  unknown  age).  We  found  209  nests  (77 
yearling,  103  adult,  and  29  unknown-age). 
The  percentage  of  females  initiating  a nest 
was  similar  ( P = 0.50)  for  yearlings  (94%) 
and  adults  (92%;  Table  1).  We  determined  fate 
for  196  of  209  (94%)  nests;  apparent  nest  suc- 
cess was  26  ± 3%  (51  of  196).  The  remaining 
nests  were  either  abandoned  (2%,  n = 5)  or 
success  could  not  be  determined  from  evi- 
dence remaining  at  the  nest  site  (4%,  n = 8). 
Nest  success  did  not  differ  across  years  (x2  — 
6.95,  df  = 5,  P = 0.22)  or  between  age  classes 
for  first  nests  (P  = 0.60)  or  renests  (P  = 0.82; 
Table  1).  An  estimated  31%  of  all  females  that 
were  unsuccessful  in  their  first  nesting  attempt 
initiated  a second  nest,  and  this  percentage  did 
not  differ  (P  = 0.85)  between  yearlings  and 


TABLE  1 . Lesser  Prairie-Chicken  nesting  statistics  (mean  ± SE),  by  nesting  attempt  and  age,  compiled  over  a 6-year  period  in  the  sand  sagebrush  prairie  of 


26 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 


Vol.  118,  No.  1,  March  2006 


— 

cn 

cn 

cn 

o 

o 

— 

NO 

o 

— 

cn 

o 

CN 

— 

CN 

+1 

+1 

+1 

+ 1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+ 1 

+ 1 

+ 1 

+ 1 

+1 

ON 

o 

in 

00 

■St 

q 

cn 

o 

NO 

CN 

__ 

cn 

CN 

r-i 

00 

cn 

— 

cn 

>n 

CN 

NO 

ON 

ON 

cn 

— 

CN 

in 

in 

CN 

r-  — — LT; 

in  no  Tf 


m o>  >o 

cn  cn  cn 


no  cn  o — no  CN  On 
Tt  h -H  h -H  O 
— LT,  m — — <N 


ON 

t-* 

3r 

CN 

o 

cn 

cn 

>n 

o 

o 

o 

>n 

CN 

O 

ON 

o 

00 

cn 

r- 

— 

>n 

CN 

in 

ON 

in 

00 

r- 

o 

o 

o 

o 

o 

o 

o 

d 

o 

o 

o 

o 

o 

d 

CN 

ON 

q 

NO 

as 

o 

NO 

in 

o 

o 

— 

00 

o 

— 

cn 

in 

o 

cn 

CN 

cn 

NO 

CN 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+ 1 

_ 

cn 

cn 

cn 

cn 

ON 

o 

r-~ 

cn 

CN 

>n 

CN 

cn 

cn 

CN 

U 

— 

CN 

U 

CN 

NO 

r- 

On 

ON 

cn 

— 

CN 

>n 

in 

NO 

CN 

00  — NO  ■'t  vO  IT)  M 

r-  oo  r-  (N  — — — 


o cn  cn 

ON  CN  CN  CN 

cn  cn 


M n 

oo  in  o 


CN 

On 

o 

0 

NO 

CN 

>n 

q 

»n 

o 

O 

CN 

— 

o 

CN 

— 

in 

d 

cn 

cn 

cn 

00 

CN 

+1 

+1 

+ 1 

+ 1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+ 1 

in 

00 

■sj- 

r- 

in 

CN 

as 

n 

cn 

<n 

''t 

cn 

cn 

i> 

NO 

— 

00 

cn 

>n 

cn 

f" 

ON 

ON 

cn 

— 

CN 

m 

in 

r» 

CN 

o — r"  o 
o o in  (N 


\r  " h o o h 
h n r,  x h r,  h 
CN 


<U  ^ 


^ <u  _ 2 


o o 
o ^ 

5 § 

U c/$ 


<u 


•?  I § 

! P 

52  3 ^ 

8 g a 

| 'a  £ -O 
M -f  ° _ 

« = | 3 

z u w ac 


Let  60 
Of)  c 


g 2 s 5 


^ cfl  c/3 


g*  s*  ^ ^ e* 


a Includes  females  of  unknown  age. 

b Females  that  attempted  a nest;  females  that  did  not  attempt  a nest  and  died  before  31  May  were  excluded. 
c n = number  of  failed  first  nests. 

d Nests  were  parasitized  by  either  Ring-necked  Pheasants  or  Northern  Bobwhites. 


Pitman  et  al.  • NESTING  ECOLOGY  OF  LESSER  PRAIRIE-CHICKENS 


27 


adults  (Table  1).  However,  success  of  renests 
(14%)  was  marginally  less  than  success  of  ini- 
tial nests  (29%;  X2  = 3.31,  df  = 1,  P = 0.090). 
No  females  were  known  to  have  initiated  a 
third  nest  in  the  same  year.  Mean  hatch  date 
(all  years  combined)  was  1 June  for  first  nest- 
ing attempts  and  22  June  for  renests  (Fig.  1), 
with  a mean  incubation  length  of  26.7  days 
(Table  1).  More  than  95%  of  all  nests  were 
initiated  and  completed  between  5 May  and  2 
July  (Fig.  1). 

Mean  clutch  size  did  not  differ  between 
yearlings  and  adults  for  either  first  nesting  or 
renesting  attempts  (Table  1).  Mean  clutch  size 
was  7.6  ± 0.4  eggs  for  renests,  significantly 
less  (f188  = 1 1.77,  P < 0.001)  than  the  mean 
clutch  size  (12.0  ± 0.1  eggs)  of  first  nests. 
Overall  hatching  success  was  74  ± 2%  and 
did  not  differ  between  yearlings  and  adults. 
Likewise,  egg  fertility  was  similar  between 
the  two  age  classes,  with  94  ± 1%  of  all  eggs 
containing  a developed  embryo  (Table  1). 

Six  of  209  (3%)  Lesser  Prairie-Chicken 
nests  were  parasitized  by  other  bird  species. 
Four  of  the  six  nests  contained  Lesser  Prairie- 
Chicken  and  Ring-necked  Pheasant  ( Phasi - 
anus  colchicus ) eggs,  and  eggs  of  both  species 
hatched  in  two  of  these  nests.  One  nest  was 
parasitized  by  a Northern  Bobwhite  ( Colinus 
virginianus\  10  prairie-chicken  eggs  and  1 
quail  egg),  and  the  remaining  nest  was  para- 
sitized by  both  Ring-necked  Pheasant  and 
Northern  Bobwhite  (3  prairie-chicken  eggs,  1 
pheasant  egg,  and  1 quail  egg).  Both  of  these 
latter  nests  were  depredated  before  hatching. 

Nest  predators. — Most  (>80%)  known  pre- 
dation events  occurred  >3  days  after  our  ini- 
tial nest  visit  (mean  = 10.2  days  ± 6.9  SD). 
We  assigned  predator  species  to  112  of  161 
(70%)  unsuccessful  Lesser  Prairie-Chicken 
nests.  Coyotes  ( Canis  latrans ) depredated  the 
majority  (64%)  of  the  nests  and  were  the  pri- 
mary cause  of  nest  predation  during  most 
years  (Table  2).  Snakes  were  responsible  for 
the  loss  of  31%  and  42%  of  the  unsuccessful 
Lesser  Prairie-Chicken  nests  in  2001  and 
2002,  respectively.  Most  of  the  snake  preda- 
tion was  probably  by  Gopher  snakes  (. Pituo - 
phis  melanoleucus ) because  they  were  the 
most  observed  snake  species  on  our  study  ar- 
eas. Other  causes  of  nest  loss  included  pre- 
dation by  ground  squirrels  ( Spermophilus 
spp.)  and  trampling  by  cattle  (Table  2). 


Renesting  probability. — The  probability  of 
a Lesser  Prairie-Chicken  renesting  was  influ- 
enced by  both  clutch  size  and  the  day  of  in- 
cubation on  which  the  initial  attempt  failed. 
An  additive  model  including  both  terms  was 
the  highest-ranking  (AAICc  = 0.00;  AICc  = 
80.90),  but  the  model  including  only  date  of 
failure  also  had  considerable  support  (AAIC( 
= 1.48).  The  model  including  only  clutch  size 
was  not  supported  (AAICc  = 15.24).  Females 
incubating  initial  nests  later  into  incubation 
tended  to  have  a lower  probability  of  renesting 
(Gdate  = -0.18,  95%  Cl  = -0.28  to  -0.08; 
Fig.  2).  Females  laying  a larger  clutch  in  the 
initial  nest  attempt  tended  to  be  more  likely 
to  renest  (Bclutch  = 0.31);  however,  the  magni- 
tude of  this  effect  was  not  clear  because  the 
confidence  interval  overlapped  zero  (95%  Cl 
= —0.01  to  0.63).  The  odds  of  a female  at- 
tempting to  renest  decreased  by  16.2%  with 
each  day  into  incubation  of  the  initial  attempt 
and  increased  20.2%  with  each  one-egg  in- 
crease in  clutch  size  (Fig.  2). 

Nest-site  location. — Between  1997  and 
2002,  we  found  28  renests  (Table  3).  Distance 
between  first  nests  and  renests  (1,271  m)  was 
not  influenced  by  age  class  (Flf23  = 1.69,  P = 
0.21)  or  year  (F4>23  = 1.65,  P = 0.21);  there 
was  no  interaction  effect  (F2i23  — 1.82,  P = 
0.19;  1998-2002  data).  Similarly,  the  distance 
between  nests  initiated  by  the  same  female  in 
subsequent  years  (mean  = 918  m,  n = 15; 
Table  3)  was  not  influenced  by  age  class  (FU4 
= 0.16,  P = 0.70)  or  success  of  the  first-year 
nest  (FU4  = 0.05,  P = 0.82);  there  was  no 
interaction  effect  (FU4  = 0.00,  P = 0.98). 

The  distance  from  a nest  to  the  nearest  lek 
(mean  = 691  m,  n = 194;  Table  4)  was  not 
influenced  by  year  (F4164  = 1.11,  P = 0.36) 
or  age  class  (F U64  = 0.00,  P = 0.99),  nor  was 
there  an  interaction  effect  (F4164  = 1.41,  P = 
0.23;  1998-2002  data).  Of  184  nests,  147 
(80%)  were  located  closer  to  a lek  other  than 
the  lek  where  the  female  was  last  captured. 
Ten  nests  (5%)  were  located  >10  km  from  the 
lek  at  which  the  incubating  female  was  cap- 
tured (median  = 20.6  km,  range  = 10.6-56.5 
km).  The  female  nesting  56.5  km  from  her  lek 
of  capture  was  successful  in  her  nesting  at- 
tempt. The  distance  from  nest  site  to  the  lek 
where  the  female  was  captured  (mean  = 3,082 
m,  n = 184;  Table  4)  was  not  influenced  by 
age  class  (FU58  = 0.12,  P = 0.73)  or  year 


Percentage  of  nests  Percentage  of  nests 


28 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


Weekly  interval 

FIG.  1.  Percentage  of  Lesser  Prairie-Chicken  first  nests  (A)  and  renests  (B)  in  southwestern  Kansas  that 
were  initiated,  incubated,  depredated,  and  hatched,  by  weekly  intervals,  1997-2002.  Mean  dates  for  each  variable 
are  listed  at  the  top  of  each  figure. 


Pitman  et  al.  • NESTING  ECOLOGY  OF  LESSER  PRAIRIE-CHICKENS 


29 


TABLE  2.  Probable  causes  of  predation  of  Lesser  Prairie-Chicken  nests  in  the  sand  sagebrush  prairie  of 
southwestern  Kansas,  1997-2002. 


Depredation  (%) 


Predator 

1997 
(n  = 24) 

1998 
(n  = 12) 

1999 
in  = 20) 

2000 
in  = 44) 

2001 

in  = 36) 

2002 
in  = 26) 

Total3 
in  = 161) 

Coyote 

71 

100 

70 

34 

22 

27 

45 

Ground  squirreP 

4 

0 

0 

11 

0 

0 

4 

Snakec 

13 

0 

5 

1 1 

31 

42 

19 

Cattle 

0 

0 

5 

2 

3 

0 

2 

Unknown 

13 

0 

20 

41 

45 

31 

30 

a Percentage  of  all  nests  destroyed  by  each  predator. 

b We  did  not  differentiate  between  thirteen-lined  ground  squirrels  and  spotted  ground  squirrels. 
c Gopher  snakes  appeared  to  be  the  most  abundant  snake  species. 

(F4158  = 1.25  P = 0.29),  and  there  was  no 
interaction  effect  (F4158  = 1.33,  P = 0.26; 

1998-2002  data). 

DISCUSSION 

Although  rainfall  during  the  primary  4- 
month  nesting  period  (April  through  July)  var- 


ied substantially  during  the  6 years  of  our 
study  (range  — 22.3-38.3  cm),  we  document- 
ed little  annual  variation  in  Lesser  Prairie- 
Chicken  nesting  activity.  Our  ability  to  detect 
annual  variation,  however,  may  have  been  hin- 
dered by  relatively  small  sample  sizes  within 
years,  especially  in  the  early  years  of  the 


FIG.  2.  Probability  of  Lesser  Prairie-Chickens  initiating  renests  after  failure  of  the  initial  nest  attempt  in 
southwestern  Kansas,  1997-2002.  Probabilities  are  plotted  for  various  clutch  sizes  (8,  10,  12,  14)  and  the  day 
of  incubation  when  the  initial  nest  attempt  failed. 


30 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


TABLE  3.  Evidence  of  nest-site  fidelity  as  shown  by  mean  distances  (m)  between  nests  for  Lesser  Prairie- 
Chickens  in  southwestern  Kansas,  1997-2002.  Within-  and  across-year  distances  are  presented  by  age  class  and 
nest  fate. 

Within-year*  Across  yearsb’c 


Category  n Distance  SE  n Distance  SE 


Age  class 
Yearling 

11 

1,071 

327 

6 

1,170 

599 

Adult 

13 

1,182 

263 

9 

750 

365 

Nest  fated 

Successful 

— 

— 

— 

6 

712 

438 

Unsuccessful 

— 

9 

1,055 

453 

Totale 

28 

1.271 

218 

15 

918 

316 

a Distance  between  the  first  nest  and  the  renest. 

bFor  two  females  that  initiated  >1  nest  within  a year,  the  mean  coordinates  of  those  nests  were  used  to  calculate  the  distance  to  the  nest  site  in 
subsequent  years. 

c Nests  for  one  female  were  located  in  non-consecutive  years;  all  other  nests  were  located  in  consecutive  years. 
d Nest  fate  refers  to  fate  of  first  nests. 
e Age  of  four  females  was  undetermined. 


study.  Additionally,  we  observed  little  age- 
specific  variation — except  that  yearlings  had 
slightly  smaller  clutches  and  marginally  later 
hatch  dates  for  first  nest  attempts  than  did 
adults. 

For  all  known  nests,  initiation  began  in  ear- 
ly May;  peak  hatching  was  1 June  for  first 
nests  and  22  June  for  renests  (Fig.  1).  Similar 
dates  of  nest  initiation  (mid-April  through  late 
May)  and  hatching  (late  May  through  mid- 
June)  have  been  reported  from  studies 
throughout  the  species’  range  (Giesen  1998, 
Patten  et  al.  2005).  Mean  incubation  length 
was  26.7  days  (this  study).  Because  nest  at- 


tentiveness of  grouse  increases  throughout  the 
laying  period  (Giesen  and  Braun  1979),  we 
may  have  overestimated  incubation  length  by 
misidentifying  the  start  of  incubation.  How- 
ever, the  time  required  to  hatch  Lesser  Prairie- 
Chicken  eggs  in  an  incubator  (24—26  days; 
Coats  1955,  Sutton  1968)  was  only  slightly 
less  than  our  estimate  for  eggs  incubated  by 
wild  birds. 

The  success  of  all  nests  averaged  26%  in 
our  study,  substantially  less  than  estimates 
from  New  Mexico  (42%)  and  Oklahoma 
(40%;  Patten  et  al.  2005),  but  similar  to  the 
28%  reported  by  Giesen  (1998)  for  10  studies 


TABLE  4.  Distances  (m)  between  Lesser  Prairie-Chicken  nest  sites  and  leks  in  southwestern  Kansas,  1997- 
2002. 


Nest  site  to  lek  of  capture  Nest  site  to  nearest  lek 


Category  n Median  Mean  ± SE  n Median  Mean  ± SE 


Year 


1997 

25 

1,528 

1,647 

± 

226 

26 

556 

557  ± 52 

1998 

14 

1,134 

1,727 

± 

529 

14 

577 

546  ± 71 

1999 

24 

2,357 

2,317 

+ 

332 

25 

726 

701  ± 55 

2000 

56 

1,282 

2,874 

-t- 

1,006 

56 

675 

742  ± 53 

2001 

37 

1,396 

3,241 

± 

983 

41 

727 

740  ± 54 

2002 

28 

2,333 

5,901 

± 

1,366 

32 

631 

703  ± 65 

Age 

Yearling 

68 

1,893 

3,580 

± 

853 

68 

633 

702  ± 48 

Adult 

91 

1,258 

3,104 

± 

591 

97 

675 

718  ± 32 

Total 

184a 

1,427 

3,082 

4- 

432 

194b 

632 

691  ± 25 

a Includes  25  nests  of  females  of  unknown  age. 
b Includes  29  nests  of  females  of  unknown  age. 


Pitman  et  al.  • NESTING  ECOLOGY  OF  LESSER  PRAIRIE-CHICKENS 


31 


conducted  throughout  the  range  of  the  Lesser 
Prairie-Chicken.  Giesen  (1998)  suggested  that 
nest  success  from  those  10  studies  was  nega- 
tively biased  due  to  observer  disturbance  at 
nest  sites.  Negative  bias  in  our  study  was  like- 
ly only  slight  because  females  were  flushed 
from  their  nests  only  once.  Westemeier  et  al. 
(1998a)  reported  that  flushing  incubating 
Greater  Prairie-Chickens  (T.  cupido ) once  did 
not  result  in  reduced  nest  success.  Also,  the 
number  of  days  between  our  initial  nest  visits 
and  predation  events  averaged  >10  days.  In 
addition,  only  2%  of  the  nests  in  our  study 
were  abandoned — a much  smaller  percentage 
than  the  25%  reported  by  Riley  et  al.  (1992) 
for  Lesser  Prairie-Chickens  in  New  Mexico. 
Further,  one  of  five  nests  abandoned  during 
our  study  was  abandoned  9 days  after  the  re- 
searcher’s visit,  indicating  that  it  probably  was 
not  due  to  human  disturbance. 

The  percentage  of  females  initiating  a sec- 
ond nest  during  our  study  (31%)  was  between 
previous  estimates  for  Lesser  Prairie-Chickens 
in  New  Mexico  (15%)  and  Oklahoma  (79%; 
Patten  et  al.  2005),  and  it  was  less  than  the 
83%  reported  for  Greater  Prairie-Chickens 
(Svedarsky  1988)  and  the  67%  estimated  for 
Sharp-tailed  Grouse  (T.  phasianellus\  Roers- 
ma  2001).  The  percentage  of  Greater  Sage- 
Grouse  ( Centrocercus  urophasianus ) initiat- 
ing a renest  was  highly  variable  (5  to  87%) 
throughout  their  range  (Schroeder  et  al.  1999), 
and  most  estimates  were  less  than  what  we 
observed  for  Lesser  Prairie-Chickens.  Our 
models  indicated  that  the  low  probability  of 
Lesser  Prairie-Chickens  renesting  in  south- 
western Kansas  was  influenced  by  the  length 
of  incubation  before  their  clutches  were  dep- 
redated (>50%  of  unsuccessful  initial  clutches 
were  incubated  >12  days  prior  to  predation). 
Similarly,  Schroeder  (1997)  reported  that 
Greater  Sage-Grouse  in  Washington  whose 
initial  nests  failed  late  in  incubation  were  less 
likely  to  renest  than  those  whose  nests  failed 
earlier  in  incubation.  Clutch  size  of  the  initial 
nesting  attempt  was  also  somewhat  associated 
with  renesting  probability  in  our  study;  how- 
ever, the  magnitude  of  this  effect  was  unclear. 
The  positive  relationship  that  we  observed 
may  have  been  due  to  increased  fitness  asso- 
ciated with  females  laying  larger  clutches  or 
the  possibility  that  we  misclassified  some  re- 
nests as  initial  nest  attempts.  We  speculate  that 


the  latter  was  not  a common  occurrence  dur- 
ing our  study,  but  our  methods  did  not  allow 
us  to  locate  nests  that  were  depredated  prior 
to  the  onset  of  incubation. 

Few  prairie  grouse  researchers  have  report- 
ed nest  success  separately  for  first  nest  at- 
tempts and  subsequent  renestings.  Bergerud 
and  Gratson  (1988)  hypothesized  that  preda- 
tion of  grouse  nests  was  density-dependent 
and  that  renests  would  be  more  successful 
than  first  nest  attempts  due  to  lower  nest  den- 
sities. They  also  believed  that  nest  success 
should  improve  as  new  vegetative  cover  ap- 
pears throughout  the  nesting  season.  Success 
of  first  and  second  nesting  attempts  of  Lesser 
Prairie-Chickens  in  Kansas,  however,  does  not 
support  Bergerud  and  Gratson’s  (1988)  hy- 
potheses, as  first  nest  attempts  were  margin- 
ally more  successful  than  renestings.  Like- 
wise, Greater  Prairie-Chicken  nests  initiated  in 
Kansas  prior  to  30  April  (presumably  first  at- 
tempts) were  more  successful  than  nests  ini- 
tiated after  1 May  (presumably  renests;  Robel 
1970).  Initial  nesting  attempts  for  Attwater’s 
Greater  Prairie-Chicken  ( T c.  attwateri)  also 
were  more  successful  than  renests  in  4 of  5 
years  (Lutz  et  al.  1994).  Similar  nest  success 
for  first  attempts  and  subsequent  renestings 
has  been  reported  for  Greater  Prairie-Chickens 
in  Colorado  (Schroeder  and  Braun  1992)  and 
Greater  Sage-Grouse  in  Washington  (Schroe- 
der 1997)  and  Alberta,  Canada  (Aldridge  and 
Brigham  2001).  The  only  support  for  Berge- 
rud and  Gratson’s  (1988)  hypothesis  comes 
from  studies  on  Sharp-tailed  Grouse  in  Min- 
nesota and  North  Dakota,  where  success  was 
higher  for  second  attempts  than  first  attempts 
(Christenson  1970,  Schiller  1973).  In  our 
study,  Lesser  Prairie-Chicken  nests  initiated 
after  15  May  were  less  successful  (11.9%,  n 
= 42)  than  earlier  nests  (31.5%,  n = 143), 
regardless  of  nesting  attempt.  We  speculate 
that  nests  initiated  after  1 5 May  were  less  suc- 
cessful due  to  an  increase  in  predator  efficien- 
cy later  in  the  nesting  season,  corresponding 
to  changes  in  the  structure  and  composition  of 
vegetation.  Cattle  grazing  began  on  our  study 
area  around  15  May,  and,  after  that  date,  grass 
cover  and  visual  obstruction  decreased  sub- 
stantially (JCP  unpubl.  data).  Grazing  coupled 
with  normal  drought  conditions  during  the 
summer  months  in  southwestern  Kansas  may 
result  in  declining  habitat  quality,  and,  there- 


32 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


fore,  the  poor  success  of  renesting  Lesser  Prai- 
rie-Chickens. Land  management  practices  that 
maintain  taller  and  denser  vegetation  structure 
later  into  the  nesting  season  may  promote  the 
overall  nesting  success  of  Lesser  Prairie- 
Chickens. 

Clutch  size  in  Kansas  averaged  11.3  eggs 
in  191  completed  clutches — greater  than  that 
reported  in  New  Mexico  (8.7)  and  Oklahoma 
(10.8;  Patten  et  al.  2005)  or  in  60  completed 
clutches  located  in  other  states  occupied  by 
Lesser  Prairie-Chickens  (10.4;  Giesen  1998). 
Our  study  is  the  first  to  document  substantially 
different  mean  clutch  sizes  for  first  nests  (12.0 
eggs)  and  renests  (7.6  eggs).  Merchant  (1982) 
reported  mean  clutch  size  for  initial  and  sec- 
ond nesting  attempts,  but  his  estimates  were 
similar  for  both  (9.8  and  10.7  eggs,  respec- 
tively). In  our  study,  the  percentage  of  eggs 
containing  a developed  embryo  was  94%  and 
hatching  success  was  74%.  Egg  fertility  has 
not  been  reported  previously  for  the  Lesser 
Prairie-Chicken,  but  hatching  success  of  eggs 
was  estimated  at  >90%  across  three  studies 
(see  Giesen  1998).  The  lower  hatching  suc- 
cess observed  in  our  study  reflects  partial  nest 
losses  that  occurred  in  32  of  48  (67%)  suc- 
cessful nests. 

Identifying  nest  predators  from  nest  re- 
mains is  difficult  because  patterns  of  egg 
breakage  overlap  among,  and  even  within, 
predator  species  (Lariviere  1999).  Uncertain- 
ties were  reduced  on  our  study  area,  however, 
because  coyotes  and  gopher  snakes  were  the 
only  common  species  capable  of  preying  on 
Lesser  Prairie-Chicken  nests.  Studies  in  New 
Mexico  and  Texas  revealed  that  Chihuahuan 
Ravens  ( Corvus  cryptoleucus ),  badgers  ( Tax- 
idea  taxus),  striped  skunks  ( Mephitis  mephi- 
tis),  and  ground  squirrels  were  the  primary 
predators  of  Lesser  Prairie-Chicken  nests  (Da- 
vis et  al.  1979,  Haukos  and  Broda  1989,  Riley 
et  al.  1992).  However,  few  corvids,  badgers, 
or  striped  skunks  were  observed  on  our  study 
area,  and,  although  ground  squirrels  were 
abundant  (estimated  from  casual  roadside  ob- 
servations), they  were  identified  as  important 
nest  predators  during  only  1 year  (2000). 

Davis  et  al.  (1979)  documented  snakes 
preying  on  Lesser  Prairie-Chicken  nests  in 
New  Mexico.  We  found  little  evidence  for 
snake  predation  of  nests  during  the  early  years 
of  our  study  (Jamison  2000),  but  snake  abun- 


dance appeared  to  increase  (estimated  from 
casual  roadside  observations),  as  did  nest  pre- 
dation by  snakes,  in  the  later  years  (Pitman 
2003).  Snakes  may  have  been  responsible  for 
most  partial-nest  depredations  because  of  the 
lack  of  eggshell  fragments  at  partly  depredat- 
ed nests.  Also,  three  incubating  Lesser  Prairie- 
Chickens  were  likely  killed  by  snakes  because 
their  intact  carcasses  were  found  with  a thin 
film  of  mucus  covering  the  heads.  In  each 
case,  it  appeared  as  if  a snake  had  tried  to 
swallow  the  bird. 

Interspecific  nest  parasitism  has  been  re- 
ported for  Greater  Prairie-Chickens  and 
Sharp-tailed  Grouse  (Leach  1994,  Westemeier 
et  al.  1998b),  but  had  not  been  reported  for 
Lesser  Prairie-Chickens  before  our  work  in 
Kansas  (Hagen  et  al.  2002).  Only  6 of  209 
(3%)  nests  were  parasitized  by  Ring-necked 
Pheasants  and/or  Northern  Bobwhites,  and  2 
of  the  6 (33%)  nests  produced  Lesser  Prairie- 
Chicken  chicks.  Hatching  success  of  eggs  in 
these  two  nests  was  72%,  similar  to  the  74% 
estimated  for  46  unparasitized  nests  (Hagen  et 
al.  2002).  Our  study  provided  no  evidence  that 
nest  parasitism  negatively  affected  nest  suc- 
cess or  hatchability  of  Lesser  Prairie-Chick- 
ens. 

Bergerud  and  Gratson  (1988)  hypothesized 
that  successful  female  grouse  would  nest  in 
the  same  area  in  the  subsequent  breeding  sea- 
son. In  southwestern  Kansas,  female  Lesser 
Prairie-Chickens  nested  within  712  m of  the 
site  of  their  previous  year’s  nest  site  (if  suc- 
cessful). This  degree  of  philopatry  is  similar 
to  that  reported  for  Greater  Sage-Grouse  in 
Wyoming  (Berry  and  Eng  1985)  and  Idaho 
(Fischer  et  al.  1993).  Greater  Sage-Grouse  in 
Washington  showed  less  philopatry  to  a pre- 
vious year’s  successful  nest  location,  moving 
an  average  of  1 ,600  m in  the  subsequent  nest- 
ing season  (Schroeder  and  Robb  2003). 

The  association  between  lek  location  and 
nest  placement  has  important  management  im- 
plications for  identifying  critical  nesting  hab- 
itat. Bradbury  (1981)  hypothesized  that  fe- 
male home  ranges  included  only  one  lek  and 
that  >50%  of  all  females  should  locate  their 
nests  nearer  to  that  lek  than  other  nearby  leks. 
Studies  of  Greater  Sage-Grouse  and  Sharp- 
tailed Grouse  have  provided  support  for  this 
hypothesis  (Bradbury  et  al.  1989,  Giesen 
1997).  In  Colorado  and  Minnesota,  however. 


Pitman  et  al.  • NESTING  ECOLOGY  OF  LESSER  PRAIRIE-CHICKENS 


33 


only  23  of  89  (26%;  Schroeder  1991)  and  7 
of  18  (39%;  Svedarsky  1988)  Greater  Prairie- 
Chickens  nested  closer  to  their  lek  of  capture 
than  to  other  leks,  respectively.  Similarly,  in 
Idaho  Wakkinen  et  al.  (1992)  found  92%  of 
Greater  Sage-Grouse  nests  within  3 km  of  a 
lek,  but  only  55%  were  within  3 km  of  the 
lek  of  capture.  Our  Lesser  Prairie-Chicken 
nesting  data  also  do  not  support  Bradbury’s 
(1981)  hypothesis:  80%  of  our  females  (147 
of  184)  nested  closer  to  a lek  other  than  that 
on  which  they  were  captured.  More  impor- 
tantly, we  located  >80%  of  all  nests  within  1 
km  of  a known  lek  site;  thus,  we  believe  that 
providing  secure  nesting  habitat  within  1 km 
of  a lek  site  is  an  important  management  strat- 
egy* 

Our  study  provides  the  first  comprehensive 
description  of  Lesser  Prairie-Chicken  nesting 
ecology  in  terms  of  age-specific  reproductive 
effort.  Our  estimates  of  Lesser  Prairie-Chick- 
en nesting  parameters  should  be  viewed  as  ap- 
proximations, however,  because  our  method- 
ology did  not  allow  us  to  locate  nests  that 
were  destroyed  during  the  laying  process. 
Nevertheless,  our  estimates  provide  a much 
better  understanding  of  Lesser  Prairie-Chick- 
en demography  in  sand  sagebrush  habitats. 
The  low  nest  success  we  observed  (26%)  is 
troubling,  especially  if  >50%  nest  success  is 
required  for  population  stability  (Westemeier 
1979).  Sensitivity  analyses  have  revealed  that 
nest  success  is  one  of  the  most  influential  de- 
mographic parameters  affecting  population 
growth  of  prairie  grouse  (Peterson  and  Silvy 
1996,  Wisdom  and  Mills  1997,  Hagen  2003). 
Thus,  habitat  management  designed  to  en- 
hance nest  success  of  Lesser  Prairie-Chickens 
in  southwestern  Kansas  should  be  a priority. 
Similar  information  on  nesting  ecology  from 
Lesser  Prairie-Chicken  populations  in  other 
states  and  habitat  types  is  needed  to  identify 
regional  and  site-specific  conservation  needs 
and  to  aid  in  the  development  of  range-wide 
population  models. 

ACKNOWLEDGMENTS 

We  thank  the  private  landowners  of  southwestern 
Kansas  and  the  Sunflower  Electric  Power  Corporation 
for  property  access.  C.  C.  Griffin,  G.  C.  Salter,  T.  G. 
Shane,  T.  L.  Walker,  Jr.,  and  T.  J.  Whyte  assisted  with 
fieldwork.  This  study  was  supported  by  Kansas  State 
University,  Division  of  Biology;  Kansas  Agricultural 
Experiment  Station  (Contribution  No.  04-41 1-J);  Kan- 


sas Department  of  Wildlife  and  Parks;  Federal  Aid  in 
Wildlife  Restoration  Projects  W-47-R  and  W-53-R; 
and  Westar  Energy,  Inc.  Finally,  we  thank  J.  W.  Con- 
nelly, M.  A.  Schroeder,  and  three  anonymous  review- 
ers for  comments  on  earlier  drafts  of  this  manuscript. 

LITERATURE  CITED 

Agresti,  A.  1996.  An  introduction  to  categorical  data 
analysis.  John  Wiley  and  Sons,  New  York. 
Aldridge,  C.  L.  and  R.  M.  Brigham.  2001.  Nesting 
and  reproductive  activities  of  Greater  Sage- 
Grouse  in  a declining  northern  fringe  population. 
Condor  103:537-543. 

Bergerud,  A.  T.  and  M.  W.  Gratson.  1988.  Adaptive 
strategies  and  population  ecology  of  northern 
grouse,  vol.  II:  theory  and  synthesis.  University  of 
Minnesota  Press,  Minneapolis. 

Berry,  J.  D.  and  R.  L.  Eng.  1985.  Interseasonal  move- 
ments and  fidelity  to  seasonal  use  areas  by  female 
Sage  Grouse.  Journal  of  Wildlife  Management  49: 
237-240. 

Bradbury,  J.  W.  1981.  The  evolution  of  leks.  Pages 
138-169  in  Natural  selection  and  social  behavior 
(R.  D.  Alexander  and  D.  W.  Tinkle,  Eds.).  Chiron 
Press,  New  York. 

Bradbury,  J.  W.,  R.  M.  Gibson,  C.  E.  McCarthy,  and 
S.  L.  Vehrencamp.  1989.  Dispersion  of  displaying 
male  Sage  Grouse:  II.  The  role  of  female  disper- 
sion. Behavioral  Ecology  and  Sociobiology  24: 
15-24. 

Burnham,  K.  P.  and  D.  R.  Anderson.  1998.  Model 
selection  and  inference:  a practical  information- 
theoretic  approach.  Springer,  New  York. 
Christenson,  C.  D.  1970.  Nesting  and  brooding  char- 
acteristics of  Sharp-tailed  Grouse  ( Pedioecetes 
phasianellus  jamesi ) in  southwestern  North  Da- 
kota. M.Sc.  thesis.  University  of  North  Dakota, 
Grand  Forks. 

Coats,  J.  1955.  Raising  Lesser  Prairie  Chickens  in 
captivity.  Kansas  Fish  and  Game  13:16-20. 
Copelin,  F.  F.  1963.  The  Lesser  Prairie  Chicken  in 
Oklahoma.  Technical  Bulletin  no.  6.  Oklahoma 
Wildlife  Conservation  Department,  Oklahoma 
City. 

Crawford,  J.  A.  1980.  Status,  problems,  and  research 
needs  of  the  Lesser  Prairie-Chicken.  Pages  1-7  in 
Proceedings  of  the  Prairie  Grouse  Symposium  (P. 
A.  Vohs  and  F.  L.  Knopf,  Eds.).  Oklahoma  State 
University,  Stillwater. 

Davis,  C.  A.,  T.  Z.  Riley,  R.  A.  Smith,  H.  R.  Suminski, 
and  M.  J.  Wisdom.  1979.  Habitat  evaluation  of 
Lesser  Prairie-Chickens  in  eastern  Chaves  County, 
New  Mexico.  New  Mexico  Agriculture  Experi- 
ment Station,  Las  Cruces. 

Fischer,  R.  A.,  A.  D.  Apa,  W.  L.  Wakkinen,  K.  P. 
Reese,  and  J.  W.  Connelly.  1993.  Nesting  area 
fidelity  of  Sage  Grouse  in  southeastern  Idaho. 
Condor  95:1038-1041. 

Giesen,  K.  M.  1997.  Seasonal  movements,  home  rang- 
es, and  habitat  use  by  Columbian  Sharp-tailed 


34 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


Grouse  in  Colorado.  Special  Report,  no.  72.  Col- 
orado Division  of  Wildlife,  Denver. 

Giesen,  K.  M.  1998.  Lesser  Prairie-Chicken  ( Tympan - 
uchus  pallidicinctus).  The  Birds  of  North  Ameri- 
ca, no.  364. 

Giesen,  K.  M.  and  C.  E.  Braun.  1979.  Nesting  be- 
havior of  female  White-tailed  Ptarmigan  in  Col- 
orado. Condor  81:215-217. 

Hagen,  C.  A.  2003.  A demographic  analysis  of  Lesser 
Prairie-Chicken  populations  in  southwestern  Kan- 
sas: survival,  population  viability,  and  habitat  use. 
Ph.D.  dissertation,  Kansas  State  University,  Man- 
hattan. 

Hagen,  C.  A.,  B.  E.  Jamison,  R.  J.  Robel,  and  R.  D. 
Applegate.  2002.  Ring-necked  Pheasant  parasit- 
ism of  Lesser  Prairie-Chicken  nests  in  Kansas. 
Wilson  Bulletin  114:522-524. 

Haukos,  D.  A.  and  G.  S.  Broda.  1989.  Northern  Har- 
rier ( Circus  cyaneus ) predation  of  Lesser  Prairie- 
Chicken  ( Tympanuchus  pallidicinctus).  Journal  of 
Raptor  Research  23:182-183. 

Haukos,  D.  A.,  L.  M.  Smith,  and  G.  S.  Broda.  1990. 
Spring  trapping  of  Lesser  Prairie-Chickens.  Jour- 
nal of  Field  Ornithology  61:20-25. 

Jamison,  B.  E.  2000.  Lesser  Prairie-Chicken  chick  sur- 
vival, adult  survival,  and  habitat  selection  and 
movements  of  males  in  fragmented  rangelands  of 
southwestern  Kansas.  M.Sc.  thesis,  Kansas  State 
University,  Manhattan. 

Jensen,  W.  E.,  D.  A.  Robinson,  Jr.,  and  R.  D.  Apple- 
gate.  2000.  Distribution  and  population  trend  of 
Lesser  Prairie-Chicken  in  Kansas.  Prairie  Natural- 
ist 32:169-175. 

Lariviere,  S.  1999.  Reasons  why  nest  predators  cannot 
be  inferred  from  nest  remains.  Condor  101:71 8— 
721. 

Leach,  S.  W.  1994.  Mallard  parasitizes  Sharp-tailed 
Grouse  nest.  Blue  Jay  52:144-146. 

Lutz,  R.  S.,  J.  S.  Lawrence,  and  N.  J.  Silvy.  1994. 
Nesting  ecology  of  Attwater’s  Prairie-Chicken. 
Journal  of  Wildlife  Management  58:230-233. 

Mayfield,  H.  F.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

Merchant,  S.  S.  1982.  Habitat  use,  reproductive  suc- 
cess, and  survival  of  female  Lesser  Prairie- 
Chickens  in  two  years  of  contrasting  weather. 
M.Sc.  thesis,  New  Mexico  State  University,  Las 
Cruces. 

Patten,  M.  A.,  D.  H.  Wolfe,  E.  Shochat,  and  S.  K. 
Sherrod.  2005.  Habitat  fragmentation,  rapid  evo- 
lution, and  population  persistence.  Evolutionary 
Ecology  Research  7:235-249. 

Peterson,  M.  J.  and  N.  J.  Silvy.  1996.  Reproductive 
stages  limiting  productivity  of  the  endangered  At- 
twater’s Prairie-Chicken.  Conservation  Biology  4: 
1264-1276. 

Pitman,  J.  C.  2003.  Lesser  Prairie-Chicken  nest  site 
selection  and  nest  success,  juvenile  gender  deter- 
mination and  growth,  and  juvenile  survival  and 
dispersal  in  southwestern  Kansas.  M.Sc.  thesis, 
Kansas  State  University,  Manhattan. 


Riley,  T.  Z.,  C.  A.  Davis,  M.  Ortiz,  and  M.  J.  Wis- 
dom. 1992.  Vegetative  characteristics  of  success- 
ful and  unsuccessful  nests  of  Lesser  Prairie-Chick- 
ens. Journal  of  Wildlife  Management  56:383-387. 

Robel,  R.  J.  1970.  Possible  role  of  behavior  in  regu- 
lating Greater  Prairie-Chicken  populations.  Jour- 
nal of  Wildlife  Management  34:306-312. 

Roersma,  S.  J.  2001.  Nesting  and  brood  rearing  ecol- 
ogy of  Plains  Sharp-tailed  Grouse  ( Tympanuchus 
phasianellus  jamesi ) in  a mixed-grass/fescue 
ecoregion  of  southern  Alberta.  M.Sc.  thesis,  Uni- 
versity of  Manitoba,  Winnipeg. 

Sargeant,  A.  B.,  M.  A.  Sovada,  and  R.  J.  Green- 
wood. 1998.  Interpreting  evidence  of  depredation 
of  duck  nests  in  the  prairie  pothole  region.  U.S. 
Geological  Survey,  Northern  Prairie  Wildlife  Re- 
search Center,  Jamestown,  North  Dakota,  and 
Ducks  Unlimited,  Memphis,  Tennessee. 

Schiller,  R.  J.  1973.  Reproductive  ecology  of  female 
Sharp-tailed  Grouse  ( Pedioecetes  phasianellus ) 
and  its  relationship  to  early  plant  succession  in 
northwestern  Minnesota.  Ph.D.  dissertation.  Uni- 
versity of  Minnesota,  St.  Paul. 

Schroeder,  M.  A.  1991.  Movement  and  lek  visita- 
tion by  female  Greater  Prairie-Chickens  in  re- 
lation to  predictions  of  Bradbury’s  female  pref- 
erence hypothesis  of  lek  evolution.  Auk  108: 
896-903. 

Schroeder,  M.  A.  1997.  Unusually  high  reproduc- 
tive effort  by  Sage  Grouse  in  a fragmented  hab- 
itat in  north-central  Washington.  Condor  99: 
933-941. 

Schroeder,  M.  A.  and  C.  E.  Braun.  1992.  Seasonal 
movement  and  habitat  use  by  Greater  Prairie- 
Chickens  in  northeastern  Colorado.  Special  Re- 
port, no.  68.  Colorado  Division  of  Wildlife,  Den- 
ver. 

Schroeder,  M.  A.  and  L.  A.  Robb.  2003.  Fidelity  of 
Greater  Sage-Grouse  Centrocercus  urophasianus 
to  breeding  areas  in  a fragmented  landscape. 
Wildlife  Biology  9:291-298. 

Schroeder,  M.  A.,  J.  R.  Young,  and  C.  E.  Braun. 
1999.  Sage  Grouse  ( Centrocercus  urophasianus). 
The  Birds  of  North  America,  no.  425. 

Sutton,  G.  M.  1968.  The  natal  plumage  of  the  Lesser 
Prairie  Chicken.  Auk  85:69. 

Svedarsky,  W.  D.  1988.  Reproductive  ecology  of  fe- 
male Greater  Prairie-Chickens  in  Minnesota.  Pag- 
es 193-239  in  Adaptive  strategies  and  population 
ecology  of  northern  grouse,  vol.  2 (A.  T.  Bergerud 
and  M.  W.  Gratson,  Eds.).  University  of  Minne- 
sota Press,  Minneapolis. 

Taylor,  M.  A.  and  F.  S.  Guthery.  1980.  Status,  ecol- 
ogy, and  management  of  the  Lesser  Prairie-Chick- 
en. General  Technical  Report  RM-77,  USDA  For- 
est Service,  Rocky  Mountain  Forest  and  Range 
Experiment  Station,  Fort  Collins,  Colorado. 

U.S.  Department  of  Commerce.  2003.  National  Oce- 
anic and  Atmospheric  Administration.  National 
Climatic  Data  Center,  www.ncdc.noaa.gov/  (ac- 
cessed 7 January  2003). 


Pitman  et  al.  • NESTING  ECOLOGY  OF  LESSER  PRAIRIE-CHICKENS 


35 


Wakkinen,  W.  L.,  K.  P.  Reese,  and  J.  W.  Connelly. 
1992.  Sage  Grouse  nest  locations  in  relation  to 
leks.  Journal  of  Wildlife  Management  56:381- 
383. 

Westemeier,  R.  L.  1979.  Factors  affecting  nest  success 
of  prairie  chickens  in  Illinois.  Proceedings  of  the 
Prairie  Grouse  Technical  Council  13:9-15. 
Westemeier,  R.  L.,  J.  E.  Buhnerkempe,  and  J.  D. 
Brawn.  1998a.  Effects  of  flushing  nesting  Greater 
Prairie-Chickens  in  Illinois.  Wilson  Bulletin  110: 
190-197. 


Westemeier,  R.  L.,  J.  E.  Buhnerkempe,  W.  R.  Ed- 
wards, J.  D.  Brawn,  and  S.  A.  Simpson.  1998b. 
Parasitism  of  Greater  Prairie-Chicken  nests  by 
Ring-necked  Pheasants.  Journal  of  Wildlife  Man- 
agement 62:854-863. 

Wisdom,  M.  J.  and  L.  S.  Mills.  1997.  Sensitivity  anal- 
ysis to  guide  population  recovery:  prairie-chick- 
ens as  an  example.  Journal  of  Wildlife  Manage- 
ment 61 :302-312. 

Zar,  J.  H.  1999.  Biostatistical  analysis,  4th  ed.  Prentice 
Hall,  Englewood  Cliffs,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  118(1  ):36 — 4 1 , 2006 


A COMPARATIVE  BEHAVIORAL  STUDY  OF  THREE  GREATER 
SAGE-GROUSE  POPULATIONS 

SONJA  E.  TAYLOR1 3 AND  JESSICA  R.  YOUNG1 2 3 


ABSTRACT. — We  compared  male  strut  behavior  of  the  genetically  distinct  Lyon,  Nevada/Mono,  California 
Greater  Sage-Grouse  ( Centrocercus  urophasianus ) population  with  that  of  two  proximal  populations:  Nye,  Ne- 
vada, and  Lassen,  California.  We  measured  strut  rates  and  nine  acoustic  components  of  the  strut  display  in  all 
three  populations.  Male  strut  rates  did  not  differ  among  populations.  Acoustic  components  of  the  Lyon/Mono 
and  Lassen  populations  were  similar,  whereas  the  Nye  population  was  distinct.  The  genetically  distinct  Lyon / 
Mono  population  was  more  similar  behaviorally  to  the  Nye  population  than  the  genetically  similar  Nye  and 
Lassen  populations  were  to  each  other.  Overall,  the  Lyon/Mono  population  did  not  exhibit  detectable  differences 
in  male  strut  behavior.  Reproductive  isolation  through  sexual  selection  does  not  appear  to  have  occurred  in  the 
Lyon/Mono  population.  Received  27  September  2004,  accepted  19  October  2005. 


Two  recent  studies  based  on  mitochondrial 
gene  sequence  (Benedict  et  al.  2003,  Oyler- 
McCance  et  al.  2005)  and  nuclear  microsat- 
ellite markers  (Oyler-McCance  et  al.  2005)  re- 
vealed a genetically  distinct  population  of 
Greater  Sage-Grouse  ( Centrocercus  urophas- 
ianus) on  the  Nevada/California  border  (Lyon, 
Nevada/Mono,  California).  Those  studies  in- 
dicated that  the  Lyon/Mono  Greater  Sage- 
Grouse  population  is  more  genetically  distinct 
from  other  Greater  Sage-Grouse  populations 
than  is  the  newly  described  (Young  et  al. 
2000)  Gunnison  Sage-Grouse  (C.  minimus) 
species.  Several  factors,  including  the  appar- 
ent genetic  and  geographic  isolation  of  Lyon/ 
Mono  sage-grouse  from  other  populations,  the 
degradation  and  loss  of  sagebrush  (Artemisia 
spp.)  habitat,  and  an  overall  population  de- 
cline, have  made  this  a population  of  interest 
from  both  evolutionary  and  conservation  per- 
spectives. 

Morphological  (Hupp  and  Braun  1991)  and 
behavioral  studies  (Young  et  al.  1994)  of  Gun- 
nison Sage-Grouse  provided  evidence  that 
sexual  selection  had  driven  speciation  in  the 
isolated  populations  of  sage-grouse  in  south- 
western Colorado  and  southeastern  Utah.  The 
use  of  both  mitochondrial  (Kahn  et  al.  1999) 
and  nuclear  markers  (Oyler-McCance  et  al. 


1 Rocky  Mountain  Center  for  Conservation  Genetics 
and  Systematics,  Dept,  of  Biological  Sciences,  Univ. 
of  Denver,  Denver,  CO  80208,  USA. 

2 Western  State  College  of  Colorado,  Dept,  of  Nat- 
ural and  Environmental  Sciences,  Gunnison,  CO 
81231,  USA. 

3 Corresponding  author;  e-mail: 
sonja_taylor@comcast.net 


1999)  supported  the  morphological  and  be- 
havioral data  and  led  to  species  designation 
for  the  Gunnison  Sage-Grouse  (American  Or- 
nithologists’ Union  2000,  Young  et  al.  2000). 
A similar  approach  would  determine  whether 
the  genetic  distinctiveness  of  the  Lyon/Mono 
population  has  been  manifested  morphologi- 
cally and/or  behaviorally  as  it  has  in  Gunnison 
Sage-Grouse.  If  so,  it  could  potentially  lead  to 
a taxonomic  reclassification. 

Male  mating  success  and  mate-choice  cues 
(Gibson  and  Bradbury  1985),  territoriality 
(Gibson  and  Bradbury  1987),  components  of 
female  choice  (Gibson  et  al.  1991),  and  male 
strutting  behavior  (Young  et  al.  1994)  have 
been  studied  previously  in  the  Mono  sage- 
grouse  population.  However,  with  the  excep- 
tion of  Young  et  al.  (1994),  there  have  been 
no  comparative  studies  among  populations. 
Young  et  al.  (1994)  compared  secondary  sex- 
ual characteristics  from  male  strut  displays 
among  three  populations — one  Gunnison 
Sage-Grouse  population  (Gunnison  Basin, 
Colorado)  and  two  Greater  Sage-Grouse  pop- 
ulations (Mono,  California,  and  Jackson,  Col- 
orado). The  structure  of  the  Gunnison  male 
strut  display  was  strikingly  different  from  that 
of  the  other  two  populations.  However,  the 
comparison  of  the  similarly  structured  strut 
display  between  males  from  Mono  and  Jack- 
son  indicated  statistically  significant  differ- 
ences in  most  of  the  acoustic  measures. 

In  light  of  the  genetic  distinctiveness  of 
Lyon/Mono  sage-grouse  and  the  behavioral 
results  of  Young  et  al.  (1994),  we  undertook 
a further  examination  of  male  strut  display  be- 
havior. We  compared  the  Lyon/Mono  popu- 


36 


Taylor  and  Young  • GREATER  SAGE-GROUSE  BEHAVIOR 


37 


FIG.  1.  Current  Greater  Sage-Grouse  distribution  in  California  and  Nevada,  and  locations  of  three  sample 
populations  (modified  from  Schroeder  et  al.  2004). 


lation  with  two  proximal  populations  of 
Greater  Sage-Grouse  (Fig.  1).  We  tested  the 
hypothesis  that  the  Lyon/Mono  population’s 
behavior  is  measurably  different  from  that  of 
other  Greater  Sage-Grouse  populations  and 
may,  in  fact,  be  considered  a separate  taxon 
given  the  genetic  differences.  Alternatively, 
although  the  Lyon/Mono  population  appears 
genetically  isolated,  behaviorally  it  may  not 
be  significantly  different  from  other  Greater 
Sage-Grouse  populations,  indicating  that  sex- 
ual selection  resulting  in  pre-mating  isolating 
mechanisms  has  not  occurred. 

METHODS 

The  three  populations  we  studied  are  from 
the  southwestern  edge  of  the  Greater  Sage- 
Grouse  range  in  Nevada  and  California  (Fig. 
1).  Behavioral  measurements  of  male  strut 


displays  were  taken  at  five  leks.  Greater  Sage- 
Grouse  in  Lyon  County,  Nevada,  and  Mono 
County,  California,  form  a connected,  inter- 
breeding population  (Lyon/Mono).  Record- 
ings were  completed  between  9 and  17  April 
2001  at  three  leks  from  the  Lyon/Mono  pop- 
ulation: Lyon  County,  Nevada  (Desert  Creek 
2 lek;  38°  42'  N,  1 19°  18'  W;  1,603  m),  south- 
ern Mono  County,  California  (Long  Valley  1 
lek;  37°  42'  N,  118°  48' W;  2,124  m),  and 
northern  Mono  County,  California  (Biedeman 
lek;  38°  12'  N,  119°6'W;  2,447  m).  Of  the 
three  recorded  Lyon/Mono  leks,  the  Desert 
Creek  and  Biedeman  leks  are  farthest  apart 
(123  km).  Lassen  County,  California  (Eastside 
lek;  40°  18'  N,  120°  0'W;  1,490  m),  is  ap- 
proximately 250  km  north  and  Nye  County, 
Nevada  (Roadside  lek;  38°  42'  N,  1 16°  47'  W; 
2,121  m),  is  approximately  215  km  east  of  the 


38 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


N 

X 

>s 

u 

c 

a> 

D 

cr 

a> 


■ 

Air  sac  pops 

jL-Whistle  minimum 

■ / 

. Whistle  peak 

* 

Whistlp  start  ^ * 

k 

• « ■ — * **  ^1 

Time  (sec) 


FIG.  2.  Typical  sonagram  of  a Greater  Sage-Grouse  male  strut  display.  The  two  air  sac  pops,  whistle  start 
frequency,  whistle  peak,  and  whistle  minimum  are  labeled.  See  Table  1 for  all  acoustic  components  (modified 
from  Young  et  al.  1994). 


Lyon/Mono  population;  recordings  at  these 
sites  were  completed  between  3 and  1 1 April 
2002.  The  number  of  males  sampled  from 
each  of  the  five  leks  was  as  follows:  Desert 
Creek  2 (n  = 6),  Long  Valley  1 ( n = 9), 
Biedeman  ( n — 9),  Eastside  ( n = 11),  and 
Roadside  ( n = 14);  therefore,  the  sample  size 
for  the  Lyon/Mono  population  was  n = 24. 

Males  perform  a ritualized  strut  display  in 
which  they  take  a few  steps  forward  and  brush 
their  wings  twice  against  their  esophageal 
pouch  producing  loud  swishing  noises  (Fig. 
2).  Following  these  wing  movements,  males 
compress  air  sacs  and  produce  syringeal 
sounds  to  complete  a single  strut  display 
(Hjorth  1970).  Male  strut  displays  were  re- 
corded and  compared  using  the  methods  of 
Young  et  al.  (1994)  with  the  following  mod- 
ifications. Only  adult  males  were  monitored, 
and  these  were  distinguished  from  juveniles  in 
the  field  by  the  presence  of  a clear  white  upper 
breast  on  adults.  Individual  males  were  iden- 
tified by  their  tail  patterns  (Wiley  1973).  At 
least  15  struts  per  male  were  recorded  using 
a Sony  DCR  TRV720  digital  camcorder  and 
a Sennheiser  MKH70-P48  microphone. 
Sounds  of  individual  struts  were  digitized  at 
22  kHz  using  Canary  1.2.4  sound  analysis 
software  (Cornell  Laboratory  of  Ornithology, 
Ithaca,  New  York). 

We  measured  nine  acoustic  components 
(Table  1,  Fig.  2)  and  calculated  population 
means  derived  from  individual  male  averages 


for  each  component.  An  estimate  of  repeat- 
ability ([r  = s2a/(s2  + s2a)];  Lessells  and  Boag 
1987)  was  used  to  measure  the  proportion  of 
within-individual  variation  within  populations 
for  each  component.  Repeatabilities  range 
from  0 (low)  to  1.0  (high).  High  repeatabilities 
indicate  that  the  measured  trait  varies  little 
within  individuals  relative  to  the  population 
variation,  suggesting  that  the  trait  could  re- 
spond to  sexual  selection. 

To  calculate  strut  display  rate,  we  timed  be- 
tween-strut  intervals  using  Etholog  2.2,  an 
ethological  transcription  tool  (Ottoni  2000). 
The  display  rate  for  each  male  was  based  on 
7—40  consecutive  struts  in  which  no  more  than 
30  sec  had  lapsed  between  struts.  Females 
were  present  on  all  leks  during  strut-rate  mea- 
surements, but  any  male  included  in  the  strut- 
rate  analyses  had  to  have  females  within  20 
m of  them  during  recording.  This  criterion 
lowered  the  sample  sizes  (number  of  males) 
for  population  strut-rate  estimation  (Fig.  3). 
At  the  Lassen  and  Lyon  leks,  measurements 
were  taken  as  one  female  moved  throughout 
the  leks.  The  southern  Mono,  northern  Mono, 
and  Nye  leks  all  had  multiple  females  visiting 
leks  over  the  various  days  that  measurements 
were  taken. 

We  used  analysis  of  variance  (ANOVA)  to 
assess  differences  among  populations  for  each 
acoustic  component  and  strut  rate.  We  then 
used  the  GT2-method  (Hochberg  1974)  to 
make  unplanned  comparisons  among  popula- 


Taylor  and  Young  • GREATER  SAGE-GROUSE  BEHAVIOR 


39 


TABLE  1.  Nine  measured  acoustic  components  of  male  Greater  Sage-Grouse  strut  display  in  three  popu- 
lations from  Nevada  and  California.  Males  were  recorded  while  strutting  during  spring  2001  and  2002. 


Lyon,  Nevada/Mono,  Lassen,  California  Nye,  Nevada 

California  ( n = 24)  (n  = 11)  (n  = 14) 

Acoustic 


Measured  variable 

component 

Mean 

SE 

Mean 

SE 

Mean 

SE 

pa 

First  pop  to  whistle  peak 
(msec) 

Whistle  peak  to  whistle 

1 

73.41 

0.37 

73.85 

0.65 

70.30 

0.52 

<0.001 

minimum  (msec) 

2 

40.21 

0.28 

39.81 

0.32 

41.69 

0.61 

0.012 

Pop  to  pop  (msec) 
Whistle  start  frequency 

3 

199.89 

0.73 

199.64 

0.97 

192.24 

0.88 

<0.001 

(Hz) 

4 

861.17 

7.61 

861.65 

10.97 

930.19 

20.19 

<0.001 

Whistle  peak  (Hz) 

5 

2,619.83 

21.06 

2,657.32 

23.09 

2,873.84 

42.85 

<0.001 

Whistle  minimum  (Hz) 
Whistle  start  to  peak  dif- 

6 

533.58 

5.89 

514.48 

7.56 

637.26 

9.63 

<0.001 

ference  (Hz) 
Whistle  peak  to  mini- 

7 

1,771.61 

20.69 

1,795.22 

23.94 

1,944.72 

35.09 

<0.001 

mum  difference  (Hz) 
Whistle  start  to  mini- 

8 

2,096.48 

21.90 

2,151.64 

17.61 

2,241.51 

39.14 

0.002 

mum  difference  (Hz) 

9 

333.90 

11.33 

353.70 

13.99 

290.38 

16.80 

0.020 

a ANOVA. 


tion  means  with  unequal  sample  sizes  for 
acoustic  components.  This  method  uses  the 
studentized  maximum  modulus  distribution  m 
to  compute  a minimum  significant  difference 
(MSD).  The  significance  level  for  the 
ANOVA  was  set  at  P — 0.05  and  for  the  GT2- 
method  it  was  lowered  from  P = 0.05  to  P = 
0.017  using  a Bonferroni  correction  (a"  = a/ 
k;  Sokal  and  Rohlf  1995)  for  multiple  tests. 
We  used  a"  = 0.01  when  referring  to  the  stu- 
dentized maximum  modulus  m critical  values 
table  (GT2-method). 

RESULTS 

All  nine  acoustic  components  of  the  strut 
display  differed  among  populations  (ANOVA, 

8.5 

^ 8.0 
j/5 
D 

2,  7.5 

Q) 

03 

| 7.0 
w 

6.5 
6.0 

Lyon/Mono  Lassen  Nye 


(ii) 

I 

F 2,31  = 3.97,  P = 0.029 

(16) 

f 

t 

(7) 

* 

all  P < 0.05;  Table  1).  The  acoustic  compo- 
nents of  the  males’  displays  were  similar  be- 
tween Lyon/Mono  and  Lassen,  whereas  those 
of  Nye  males’  displays  were  consistently  dis- 
tinct from  those  of  the  other  two  populations. 
Nye  differed  from  both  Lyon/Mono  and  Las- 
sen for  acoustic  components  1 and  3-7  (GT2- 
test,  all  P < 0.01).  For  component  8,  Nye  dif- 
fered only  from  Lyon/Mono  (GT2-test,  P < 
0.010).  All  other  pairwise  population  compar- 
isons for  minimum  significant  differences 
were  not  significant  (GT2-test,  all  P > 0.01). 

Repeatability  estimates  of  the  acoustic  com- 
ponents ranged  from  0.41  to  0.84  in  Lassen, 
0.57  to  0.96  in  Nye,  and  0.35  to  0.91  in  Lyon/ 
Mono  (Table  2).  The  highest  repeatability  es- 
timate for  all  three  populations  was  for  whistle 
peak  (component  5). 

Strut  rates  (struts/min)  differed  (F2<31  = 
3.97,  P = 0.029)  among  populations  (Fig.  3). 
However,  pairwise  comparisons  between  pop- 
ulations indicated  that  none  were  significant 
(GT2-test,  all  P > 0.01).  Lassen  males  had  the 
highest  strutting  rate  (7.84  struts/min),  where- 
as males  from  Nye  had  the  lowest  strutting 
rate  (6.92  struts/min).  Lyon/Mono  males  had 
an  intermediate  strutting  rate  (7.21  struts/min). 


FIG.  3.  Means  (with  standard  error  bars)  and 
ANOVA  result  for  strut  rates  of  male  Greater  Sage- 
Grouse  from  three  populations:  Lyon,  Nevada/Mono, 
California;  Lassen,  California;  and  Nye,  Nevada.  Sam- 
ple sizes  (number  of  males)  are  in  parentheses. 


DISCUSSION 

We  measured  behavioral  traits  and  second- 
ary sexual  characteristics  that  are  related  to 
sexual  selection  in  sage-grouse,  which  could 


40 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


TABLE  2.  Repeatability  estimates  of  strut  display 
acoustic  components  within  individual  males  from 
three  Greater  Sage-Grouse  populations  in  California 
and  Nevada.  Males  were  recorded  while  strutting  dur- 
ing spring  2001  and  2002. 


Acoustic 

component 

Lyon,  Nevada/ 
Mono,  California 
n = 24 

Lassen, 
California 
n = 11 

Nye,  Nevada 
n = 14 

1 

0.51 

0.78 

0.65 

2 

0.35 

0.44 

0.62 

3 

0.64 

0.74 

0.65 

4 

0.57 

0.67 

0.79 

5 

0.91 

0.84 

0.96 

6 

0.57 

0.68 

0.79 

7 

0.53 

0.80 

0.88 

8 

0.74 

0.49 

0.87 

9 

0.41 

0.41 

0.57 

therefore  lead  to  divergence.  Based  on  behav- 
ioral differences  in  male  strut  displays,  our 
study  did  not  support  the  idea  that  the  genet- 
ically distinct  Lyon/Mono  population  should 
be  considered  for  separate  taxonomic  status. 
The  Lyon/Mono  and  Lassen  populations  were 
similar  to  each  other,  while  the  Nye  popula- 
tion was  the  most  unique  across  nine  acoustic 
components  of  male  mating  displays.  How- 
ever, across  six  components  (1-4,  6,  9),  the 
Nye  versus  Lassen  populations  were  either 
more  different  or  as  different  as  Nye  versus 
Lyon/Mono  populations  (Table  1).  Even 
though  the  Lyon/Mono  population  is  geneti- 
cally distinct,  male  mating  behaviors  are  more 
similar  to  those  of  the  Nye  population  than 
those  of  the  genetically  similar  Nye  and  Las- 
sen populations  are  to  each  other  (Table  1). 

The  repeatability  estimates  generally  varied 
widely  across  populations.  However,  three 
acoustic  components  (3,  5,  and  9)  were  rela- 
tively comparable  among  the  three  popula- 
tions. The  high  repeatability  estimates  for 
components  3 (pop  to  pop)  and  5 (whistle 
peak)  indicate  that  these  traits  vary  little  with- 
in individual  males  relative  to  the  variation 
within  populations  and  could  potentially  re- 
spond to  selection.  Young  et  al.  (1994)  also 
found  high  repeatability  estimates  for  whistle 
peak,  which  has  been  shown  to  be  related  to 
female  mate  choice  (Gibson  and  Bradbury 
1985,  but  see  Gibson  et  al.  1991).  A low  re- 
peatability for  component  9 (whistle  start  to 
minimum  difference)  is  most  likely  the  result 
of  high  levels  of  variability  within  individuals 


rather  than  a lack  of  genetic  variation  or  in- 
accuracies in  measurement  (Boake  1989).  Nye 
had  the  highest  repeatability  estimates  for  sev- 
en of  the  nine  acoustic  components,  suggest- 
ing low  variation  in  the  acoustic  measure- 
ments, despite  samples  being  taken  across 
several  days  with  multiple  females  being  pres- 
ent. 

Although  strut  rates  did  differ  among  pop- 
ulations, pairwise  comparisons  of  strut  rate 
did  not  differ  statistically  between  popula- 
tions. This  result  agrees  with  the  observations 
of  Young  et  al.  (1994),  who  found  that  strut 
rates  did  not  differ  between  two  Greater  Sage- 
Grouse  populations — Mono,  California,  and 
Jackson,  Colorado.  Strut  rates  may  vary  with 
time  of  day,  time  of  season,  and  proximity  of 
females  (R.  M.  Gibson  pers.  comm.);  there- 
fore, variation  in  strut  rate  within  and  between 
males  may  outweigh  differences  in  strut  rates 
among  populations  except  in  strong  cases  of 
population  divergence. 

Our  results  suggest  that  the  Lyon/Mono 
population  does  not  exhibit  any  appreciable 
behavioral  differences  in  male  mating  displays 
from  other  Greater  Sage-Grouse  populations. 
The  Lyon/Mono  population  is  significantly 
different  genetically  from  the  Lassen  popula- 
tion (Benedict  et  al.  2003,  Oyler-McCance  et 
al.  2005),  yet  behaviorally,  the  Lyon/Mono 
and  Lassen  populations  have  similar  acoustic 
strut  components  and  strut  rates.  The  impli- 
cations of  the  slight  behavioral  differences  ob- 
served in  the  Nye  population  on  female  mate 
choice  may  be  determined  upon  further  be- 
havioral observations  that  include  additional 
leks,  years,  and  populations.  It  is  possible  that 
there  are  measurable  differences  in  acoustic 
components  of  the  strut  display  between  most 
populations,  but  these  differences  are  gener- 
ally minimized  by  gene  flow. 

The  Lyon/Mono  population  is  genetically 
more  diverse  and  distinct  than  the  Gunnison 
Sage-Grouse  species  (Kahn  et  al.  1999,  Oyler- 
McCance  et  al.  1999,  Benedict  et  al.  2003, 
Oyler-McCance  et  al.  2005).  Using  mitochon- 
drial DNA  sequence,  Benedict  et  al.  (2003) 
estimated  that  the  Lyon/Mono  population  has 
been  isolated  from  other  Greater  Sage-Grouse 
populations  for  tens  of  thousands  of  years. 
Yet,  neither  local  adaptation  to  ecological  or 
environmental  factors,  nor  genetic  drift,  nor 
sexual  selection  has  led  to  detectable  pheno- 


Taylor  and  Young  • GREATER  SAGE-GROUSE  BEHAVIOR 


41 


typic  (behavioral)  differences  in  this  popula- 
tion. Reproductive  isolation  does  not  appear 
to  have  occurred  through  sexual  selection  in 
the  Lyon/Mono  population  as  it  has  in  the 
Gunnison  Sage-Grouse  species. 

ACKNOWLEDGMENTS 

Funding  and  support  for  this  project  was  provided 
by  the  California  Department  of  Fish  and  Game,  Quail 
Unlimited,  the  National  Fish  and  Wildlife  Foundation, 
Western  State  College  of  Colorado,  and  the  Bureau  of 
Land  Management.  We  are  grateful  to  D.  S.  Blanken- 
ship, F.  A.  Hall,  T.  L.  Russi,  J.  Fatooh,  S.  L.  Nelson, 

R.  M.  Gibson,  W.  F.  Mandeville,  and  T.  Slatauski  for 
logistical  and  field  support.  We  appreciate  M.  K.  Bollig 
and  M.  D.  Kascak  for  their  assistance  with  graphics. 
We  thank  C.  E.  Braun,  J.  W.  Connelly,  R.  M.  Gibson, 

S.  J.  Oyler-McCance,  K.  P.  Reese,  and  J.  St.  John  for 
helpful  comments  on  the  manuscript.  Finally,  we  thank 
S.  L.  Thode  and  R.  D.  Taylor  for  patience  and  support. 

LITERATURE  CITED 

American  Ornithologists’  Union.  2000.  Forty-sec- 
ond supplement  to  the  American  Ornithologists’ 
Union  check-list  of  North  American  birds.  Auk 
117:847-858. 

Benedict,  N.  G.,  S.  J.  Oyler-McCance,  S.  E.  Taylor, 
C.  E.  Braun,  and  T.  W.  Quinn.  2003.  Evaluation 
of  the  eastern  ( Centrocercus  urophasianus  uro- 
phasianus ) and  western  ( Centrocercus  urophasi- 
anus phaios)  subspecies  of  sage-grouse  using  mi- 
tochondrial control-region  sequence  data.  Conser- 
vation Genetics  4:301-310. 

Boake,  C.  R.  B.  1989.  Repeatability:  its  role  in  evo- 
lutionary studies  of  mating  behaviour.  Evolution- 
ary Ecology  3:173-182. 

Gibson,  R.  M.  and  J.  W.  Bradbury.  1985.  Sexual  se- 
lection in  lekking  Sage  Grouse:  phenotypic  cor- 
relates of  male  mating  success.  Behavioral  Ecol- 
ogy and  Sociobiology  18:117-123. 

Gibson,  R.  M.  and  J.  W.  Bradbury.  1987.  Lek  orga- 
nization in  Sage  Grouse:  variations  on  a territorial 
theme.  Auk  104:77-84. 

Gibson,  R.  M.,  J.  W.  Bradbury,  and  S.  L.  Vehren- 
camp.  1991.  Mate  choice  in  lekking  Sage  Grouse 
revisited:  the  roles  of  vocal  display,  female  site 
fidelity,  and  copying.  Behavioral  Ecology  2:165- 
180. 

Hjorth,  I.  1970.  Reproductive  behavior  in  Tetraoni- 


dae,  with  special  reference  to  males.  Viltrevy  7: 
183-596. 

Hochberg,  Y.  1974.  Some  generalizations  of  the  T- 
method  in  simultaneous  inference.  Journal  of  Mul- 
tivariate Analysis  4:224-234. 

Hupp,  J.  W.  and  C.  E.  Braun.  1991.  Geographic  var- 
iation among  Sage  Grouse  in  Colorado.  Wilson 
Bulletin  103:255-261. 

Kahn,  N.  W.,  C.  E.  Braun,  J.  R.  Young,  S.  Wood,  D. 
R.  Mata,  and  T.  W.  Quinn.  1999.  Molecular  anal- 
ysis of  genetic  variation  among  large-  and  small- 
bodied Sage  Grouse  using  mitochondrial  control 
region  sequences.  Auk  116:819-824. 

Lessells,  C.  M.  and  P.  T.  Boag.  1987.  Unrepeatable 
repeatabilities:  a common  mistake.  Auk  104:1 16- 
121. 

Ottoni,  E.  B.  2000.  EthoLog  2.2:  a tool  for  the  tran- 
scription and  timing  of  behavior  observation  ses- 
sions. Behavior  Research  Methods,  Instruments, 
& Computers  32:446-449. 

Oyler-McCance,  S.  J.,  N.  W.  Kahn,  K.  P.  Burnham, 
C.  E.  Braun,  and  T.  W.  Quinn.  1999.  A popula- 
tion genetic  comparison  of  large-  and  small-bod- 
ied Sage  Grouse  in  Colorado  using  microsatellite 
and  mitochondrial  markers.  Molecular  Ecology  8: 
1457-1465. 

Oyler-McCance,  S.  J.,  S.  E.  Taylor,  and  T.  W. 
Quinn.  2005.  A multilocus  population  genetic  sur- 
vey of  Greater  Sage-Grouse  across  their  range. 
Molecular  Ecology  14:1 293- 1310. 

Schroeder,  M.  J.,  C.  A.  Aldridge,  A.  D.  Apa,  J.  R. 
Bohne,  C.  E.  Braun,  S.  D.  Bunnell,  J.  W.  Con- 
nelly, et  al.  2004.  Distribution  of  Sage  Grouse 
in  North  America.  Condor  106:363-376. 

Sokal,  R.  R.  and  F.  J.  Rohlf.  1995.  Biometry:  the 
principles  and  practice  of  statistics  in  biological 
research.  W.  H.  Freeman  and  Company,  New 
York. 

Wiley,  R.  H.  1973.  Territoriality  and  non-random  mat- 
ing in  Sage  Grouse,  Centrocercus  urophasianus. 
Animal  Behavior  Monographs  6:85-169. 

Young,  J.  R.,  C.  E.  Braun,  S.  J.  Oyler-McCance,  J. 
W.  Hupp,  and  T.  W.  Quinn.  2000.  A new  species 
of  Sage-Grouse  (Phasianidae:  Centrocercus ) from 
southwestern  Colorado.  Wilson  Bulletin  112:445- 
453. 

Young,  J.  R.,  J.  W.  Hupp,  J.  W.  Bradbury,  and  C.  E. 
Braun.  1994.  Phenotypic  divergence  of  second- 
ary sexual  traits  among  Sage  Grouse,  Centrocer- 
cus urophasianus , populations.  Animal  Behaviour 
47:1353-1362. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):42— 52,  2006 


FIRST  KNOWN  SPECIMEN  OF  A HYBRID  BUTEO:  SWAINSON’S 
HAWK  ( BUTEO  SWAINSONI)  X ROUGH-LEGGED  HAWK 
(B.  LAGOPUS)  FROM  LOUISIANA 

WILLIAM  S.  CLARK1 2 3  ’ AND  CHRISTOPHER  C.  WITT- 


ABSTRACT. — We  report  a specimen  that  appears  to  be  a hybrid  between  Swainson’s  Hawk  ( Buteo  swainsoni) 
and  Rough-legged  Hawk  ( B . lagopus ),  which,  to  our  knowledge,  is  the  first  hybrid  specimen  for  the  genus. 
There  are  few  reports  of  hybridization  between  Buteo  species,  most  of  which  have  been  observations  of  inter- 
specific nesting  pairs.  The  specimen  described  herein  was  collected  in  Louisiana  and  initially  identified  as  a 
Rough-legged  Hawk  because  of  its  feathered  tarsi  and  the  dark  bellyband  and  carpals.  A DNA  sequence  from 
the  maternally  inherited  mitochondrial  ND6  gene  was  identical  to  a published  sequence  for  Swainson’s  Hawk. 
Nuclear  DNA  sequences  from  two  introns  contained  only  five  variable  sites  among  a panel  of  five  potential 
parental  taxa,  but  the  hybrid  sequence  was  most  consistent  with  parentage  by  Rough-legged  and  Swainson’s 
hawks.  The  feathered  tarsi  of  the  hybrid  strongly  suggested  that  the  father  was  either  a Rough-legged  or  Fer- 
ruginous hawk  ( B . regalis ),  the  only  North  American  raptors  other  than  Golden  Eagle  ( Aquila  chrysaetos ) that 
have  feathered  tarsi.  Plumage  and  size  characters  were  inconsistent  with  those  of  Ferruginous  Hawk,  and,  other 
than  the  darkly  pigmented  leg  feathers,  were  intermediate  between  the  light  morphs  of  Swainson’s  and  Rough- 
legged hawks.  The  breeding  range  of  Swainson’s  Hawk  in  Alaska  and  northern  Canada  is  poorly  known,  but  it 
overlaps  that  of  the  Rough-legged  Hawk  in  at  least  a few  locations,  albeit  at  low  densities,  which  may  be  a 
factor  in  hybridization.  The  occurrence  of  this  hybrid  is  evidence  of  the  potential  for  interbreeding  between 
North  American  members  of  the  genus  Buteo , most  of  which  are  genetically  closely  related.  Such  hybridization 
could  have  implications  for  genetic  diversity,  adaptation,  or  the  evolution  of  reproductive  barriers.  In  any  case, 
such  hybrids  present  field  and  museum  identification  problems.  Received  6 December  2004,  accepted  3 October 
2005. 


Few  documented  cases  of  hybridization  ex- 
ist between  any  2 of  the  27  or  so  species  in 
the  genus  Buteo.  Hybrid  combinations  have 
been  reported  for  Long-legged  Buzzard  ( B . 
rufinus ) and  Upland  Buzzard  (B.  hemilasius ) 
in  Asia  (Pfander  and  Schmigalew  2001), 
Common  Buzzard  (B.  buteo ) and  Long-legged 
Buzzard  in  Europe  (Dudas  et  al.  1999),  and 
Red-shouldered  Hawk  (B.  lineatus)  and  Gray 
Hawk  ( Asturina  nitidus ) in  North  America 
(Lasley  1989).  Additionally,  an  adult  Swain- 
son’s Hawk  (B.  swainsoni ) bred  for  more  than 
8 years  with  a presumably  escaped  South 
American  Red-backed  Hawk  (Red-backed 
Buzzard,  B.  polyosoma ) in  Colorado,  USA, 
and  produced  offspring  in  some  years  (Allen 
1988,  Wheeler  1988);  a Red-tailed  Hawk  ( B . 
jamaicensis ) that  escaped  from  a falconer  bred 
with  a Common  Buzzard  in  Scotland  (Murray 
1970).  However,  to  our  knowledge,  there  are 


1 2301  S.  Whitehouse  Cir.,  Harlingen,  TX  78550, 
USA. 

2 Dept,  of  Biological  Sciences  and  Museum  of  Nat- 
ural Science,  Louisiana  State  Univ.,  119  Foster  Hall, 
Baton  Rouge,  LA  70803,  USA. 

3 Corresponding  author;  e-mail; 
raptours@earthlink.net 


no  museum  specimens  of  the  offspring  of  such 
unions.  Thus,  it  was  with  great  interest  that 
we  found  a specimen  of  an  apparent  hybrid  in 
the  Louisiana  State  University  Museum  of 
Natural  Science  (LSUMNS),  Baton  Rouge.  It 
is  a juvenile  male,  has  feathered  tarsi  and 
mostly  dark  carpal  patches,  was  collected  near 
Baton  Rouge,  Louisiana,  and  was  identified  as 
a Rough-legged  Hawk  (B.  lagopus ).  Its  plum- 
age appears  almost  the  same  as  that  of  a prob- 
able hybrid  between  the  same  two  species, 
first  seen  and  photographed  in  November 
2002  by  Martin  Reid  near  Ft.  Worth,  Texas; 
WSC  observed  and  took  photos  of  that  bird 
in  January  2003. 

Herein  we  present  a description  of  the  pu- 
tative hybrid  Buteo  based  on  its  morphology, 
plumage,  and  mitochondrial  and  nuclear  DNA 
sequences.  A comparison  of  the  hybrid  to  a 
set  of  potential  parental  Buteo  taxa  led  to  the 
conclusion  that  it  descended  from  the  mating 
of  a female  Swainson’s  Hawk  with  a male 
Rough-legged  Hawk.  Although  not  shown  on 
some  published  range  maps,  Swainson’s 
Hawks  breed  sparsely  throughout  at  least  a 
part  of  the  Rough-legged  Hawk’s  breeding 
range  in  far-northern  North  America. 


42 


Clark  and  Witt  • HYBRID  BUTEO  SPECIMEN 


43 


METHODS 

WSC  noted  that  the  specimen,  LSUMZ 
159785,  which  was  stored  with  a handful  of 
juvenile  light-morph  Rough-legged  Hawks, 
differed  from  them  and  was  much  like  a pre- 
sumed hybrid  he  had  seen  and  photographed 
near  Ft.  Worth,  Texas  in  January  2003.  After 
a comparison  of  this  specimen  with  those  of 
juvenile  Rough-legged  and  Swainson’s  hawks, 
he  determined  that  it  might  be  a hybrid.  The 
specimen  had  been  collected  on  4 November 
1994  in  East  Baton  Rouge  Parish,  Highway 
30  at  Burtville,  Louisiana,  by  S.  W.  Cardiff 
and  D.  L.  Dittmann.  A tissue  sample  was  de- 
posited in  the  LSUMNS  Collection  of  Genetic 
Resources  (catalog  No.  B23743).  The  speci- 
men was  sexed  internally  as  a male  (left  testis 
7X11  mm)  and  was  in  juvenal  plumage;  the 
skull  was  75%  ossified. 

We  used  a DNEasy  tissue  kit  (Qiagen,  Va- 
lencia, California)  to  extract  DNA  from  frozen 
muscle  tissue  of  the  putative  hybrid  specimen, 
and  one  specimen  of  each  of  the  following 
taxa:  Rough-legged  Hawk,  Swainson’s  Hawk, 
Red-tailed  Hawk,  Harlan’s  Red-tailed  Hawk 
( B . jamaicensis  harlani),  and  Ferruginous 
Hawk.  We  amplified  the  mitochondrial  ND6 
gene  for  the  hybrid  specimen  in  25  jjlI  PCR 
reactions  using  Amplitaq  Gold  (Applied  Bio- 
systems [ABI],  Foster  City,  California)  with 
the  primers  tPROfwd  and  tGLUrev  (Haring  et 
al.  1999).  For  all  six  specimens,  we  amplified 
two  nuclear  loci,  as  follows:  (1)  intron  5 and 
flanking  exon  regions  of  the  cytosolic  ade- 
nylate kinase  gene  (AK1)  using  the  primers 
AK5b  + and  AK6c-  (Shapiro  and  Dumbacher 
2001),  and  (2)  intron  3 and  flanking  exon  re- 
gions of  the  Z-chromosome-linked  muscle- 
specific  receptor  tyrosine  kinase  gene 
(MUSK)  using  primers  designed  by  F.  K. 
Barker:  MUSK-E3F  (CTTCCATGCACTAC 
AATGGGAAA)  and  MUSK-E4R  (CTCTGA 
ACATTGTGGATCCTCAA).  Standard  PCR 
reactions  were  run  on  an  MJ  Research  PTC- 
200  thermal-cycler  under  the  following  tem- 
perature regime:  initial  denaturation  at  95°  C 
for  8 min;  35  cycles  of  92°  C for  20  sec,  55° 
C for  60  sec,  72°  C for  60  sec;  and  a final 
extension  at  72°  C for  10  min.  For  MUSK,  the 
annealing  temperature  was  adjusted  to  50°  C. 
Negative  control  reactions  were  used  for  all 
extractions  and  PCR  to  insure  against  contam- 


ination. PCR  products  were  purified  using  a 
Qiagen  Gel  Extraction  Kit  (Qiagen,  Valencia, 
California).  Cycle-sequencing  reactions  were 
carried  out  in  both  directions  using  the  prim- 
ers described  above  in  quarter-  or  sixteenth- 
volume  reactions  with  a Big  Dye  Terminator 
Cycle  Sequencing  Kit  (ver.  2 or  3.1,  ABI).  Cy- 
cle-sequencing products  were  purified  using 
Sephadex  columns.  Purified  samples  were 
electrophoresed  on  an  ABI  377  or  3100  au- 
tomated sequencer.  Sequences  were  assem- 
bled and  edited  using  Sequencher  4.2.2 
(Gene  Codes  Corporation,  Ann  Arbor,  Mich- 
igan). The  ND6  sequence  was  compared  with 
published  sequences  for  various  Buteo  species 
(Riesing  et  al.  2003). 

We  compared  morphology  and  plumage  of 
the  hybrid  to  a panel  of  five  potential  parental 
taxa.  We  followed  the  “contradictory  charac- 
ters” approach  of  Rohwer  (1994)  to  eliminate 
potential  pairs  of  parental  taxa  for  which  char- 
acters of  the  presumed  hybrid  fall  outside  of 
the  range  of  variation.  We  assembled  standard 
measurements  of  body  mass,  wing  chord  (un- 
flattened), exposed  culmen,  and  hallux  (Bald- 
win et  al.  1931)  for  juvenile  males  of  potential 
paternal  taxa  from  banding  data  for  Swain- 
son’s, Rough-legged,  and  eastern  Red-tailed 
hawks  ( B . j.  borealis),  and  from  museum 
specimen  data  for  western  Red-tailed  ( B . j. 
calurus ),  Harlan’s  Red-tailed,  and  Ferruginous 
hawks.  We  performed  two  stepwise  discrimi- 
nant function  analyses  with  these  four  mor- 
phological variables  using  SPSS  ver.  11.5 
(SPSS,  Inc.  2002).  In  both  stepwise  analyses, 
we  used  0.05  probability  of  F for  entry  and 
0.10  probability  of  F for  removal  of  each  var- 
iable, set  equal  prior  probabilities  of  group 
membership,  and  used  within-group  covari- 
ance matrices.  The  three  Ferruginous  Hawk 
specimens  were  not  included  in  the  analysis 
due  to  small  sample  size,  and  the  single  Har- 
lan’s Red-tailed  Hawk  individual  was  includ- 
ed in  the  western  Red-tailed  Hawk  group.  The 
first  discriminant  function  analysis  included 
Rough-legged,  Swainson’s,  eastern  Red-tailed, 
and  western  Red-tailed  hawks  as  groups.  All 
four  morphological  variables  were  significant 
and  included  in  the  final  model,  and  three  sig- 
nificant discriminant  functions  were  generat- 
ed. The  putative  hybrid  individual  and  the 
three  Ferruginous  Hawks  were  then  classified 
using  these  discriminant  functions.  In  the  sec- 


44 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


ond  discriminant  function  analysis,  we  only 
included  Rough-legged  and  Swainson’s  hawks 
as  groups.  Only  mass,  wing  chord,  and  cul- 
men  were  significant  and  included  in  the  final 
model,  and  only  one  discriminant  function  ex- 
plained 100%  of  the  variation  between  the  two 
groups.  The  putative  hybrid  was  then  again 
classified  according  to  this  discriminant  func- 
tion. To  account  for  possible  shrinkage  of  mu- 
seum specimens  relative  to  live  birds  (Winker 
1993),  we  repeated  all  analyses  under  the  as- 
sumption of  a 3%  reduction  in  size  due  to  dry- 
ing. The  adjustment  for  shrinkage  had  no  sub- 
stantive effect  on  the  results.  Finally,  with  re- 
spect to  plumage  characters,  we  compared  the 
specimen  with  juvenile  male  Swainson’s  and 
Rough-legged  hawks,  including  pigmentation 
of  the  head,  upperparts,  breast,  belly,  tail,  and 
legs,  and  emargination  of  the  seventh  primary 
(P7). 

RESULTS 

The  mitochondrial  DNA  sequence  of  the 
putative  hybrid,  totaling  558  bp,  was  an  iden- 
tical match  to  a published  sequence  from  a 
Swainson’s  Hawk  collected  in  New  Mexico 
(Table  1;  GenBank  accession  No.  AY2 13028). 
The  sequence  was  0.76%  divergent  from  its 
nearest  relative,  the  Galapagos  Hawk  (B.  ga- 
lapagoensis ),  and  3.23-3.58%  divergent  from 
the  only  sympatric  congeners:  Red-tailed,  Fer- 
ruginous, and  Rough-legged  hawks  (Clark  and 
Wheeler  2001,  Riesing  et  al.  2003;  Table  1). 
Mitochondrial  haplotypes  are  shared  between 
mothers  and  their  offspring  because  the  mi- 
tochondrial genome  is  non-recombining  and 
maternally  inherited  (Lansman  et  al.  1983). 
The  identical  mtDNA  sequences  of  the  spec- 
imen and  a known  Swainson’s  Hawk  strongly 
suggests  that  the  maternal  parent  was  a Swain- 
son’s Hawk. 

The  nuclear  AK1  sequence  of  the  putative 
hybrid,  totaling  542  bp,  was  identical  to  se- 
quences from  the  Swainson’s,  Rough-legged, 
eastern  Red-tailed,  Harlan’s,  and  Ferruginous 
hawks.  The  complete  lack  of  variation  at  this 
locus  prevents  the  elimination  of  any  of  these 
taxa  as  potential  parents.  The  nuclear  MUSK 
sequence,  totaling  599  bp,  contained  five  var- 
iable sites  for  the  six  taxa  included  in  this 
study  (Table  2).  Among  the  five  variable  sites 
was  a substitution  unique  to  the  Ferruginous 
Hawk  sample  (T;  site  no.  480),  and  another 


Louisiana  State  University  Museum  of  Natural  Science,  Baton  Rouge. 


Clark  and  Witt  • HYBRID  BUTEO  SPECIMEN 


45 


TABLE  2.  Variable  sites  on  the  599  bp  MUSK  gene  sequence  for  the  presumed  Buteo  hybrid  and  five  other 
buteos.  The  sites  span  part  of  exon  3,  the  entire  intron  3,  and  part  of  exon  4,  corresponding  to  positions  131 1922- 
1312509  of  the  Gallus  gallus  chromosome  Z genomic  contig  (GenBank  NW  060751).  Both  states  (i.e.,  A/T  and 
A/G)  are  reported  for  heterozygous  sites,  as  inferred  by  unambiguous  double  peaks  on  chromatograms. 


Variable  position 

65 

113 

157 

452 

480 

Hybrid 

A/T 

A/G 

c 

A/G 

c 

Swainson’s  Hawk 

T 

A 

c 

G 

c 

Rough-legged  Hawk 

A/T 

A/G 

c 

G 

c 

Ferruginous  Hawk 

T 

A 

c 

G 

T 

Eastern  Red-tailed  Hawk 

T 

A 

T 

G 

c 

Harlan’s  Red-tailed  Hawk 

T 

A 

T 

G 

c 

that  was  shared  only  by  the  eastern  Red-tailed 
and  Harlan’s  Red-tailed  hawks  (T;  site  no. 
157).  At  two  other  sites  (nos.  65  and  1 13),  the 
Rough-legged  Hawk  and  the  hybrid  were  both 
heterozygous  (A/T  and  A/G),  with  one  exclu- 
sively shared  state  and  one  state  in  common 
with  all  other  taxa  (Table  2).  The  fifth  variable 
site  (no.  452)  was  heterozygous  in  the  hybrid 
specimen  only.  Heterozygotes  were  inferred 
when  chromatograms  showed  strong  signal 
and  unambiguous  double  peaks  of  nearly 
equal  height. 

We  identified  the  paternal  parent  using  phe- 
notypic characters.  Red-tailed  Hawk,  includ- 
ing Harlan’s  Hawk,  can  be  eliminated  as  the 
putative  father  because  it  always  has  unfeath- 
ered tarsi.  It  seems  unlikely  that  two  species 
with  bare  tarsi  would  produce  a hybrid  with 
feathered  tarsi.  Further,  the  Red-tailed  Hawk’s 
culmen  is  considerably  larger  than  that  of  the 
hybrid  (Table  3).  Finally,  juvenile  Red-tailed 


Hawks  share  few  plumage  characters  with  the 
hybrid  (Wheeler  and  Clark  1995,  Clark  and 
Wheeler  2001);  we  would  not  expect,  for  ex- 
ample, a hybrid  Red-tailed  Hawk  X Swain- 
son’s  Hawk  juvenile  to  have  the  heavy,  dark 
bellyband  (Fig.  1)  or  the  dark  carpal  patches 
of  the  hybrid. 

Both  Ferruginous  and  Rough-legged  hawks 
have  feathered  tarsi  and  are  the  most  likely 
paternal  candidates  of  the  hybrid  specimen. 
However,  Ferruginous  Hawks  have  noticeably 
wider  gapes  (Bechard  and  Schmutz  1995)  and 
longer  bills,  wings,  and  halluces  than  the  hy- 
brid (Table  3).  The  measurements  of  the  hy- 
brid are  far  closer  to  those  of  Swainson’s 
Hawk  than  to  Ferruginous  Hawk,  suggesting 
that  the  bird  is  not  intermediate  in  size  as 
would  be  expected  in  an  FI  hybrid  between 
these  two  species.  In  contrast,  the  measure- 
ments for  body  mass  and  wing  chord  are  in- 
termediate between  juvenile  male  Swainson’s 


TABLE  3.  Comparison  of  measurements  (mean  ± SE)  of  the  hybrid  Buteo  specimen  with  juvenile  male 
Rough-legged,  Swainson’s,  Ferruginous,  eastern  Red-tailed,  western  Red-tailed,  and  Harlan’s  Red-tailed  hawks. 
Body  mass  and  wing  chord  of  the  hybrid  are  intermediate  between  Rough-legged  and  Swainson’s  hawks.  Culmen 
and  hallux  are  closest  to  Swainson’s  Hawk. 


n 

Body  mass  (g) 

Wing  chord  (mm) 

Culmen  (mm) 

Hallux  (mm) 

Hybrid 

i 

702.0 

381.0 

19.3 

21.4 

Swainson’s  Hawk3 

20 

638.3  ± 16.8 

378.5  ± 2.4 

21.4  ± 0.3 

21.7  ± 0.4 

Rough-legged  Hawkb 

39 

860.8  ± 12.6 

398.2  ± 1.6 

21.5  ± 0.1 

23.9  ± 0.2 

Ferruginous-Hawkc 

3 

1,091.4  ± 14.3 

413.7  ± 1.8 

25.0  ± 0.3 

25.6  ± 0.3 

Eastern  Red-tailed  Hawkd 

24 

825.4  ± 15.8 

351.8  ± 1.9 

27.2  ± 0.2 

24.1  ± 0.2 

Western  Red-tailed  Hawke 

12 

905.5  ± 30.3 

374.4  ± 2.9 

24.2  ± 0.3 

27.7  ± 0.4 

Harlan’s  Red- tailed  Hawkf 

1 

932.0 

365.0 

23.5 

26.0 

3 Unpublished  banding  data  from  Texas  and  New  Jersey,  sex  determined  by  size. 
b Unpublished  banding  data  from  New  York,  sex  determined  by  size. 

e MVZ  (Museum  of  Vertebrate  Zoology,  University  of  California,  Berkeley)  specimen  data  from  California. 
d Unpublished  banding  data  from  New  Jersey,  sex  determined  by  size. 

e MVZ  specimen  data  from  British  Columbia,  California,  Arizona,  New  Mexico,  and  Nevada. 
f MVZ  specimen  data  from  British  Columbia. 


46 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


FIG.  1 . Specimens  showing  ventral  view  of  the  hybrid  Buteo  (center),  compared  with  juvenile  male  Rough- 
legged Hawk  (left)  and  juvenile  male  Swainson’s  Hawk  (right).  Characters  of  the  hybrid  are  intermediate. 


and  Rough-legged  hawks  (Table  3).  Finally, 
the  plumage  characters  of  both  light-  and 
dark-morph  juvenile  Ferruginous  Hawks  do 
not  match  those  of  the  specimen  (Wheeler  and 
Clark  1995,  Clark  and  Wheeler  2001);  a hy- 
brid Ferruginous  Hawk  X Swainson’s  Hawk 
juvenile,  for  example,  would  not  be  expected 
to  have  the  dark  bellyband  (Fig.  1)  nor  the 
dark  carpal  patches  of  the  hybrid. 

Most  plumage  characters  of  the  hybrid 
specimen  are  similar  to  those  of  juvenile  male 
Swainson’s  or  Rough-legged  hawks,  or  inter- 
mediate between  them  (Figs.  1-2,  Table  4). 


The  notching  of  P7  is  also  intermediate  (Fig. 
3).  This  feather  has  a noticeable  abrupt  wid- 
ening or  “notch”  on  the  trailing  edge  for 
Rough-legged  Hawk  (same  for  Ferruginous 
and  Red-tailed  hawks)  but  not  for  Swainson’s 
Hawk.  The  widening  begins  93  mm  from  the 
tip  on  a juvenile  male  specimen  Rough-legged 
Hawk  (Fig.  3 A),  widening  about  15  mm  at  an 
angle  of  70°  to  the  feather  shaft.  P7  on  a ju- 
venile male  Swainson’s  Hawk  specimen  be- 
gan widening  gradually  47  mm  from  the  tip 
and  lacked  a distinctive  notch  (Fig.  3B).  The 
hybrid’s  P7  began  widening  59  mm  from  the 


Clark  and  Witt  • HYBRID  BUTEO  SPECIMEN 


47 


FIG.  2.  Specimens  showing  dorsal  view  of  the  hybrid  Buteo  (center),  compared  with  juvenile  male  Rough- 
legged Hawk  (left),  and  juvenile  male  Swainson’s  Hawk  (right).  Characters  of  the  hybrid  are  intermediate. 


tip  with  a notch  and  widened  about  9 mm  at 
a 60°  angle  (Fig.  3C). 

In  the  first  discriminant  function  analysis, 
which  included  Rough-legged,  Swainson’s, 
eastern  Red-tailed,  and  western  Red-tailed 
hawks  as  groups,  the  first  two  discriminant 
functions  explained  96.2%  of  the  variation  be- 
tween the  groups  (Fig.  4A).  The  first  function 
correlated  strongly  with  culmen  (r  = 0.651) 
and  wing  chord  (r  = —0.513)  and  explained 
80.1%  of  the  variance.  The  second  function 
correlated  strongly  with  hallux  (r  = 0.814) 
and  body  mass  (r  = 0.646)  and  explained 


16.1%  of  the  variance.  Using  both  functions, 
the  hybrid  was  classified  as  a Rough-legged 
Hawk  with  3 1 .2%  probability,  as  a Swainson’s 
Hawk  with  68.8%  probability,  and  as  an  east- 
ern or  western  Red-tailed  Hawk  with  0% 
probability.  In  the  second  discriminant  func- 
tion analysis,  which  included  only  Rough-leg- 
ged and  Swainson’s  hawks  as  groups,  one  dis- 
criminant function  explained  100%  of  the  var- 
iation between  the  groups  (Fig.  4B).  This 
function  correlated  strongly  with  mass  (r  = 
0.875)  and  wing  chord  (r  = -0.580).  Using 
this  function,  the  hybrid  was  classified  as  a 


48 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


TABLE  4.  Comparison  of  plumage  characters  of  the  hybrid  Buteo  specimen  with  juvenile  male  Rough- 
legged and  Swainson’s  hawks.  Characters  of  the  hybrid  are  intermediate  or  like  one  or  the  other  of  the  parent 
species. 

Character 

Rough-legged  Hawk 

Swainson’s  Hawk 

Hybrid 

Crown 

Pale 

Dark 

Dark,  pale  streaks 

Superciliary 

None 

Rufous 

Buffy 

Malar 

Narrow 

Wide 

Wide 

Back  feathers 

Brown,  pale  sides 

Dark  brown,  pale 
tips 

Dark  brown,  pale  tips  and  sides 

Breast 

Lightly  streaked 

Heavily  streaked 

Heavily  streaked 

Belly 

Solidly  dark 

Buffy 

Dark  with  pale  edges 

Legs 

Feathered,  lightly  marked 

Bare 

Feathered,  darkly  marked 

Uppertail 

White  base,  dusky  tip,  no 
bands 

Gray-brown,  dark 
bands 

Narrow  white  base,  gray-brown, 
dark  bands 

Primary,  outer  web 

Grayish  cast 

Dark 

Grayish  cast 

Primary,  inner 

web 

Pale,  no  barring 

Darker,  barring 

Pale,  barring 

Rough-legged  Hawk  with  45.4%  probability 
and  as  a Swainson’s  Hawk  with  54.5%  prob- 
ability. 

DISCUSSION 

Based  on  mtDNA,  we  conclude  that  the 
mother  of  this  putative  hybrid  is  a Swainson’s 
Hawk.  The  most  likely  paternal  candidates  are 
raptors  with  feathered  tarsi.  Rough-legged  and 


Ferruginous  hawks.  The  latter  was  eliminated 
because  of  its  plumage  characters,  much  larg- 
er size,  and  unique  MUSK  intron  haplotype. 

Independent  lines  of  evidence  converged  on 
the  identification  of  the  specimen  as  a hybrid 
between  Swainson’s  and  Rough-legged  hawk. 
The  combination  of  morphological  and  mo- 
lecular characters,  as  in  the  diagnosis  of  a hy- 
brid manakin  ( Ilicura  X Chiroxiphia)  by  Ma- 


FIG.  3.  Notching  of  primary  7.  (A)  Rough-legged  Hawk,  (B)  hybrid,  and  (C)  Swainson’s  Hawk.  The  pos- 
terior margin  of  each  P7  is  highlighted  in  white.  (Scale  is  not  the  same  on  each  figure.)  The  shape  of  P7  of  the 
hybrid  is  intermediate  and  unlike  those  of  any  Buteo  species. 


Clark  and  Witt  • HYBRID  BUTEO  SPECIMEN 


49 


• Eastern  Red-tailed  Hawk 

a Western  Red-tailed  Hawk  o Rough-legged  Hawk  (RLHA) 

O Harlan's  Hawk  & Swainson's  Hawk  (SWHA) 


Discriminant  function  1 


Discriminant  function  1 

FIG.  4.  Discriminant  function  analyses  comparing 
juvenile  males  of  Buteo  species.  In  panel  (A),  plots  of 
points  along  the  first  two  significant  discriminant  func- 
tions are  from  an  analysis  that  included  Rough-legged, 
Swainson’s,  eastern  Red-tailed,  and  western  Red-tailed 
hawks  as  groups.  These  two  discriminant  functions  ex- 
plained 96.2%  of  the  variation  between  the  groups. 
The  Harlan’s  Hawk  was  included  in  the  western  Red- 
tailed Hawk  group,  but  was  plotted  with  a unique  sym- 
bol. The  hybrid  individual  and  three  Ferruginous 
Hawks  were  then  classified  and  plotted  using  these  dis- 
criminant functions.  In  panel  (B),  points  are  plotted 
according  to  a discriminant  function  from  an  analysis 
that  only  included  Rough-legged  and  Swainson’s 
hawks  as  groups.  One  discriminant  function  explained 
100%  of  the  variation  between  the  two  groups.  The 
hybrid  was  classified  and  plotted  according  to  this  dis- 
criminant function. 


rini  and  Hackett  (2002),  is  a powerful  method 
for  the  identification  of  avian  hybrids.  In  par- 
ticular, the  comparison  of  a single  mtDNA  se- 
quence to  the  growing  database  of  published 
sequences  is  an  outstanding  tool  for  identifi- 
cation of  the  maternal  parent.  In  this  case,  the 
mtDNA  sequence  of  the  hybrid  strongly  sug- 


gests that  its  maternal  parent  was  a Swain- 
son’s Hawk.  The  mother  could  have  been  a 
species  other  than  Swainson’s  Hawk  only  if 
the  mitochondrial  identity  were  a mere  artifact 
of  incomplete  lineage  sorting.  We  consider 
this  possibility  unlikely  because  the  mitochon- 
drial study  of  Riesing  et  al.  (2003)  demon- 
strated that  geographically  heterogeneous 
samples  of  five  Rough-legged,  two  Ferrugi- 
nous, nine  Red-tailed,  and  three  Swainson’s 
hawks  are  each  reciprocally  monophyletic, 
and  the  divergence  levels  between  Swainson’s 
Hawk  and  each  of  its  sympatric  congeners  are 
greater  than  3%. 

The  paucity  of  variation  in  the  two  nuclear 
introns  illustrates  the  difficulty  of  using  nu- 
clear DNA  to  diagnose  hybrids  among  closely 
related  species.  Intraspecific  variation  and  lack 
of  lineage  sorting  pose  significant  challenges 
to  the  conclusive  identification  of  hybrid  in- 
dividuals, and  these  problems  are  compound- 
ed when  potential  parental  taxa  cannot  be 
thoroughly  sampled  at  the  population  level. 
Despite  these  difficulties,  our  sample  of  a sin- 
gle individual  for  each  potential  parental  tax- 
on yielded  some  variation  that  was  consistent 
with  the  identification  of  Rough-legged  Hawk 
as  the  paternal  species.  The  eastern  Red- 
tailed, Harlan’s  Red-tailed,  and  Ferruginous 
hawk  samples  each  contained  single  substi- 
tutions on  the  MUSK  intron  that  were  not 
found  in  the  hybrid.  In  contrast,  only  the 
Swainson’s  and  Rough-legged  hawk  samples 
were  completely  compatible  with  parentage  of 
the  hybrid.  Importantly,  two  heterozygous  po- 
sitions in  the  hybrid  each  contained  a state 
that  was  shared  exclusively  with  the  Rough- 
legged Hawk  sample. 

Plumage  and  morphological  characters  of 
the  hybrid  specimen  were  generally  interme- 
diate between  those  of  juvenile  males  of  the 
parent  species.  This  pattern  is  born  out  by  the 
discriminant  function  analyses  and  is  consis- 
tent with  the  characters  of  hybrids  between 
other  species  of  birds  (e.g..  Graves  1990,  Roh- 
wer  1994,  Marini  and  Hackett  2002).  How- 
ever, the  coloration  of  the  tarsi  feathers  was 
not  intermediate.  Juvenile  male  Rough-legged 
Hawks  have  buffy  tarsal  feathers  with  sparse, 
dark  markings,  whereas  Swainson’s  Hawks 
have  bare  tarsi.  The  hybrid  specimen  has  tar- 
sal feathers  with  heavy,  dark  barring,  clearly 
not  intermediate.  The  expectation  that  hybrid 


50 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


traits  fall  within  the  range  of  traits  expressed 
by  the  parental  taxa  is  based  on  the  assump- 
tion that  most  traits  are  additive  and  polygenic 
(Falconer  1989)  and  is  implicit  in  most  hybrid 
diagnoses.  Nonetheless,  hybrids  can  also  ex- 
press traits  that  are  extreme  relative  to  those 
of  the  parental  taxa  (Rieseberg  et  al.  1999).  It 
is  possible  that  the  darkly  pigmented  tarsal 
feathers  could  be  one  such  transgressive  trait, 
caused  by  complementary  gene  action,  over- 
dominance, or  epistasis.  Swainson’s  and 
Rough-legged  hawk  populations  are  known  to 
possess  genetic  variation  that  results  in  differ- 
ences in  the  quantity  and  distribution  of  mel- 
anin-based plumage  pigments  (Clark  and 
Wheeler  2001).  Rohwer  (1994)  reported  other 
examples  of  characters  that  were  not  inter- 
mediate between  those  of  the  parental  species. 
The  culmen,  and,  to  a lesser  degree,  the  hallux 
of  the  hybrid  were  slightly  smaller  than  our 
Swainson’s  and  Rough-legged  hawk  measure- 
ments for  those  characters,  providing  another 
potential  example  of  a non-intermediate  char- 
acter. However,  specimen  shrinkage  could  at 
least  partly  account  for  this  difference. 

The  Swainson’s  Hawk  breeds  in  an  un- 
known amount  of  the  breeding  range  of  the 
Rough-legged  Hawk  in  far  northwestern 
North  America.  This  is  the  extreme  northern 
periphery  of  their  distribution,  and  they  occur 
at  very  low  densities  in  taiga  habitat  where 
they  are  sympatric  with  the  Rough-legged 
Hawk  (England  et  al.  1997,  Bechard  and 
Swem  2002,  Sinclair  et  al.  2003).  This  could 
increase  the  possibility  that  a female  Swain- 
son’s Hawk  could  fail  to  find  a conspecific 
mate.  Given  the  broad  overlap  in  distribution 
between  Swainson’s,  Red-tailed,  and  Ferrugi- 
nous hawks,  the  lack  of  documented  instances 
of  hybridization  or  interspecific  pairings  be- 
tween any  two  of  these  three  species  suggests 
behavioral  barriers  to  reproduction.  Such  bar- 
riers may  not  exist  between  Swainson’s  and 
Rough-legged  hawks,  which  overlap  only 
marginally  and  may  have  come  into  sympatry 
only  recently.  This  hybrid  pairing  is  consistent 
with  the  model  of  Short  (1969),  who  proposed 
that  hybridization  is  most  likely  to  occur  at 
the  edges  of  a species’  range. 

Swainson’s  Hawks  are  rare  during  Novem- 
ber in  the  area  where  the  hybrid  individual 
was  found;  there  is  only  one  November  record 
for  East  Baton  Rouge  Parish,  despite  intensive 


coverage  by  birdwatchers  and  collectors 
(LSUMNS  data).  Although  Lowery  (1974)  in- 
dicated that  Rough-legged  Hawk  is  a regular 
winter  visitor  to  Louisiana,  and  several  sub- 
sequent sight-based  reports  lacking  photos 
have  been  accepted  by  the  Louisiana  Bird  Re- 
cords Committee,  the  only  physical  evidence 
substantiating  the  occurrence  of  a Rough-leg- 
ged Hawk  in  Louisiana  is  a specimen  collect- 
ed on  12  March  1933  at  Grand  Isle  (LSUMZ 
4803).  The  present  hybrid  occurred  at  a place 
(and  time)  unexpected  for  either  species — 
Rough-legged  Hawks  should  occur  farther 
north  and  Swainson’s  Hawks  farther  south. 
This  intermediate  migratory  behavior,  as  well 
as  a myriad  of  other  ecological  differences  be- 
tween Swainson’s  and  Rough-legged  hawks, 
suggests  potential  sources  of  reduced  fitness 
in  hybrids.  Hybridization  can  provide  a mech- 
anism for  gene  flow  between  species,  partic- 
ularly if  hybrids  are  interfertile  with  parental 
species  and  do  not  suffer  reduced  fitness  (Ar- 
nold 1992).  Alternatively,  hybrid  unfitness  can 
reinforce  behavioral  pre-mating  barriers 
through  natural  selection  (Saetre  et  al.  1997), 
particularly  in  taxa  such  as  Swainson’s  and 
Rough-legged  hawks  that  may  have  recently 
come  into  secondary  contact. 

Hybrids  between  raptor  species  are  reported 
infrequently,  most  likely  because  they  are 
rare,  but  also  because  they  are  difficult  to  di- 
agnose in  the  field  and  are  underrepresented 
in  collections.  That  this  specimen  went  unrec- 
ognized for  9 years  after  being  collected  un- 
derscores the  field  and  museum  identification 
problems  posed  by  hybrids.  Hybrids  have 
been  reported  between  Red  Kite  ( Milvus  mil- 
vus)  and  Black  Kite  (A/,  migrans)  in  Sweden 
(Sylven  1977),  a possible  hybrid  Rueppell’s 
Vulture  ( Gyps  rueppellii ) and  Cape  Vulture 
( G . coprotheres ) in  Botswana  (Borello  2001), 
Brown  Goshawk  {Accipiter  fasciatus ) and 
Grey  Goshawk  (A.  novaehollandiae ) in  Aus- 
tralia (Olsen  1995),  Shikra  (A.  badius ) and  Le- 
vant Sparrowhawk  (A.  brevipes)  in  Israel 
(Yosef  et  al.  2001),  Pallid  Harrier  ( Circus  ma- 
crourus ) and  Montagu’s  Harrier  (C.  pygargus ) 
in  Finland  (Forsman  1995),  Western  Marsh 
Harrier  (C.  aeruginosus ) and  Eastern  Marsh 
Harrier  (C.  spilonotus ) in  Siberia  (Fefelov 
2001),  and  Greater  Spotted  Eagle  ( Aquila 
clanga)  and  Lesser  Spotted  Eagle  (A.  poma- 
rina)  in  Latvia  (Bergmanis  et  al.  1996).  We 


Clark  and  Witt  • HYBRID  BUTEO  SPECIMEN 


51 


were  unable  to  locate  a copy  of  Suchelet 
(1897),  who  apparently  reported  a hybrid  be- 
tween Common  Buzzard  and  Rough-legged 
Hawk.  Most  unusual  were  intergeneric  hy- 
brids reported  between  Black  Kite  and  Com- 
mon Buzzard  near  Rome,  Italy,  that  produced 
rather  strange-looking  offspring  (Corso  and 
Glidi  1998).  Equally  unusual  was  a pairing 
between  Gyrfalcon  ( Falco  rusticolus ) and  Per- 
egrine Falcon  (F.  peregrinus),  in  which  both 
members  of  the  pair  were  females  (Gjershaug 
et  al.  1998).  The  hybrid  Turkey  Vulture  X 
Black  Vulture  reported  by  Mcllhenny  (1937) 
was  later  determined  to  be  a practical  joke 
(Jackson  1988).  Most  instances  of  hybridiza- 
tion listed  above  were  determined  at  the  nests 
by  observing  that  the  adults  were  different 
species,  although  one  was  a hybrid  captured 
for  banding  (Yosef  et  al.  2001)  and  another 
was  identified  using  field  observations  and 
photographs  (Corso  and  Glidi  1998). 

To  our  knowledge,  our  report  is  the  first  of 
a hybrid  specimen  arising  from  two  Buteo 
species,  and,  perhaps,  the  first  hybrid  speci- 
men for  any  raptor.  It  provides  the  first  con- 
clusive documentation  of  hybridization  be- 
tween two  native  North  American  members  of 
the  genus  Buteo.  A pairing  of  a Red-shoul- 
dered Hawk  with  a Gray  Hawk  (Lasley  1989) 
produced  a downy  chick,  but  it  did  not  fledge, 
and  there  were  neither  photographs  nor  spec- 
imens from  this  union. 

ACKNOWLEDGMENTS 

This  study  was  facilitated  by  the  Collections  of 
Birds  and  the  Collection  of  Genetic  Resources  at  the 
LSU  Museum  of  Natural  Science,  and  the  Collection 
of  Birds  at  the  Museum  of  Vertebrate  Zoology,  Uni- 
versity of  California,  Berkeley.  We  thank  S.  W.  Cardiff 
and  D.  L.  Dittmann  for  finding  and  collecting  this  un- 
usual specimen  and  P.  Bloom,  A.  Hinde,  Braddock  Bay 
Raptor  Research,  and  Cape  May  Raptor  Banding  Pro- 
ject for  sharing  measurements  of  juvenile  male  Swain- 
son’s,  Rough-legged,  and  Red-tailed  hawks  with  us.  F. 
K.  Barker  provided  previously  unpublished  primers  for 
the  MUSK  gene.  J.  V.  Remsen,  Jr.,  J.  Schmutz,  T. 
Swem,  S.  M.  Witt,  and  an  anonymous  reviewer  made 
helpful  comments  on  previous  drafts. 

LITERATURE  CITED 

Allen,  S.  1988.  Some  thoughts  on  the  identification 
of  Gunnison’s  Red-backed  Hawk  ( Buteo  polyo- 
soma)  and  why  it’s  not  a natural  vagrant.  Colorado 
Field  Ornithologist  Journal  22:9-14. 

Arnold,  M.  L.  1992.  Natural  hybridization  as  an  evo- 


lutionary process.  Annual  Review  of  Ecology  and 
Systematics  23:237-261. 

Baldwin,  S.  P,  H.  C.  Oberholser,  and  L.  G.  Worley. 
1931.  Measurements  of  birds.  Scientific  Publica- 
tions of  the  Cleveland  Museum  of  Natural  Histo- 
ry, no.  2.  Cleveland,  Ohio. 

Bechard,  M.  J.  and  J.  K.  Schmutz.  1995.  Ferruginous 
Hawk  ( Buteo  regalis).  The  Birds  of  North  Amer- 
ica, no.  172. 

Bechard,  M.  J.  and  T.  R.  Swem.  2002.  Rough-legged 
Hawk  ( Buteo  lagopus).  The  Birds  of  North  Amer- 
ica, no.  641. 

Bergmanis,  U.,  A.  Pertins,  M.  Strads,  and  I.  Krams. 

1996.  Possible  case  of  hybridization  of  the  Lesser 
Spotted  Eagle  and  the  Greater  Spotted  Eagle  in 
eastern  Latvia.  Putni  daba  6.3: 1-6.  [In  Latvian, 
with  an  English  summary.] 

Borello,  W.  2001.  Possible  hybrid  vulture  at  Man- 
nyelanong  Cape  Vulture  Gyps  coprotheres  colony, 
southeastern  Botswana.  Babbler  38:19-21. 

Clark,  W.  S.  and  B.  K.  Wheeler.  2001.  Hawks  of 
North  America,  2nd  ed.  Peterson  Field  Guide  Se- 
ries, no.  35.  Houghton  Mifflin,  Boston,  Massachu- 
setts. 

Corso,  A.  and  R.  Glidi.  1998.  Hybrids  between  Black 
Kite  and  Common  Buzzard  in  Italy  in  1996.  Dutch 
Birding  20:226-233. 

Dudas,  M.,  J.  Tar,  and  I.  Toth.  1999.  Natural  hy- 
bridization of  Long-legged  Buzzard  ( Buteo  rufi- 
nus)  and  Common  Buzzard  ( B . buteo)  in  the  Hor- 
tobagy  National  Park.  Temeszet  5-6:8-10.  [In 
Hungarian] 

England,  A.  S.,  M.  J.  Bechard,  and  C.  S.  Houston. 

1997.  Swainson’s  Hawk  ( Buteo  swainsoni).  The 
Birds  of  North  America,  no.  265. 

Falconer,  D.  S.  1989.  Introduction  to  quantitative  ge- 
netics, 3rd  ed.  Longman  Scientific  Technical,  Es- 
sex, United  Kingdom. 

Fefelov,  I.  2001.  Comparative  breeding  ecology  and 
hybridization  of  Eastern  and  Western  Marsh  Har- 
riers Circus  spilonotus  and  C.  aeruginosus  in  the 
Baikal  region  of  eastern  Siberia.  Ibis  143:587- 
592. 

Forsman,  D.  1995.  Male  Pallid  and  female  Montagu’s 
Harrier  raising  hybrid  young  in  Finland  in  1993. 
Dutch  Birding  17:102-106. 

Gjershaug,  J.  O.,  A.  O.  Folkestad,  and  L.  O. 
Gok0yr.  1998.  Female-female  pairing  between  a 
Peregrine  Falcon  Falco  peregrinus  and  a Gyrfal- 
con F.  rusticolus  in  two  successive  years.  Fauna 
Norvegica  Series  C,  Cinclus  21:87-91. 

Graves,  G.  R.  1990.  Systematics  of  the  “Green-throat- 
ed Sunangels”  (Aves:  Trochilidae):  valid  taxa  or 
hybrids?  Proceedings  of  the  Biological  Society  of 
Washington  103:6-25. 

Haring,  E.,  M.  J.  Riesing,  W.  Pinsker,  and  A.  Ga- 
mauf.  1999.  Evolution  of  a pseudo-control  region 
in  the  mitochondrial  genome  of  Palearctic  buz- 
zards (genus  Buteo).  Journal  of  Zoological  Sys- 
tematic and  Evolutionary  Research  37:185-194. 

Jackson,  J.  A.  1988.  Turkey  Vulture.  Page  27  in  Hand- 


52 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


book  of  North  American  birds,  vol.  4 (R.  S.  Palm- 
er, Ed.).  Yale  University  Press,  New  Haven,  Con- 
necticut. 

Lansman,  R.  A.,  J.  C.  Avise,  and  M.  D.  Huetel.  1983. 
Critical  experimental  test  of  the  possibility  of  “pa- 
ternal leakage”  of  mitochondrial  DNA.  Proceed- 
ings of  the  National  Academy  of  Sciences  USA 
80:1969-1971. 

Lasley,  G.  1989.  Texas.  American  Birds  43:505. 

Lowery,  G.  H.,  Jr.  1974.  Louisiana  birds.  Louisiana 
State  University  Press,  Baton  Rouge. 

Marini,  M.  A.  and  S.  J.  Hackett.  2002.  A multifac- 
eted approach  to  the  characterization  of  an  inter- 
generic hybrid  manakin  (Pipridae)  from  Brazil. 
Auk  119:1114-1120. 

McIlhenny,  E.  A.  1937.  Hybrid  between  Turkey  Vul- 
ture and  Black  Vulture.  Auk  54:384. 

Murray,  J.  B.  1970.  Escaped  American  Red-tailed 
Hawk  nesting  with  a Buzzard  in  Midlothian.  Scot- 
tish Birds  6:34-37. 

Olsen,  P.  1995.  Australian  birds  of  prey.  The  Johns 
Hopkins  University  Press,  Baltimore,  Maryland. 

Pfander,  P.  and  S.  Schmigalew.  2001.  Extensive  hy- 
bridization of  Long-legged  Buzzard  Buteo  rufinus 
and  Upland  Buzzard  B.  hemilasius.  Omithologis- 
che  Mitteilung  53:344-349.  [In  German] 

Rieseberg,  L.  H..  M.  A.  Archer,  and  R.  K.  Wayne. 
1999.  Transgressive  segregation,  adaptation  and 
speciation.  Heredity  83:363-372. 

Riesing,  M.  J.,  L.  Kruckenhauser,  A.  Gamauf,  and 
E.  Haring.  2003.  Molecular  phylogeny  of  the  ge- 
nus Buteo  (Aves:  Accipitridae)  based  on  mito- 
chondrial marker  sequences.  Molecular  Phyloge- 
netics and  Evolution  27:328-342. 

Rohwer,  S.  1994.  Two  new  hybrid  Dendroica  war- 


blers and  new  methodology  for  inferring  parental 
species.  Auk  111:441-449. 

Saetre,  G.  P,  T.  Moum,  S.  Bures,  M.  Kral,  M.  Ada- 
mjan,  and  J.  Moreno.  1997.  A sexually  selected 
character  displacement  in  flycatchers  reinforces 
premating  isolation.  Nature  387:589-592. 

Shapiro,  L.  H.  and  J.  P.  Dumbacher.  2001.  Adenylate 
kinase  intron  5:  a new  nuclear  locus  for  avian  sys- 
tematics.  Auk  118:248-255. 

Short,  L.  L.  1969.  Taxonomic  aspects  of  avian  hy- 
bridization. Auk  86:84-105. 

Sinclair,  P.  H.,  W.  A.  Nexon,  C.  D.  Eckert,  and  N. 
L.  Hughes  (Eds.).  2003.  Birds  of  the  Yukon  Ter- 
ritory. University  of  British  Columbia  Press,  Van- 
couver, British  Columbia,  Canada. 

SPSS,  Inc.  2002.  SPSS  for  Window,  ver.  1 1.5.0.  SPSS 
Inc.,  Chicago,  Illinois. 

Suchelet,  A.  1897.  Des  hybrids  a L’etat  sauvage,  part 
4:  birds  of  prey.  Librairie.  J.  B.  Bailliere  & fils, 
Paris,  France.  [In  French] 

Sylven,  M.  1977.  Hybridization  between  Red  Kite 
Milvus  milvus  and  Black  Kite  M.  migrans  in  Swe- 
den in  1976.  Var  Fagelvarld  36:38-44.  [In  Swed- 
ish, with  an  English  summary.] 

Wheeler,  B.  K.  1988.  A Red-backed  Hawk  in  Colo- 
rado. Colorado  Field  Ornithologist  Journal  22:5-8. 

Wheeler,  B.  K.  and  W.  S.  Clark.  1995.  A photo- 
graphic guide  to  North  American  raptors.  Aca- 
demic Press,  London,  United  Kingdom. 

Winker,  K.  1993.  Specimen  shrinkage  in  Tennessee 
Warblers  and  “Traill’s”  Flycatchers.  Journal  of 
Field  Ornithology  64:331-336. 

Yosef,  R.,  A.  J.  Helbig,  and  W.  S.  Clark.  2001.  An 
intrageneric  Accipiter  hybrid  from  Eilat,  Israel. 
Sandgrouse  23:141-143. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):53 — 58,  2006 


NOCTURNAL  HUNTING  BY  PEREGRINE  FALCONS  AT  THE 
EMPIRE  STATE  BUILDING,  NEW  YORK  CITY 

ROBERT  DeCANDIDO1  34  AND  DEBORAH  ALLEN1 2 3 4 


ABSTRACT. — We  report  on  nocturnal  hunting  by  Peregrine  Falcons  ( Falco  peregrinus)  at  the  Empire  State 
Building  in  Manhattan,  New  York  City.  From  4 August  through  13  November  2004,  we  saw  Peregrine  Falcons 
on  41  of  77  nights  of  observation.  During  this  period,  they  hunted  migrating  birds  on  25  evenings,  with  the  first 
hunting  attempt  occurring  an  average  of  119  min  after  sunset.  Peregrine  Falcons  made  111  hunting  attempts 
and  captured  37  birds  (33%  success).  Hunting  success  was  highest  in  September,  but  was  most  often  observed 
in  October.  Peregrines  hunted  migratory  birds  at  night  more  frequently  in  autumn  than  in  spring.  Peregrines 
were  significantly  more  likely  to  be  present  on  autumn  nights  when  >50  migrants  were  passing  by  the  Empire 
State  Building.  Although  the  lights  associated  with  skyscrapers  are  believed  to  disorient  migrating  birds  and 
result  in  many  bird-to-skyscraper  collisions  each  year.  Peregrine  Falcons  are  able  to  take  advantage  of  the 
situation.  Skyscrapers  provide  hunting  perches  at  altitudes  often  flown  by  nocturnal  migrants,  and  disorientation 
caused  by  the  lights  sometimes  results  in  birds  circling  skyscrapers  and  possibly  becoming  more  vulnerable  to 
predation  by  falcons.  Received  26  January  2005,  accepted  11  October  2005. 


Several  diurnal  raptor  species,  including 
Black-shouldered  Kite  ( Elanus  axillaris ),  Bald 
Eagle  ( Haliaeetus  leucocephalus),  and  Lesser 
Kestrel  ( Falco  naumanni ),  forage  at  night  (see 
Kaiser  1989,  McLaughlin  1989,  Negro  et  al. 
2000).  Others,  such  as  Turkey  Vulture  ( Ca - 
thartes  aura).  Osprey  ( Pandion  haliaetus ), 
Northern  Harrier  ( Circus  cyaneus ),  and  Le- 
vant Sparrowhawk  ( Accipiter  brevipes),  have 
been  observed  flying  or  migrating  at  night 
(Tabor  and  McAllister  1988,  Russell  1991, 
Yosef  2003,  DeCandido  et  al.  2006). 

Peregrine  Falcons  ( Falco  peregrinus)  are 
considered  nocturnal  migrants  in  some  parts 
of  the  world  (Cochran  1985,  Ellis  et  al.  1990), 
and  they  are  known  to  hunt  at  night  (Clunie 
1976,  Russell  1998).  With  increased  numbers 
of  peregrines  nesting  and  wintering  in  cities, 
biologists  are  beginning  to  document  noctur- 
nal activity  by  these  falcons  in  all  seasons. 
Recently,  there  have  been  reports  of  urban 
peregrines  feeding  young  and/or  hunting  at 
night  in  North  America  (Cade  and  Bird  1990, 
Wendt  et  al.  1991,  Cade  et  al.  1996),  England 
(Crick  et  al.  2003),  France  (Marconot  2003), 
Germany  (Schneider  and  Wilden  1994,  Klad- 


1 Hawk  Mountain  Sanctuary,  Acopian  Center  for 
Conservation  Learning,  410  Summer  Valley  Rd.,  Or- 
wigsburg,  PA  17961,  USA. 

2 P.O.  Box  1452,  Peter  Stuyvesant  Station,  New 
York,  NY  10009,  USA. 

3 Current  address:  1831  Fowler  Ave.,  The  Bronx, 
NY  10462,  USA. 

4 Corresponding  author;  e-mail:  rdcny@earthlink.net 


ny  2001),  Netherlands  (van  Dijk  2000,  van 
Geneijgen  2000),  Poland  (Rejt  2000,  2001, 
2004a),  Hong  Kong  (Feare  et  al.  1995),  and 
Taiwan  (K.  Y.  Huang  and  L.  L.  Severinghaus 
unpubl.  data).  However,  direct  observation 
and  analysis  of  nocturnal  hunting  by  Peregrine 
Falcons,  particularly  during  migration,  is  rare 
in  the  literature. 

In  New  York  City,  New  York,  the  number 
and  distribution  of  Peregrine  Falcons  has 
changed  considerably  since  such  observations 
were  first  recorded  in  the  late  1920s.  Before 
the  era  of  DDT  (until  1946),  from  autumn 
through  early  spring,  lone  female  peregrines 
were  much  more  common  at  skyscrapers  than 
males  (Herbert  and  Herbert  1965).  Peregrine 
Falcons  rarely  nested  in  the  city,  and  nocturnal 
activity  by  these  falcons  was  not  reported  in 
any  season  (Herbert  and  Herbert  1965).  Be- 
ginning in  the  mid-1990s,  however,  more  pairs 
of  Peregrine  Falcons  have  begun  residing 
year-round  in  Manhattan  (and  the  metropoli- 
tan area)  than  previously  noted  (B.  A.  Loucks 
pers.  comm.,  C.  Nadareski  unpubl.  data.).  To- 
day, most,  if  not  all,  of  the  seven  pairs  of  per- 
egrines that  nest  in  Manhattan  remain  on  ter- 
ritory year-round.  Here,  we  report  our  obser- 
vations of  Peregrine  Falcon  activity  at  night 
during  the  2004  southbound  bird  migration  at 
one  location  in  New  York  City. 

METHODS 

Most  of  our  observations  of  Peregrine  Fal- 
cons and  nocturnal  migrants  occurred  during 


53 


54 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


the  southbound  migration,  from  4 August  to 
13  November  2004;  we  made  observations  on 
77  of  102  evenings  during  that  period.  In 
spring  2004,  we  observed  northbound  mi- 
grants on  33  evenings  from  19  April  through 
25  May.  In  spring  2002,  we  made  observa- 
tions on  only  2 evenings  (8  May  and  15  May). 

We  made  our  observations  from  the  outside 
observation  deck  (elevation  —325  m above 
ground  level)  of  the  Empire  State  Building 
(ESB),  located  in  midtown  Manhattan  in  New 
York  City.  We  arrived  each  evening  approxi- 
mately 15-30  min  prior  to  sunset.  Bird  mi- 
gration, on  average,  began  30-90  min  after 
sunset.  Any  Peregrine  Falcon  activities  de- 
fined as  nocturnal  occurred  after  nautical  twi- 
light (1  hr  after  sunset).  We  were  able  to  con- 
duct our  study  until  22:45  EST  each  evening 
(August  through  October)  and  until  23:45  in 
November;  the  observation  deck  of  the  build- 
ing was  closed  to  all  visitors  after  these  times. 
In  spring  2004,  we  observed  from  just  before 
sunset  until  22:45  each  evening,  and  in  spring 
2002,  we  observed  from  19:00  until  21:00. 
During  fall  migration,  the  northwest  corner  of 
the  building  provided  the  best  vantage  point 
to  count  the  greatest  number  of  migrating 
birds,  and  in  spring,  we  observed  migrants 
from  the  southwest  corner  of  the  observation 
deck.  These  locations  afforded  unobstructed 
views  to  the  horizon  and  the  sky  above.  We 
used  10X  binoculars  to  follow  peregrines 
when  they  made  long  flights  in  pursuit  of 
prey.  It  was  possible  to  observe  migrating 
birds  and  the  activities  of  peregrines  because 
the  upper  floors  of  the  building  were  illumi- 
nated with  (external)  upward-directed  halogen 
lights,  and  the  spire  above  us  was  illuminated 
with  (internal)  florescent  lights.  We  could  not 
identify  the  majority  of  migrants  to  species 
because  the  external  halogen  lights  washed 
out  most  plumage  details.  However,  this  light- 
ing array  permitted  us  to  count  migrants  up  to 
—30-60  m above  the  highest  point  (445  m 
agl)  of  the  ESB,  and  up  to  30  m (perpendic- 
ular) from  the  observation  deck.  We  estimated 
that  the  building’s  lights  allowed  us  to  see  per- 
egrines chasing  small  birds  in  flight  up  to  60- 
80  m distant. 

Count  protocols  to  assess  nocturnal  bird  mi- 
gration in  2004  followed  those  described  in 
Bildstein  and  Zalles  (1995)  for  migrating  rap- 
tors. An  individual  was  considered  a migrant 


if  it  passed  south-to-north  (or  north-to- south) 
across  an  imaginary  east-west  line  at  the  site, 
and  continued  north  (or  south)  out  of  sight. 
On  2 evenings  during  southbound  migration, 
when  >100  birds  simultaneously  circled  the 
ESB,  we  estimated  the  maximum  number  of 
birds  circling  per  hour  and  recorded  it  as  the 
number  of  migrants  seen  for  that  hour.  We  de- 
fined the  peak  of  migration  as  the  several-day 
period  in  which  we  counted  the  highest  num- 
ber of  migrants.  For  both  northbound  and 
southbound  migration,  total  counts  presented 
here  do  not  include  migrating  waterfowl,  her- 
ons, or  gulls. 

We  defined  a hunting  attempt  as  one  in 
which  a Peregrine  Falcon  approached  to  with- 
in 1 m of  its  intended  prey.  On  a few  occa- 
sions, peregrines  made  repeated  stoops  at  the 
same  prey,  but  did  not  capture  or  gain  control 
of  it.  Each  of  these  stoops  was  considered  a 
separate  hunting  attempt.  Several  times,  we 
observed  a peregrine  strike  a bird  but  fail  to 
seize  it.  We  classified  these  as  unsuccessful 
hunting  attempts. 

We  defined  the  peak  period  of  Peregrine 
Falcon  activity  as  that  during  which  we  ob- 
served falcons  at  the  ESB  during  the  greatest 
number  of  consecutive  nights.  We  used  cor- 
relation statistics  (Microsoft  Excel  2003)  to 
analyze  data  collected  during  this  peak  period. 
We  compared  (a)  the  time  of  arrival  of  the  first 
migrant  after  sunset  with  the  arrival  of  the  first 
Peregrine  Falcon,  and  (b)  the  time  of  arrival 
of  the  first  migrant  with  the  time  of  the  first 
peregrine  hunting  attempt.  Means  are  present- 
ed as  ± SD. 

RESULTS 

During  southbound  migration  in  2004,  we 
saw  the  first  Peregrine  Falcon  at  night  on  4 
August  and  the  last  one  on  the  evening  of  9 
November.  During  this  time,  at  least  two  adult 
peregrines  (male  and  female),  as  well  as  im- 
mature^), used  the  ESB  as  a hunting  perch. 
Peregrines  were  seen  hunting  or  flying  at  night 
on  53%  (41  of  77)  of  the  evenings  we  spent 
at  the  ESB  (Table  1).  Falcons  were  signifi- 
cantly more  likely  to  be  present  on  evenings 
when  >50  migrants  were  counted  in  migra- 
tion (x2  = 14.7,  df  = 1,  P = 0.001;  Table  1). 
Of  the  67  nights  we  observed  migrating  birds, 
peregrines  hunted  migrants  on  25  nights 
(37%),  made  111  hunting  attempts,  and  cap- 


DeCandido  and  Allen  • NOCTURNAL  HUNTING  BY  PEREGRINE  FALCONS 


55 


TABLE  1.  Summary  of  nocturnal  hunting  behavior  by  Peregrine 
migrants  present  after  sunset  in  autumn  2004  at  the  Empire  State  Build 

Falcons  in  relation 
ing.  New  York. 

to  the  number  of 

Number  classes  of  migrant  passerines 

Total 

0 

1-10 

11-50 

51-100 

101-250 

251  + 

No.  nights  migrants  counted 

10 

9 

23 

10 

13 

12 

77 

No.  nights  peregrines  present 

1 

1 

12 

8 

9 

10 

41 

No.  nights  peregrines  hunted 

— 

0 

8 

3 

7 

7 

25 

No.  hunting  attempts 

— 

0 

29 

17 

15 

50 

111 

No.  successful  hunts 

— 

0 

8 

7 

8 

14 

37 

Hunting  success 

— 

— 

28% 

41% 

53% 

28% 

33% 

No.  nights  male  observed  hunting 

— 

0 

5 

2 

5 

6 

18 

No.  nights  female  observed  hunting 

— 

0 

2 

1 

1 

1 

5 

No.  nights  unknown  sex  observed  hunting 

— 

0 

1 

0 

1 

1 

3 

tured  prey  37  times  (33%  success).  All  of  the 
migrants  we  observed  being  captured  or 
chased  were  in  the  warbler-to-oriole  size  class. 

The  peak  of  Peregrine  Falcon  activity  oc- 
curred from  26  September  through  14  October 
2004.  During  that  time,  we  conducted  obser- 
vations on  17  nights;  on  16  of  those  nights  we 
observed  Peregrine  Falcons,  and  on  1 1 nights 
we  observed  them  hunting  (70  total  hunts,  21 
prey  captures,  30%  success).  During  this  pe- 
riod, the  first  migrant  birds  were  observed  65 
± 20  min  after  sunset  (range  = 42-1 14  min); 
Peregrine  Falcons  arrived  91  ±41  min  after 
sunset  (range  = 47-190  min),  and  made  their 
first  hunting  attempt  45  ± 59  min  later  (range 
= 61-284  min),  or  approximately  136  min  af- 
ter sunset.  There  was  no  correlation  between 
passage  of  the  evening’s  first  migrant  and  the 
arrival  of  a Peregrine  Falcon  at  the  ESB  (r2  = 
0.10,  P = 0.73)  or  between  passage  of  the  first 
migrant  and  the  time  of  a peregrine’s  first 
hunting  attempt  (r2  = 0.15,  P — 0.24). 

Nocturnal  hunting  success  was  greatest  in 
September  (12  of  27,  44%)  and  lowest  in  No- 


vember (1  of  8,  13%;  Table  2).  On  10  October 
from  20:12  to  20:42,  a male  Peregrine  Falcon 
made  25  hunting  attempts  and  captured  9 
birds  (36%),  caching  the  birds  on  the  ESB 
tower  after  each  kill.  Throughout  the  autumn, 
we  observed  Peregrine  Falcons  capture  only 
migratory  birds,  although  a few  Rock  Pigeons 
( Columba  livia),  and  at  least  two  bat  species. 
Little  Brown  ( Myotis  lucifugus ) and  Red  ( Las - 
iurus  borealis ) bats,  were  present  on  some 
evenings.  We  could  identify  only  two  prey 
species:  a Baltimore  Oriole  (. Icterus  galbula) 
captured  on  23  August,  and  a Yellow-billed 
Cuckoo  ( Coccyzus  americanus)  taken  on  9 
October.  On  3 and  9 November,  despite  high 
numbers  of  American  Woodcocks  ( Scolopax 
minor ) migrating  past  the  ESB  tower  (36 
counted  each  night),  no  peregrines  were  ob- 
served. 

In  autumn  2004,  most  bird  migration  oc- 
curred at  eye-level  and  above  the  observation 
deck.  We  counted  10,826  migrating  birds,  and 
the  peak  of  the  migration  occurred  from  5 to 
11  October  when  3,871  migrants  (36%  of  the 


TABLE  2.  Summary  of  nocturnal  hunting  behavior  and  success  by  Peregrine  Falcons  during  four  autumn 
months  in  2004  at  the  Empire  State  Building,  New  York. 

Aug 

Sep 

Oct 

Nov 

Total 

No.  hunting  attempts 

16 

27 

60 

8 

111 

No.  successful  hunts 

6 

12 

18 

1 

37 

Hunting  success 

38% 

44% 

30% 

13% 

33% 

No.  nights  one  peregrine  present 

10 

11 

10 

3 

34 

No.  nights  ^2  peregrines  present 

0 

3 

4 

0 

7 

No.  nights  hunting  observed 

5 

9 

10 

1 

25 

No.  nights  male  made  a hunting  attempt 

5 

7 

5 

1 

18 

No.  nights  female  made  a hunting  attempt 

0 

2 

3 

0 

5 

No.  nights  unknown  sex  made  a hunting  attempt 

— 

1 

2 

— 

3 

56 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


fall  flight)  were  counted,  averaging  1 14  birds/ 
hr  on  these  7 evenings.  In  spring  2004,  we 
counted  3,359  migrants  during  33  nights  of 
observation.  The  peak  of  the  migration  oc- 
curred from  6 to  15  May  when  1,752  migrants 
(52%  of  the  spring  flight)  were  counted,  av- 
eraging 51  birds/hr  on  these  10  evenings. 
Lone  Peregrine  Falcons  were  observed  on  2 
evenings:  24  April  (0  migrants  counted)  and 
22  May  (79  counted),  but  no  hunting  attempts 
were  observed  on  either  night.  On  15  May 
2002,  we  observed  an  adult  female  peregrine 
make  10  unsuccessful  hunting  attempts  on  mi- 
grants from  20:15  until  21:00. 

In  the  breeding  season  of  2004,  a pair  of 
Peregrine  Falcons  may  have  attempted  to  nest 
on  the  ESB  (B.  A.  Loucks  pers.  comm.).  It  is 
possible  that  this  pair  executed  many  of  the 
hunting  attempts  we  observed  in  autumn 
2004.  During  5 evenings  between  26  Septem- 
ber and  7 October,  we  saw  an  adult  male  and 
an  adult  female  peregrine  perched  near  one 
another,  each  vocalizing  with  the  “eechup”  or 
“creaking”  call,  and  the  “wailing”  calls  (see 
Ratcliffe  1980).  On  3 October,  we  observed 
three  adults  (a  male,  his  mate,  and  a second 
female)  perched  for  <5  min  within  —20  m of 
one  another  on  the  ESB  tower  until  the  second 
female  was  chased  away — mostly  by  the  fe- 
male of  the  pair.  An  immature  peregrine  was 
present  on  3 evenings:  9 and  14  October,  and 
9 November  2004,  although  we  could  not  be 
sure  if  it  was  the  same  bird  on  all  3 evenings. 
On  5 October,  a Peregrine  Falcon  passed  high 
overhead  flying  south  on  moderate  northerly 
winds  while  an  adult  female  flew  back  and 
forth  near  the  ESB.  It  was  not  uncommon  to 
see  peregrines  flying  high  above  (25-75  m) 
the  top  of  the  ESB  tower  at  night  in  Septem- 
ber and  October. 

DISCUSSION 

Tall,  lighted,  man-made  structures  present 
opportunities  for  biologists  to  study  nocturnal 
hunting  by  Peregrine  Falcons  that  may  not  be 
observed  readily  in  remote  locations.  Urban 
skyscrapers  provide  hunting  platforms  that 
permit  these  raptors  to  perch  at  or  above  the 
elevation  of  nocturnal  migrants,  and  the  lights 
used  to  illuminate  tall  buildings  can  disorient 
migrating  birds  that  may  then  circle  these 
structures,  especially  on  evenings  with  over- 
cast skies  and  light  winds.  These  migrants 


constitute  an  abundant,  easily  accessible  re- 
source for  resident  Peregrine  Falcons,  and  for 
peregrines  migrating  through  the  area  as  well. 

In  New  York  City  in  2004,  Peregrine  Fal- 
cons were  more  likely  to  be  present  and  hunt- 
ing at  the  ESB  on  autumn  nights  when  >50 
migrants  were  observed.  The  peak  of  pere- 
grine activity  at  the  ESB  corresponded  to  the 
peak  of  the  southbound  bird  migration  from 
late  September  through  mid-October.  During 
this  time,  two  adult  peregrines  occasionally 
perched  near  one  another  and  used  the  ESB 
as  a hunting  platform.  More  night  migrants 
were  attracted  to  the  building’s  lights  during 
autumn  rather  than  spring  migration,  and 
many  more  circled  the  tower  for  longer  time 
periods  from  August  through  late  October.  In 
spring,  there  are  fewer  nocturnal  migrants, 
and  these  mostly  pass  higher  above  New  York 
City  on  warm  air  currents  that  override  heavi- 
er, cooler  air  near  the  ground  (see  Kerlinger 
and  Moore  1989).  Each  of  these  factors  likely 
influences  a peregrine’s  decision  to  hunt  mi- 
grants more  frequently  at  night  during  au- 
tumn. On  the  only  spring  night  (15  May  2002) 
during  which  we  did  see  several  peregrine 
hunting  attempts,  winds  were  —24-32  km/hr 
from  the  northwest,  and  many  migrants  passed 
at  or  just  above  the  level  of  the  observation 
deck. 

Peregrine  Falcons  hunted  migrants  in  two 
ways:  pursuit  and  “still  hunting”  ( sensu  Cade 
1982).  At  the  ESB,  greater  success  occurred 
when  they  pursued  prey  in  level  flight  from 
behind;  however,  peregrines  more  often  em- 
ployed still  hunting  from  a west-  or  north-fac- 
ing perch  on  the  spire  above  the  observation 
deck.  When  still  hunting,  they  launched  their 
attacks  at  a 5 to  15°  angle  down  toward  in- 
coming migrants  flying  along  a northwest-to- 
southwest  route  past  the  ESB.  Such  direct  at- 
tacks were  often  unsuccessful,  and  peregrines 
had  to  make  additional  short  stoops  to  secure 
the  prey.  If  the  intended  prey  was  able  to 
dodge  the  initial  attack,  it  would  then  fly 
straight  down  toward  the  ground,  and  pere- 
grines often  made  no  further  pursuits.  We  nev- 
er observed  targeted  prey  attempt  to  escape  by 
“ringing  up,”  nor  did  we  ever  observe  birds 
mass  together  in  a flock  when  a Peregrine  Fal- 
con flew  among  them.  On  some  nights  (e.g., 
10  October),  when  many  migrants  passed  the 
ESB  and  peregrines  captured  several  birds,  we 


DeCandido  and  Allen  • NOCTURNAL  HUNTING  BY  PERLGRINE  FALCONS 


57 


also  observed  unsuccessful  hunting  attempts 
that  were  considerably  less  intense  than  others 
made  on  the  same  evening.  Such  behavior 
may  account  for  the  low  hunting  success  rate 
on  nights  when  >250  migrants  were  counted. 

As  camera  use  increases  for  24-hr  nest  sur- 
veillance, it  may  become  possible  to  deter- 
mine whether  Peregrine  Falcons  frequently 
hunt  at  night  during  the  nesting  season,  and 
whether  this  varies  from  year  to  year  (see  Rejt 
2004b).  Future  studies  at  the  ESB  may  also 
determine  whether  nocturnal  flights  made  to- 
ward conspecifics  are  directed  at  neighboring 
Peregrine  Falcons  or  at  night-migrating  fal- 
cons simply  passing  through  the  area. 

ACKNOWLEDGMENTS 

We  thank  B.  A.  Loucks  for  information  and  kind 
words  of  encouragement.  We  also  acknowledge  C.  A. 
Nadareski’s  long-term  work  with  Peregrine  Falcons  in 
the  metropolitan  area.  A.  Braunlich  and  E.  J.  A.  Drew- 
itt  directed  us  to  recent  literature  about  nocturnal  Per- 
egrine Falcon  activity  in  Europe  and  Asia.  We  thank 
H.  Q.  P.  Crick  for  providing  information  about  Pere- 
grine Falcons  in  England.  K.  L.  Bildstein,  J.  B.  Buch- 
anan, and  D.  Panko  each  critically  read  the  manuscript 
and  made  many  helpful  suggestions  regarding  how  to 
interpret  the  data.  We  also  wish  to  thank  two  anony- 
mous reviewers  for  their  comments  and  ideas.  In  New 
York  City,  S.  Critelli,  C.  R.  Howard,  M.  W.  Kola- 
kowski,  W.  J.  Paulson,  B.  J.  Saunders,  E.  Shapiro,  S. 
J.  Wiley,  and  C.  A.  Wood  observed  Peregrine  Falcons 
with  us.  We  thank  the  staff  of  the  Empire  State  Build- 
ing, including  L.  A.  Ruth  and  all  security  personnel, 
for  facilitating  our  research  and  making  our  evenings 
at  the  building  much  more  pleasant.  We  dedicate  this 
paper  to  Rev.  M.  A.  Hegyi  who  encouraged  the  senior 
author  to  study  New  York  City’s  fauna  and  flora.  This 
is  Hawk  Mountain  Sanctuary’s  contribution  to  conser- 
vation science  number  124. 

LITERATURE  CITED 

Bildstein,  K.  L.  and  J.  I.  Zalles  (Eds.).  1995.  Raptor 
migration  watchsite  manual.  Hawk  Mountain 
Sanctuary,  Kempton,  Pennsylvania. 

Cade,  T.  J.  1982.  The  falcons  of  the  world.  Comstock/ 
Cornell  University  Press,  Ithaca,  New  York. 
Cade,  T.  J.  and  D.  M.  Bird.  1990.  Peregrine  Falcons, 
Falco  peregrinus,  nesting  in  an  urban  environ- 
ment: a review.  Canadian  Field-Naturalist  104: 
209-218. 

Cade,  T.  J.,  M.  Martell,  P.  Redig,  G.  Septon,  and  H. 
Tordoff.  1996.  Peregrine  Falcons  in  urban  North 
America.  Pages  3-13  in  Raptors  in  human  land- 
scapes (D.  M.  Bird,  D.  E.  Varland,  and  J.  J.  Negro, 
Eds.).  Academic  Press,  London,  United  Kingdom. 
Clunie,  F.  1976.  A Fiji  Peregrine  ( Falco  peregrinus) 


in  an  urban-marine  environment.  Notornis  23:8- 
28. 

Cochran,  W.  W.  1985.  Ocean  migration  of  Peregrine 
Falcons:  is  the  adult  male  pelagic?  Pages  223-227 
in  Proceedings  of  Hawk  Migration  Conference  IV 
(M.  Harwood,  Ed.).  Hawk  Migration  Association 
of  North  America,  Rochester,  New  York. 

Crick,  H.,  A.  Banks,  and  R.  Coombes.  2003.  The  na- 
tional peregrine  survey  2002:  results.  British  Trust 
for  Ornithology  News  248:8-9. 

DeCandido,  R.,  R.  O.  Bierregaard,  Jr.,  M.  S.  Mar- 
tell, and  K.  L.  Bildstein.  2006.  Evidence  of  noc- 
turnal migration  by  Osprey  ( Pandion  haliaetus)  in 
North  America  and  Western  Europe.  Journal  of 
Raptor  Research.  In  press. 

Ellis,  D.  H.,  A.  K.  Kepler,  and  C.  B.  Kepler.  1990. 
Evidence  for  a fall  raptor  migration  pathway 
across  the  South  China  Sea.  Journal  of  Raptor  Re- 
search 24:12-18. 

Feare,  C.  J.,  D.  J.  Haskell,  and  J.  R.  Allan.  1995. 
Peregrine  and  Tree  Sparrow  feeding  at  night. 
Hong  Kong  Bird  Report  1994:218-219. 

Herbert,  R.  A.  and  K.  G.  S.  Herbert.  1965.  Behavior 
of  Peregrine  Falcons  in  the  New  York  City  region. 
Auk  69:246-253. 

Kaiser,  G.  W.  1989.  Nightly  concentration  of  Bald  Ea- 
gles at  an  auklet  colony.  Northwestern  Naturalist 
70:12-13. 

Kerlinger,  P.  and  F.  R.  Moore.  1989.  Atmospheric 
structure  and  avian  migration.  Current  Ornitholo- 
gy 6:109-142. 

Kladny,  M.  2001.  Slechtvalk  jaagt  ’s  nachts  op  kok- 
meeuwen.  [Peregrine  Falcon  hunts  Black-headed 
Gulls  at  night.]  Werkgroep  Slechtvalk  Nederland 
7(2):  11.  [In  Dutch] 

Marconot,  B.  2003.  Comportement  de  chasse  noc- 
turne du  Faucon  Pelerin  Falco  peregrinus  a Bel- 
fort. [Nocturnal  hunting  behavior  of  the  Peregrine 
Falcon  Falco  peregrinus  at  Belfort.]  Ornithos  10: 
207-211.  [In  French] 

McLaughlin,  J.  1989.  Black-shouldered  Kites  Elanus 
notatus  active  at  night.  Australian  Bird  Watcher 
13:133. 

Negro,  J.  J.,  J.  Bustamante,  C.  Melguizo,  J.  L.  Ruiz, 
and  J.  M.  Grande.  2000.  Nocturnal  activity  of 
Lesser  Kestrels  under  artificial  lighting  conditions 
in  Seville,  Spain.  Journal  of  Raptor  Research  34: 
327-329. 

Ratcliffe,  D.  A.  1980.  The  Peregrine  Falcon,  2nd  ed. 
T & AD  Poyser,  London,  United  Kingdom. 

Rejt,  L.  2000.  Sklad  pokarmu  sokola  wedrownego 
Falco  peregrinus  w Warszawie.  [Preliminary 
studies  on  the  Peregrine  Falcon  Falco  peregrinus 
diet  in  Warsaw.]  Notatki  Ornitologiczne  41:161  — 
166.  [In  Polish] 

Rejt,  L.  2001.  Feeding  activity  and  seasonal  changes 
in  prey  composition  of  urban  Peregrine  Falcons 
Falco  peregrinus.  Acta  Ornithologica  36:165— 
169. 

Rejt,  L.  2004a.  Nocturnal  behaviour  of  adult  pere- 


58 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


grines  at  the  nest  during  nestling  period.  Vestnik 
Zoologii  38:87-90. 

Rejt,  L.  2004b.  Nocturnal  feeding  of  young  by  urban 
Peregrine  Falcons  ( Falco  peregrinus ) in  Warsaw 
(Poland).  Polish  Journal  of  Ecology  52:63-68. 

Russell,  R.  W.  1991.  Nocturnal  flight  by  migrant  “di- 
urnal” raptors.  Journal  of  Field  Ornithology  62: 
505-508. 

Russell,  R.  W.  1998.  More  peregrine  adventures  from 
the  Gulf,  www.learner.org/jnorth/falll998/jsouth/ 
Updatel02398.html  (accessed  10  December  2004). 

Schneider,  R.  and  I.  Wilden.  1994.  Choice  of  prey 
and  feeding  activity  of  urban  Peregrine  Falcons 
Falco  peregrinus  during  the  breeding  season.  Pag- 
es 203-209  in  Raptor  conservation  today  (B.-U. 
Meyburg  and  R.  D.  Chancellor,  Eds.).  World 
Working  Group  on  Birds  of  Prey  and  Owls,  Lon- 
don, United  Kingdom,  and  Pica  Press,  Shipman, 
Virginia. 


Tabor,  S.  P.  and  C.  T.  McAllister.  1988.  Nocturnal 
flight  by  Turkey  Vultures  ( Cathartes  aura)  in 
southcentral  Texas.  Journal  of  Raptor  Research 
22:91. 

van  Dijk,  J.  2000.  Zwolse  slechtvalken  op  middlebare 
leeftijd.  [Wintering  peregrines  from  juvenile  to 
middle  age.]  Slechtvalk  Nieeuwsbrief.  Werkgroep 
Slechtvalk  Nederland  6(2):6-10.  [In  Dutch] 
van  Geneijgen,  P.  2000.  Slechtvalken  jagen  op  na- 
chtelijke  trekvogels.  [Peregrines  prey  on  nightly 
migrants.]  Slechtvalk  Nieeuwsbrief.  Werkgroep 
Slechtvalk  Nederland  6(1  ):6.  [In  Dutch] 

Wendt,  A.,  G.  Septon,  and  J.  Moline.  1991.  Juvenile 
urban-hacked  Peregrine  Falcons  ( Falco  peregri- 
nus) hunt  at  night.  Journal  of  Raptor  Research  25: 
94-95. 

Yosef,  R.  2003.  Nocturnal  arrival  at  a roost  by  mi- 
grating Levant  Sparrowhawks.  Journal  of  Raptor 
Research  37:64-67. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):59 — 63,  2006 


FIELD  EXPERIMENTS  ON  EGGSHELL  REMOVAL  BY 
MOUNTAIN  PLOVERS 

TEX  A.  SORDAHL1 


ABSTRACT. — I conducted  18  eggshell  removal  trials  at  six  Mountain  Plover  ( Charadrius  montanus ) nests 
in  the  Pawnee  National  Grassland,  Weld  County,  Colorado,  during  June  1994.  Eggshell  fragments  were  placed 
at  various  distances  (10  cm  to  10  m)  from  active  nests.  Attending  adult  plovers  removed  eggshells  throughout 
the  incubation  period.  When  eggshells  were  placed  within  2 m of  the  nest,  plovers  usually  removed  them 
immediately  upon  their  return  to  the  nest.  Shells  placed  farther  away — up  to  10  m — were  removed  after  longer 
time  intervals.  Plovers  removed  shells  by  picking  them  up  with  their  bills  and  running  or  flying  away  with  them 
before  dropping  them  6 to  100  m from  the  nest.  When  returning  to  their  nests,  plovers  approached  by  ground. 
Of  the  five  hypotheses  proposed  in  the  literature  to  explain  the  function  of  eggshell  removal  behavior  in  birds, 
only  one  (reducing  cues  predators  might  use  for  finding  nests)  predicts  removal  of  shells  already  outside  the 
nest  and  disposal  of  shells  far  from  the  nest.  Thus,  my  results  support  an  anti-predator  function  for  eggshell 
removal  in  Mountain  Plovers.  Received  3 November  2004,  accepted  1 October  2005. 


Shortly  after  their  young  hatch,  many  birds 
remove  the  empty  eggshells  and  dispose  of 
them  away  from  the  nest  (Nethersole-Thomp- 
son  and  Nethersole-Thompson  1942,  Skutch 
1976).  This  behavior  is  well  developed  in 
charadriiform  birds,  including  shorebirds  and 
gulls.  In  their  classic  paper,  Tinbergen  et  al. 
(1962)  suggested  five  possible  hypotheses  for 
the  adaptive  value  of  eggshell  removal  behav- 
ior: (1)  eggshells  might  provide  cues  that 
would  attract  predators  to  the  nest;  (2)  later- 
hatching  eggs  might  become  encapsulated,  the 
young  in  hatching  eggs  thus  becoming  trapped 
inside  a double  shell  (termed  “egg-capping” 
by  Derrickson  and  Warkentin  1991);  (3)  sharp 
edges  of  shells  might  injure  chicks  in  the  nest; 
(4)  organic  material  associated  with  eggshells 
might  promote  growth  of  pathogenic  bacteria 
and  mold  in  the  nest;  and  (5)  hatched  shells 
could  interfere  with  brooding  chicks  in  the 
nest.  Tinbergen’s  field  experiments  with  gull 
eggs,  which  are  cryptically  colored  externally 
but  conspicuously  white  inside,  supported  the 
first  hypothesis  by  showing  that  artificial  nests 
with  eggshells  nearby  experienced  greater  pre- 
dation rates  than  those  without  nearby  egg- 
shells (Tinbergen  et  al.  1962,  Tinbergen 
1963).  Tinbergen,  however,  did  not  rule  out 
the  remaining  hypotheses.  Subsequent  litera- 
ture has  tended  to  support  the  predation  (Sor- 
dahl  1994,  Sandercock  1996)  and  egg-capping 
hypotheses  (Derrickson  and  Warkentin  1991, 


1 Dept,  of  Biology,  Luther  College,  Decorah,  I A 
52101,  USA;  e-mail:  sordahlt@luther.edu 


Sandercock  1996,  Verbeek  1996,  Hauber 
2003). 

Hypotheses  3,  4,  and  5 seem  unlikely  ex- 
planations of  the  evolution  of  eggshell  remov- 
al behavior  in  shorebirds  because  their  eggs 
usually  hatch  synchronously  and  the  precocial 
young  leave  the  nest  within  24  hr  of  hatching. 
Sandercock  (1996)  reported  observations  of 
egg-capping  in  two  sandpiper  species,  sup- 
porting hypothesis  2.  However,  he  recognized 
that  egg-capping  alone  could  not  account  for 
the  form  of  removal  behavior  typically  seen 
in  shorebirds — specifically,  the  disposal  of 
eggshells  far  from  the  nest — and  concluded 
that  both  egg-capping  and  predation  have  con- 
tributed to  the  evolution  of  eggshell  removal 
behavior  in  these  birds. 

Here,  I report  the  results  of  field  trials  on 
eggshell  removal  behavior  of  Mountain  Plo- 
vers ( Charadrius  montanus).  Mountain  Plo- 
vers nest  on  the  ground  in  very  open  habitat, 
where  predation  is  the  major  cause  of  egg  and 
chick  losses  (Graul  1975,  McCaffery  et  al. 
1984,  Sordahl  1991,  Miller  and  Knopf  1993, 
Knopf  1996,  Knopf  and  Rupert  1996).  Gen- 
eral aspects  of  eggshell  removal  in  this  species 
were  described  by  Graul  (1975).  My  experi- 
ments enabled  me  to  provide  a quantitative 
description  of  the  behavior  and  to  evaluate  its 
function. 

METHODS 

I performed  field  trials  on  eggshell  removal 
by  Mountain  Plovers  from  9 to  18  June  1994 
at  Pawnee  National  Grassland,  Weld  County, 


59 


60 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


TABLE  1.  Results  of  18  field  trials  on  eggshell  removal  behavior  at  six  Mountain  Plover  nests.  Pawnee 
National  Grassland,  Colorado,  9-18  June  1994.  In  each  trial,  one-third  of  a complete  eggshell  (of  Mountain 
Plover  or  Japanese  Quail)  was  placed  near  the  nest  and  the  behavior  of  the  adult  was  observed  upon  its  return 
to  the  nest. 

Nest3 

Incubation  day 

Shell  type 

Nest-shell 
distance  (m) 

Time  until 
removal  (min) 

Removal 

method 

Disposal 
distance  (m) 

SI 

5 

Quail 

0.5 

0.08 

Fly 

70 

SI 

6 

Quail 

0.7 

10 

b 

— 

SI 

7 

Quail 

0.6 

0 

Fly 

60 

K1 

7 

Quail 

1.0 

0 

Run-fly 

30 

K1 

8 

Quail 

0.5 

0.17 

Run 

6 

K1 

8 

Quail 

2.5 

97 

Run 

17 

K1 

8 

Quail 

5.0 

105 

— 

— 

K1 

8 

Quail 

10.0 

— 

— 

22 

K1 

9 

Quail 

1.5 

26 

— 

— 

S2C 

8 

Quail 

0.2 

0 

Run 

20 

S2 

8 

Quail 

0.5 

0 

Run 

30 

K2 

15 

Plover 

2.0 

3 

Run-fly 

100 

K2 

16 

Quail 

3.0 

0 

Run 

18 

K2 

16 

Quail 

4.0 

69 

— 

- — 

R1 

20 

Plover 

0.1 

0 

Run 

12 

R1 

23 

Plover 

0.3 

0 

Fly 

90 

R1 

25 

Quail 

0.7 

0 

Run 

15 

K3 

27 

Plover 

1.5 

0 

Run 

18 

a Mountain  Plovers  typically  exhibit  uniparental  care;  therefore,  egg  removals  were  assumed  to  represent  the  behavior  of  one  adult  per  nest. 
b Missing  data  in  the  table  indicate  that  shell  removal  was  not  observed  (see  text)  or  that  the  disposed  shell  was  not  found. 
c Nest  S2  contained  four  eggs;  all  other  nests  contained  three. 


Colorado  (40°  45'  N,  104°  00'  W).  This  short- 
grass  prairie  site  has  been  well  described  else- 
where (Graul  1973,  1975;  McCaffery  et  al. 
1984).  Its  vegetation  was  very  short  and 
sparse,  and  it  was  grazed  by  cattle. 

I studied  eggshell  removal  at  six  Mountain 
Plover  nests.  Five  nests  contained  three-egg 
clutches  (normal  for  Mountain  Plovers)  and 
one  nest  contained  four  eggs.  The  attending 
adults  were  not  marked  for  identification,  but 
since  uniparental  care  is  typical  in  this  species 
(Knopf  1996),  it  is  likely  that  I tested  six  dif- 
ferent individuals.  Mountain  Plovers  are  sex- 
ually monomorphic  (Hayman  et  al.  1986, 
Knopf  1996),  so  I was  unable  to  determine  the 
sex  of  the  birds.  Trials  entailed  placing  ap- 
proximately one-third  of  a complete  eggshell 
on  the  ground  (interior — or  white — side  up)  at 
various  distances  (ranging  from  10  cm  to  10 
m)  from  the  nest  and  then  observing  the  be- 
havior of  the  adult  when  it  returned  to  its  nest. 
I conducted  18  trials,  14  with  Japanese  Quail 
{Coturnix  japonica)  eggshells  obtained  com- 
mercially and  4 with  Mountain  Plover  egg- 
shells that  I found  opportunistically  in  the 
field.  The  two  species’  shells  are  similar  in 
size  and  appearance,  both  having  earth-tone 


background  colors  and  dark,  irregular  mark- 
ings. Adult  plovers  responded  similarly  to  the 
two  kinds  of  shells;  therefore,  I pooled  the 
results. 

Observations  were  made  from  a vehicle 
about  100  m from  nests  with  7 X 35  binoc- 
ulars. For  each  trial,  I recorded  the  nest-to- 
shell  distance,  the  amount  of  time  elapsed  be- 
tween the  adult’s  return  to  the  nest  and  re- 
moval of  the  shell,  the  removal  method  (run 
or  fly),  the  disposal  distance,  and  the  method 
(run  or  fly)  of  returning  to  the  nest  after  shell 
disposal.  At  least  one  egg  hatched  in  every 
nest  and,  assuming  that  incubation  begins 
when  the  clutch  is  complete  and  the  average 
incubation  period  is  29  days  (Knopf  1996),  I 
used  backdating  to  determine  days  since  in- 
cubation began.  I measured  the  distances  of 
eggshells  from  nests  with  a tape  measure,  and 
disposal  distances  of  shells  that  I was  able  to 
relocate  by  pacing. 

RESULTS 

The  number  of  trials  conducted  at  each  of 
the  six  nests  was  6,  3,  3,  3,  2,  and  1 (Table 
1).  The  attending  adult  Mountain  Plover  re- 
moved shells  at  all  six  nests.  Nine  of  18  shells 


Sordahl  • EGGSHELL  REMOVAL  BY  MOUNTAIN  PLOVERS 


61 


FIG.  1.  Relationship  between  the  distance  egg- 
shells were  placed  from  Mountain  Plover  nests  and  the 
time  elapsed  before  the  adult  removed  the  shell,  9-18 
June  1994,  Pawnee  National  Grassland,  Weld  County, 
Colorado.  Three  of  the  17  points  overlap  at  0.5  m (2 
are  hidden). 

were  removed  immediately  upon  the  adult’s 
return  to  the  nest;  2 more  were  removed  with- 
in 10  sec.  Two  other  shells  were  removed  3 
min  and  10  min  after  the  adults  had  returned. 
Four  of  the  remaining  five  shells  were  re- 
moved in  less  than  2 hr.  The  final  shell,  placed 
10  m from  the  nest,  was  not  removed  during 
15  min  of  observation,  at  which  time  I de- 
parted the  nest  site;  the  following  morning  I 
found  the  shell  22  m from  the  nest.  Although 
it  is  possible  that  the  wind  or  another  animal 
moved  this  shell,  it  seems  most  likely  that  the 
adult  plover  moved  it.  Overall,  shells  placed 
within  2 m of  the  nest  were  removed  promptly 
(most  of  them  immediately),  whereas  shells 
placed  farther  away  were  removed  after  longer 
intervals  (Fig.  1).  Eggshell  removal  was  doc- 
umented on  incubation  days  5-9,  15,  16,  20, 
23,  25,  and  27  (Table  1). 

I recorded  eggshell  removal  and  adult  re- 
turn to  the  nest  for  13  of  18  trials  (Table  1). 
During  the  remaining  five  trials,  which  had 
long  eggshell  removal  times,  my  vigilance 
was  intermittent  and  I did  not  observe  the  ac- 
tual removal.  However,  by  checking  for  the 
eggshell  as  soon  as  I noticed  that  the  bird  was 
off  the  nest,  I was  able  to  record  removal 
times  with  only  a small  margin  of  error  (ex- 
cept in  the  case  described  above,  where  I left 
the  site  before  removal  occurred).  When  a 
Mountain  Plover  removed  an  eggshell,  it  pick- 
ed the  shell  up  with  its  bill  and  ran  away  with 


it  (8  of  13  observations),  flew  off  with  it  (3 
of  13  observations),  or  ran  2-3  m before  fly- 
ing off  with  it  (2  of  13  observations).  On  14 
occasions  I was  able  to  recover  shells  where 
they  were  dropped;  disposal  distances  ranged 
from  6 to  100  m from  the  nest  (Table  1).  Plo- 
vers tended  to  dispose  of  shells  at  greater  dis- 
tances when  they  flew  (mean  = 70.0  m,  range 
— 30—100,  n = 5)  than  when  they  ran  (mean 
= 17.0  m,  range  = 6-30,  n = 8).  On  four 
occasions  I recorded  which  facet  (inside  or 
outside)  of  a recovered  shell  was  exposed;  two 
shells  were  lying  with  the  cryptic  outside  fac- 
ing up  and  two  were  lying  with  the  conspic- 
uous inside  of  the  shell  facing  upward.  After 
disposing  of  the  shells,  adults  always  returned 
to  their  nests  by  a ground  approach  (13  of  13 
observations),  which  is  typical  of  plovers 
(TAS  pers.  obs.). 

DISCUSSION 

My  field  experiments  demonstrated  that 
Mountain  Plovers  remove  eggshells  through- 
out the  incubation  period.  This  may  be  true 
for  most  birds,  and  the  expression  of  the  be- 
havior long  before  hatching  occurs  likely  has 
been  selected  for  in  the  context  of  removal  of 
damaged  eggs  (Nethersole-Thompson  and 
Nethersole-Thompson  1942,  Montevecchi 
1976,  Kemal  and  Rothstein  1988,  Sordahl 
1994).  Removal  of  dead  chicks  from  the  nest 
also  has  been  reported  (Nethersole-Thompson 
1951:183,  Skutch  1976:284,  Sordahl  1994). 

Because  it  had  already  been  demonstrated 
that  Mountain  Plovers  remove  eggshells  lo- 
cated in  their  nests  (Graul  1975,  Knopf  1996; 
TAS  pers.  obs.),  I designed  my  experiments 
to  determine  whether  they  would  remove 
shells  placed  outside  the  nest  and,  if  so,  how 
far  from  the  nest  they  would  go  to  remove 
shells.  I observed  adults  immediately  remove 
shells  that  had  been  placed  up  to  3 m from 
their  nests  (Table  1,  Fig.  1).  They  also  even- 
tually removed  shells  at  distances  of  4,  5,  and 
probably  10  m,  as  well.  Because  the  average 
disposal  distance  was  only  17  m for  birds  that 
removed  eggshells  by  running,  it  seems  un- 
likely that  Mountain  Plovers  would  remove 
shells  located  much  farther  from  their  nests 
than  10  m. 

The  closer  a shell  was  placed  to  the  nest, 
the  more  quickly  it  was  removed  (Fig.  1).  The 
proximate  explanation  for  this  probably  is  that 


62 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  1,  March  2006 


adults  were  less  likely  to  detect  eggshells  that 
were  farther  from  the  nest.  Even  though 
Mountain  Plover  nesting  habitat  is  shortgrass 
prairie,  the  line  of  sight  a plover  has  when 
making  a ground  approach  to  its  nest  is  low 
enough  that  even  small  obstructions  could  in- 
terfere with  its  ability  to  notice  a distant  shell. 
An  ultimate  explanation  for  this  finding  would 
be  that  the  risk  of  predation  due  to  the  pres- 
ence of  eggshells  diminishes  with  distance 
from  the  nest,  as  shown  by  Tinbergen  et  al. 
(1962)  for  Black-headed  Gull  (. Larus  ridibun- 
dus ) eggs.  Tinbergen  et  al.  (1962)  found  that 
a broken  eggshell  ^1  m from  an  artificial 
clutch  increased  the  predation  rate,  but  an 
eggshell  2 m away  did  not.  If  the  radius  of 
increased  risk  is  similar  for  Mountain  Plovers, 
one  might  expect  them  to  be  less  diligent 
about  removing  shells  >2  m from  the  nest. 
My  results  are  consistent  with  this  because  the 
birds  did  not  immediately  remove  shells  that 
were  >2-3  m away.  Nevertheless,  they  even- 
tually did  remove  those  shells,  which  suggests 
that  such  shells  pose  at  least  some  risk  to  the 
clutch. 

Although  eggshell  removal  and  disposal 
distances  have  not  been  investigated  system- 
atically in  birds,  these  distances  most  likely 
represent  a compromise  between  the  benefits 
of  removal  and  the  costs  of  leaving  the  nest 
when  young  are  hatching.  Factors  that  prob- 
ably influence  these  distances  are  habitat  (es- 
pecially open  habitats  in  the  case  of  Mountain 
Plovers),  the  degree  of  nest  dispersion  (widely 
spaced  in  Mountain  Plovers),  and  which  spe- 
cies of  egg  and  chick  predators  inhabit  the 
area  (mammals  and  snakes  are  thought  to  be 
important  predators  of  Mountain  Plovers; 
Knopf  1996). 

Of  the  five  hypotheses  explaining  the  adap- 
tive value  of  eggshell  removal,  the  only  one 
that  predicts  removal  of  eggshells  already  out- 
side the  nest  is  the  predation  hypothesis.  It 
also  is  the  only  hypothesis  that  predicts  dis- 
posal far  from  the  nest.  Thus  my  results  sup- 
port an  anti-predator  function  for  eggshell  re- 
moval in  Mountain  Plovers.  Similarly,  fecal 
sac  removal  by  many  nidicolous  birds  (which 
is  analogous  to  eggshell  removal)  involves 
disposal  of  fecal  sacs  far  from  the  nest  (Petit 
et  al.  1989  and  references  therein),  and  this 
behavior  also  seems  best  explained  as  a means 
of  reducing  cues  that  could  lead  predators  to 


nests  (Petit  et  al.  1989,  Lang  et  al.  2002). 
However,  I cannot  rule  out  the  possibility  that 
eggshell  removal  serves  functions  other  than 
predation  avoidance.  For  example,  if  there  is 
a risk  that  wind  may  blow  shells  back  into  the 
nest,  it  may  be  adaptive  to  dispose  of  them 
far  away  so  they  do  not  threaten  the  chicks 
with  encapsulation  or  injury.  Further  research 
is  needed  to  examine  these  alternative  expla- 
nations of  eggshell  removal  behavior. 

ACKNOWLEDGMENTS 

I especially  thank  C.  A.  Ristau  for  enabling  me  to 
study  Mountain  Plovers  in  Colorado.  She  and  F.  L. 
Knopf  generously  shared  information  about  nests  they 
had  found.  The  manuscript  benefited  from  comments 
by  D.  I.  Bishop,  S.  J.  Dinsmore,  M.  B.  Wunder,  and 
two  anonymous  referees.  This  research  was  conducted, 
in  part,  on  the  Central  Plains  Experimental  Range  op- 
erated by  the  USDA  Agricultural  Research  Service  and 
made  available  to  the  Long  Term  Ecological  Research 
Program  administered  by  Colorado  State  University.  I 
thank  M.  Lindquist  for  courtesies  extended  during  my 
stay.  I was  supported  by  a Luther  College  Faculty  Re- 
search Grant. 

LITERATURE  CITED 

Derrickson,  K.  C.  and  I.  G.  Warkentin.  1991.  The 
role  of  egg-capping  in  the  evolution  of  eggshell 
removal.  Condor  93:757-759. 

Graul,  W.  D.  1973.  Adaptive  aspects  of  the  Mountain 
Plover  social  system.  Living  Bird  12:69-94. 
Graul,  W.  D.  1975.  Breeding  biology  of  the  Mountain 
Plover.  Wilson  Bulletin  87:6-31. 

Hauber,  M.  E.  2003.  Egg-capping  is  a cost  paid  by 
hosts  of  interspecific  brood  parasites.  Auk  120: 
860-865. 

Hayman,  P,  J.  M archant,  and  T.  Prater.  1986. 
Shorebirds:  an  identification  guide  to  the  waders 
of  the  world.  Houghton  Mifflin,  Boston,  Massa- 
chusetts. 

Kemal,  R.  E.  and  S.  I.  Rothstein.  1988.  Mechanisms 
of  avian  egg  recognition:  adaptive  responses  to 
eggs  with  broken  shells.  Animal  Behaviour  36: 
175-183. 

Knopf,  F.  L.  1996.  Mountain  Plover  ( Charadrius  mon- 
tanus).  The  Birds  of  North  America,  no.  211. 
Knopf,  F.  L.  and  J.  R.  Rupert.  1996.  Reproduction 
and  movements  of  Mountain  Plovers  breeding  in 
Colorado.  Wilson  Bulletin  108:28—35. 

Lang,  J.  D.,  C.  A.  Straight,  and  P.  A.  Gowaty.  2002. 
Observations  of  fecal  sac  disposal  by  Eastern 
Bluebirds.  Condor  104:205-207. 

McCaffery,  B.  J.,  T.  A.  Sordahl,  and  P.  Zahler. 
1984.  Behavioral  ecology  of  the  Mountain  Plover 
in  northeastern  Colorado.  Wader  Study  Group 
Bulletin  40:18-21. 

Miller,  B.  J.  and  F.  L.  Knopf.  1993.  Growth  and  sur- 


Sordahl  • EGGSHELL  REMOVAL  BY  MOUNTAIN  PLOVERS 


63 


vival  of  Mountain  Plovers.  Journal  of  Field  Or- 
nithology 64:500-506. 

Montevecchi,  W.  A.  1976.  Eggshell  removal  by 
Laughing  Gulls.  Bird-Banding  47:129—135. 

Nethersole-Thompson,  C.  and  D.  Nethersole- 
Thompson.  1942.  Egg-shell  disposal  by  birds. 
British  Birds  35:162-169,  190-200,  214-223, 
241-250. 

Nethersole-Thompson,  D.  1951.  The  Greenshank. 
Collins,  London,  Great  Britain. 

Petit,  K.  E.,  L.  J.  Petit,  and  D.  R.  Petit.  1989.  Fecal 
sac  removal:  do  the  pattern  and  distance  of  dis- 
persal affect  the  chance  of  nest  predation?  Condor 
91:479-482. 

Sandercock,  B.  K.  1996.  Egg-capping  and  eggshell 
removal  by  Western  and  Semipalmated  sandpip- 
ers. Condor  98:431-433. 


Skutch,  A.  F.  1976.  Parent  birds  and  their  young.  Uni- 
versity of  Texas  Press,  Austin. 

Sordahl,  T.  A.  1991.  Antipredator  behavior  of  Moun- 
tain Plover  chicks.  Prairie  Naturalist  23:109-1  15. 

Sordahl,  T.  A.  1994.  Eggshell  removal  behavior  of 
American  Avocets  and  Black-necked  Stilts.  Jour- 
nal of  Field  Ornithology  65:461—465. 

Tinbergen,  N.  1963.  The  shell  menace.  Natural  His- 
tory 72(7):28-35. 

Tinbergen,  N.,  G.  J.  Broekhuysen,  F.  Feekes,  J.  C. 
W.  Houghton,  H.  Kruuk,  and  E.  Szulc.  1962. 
Egg  shell  removal  by  the  Black-headed  Gull,  La- 
rus  ridibundus  L.:  a behaviour  component  of  cam- 
ouflage. Behaviour  19:74-117. 

Verbeek,  N.  A.  M.  1996.  Occurrence  of  egg-capping 
in  birds’  nests.  Auk  113:703-705. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):64 — 69,  2006 


SEED-SIZE  SELECTION  IN  MOURNING  DOVES  AND 
EURASIAN  COLLARED-DOVES 

STEVEN  E.  HAYSLETTE1 


ABSTRACT. — I studied  seed-size  selection  among  Mourning  Doves  ( Zenaida  macroura ) and  Eurasian  Col- 
lared-Doves  ( Streptopelia  decaocto ),  two  newly  sympatric  species  for  which  mechanisms  of  seed  selection  are 
not  well  understood.  I measured  and  compared  mean  length,  breadth,  and  thickness  of  seeds  available  to,  and 
consumed  by,  these  species  in  feeding  trials  of  penned  birds.  Both  species  selected  com  ( Zea  mays ) seeds  that 
were  shorter  and  narrower  than  average,  but  Eurasian  Collared-Doves  selected  com  that  was  thicker  than  average 
and  sunflower  ( Helianthus  annuus ) seeds  that  were  broader  and  thicker  than  average.  Mourning  Doves  consumed 
com  of  average  thickness,  and  wheat  ( Triticum  aestivum ) and  sunflower  seeds  of  average  size  with  respect  to 
all  dimensions.  Corn  consumption  by  both  species  seems  limited  by  seed  length  and  breadth,  but  Mourning 
Dove  consumption  of  smaller  seed  types  (wheat  and  milo  [Sorghum  vulgare ])  appears  largely  unaffected  by 
seed  size.  Among  larger  seed  types  (com  and  sunflower),  Eurasian  Collared-Doves  may  select  thicker-  and/or 
broader-than-average  seeds  to  maximize  foraging  efficiency.  Sunflower  and  com  seeds  consumed  did  not  vary 
between  species  with  respect  to  any  dimension,  but  Eurasian  Collared-Doves  seemed  willing  to  select,  and  able 
to  eat,  broader  and  thicker  seeds  than  Mourning  Doves,  which  may  limit  foraging  competition  between  these 
species.  Received  7 February  2005,  accepted  23  November  2005. 


Seed  selection  by  granivorous  birds  is  a 
complex  phenomenon  potentially  affected  by 
a number  of  factors  (Ramos  1996),  the  rela- 
tive contributions  of  which  remain  poorly  un- 
derstood in  many  avian  granivores.  In  partic- 
ular, seed  selection  by  doves  and  other  species 
that  do  not  husk  seeds  before  swallowing  is 
not  well  understood;  most  studies  of  seed  se- 
lection in  birds  have  focused  on  finches  and 
other  species  that  husk  seeds  during  the  course 
of  foraging.  Generally,  these  studies  have  re- 
vealed that  the  physical  characteristics  of 
seeds  affecting  handling  time,  such  as  size, 
shape,  and  hardness,  are  important  determi- 
nants of  preference,  and  that  nutritional  com- 
position of  foods  appears  relatively  unimpor- 
tant (Willson  1971,  Willson  and  Harmeson 
1973,  Goldstein  and  Baker  1984,  De  Nagy 
Koves  Hrabar  and  Perrin  2002),  especially 
without  consideration  of  the  overall  econom- 
ics of  nutrient  intake  and  the  factors  affecting 
it  (Greig-Smith  and  Wilson  1985). 

One  approach  to  understanding  the  effect  of 
seed  size  on  selection  has  been  to  examine 
size  selection  by  one  or  more  species  for  a 
single  seed  type  (Hespenheide  1966,  Myton 
and  Ficken  1967,  Willson  1972,  Abbott  et  al. 
1975,  Greig-Smith  and  Crocker  1986,  van  der 
Meij  and  Bout  2000).  A number  of  these  stud- 
ies have  indicated  seed-size  preference  within 


1 Dept,  of  Biology,  Tennessee  Tech  Univ.,  Cooke- 
ville, TN  38505,  USA;  e-mail:  shayslette@tntech.edu 


a species  (Greig-Smith  and  Crocker  1986,  van 
der  Meij  and  Bout  2000),  and/or  correspon- 
dence between  size  selection  and  bill  size 
among  multiple  species  (Hespenheide  1966, 
Myton  and  Ficken  1967,  Willson  1972).  One 
study  indicated  no  size  preferences  and/or  no 
seed  size/bill  size  correspondence  (Abbott  et 
al.  1975),  but  this  study  focused  on  a seed- 
husking  species.  Because  doves  and  pigeons 
have  relatively  long  slender  bills  and  pecking 
behaviors  that  maximize  speed  of  seed  intake 
without  husking  seeds  (De  Nagy  Koves  Hra- 
bar and  Perrin  2002),  seed  size  may  be  ex- 
pected to  affect  seed  handling  and  preferences 
differently  than  in  most  species  studied  pre- 
viously. 

The  overall  goal  of  this  project  was  to  de- 
termine the  effect  of  seed  size  on  food  selec- 
tion by  Mourning  Doves  {Zenaida  macroura) 
and  Eurasian  Collared-Doves  {Streptopelia 
decaocto).  Eurasian  Collared-Doves  are  recent 
exotic  invaders  of  North  America,  and  may 
compete  for  food  or  other  resources  with  na- 
tive species,  such  as  Mourning  Doves,  to  the 
detriment  of  native  species  (Romagosa  2002). 
Bill-size-related  differences  in  seed-size  selec- 
tion between  Eurasian  Collared-Doves  and 
Mourning  Doves  may  mitigate  competition 
between  these  species  for  food  resources, 
however  (Poling  and  Hayslette  2006).  Subdi- 
vision of  food  resources  among  sympatric  avi- 
an granivores  often  is  based  on  bill-size-relat- 


64 


Hayslette  • SEED-SIZE  SELECTION  IN  DOVES 


65 


ed  differences  in  seed-size  selection  (Grant 
1986,  Faaborg  1988,  Ricklefs  2001).  Mourn- 
ing Dove  bill  length  averages  12.83-14.53  ± 
0.97-1.01  mm  (Mirarchi  and  Baskett  1994), 
whereas  collared-dove  bill  length  averages 
16.9  ± 0.71  mm  for  males  and  16.6  ± 1.02 
mm  for  females  (Romagosa  2002).  Corre- 
sponding differences  in  seed-size  selection 
patterns  between  these  species  may  have  im- 
portant implications  regarding  dietary  overlap 
and  competition,  and  ultimately,  coexistence 
of  the  two  species. 

In  cafeteria  trials,  previous  work  has  indi- 
cated that  Mourning  Doves  and  other  dove 
species  prefer  small,  round  seeds  such  as 
white  proso  millet  ( Panicum  miliaceum ; 
Hayslette  and  Mirarchi  2001,  De  Nagy  Koves 
Hrabar  and  Perrin  2002),  and  consume  rela- 
tively few  large-seeded  species  such  as  com 
{Zea  mays ) and  sunflower  ( Helianthus  an- 
nuus’,  LeBlanc  and  Otis  1998,  Hayslette  and 
Mirarchi  2001).  These  results  appear  enig- 
matic, as  wild  Mourning  Doves  are  known  to 
exploit  com  and  sunflower  as  important  food 
sources  (Lewis  1993).  De  Nagy  Koves  Hrabar 
and  Perrin  (2002)  concluded  that  among  Di- 
amond Doves  ( Geopelia  cuneata ),  seed  size 
becomes  a limiting  factor  above  a threshold 
size,  but  below  that  threshold,  size  is  of  little 
importance  in  food  handling  and  selection.  I 
hypothesized  that  seed  size  has  little  influence 
on  Mourning  Dove  selection  of  small  seeds, 
but  that  preferences  for,  and  consumption  of, 
larger  seeds,  such  as  com  and  sunflower,  are 
limited  by  seed  size.  Based  on  this  hypothesis, 
I predicted  that  Mourning  Doves  would  pref- 
erentially select  smaller  than  average  corn 
seeds,  and  that  this  within-seed-type  selectiv- 
ity for  size  would  decrease  with  progressively 
smaller  seed  types.  I also  hypothesized  that 
Eurasian  Collared-Doves  are  less  limited  than 
Mourning  Doves  by  size  of  com  and  sunflow- 
er seeds  because  their  bills  are  larger;  thus, 
they  are  able  to  exploit  com  and  sunflower 
food  sources  to  a greater  extent  than  Mourn- 
ing Doves.  Previous  research  has  shown  that 
Eurasian  Collared-Doves  consume  more  com 
than  Mourning  Doves  in  cafeteria  trials  (Pol- 
ing and  Hayslette  2006).  Based  on  this,  I pre- 
dicted that  Eurasian  Collared-Doves  would 
show  less  within-seed-type  selectivity  for 
smaller  com  and  sunflower  seeds  than  Mourn- 


ing Doves,  and  would  select  larger  corn  and 
sunflower  seeds  than  Mourning  Doves. 

METHODS 

The  first  phase  of  this  research  was  con- 
ducted at  the  captive  Mourning  Dove  research 
facility  at  Auburn  University,  Alabama,  from 
June  to  August  2000.  I used  15  2nd-year 
Mourning  Doves,  captured  as  immatures  on 
the  university  campus  during  the  previous 
breeding  season;  doves  were  initially  housed 
and  cared  for  according  to  Mirarchi  (1993). 
Prior  to  feeding  trials,  doves  were  fed  an  equal 
mixture  (by  volume)  of  the  four  foods  used  in 
feeding  trials  (described  below)  plus  proso 
millet  and  browntop  millet  ( Panicum  fasci- 
culatum).  I randomly  assigned  doves  to  five 
flocks  of  three  birds  each,  and  each  flock  was 
used  in  a 20-hr  feeding  trial  in  a 3.7  X 7.3  X 
2.0-m  outdoor  aviary.  During  each  trial,  doves 
were  offered  200  seeds  of  each  of  four  spe- 
cies— corn,  black-oil  sunflower,  milo  ( Sor- 
ghum vulgare),  and  wheat  ( Triticum  aesti- 
vum).  Trials  were  preceded  by  >24-hr  accli- 
mation periods,  during  which  doves  were  fed 
an  equal  mixture  (by  volume)  of  test  seeds 
only,  followed  by  24-hr  fasting  periods.  Prior 
to  each  trial,  I estimated  size  of  seeds  offered 
using  a sample  of  20  seeds  drawn  at  random 
from  each  200-seed  batch.  I attempted  to  in- 
sure that  sizes  of  seeds  in  samples  were  rep- 
resentative of  those  in  feeding  batches  by 
comparing  the  mass  of  each  sample  to  the 
mass  of  the  batch  from  which  it  was  drawn. 
Because  I sampled  10%  of  each  batch  (20  out 
of  200  seeds),  mass  of  a representative  sample 
would  be  10%  of  the  mass  of  the  batch  from 
which  it  was  drawn.  Thus,  a sample  was 
deemed  representative  and  used  if  the  ratio  of 
sample  mass  to  batch  mass  was  0.100  ± 
0.003.  If  not,  the  sample  was  returned  to  the 
batch  and  redrawn.  Using  digital  calipers,  I 
measured  length,  breadth,  and  thickness  (cor- 
responding to  the  longest,  intermediate-most, 
and  shortest  dimensions,  respectively;  Greig- 
Smith  and  Crocker  1986)  of  each  seed  in  each 
sample  to  the  nearest  0.1  mm.  Seeds  of  each 
species  were  then  hand-scattered  on  a separate 
wooden  seed  tray  (41  X 41  X 4 cm)  filled 
with  commercially  available  topsoil;  trays 
were  randomly  arranged  in  a 2 X 2 arrange- 
ment on  the  floor  of  the  aviary,  with  1.8  m 
between  adjoining  trays.  After  allowing  doves 


66 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


to  forage  from  the  trays  for  20  hr,  I removed 
the  trays  and  removed  and  counted  the  seeds 
that  remained  on  the  trays.  If  >20  seeds  re- 
mained, I estimated  size  of  seeds  remaining  in 
the  batch  using  a sample  of  20  seeds  as  before 
the  trial.  Analogous  to  pre-trial  sampling, 
post-trial  samples  were  deemed  acceptable 
based  on  comparisons  of  sample  mass  and 
batch  mass.  A sample  was  used  if  the  ratio  of 
sample  mass  to  batch  mass  was  within  0.003 
of  the  ratio  of  sample  seed  number  (20)  to 
batch  seed  number.  If  not,  the  sample  was  re- 
turned to  the  batch  and  redrawn.  I compared 
number  of  seeds  consumed  among  seed  types 
using  one-way  analysis  of  variance  (ANOVA) 
and  Tukey’s  procedure.  I calculated  mean  size 
of  each  pre-  and  post-trial  sample  with  respect 
to  all  three  size  dimensions.  I calculated  the 
average  size  of  seeds  consumed  with  respect 
to  each  dimension  for  each  seed  type  in  each 
trial  based  on  number  of  seeds  consumed  and 
pre-  versus  post-trial  differences  in  average 
available  seed  size.  The  formula  for  this  cal- 
culation is 

Se  = {(200  X Sh ) - [(200  - Ne)  X Sa]  }/Ne, 

where  Se  = average  size  of  seeds  eaten,  Sb  = 
average  size  of  seeds  prior  to  foraging  (initial 
sample),  Sa  = average  size  of  seeds  not  con- 
sumed, and  Ne  = number  of  seeds  consumed. 
I then  used  a paired  Mest — with  trials  as  rep- 
licates— to  compare  mean  size  consumed  with 
mean  size  available  for  each  size  dimension 
and  each  seed  type. 

The  second  phase  of  this  research  was  con- 
ducted at  the  captive  avian  research  facility  at 
Tennessee  Tech  University  during  January- 
April  2004.  I used  14  Mourning  Doves  and 
13  Eurasian  Collared-Doves  captured  during 
July— September  2003  in  Coffee  County,  Ten- 
nessee. Methods  generally  followed  those 
used  previously,  except  as  noted  below.  Indi- 
viduals of  each  species  were  tested  in  a se- 
quential manner  in  2.4  X 1.8  X 1.8-m  pens. 
Corn  and  sunflower  seeds  ( n = 200)  were  pre- 
sented to  each  individual  in  separate  5.5-  and 
4-hr  trials,  respectively;  trials  involving  corn 
were  longer  due  to  the  slower  consumption  of 
corn  by  both  species.  Trial  order  (i.e.,  seed 
type)  was  determined  randomly  for  each  dove, 
so  that  approximately  half  the  individuals  re- 
ceived corn  first,  and  half  received  sunflower 
first.  The  two  trials  for  each  dove  were  inter- 


ceded by  24-hr  acclimation  and  fasting  peri- 
ods. Seeds  were  scattered  in  trays  without  top- 
soil. Pre-  and  post-trial  seed  sampling  and 
measurements  were  conducted  as  in  the  pre- 
vious trials,  and  similar  analyses  were  con- 
ducted separately  for  each  species,  with  indi- 
viduals serving  as  replicates.  Additionally,  I 
used  two-sample  f-tests  to  compare  the  aver- 
age length,  breadth,  and  thickness  of  seeds,  by 
seed  type,  that  each  dove  species  consumed. 
Trials  in  which  doves  consumed  <25  seeds 
were  omitted  from  analyses  because  I sus- 
pected that  dove  foraging  during  these  trials 
was  too  limited  to  allow  for  sufficient  discrim- 
ination among  available  seeds.  I did  not  con- 
duct statistical  comparisons  of  corn  and  sun- 
flower consumption  because  trial  length  dif- 
fered between  seed  types.  I conducted  all 
analyses  using  SAS/STAT  (SAS  Institute,  Inc. 
1990)  and  set  a = 0.05.  All  means  are  pre- 
sented ± SE. 

RESULTS 

In  the  first  phase  (2000  Alabama  study), 
consumption  of  seeds  during  trials  varied 
among  seed  types  (F3 19  = 11.9,  P < 0.001). 
Doves  consumed  more  milo,  wheat,  and  sun- 
flower (161.2  ± 24.3,  142.2  ± 20.4,  and  103 
± 23.4  seeds,  respectively),  than  corn  (8.4  ± 
3.0  seeds).  Doves  ate  nearly  all  (198  + ) of  the 
200  milo  seeds  offered  in  three  of  five  trials, 
and  consumed  almost  no  (<2)  corn  seeds  in 
two  of  five  trials,  so  these  seed  types  were 
excluded  from  further  analyses.  One  trial,  in 
which  doves  ate  198  wheat  seeds,  was  omitted 
from  analysis  of  wheat  size  consumption. 
Wheat  and  sunflower  seeds  consumed  were 
average  in  size  with  respect  to  all  dimensions 
(Table  1). 

In  the  second  phase  (2004  Tennessee 
study).  Mourning  Doves  consumed  24.9  ±4.1 
corn  and  50.1  ± 10.3  sunflower  seeds,  and 
Eurasian  Collared-Doves  consumed  an  aver- 
age of  47.0  ± 4.9  corn  and  52.5  ± 8.7  sun- 
flower seeds  during  trials.  Two  collared-doves 
and  eight  Mourning  Doves  ate  <25  corn 
seeds,  and  one  collared-dove  and  four  Mourn- 
ing Doves  ate  <25  sunflower  seeds;  these 
doves  were  not  included  in  seed-size  analyses. 
Both  Mourning  Doves  and  Eurasian  Collared- 
Doves  selected  smaller-than-average  corn 
seeds  with  respect  to  length  and  breadth  (Ta- 
ble 1).  Mourning  Doves  consumed  corn  seeds 


Hayslette  • SEED-SIZE  SELECTION  IN  DOVES 


67 


TABLE  1.  Measurements  (mm)  of  seeds  initially  available  to,  and  consumed  by.  Mourning  Doves  (MODO; 
2000  and  2004)  and  Eurasian  Collared-Doves  (EUCD;  2004)  in  seed-size  selection  trials  on  captive  birds  in 
Tennessee  and  Alabama. 


Initially  available  Consumed 


Year  Food  Dimension 

Species 

/ta 

Mean 

SE 

Mean 

SE 

2000  Wheat  Length 

MODO 

4 

6.3 

0.1 

6.4 

0.2 

-0.9 

Breadth 

MODO 

4 

3.1 

0.0 

3.1 

0.0 

0.3 

Thickness 

MODO 

4 

2.6 

0.0 

2.5 

0.0 

1.3 

Sunflower  Length 

MODO 

5 

9.8 

0.1 

10.3 

0.5 

-1.1 

Breadth 

MODO 

5 

5.2 

0.1 

5.0 

0.2 

2.6 

Thickness 

MODO 

5 

3.1 

0.1 

2.7 

0.3 

2.0 

2004  Corn  Length 

MODO 

6 

12.7 

0.1 

12.3 

0.1 

4.3C 

EUCD 

11 

12.7 

0.1 

12.1 

0.3 

2.7C 

Breadth 

MODO 

6 

8.4 

0.1 

7.9 

0.2 

6.  lc 

EUCD 

11 

8.4 

0.0 

7.6 

0.2 

5.3C 

Thickness 

MODO 

6 

4.5 

0.1 

4.6 

0.3 

-0.3 

EUCD 

1 1 

4.6 

0.1 

5.2 

0.2 

-3.0C 

Sunflower  Length 

MODO 

10 

10.4 

0.1 

10.5 

0.3 

-0.4 

EUCD 

12 

10.2 

0.1 

10.2 

0.4 

0.2 

Breadth 

MODO 

10 

4.9 

0.0 

5.0 

0.3 

-0.1 

EUCD 

12 

4.9 

0.1 

5.8 

0.3 

— 3.0C 

Thickness 

MODO 

10 

2.9 

0.0 

3.0 

0.3 

-0.3 

EUCD 

12 

2.8 

0.0 

3.5 

0.2 

— 3.5C 

a Sample  size  (n)  equals  number  of  three-bird  flocks  in 

2000,  individual  doves  in  2004. 

b r-values  from  paired  r-tests  comparing  sizes  of  seeds 

initially  available 

and  seeds  consumed;  df 

= n - 1. 

c P < 0.05. 

of  average  thickness,  but  collared-doves  se- 

in 

determining  ease 

with 

which 

seeds  are 

lected  corn  seeds  that  were  thicker  than  av- 
erage. Mean  size  of  corn  seeds  consumed  by 
the  two  species  did  not  differ  with  respect  to 
length,  breadth,  or  thickness  (—1.0  < tl5  < 
2.8,  all  P > 0.1 1).  Mourning  Doves  consumed 
sunflower  seeds  that  were  of  average  size  with 
respect  to  all  dimensions,  and  average  length. 


swallowed.  Thickness  seems  important,  how- 
ever, in  species  that  husk  seeds  before  swal- 
lowing. In  a similar  study  of  Eurasian  Bull- 
finches ( Pyrrhula  pyrrhula),  sunflower  seeds 
consumed  were  4.6%  thinner,  3.1%  narrower, 
and  2.1%  shorter  than  average  (Greig-Smith 
and  Crocker  1986).  Willson  (1972)  concluded 


breadth,  and  thickness  of  sunflower  seeds  con- 


that  Purple  Finches  ( Carpodacus  purpureus ) 


sumed  were  similar  (within  0.3  mm)  to  aver- 
age sizes  consumed  by  Mourning  Doves  in 
the  2000  study.  Eurasian  Collared-Doves  con- 
sumed sunflower  seeds  of  average  length,  but 
selected  seeds  of  larger  than  average  breadth 
and  thickness.  Mean  size  of  sunflower  seeds 
consumed  did  not  differ  (—0.7  < t20  ^ 1.6,  all 
P ^ 0.12)  between  dove  species  with  respect 
to  any  dimension. 


selected  seeds  based  on  thickness  rather  than 
length.  The  sunflower  seeds  we  used  seemed 
to  be  below  the  size  threshold  suggested  by 
De  Nagy  Koves  Hrabar  and  Perrin  (2002), 
above  which  size  becomes  a factor  affecting 
dove  seed  handling  and  selection.  We  used 
seeds  of  black  oil  sunflower,  a relatively 
small-seeded  sunflower  variety  commonly 
used  in  food  plantings  for  Mourning  Doves  in 


DISCUSSION 

Results  suggest  that  seed  length  and  breadth 
limit  Mourning  Dove  consumption  of  corn; 
seed  thickness  seems  less  important  in  selec- 
tion. If  seeds  such  as  corn  are  oriented  length- 
wise as  they  are  swallowed,  it  seems  logical 
that  the  smaller  of  the  two  cross-sectional  di- 
mensions (thickness)  would  be  less  important 


the  southeastern  U.S.  These  seeds  were  small- 
er than  sunflower  seeds  used  in  previous  stud- 
ies (Hespenheide  1966,  Willson  1972,  Greig- 
Smith  and  Crocker  1986,  Diaz  1990).  As  pre- 
dicted, consumption  of  smaller  seeds,  such  as 
milo  and  wheat,  seemed  unaffected  by  seed 
size.  Preference  by  Mourning  Doves  for  milo 
in  this  study  agrees  with  preference  patterns 
documented  elsewhere  (Poling  and  Hayslette 


68 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


2006),  and  suggests  that  seed  preference  in 
this  species  may  be  based,  at  least  in  part,  on 
seed  size  and/or  shape.  Selection  patterns  fa- 
voring small  seeds  have  been  documented  in 
several  studies  of  seed-husking  species  (Hes- 
penheide  1966;  Willson  1971,  1972;  Keating 
et  al.  1992),  but  De  Nagy  Koves  Hrabar  and 
Perrin  (2002)  suggested  that  dove  food  pref- 
erences are  influenced  more  by  seed  shape 
than  size;  round  seeds  are  handled  more  rap- 
idly than,  and  preferred  to,  elongate  seeds.  Be- 
cause milo  was  both  the  smallest  and  roundest 
of  the  seeds  we  tested,  it  is  impossible  to  tell 
which  of  these  seed  characteristics  actually  in- 
fluenced seed  selection. 

As  with  Mourning  Doves,  consumption  of 
corn  by  Eurasian  Collared-Doves  seems  lim- 
ited by  seed  length  and  breadth.  In  contrast  to 
Mourning  Doves,  however,  Eurasian  Collared- 
Doves  were  influenced  by  corn  seed  thickness, 
choosing  thicker-than-average  seeds.  If  seed 
thickness  is  relatively  unimportant  in  deter- 
mining handling  and/or  swallowing  efficiency, 
as  postulated  earlier  for  Mourning  Doves,  per- 
haps the  Eurasian  Collared-Dove’s  selection 
of  thicker  seeds  increases  foraging  efficiency 
by  increasing  nutrient  intake  (benefit)  per  unit 
handling  time  (cost)  (Stephens  and  Krebs 
1986).  Likewise,  selection  of  broader-  and 
thicker-than-average  sunflower  seeds  may  in- 
crease foraging  profitability  of  Eurasian  Col- 
lared-Doves. Selection  for  large  foods,  pre- 
sumably to  maximize  foraging  profitability, 
has  been  reported  in  other  seed-eating  birds 
(Myton  and  Ficken  1967,  Ramos  1996). 

Selection  of  larger  than  average  corn  and 
sunflower  seeds  with  respect  to  certain  dimen- 
sions by  Eurasian  Collared-Doves,  but  not  by 
Mourning  Doves  (paired  Mests),  suggests  that 
collared-doves  select  thicker  corn  seeds,  and 
broader  and  thicker  sunflower  seeds,  than  do 
Mourning  Doves.  Although  direct  between- 
species  comparisons  of  average  seed  sizes 
consumed  failed  to  reveal  any  such  differenc- 
es, the  average  corn  seed  consumed  by  col- 
lared-doves was  0.5  mm  thicker  than  that  eat- 
en by  Mourning  Doves,  and  the  average  sun- 
flower seed  consumed  by  collared-doves  was 
0.8  mm  broader  and  0.5  mm  thicker.  Selection 
for  larger  seeds  by  Eurasian  Collared-Doves 
than  by  Mourning  Doves  likely  is  related  to 
differences  in  their  bill  sizes  (Hespenheide 
1966,  Myton  and  Ficken  1967,  Willson  1972). 


Previous  authors  have  expressed  concern 
about  how  the  recent  Eurasian  Collared-Dove 
invasion  of  North  America  may  affect  native 
species,  particularly  ecologically  similar  spe- 
cies, such  as  the  Mourning  Dove,  with  which 
Eurasian  Collared-Doves  may  compete  (Rom- 
agosa  and  McEneaney  1999,  Romagosa  and 
Labisky  2000,  Romagosa  2002).  A recent 
study  of  food-selection  patterns  of  these  two 
dove  species  revealed  a high  degree  (95%)  of 
dietary  overlap  between  them  (Poling  and 
Hayslette  2006),  although  this  was  based  on 
relative  consumption  of  different  seed  types  in 
a cafeteria  experiment,  rather  than  on  size 
preferences  within  seed  types.  If  genuine,  the 
Eurasian  Collared-Dove’s  preferences  for 
seeds  that  are  broader  and/or  thicker  than 
those  selected  by  Mourning  Doves  may  result 
in  differential  exploitation  of  larger  seeds, 
such  as  corn  and  sunflower,  and  concomitant 
mitigation  of  foraging  competition  between 
them.  Partitioning  of  food  resources  among 
sympatric  species  based  on  seed  size  has  been 
documented  in  a number  of  avian  granivore 
communities  (Grant  1986,  Faaborg  1988, 
Ricklefs  2001).  Similarly,  limited  foraging 
competition  may  mean  greater  potential  for 
coexistence  of  Mourning  Doves  and  Eurasian 
Collared-Doves  than  previously  believed. 

ACKNOWLEDGMENTS 

I thank  M.  P.  Cook,  A.  Goodman,  T.  D.  Poling,  C. 
Smith,  D.  Smith,  C.  Spiller,  J.  Turner,  and  T.  M.  Un- 
deutsch  for  assistance  with  data  collection.  Funding 
and  other  support  for  this  project  was  provided  by  the 
Alabama  Department  of  Conservation  and  Natural  Re- 
sources (Division  of  Wildlife  and  Freshwater  Fisher- 
ies), Auburn  University  School  of  Forestry  and  Wild- 
life Sciences,  Tennessee  Tech  University  Department 
of  Biology,  and  Tennessee  Wildlife  Resources  Agency. 
This  research  was  approved  by,  and  conducted  in  ac- 
cordance with,  the  Auburn  University  Institutional  An- 
imal Care  and  Use  Committee  (protocol  review  num- 
ber 0009-R-0786),  and  Tennessee  Tech  University  In- 
stitutional Committee  for  the  Care  and  Use  of  Lab  An- 
imals in  Experimentation  (protocol  review  number 
1 [02-03]).  I thank  M.  R.  Perrin  and  two  anonymous 
referees  for  their  comments  on  an  earlier  draft  of  the 
manuscript. 

LITERATURE  CITED 

Abbott,  I.,  L.  K.  Abbott,  and  P.  R.  Grant.  1975. 
Seed  selection  and  handling  ability  of  four  species 
of  Darwin’s  Finches.  Condor  77:332-335. 

De  Nagy  Koves  Hrabar,  H.  and  M.  Perrin.  2002. 


Hayslette  • SEED-SIZE  SELECTION  IN  DOVES 


69 


The  effect  of  bill  structure  on  seed  selection  by 
granivorous  birds.  African  Zoology  37:67-80. 

Diaz,  M.  1990.  Interspecific  patterns  of  seed  selection 
among  granivorous  passerines:  effects  of  seed  nu- 
tritive value  and  bird  morphology.  Ibis  132:467- 
476. 

Faaborg,  J.  1988.  Ornithology:  an  ecological  ap- 
proach. Prentice-Hall,  Englewood  Cliffs,  New  Jer- 
sey. 

Goldstein,  G.  B.  and  M.  C.  Baker.  1984.  Seed  se- 
lection by  juncos.  Wilson  Bulletin  96:458-463. 

Grant,  P.  R.  1986.  Ecology  and  evolution  of  Darwin’s 
finches.  Princeton  University  Press,  Princeton, 
New  Jersey. 

Greig-Smith,  P.  W.  and  D.  R.  Crocker.  1986.  Mech- 
anisms of  food  size  selection  by  Bullfinches  ( Pyr - 
rhula  pyrrhula  L.)  feeding  on  sunflower  seeds. 
Animal  Behaviour  34:843-859. 

Greig-Smith,  P.  W.  and  M.  F.  Wilson.  1985.  Influence 
of  seed  size,  nutrient  composition  and  phenolic 
content  on  the  preferences  of  bullfinches  feeding 
in  ash  trees.  Oikos  44:47-54. 

Hayslette,  S.  E.  and  R.  E.  Mirarchi.  2001.  Patterns 
of  food  preferences  in  Mourning  Doves.  Journal 
of  Wildlife  Management  65:816-827. 

Hespenheide,  H.  A.  1966.  The  selection  of  seed  size 
by  finches.  Wilson  Bulletin  78:191-197. 

Keating,  J.  E,  R.  J.  Robel,  A.  W.  Adams,  K.  C.  Behn- 
ke,  and  K.  E.  Kemp.  1992.  Role  of  handling  time 
in  selection  of  extruded  food  morsels  by  two  gra- 
nivorous bird  species.  Auk  109:863-868. 

LeBlanc,  D.  K.  and  D.  L.  Otis.  1998.  Preferences  by 
Mourning  Doves  and  two  granivorous  songbirds 
for  selected  seeds.  Proceedings  of  the  Annual 
Conference  of  the  Southeastern  Association  of 
Fish  and  Wildlife  Agencies  52:324-335. 

Lewis,  J.  C.  1993.  Foods  and  feeding  ecology.  Pages 
181-204  in  Ecology  and  management  of  the 
Mourning  Dove  (T.  S.  Baskett,  M.  W.  Sayre,  R. 
E.  Tomlinson,  and  R.  E.  Mirarchi,  Eds.).  Stack- 
pole  Books,  Harrisburg,  Pennsylvania. 

Mirarchi,  R.  E.  1993.  Care  and  propagation  of  captive 
Mourning  Doves.  Pages  409-428  in  Ecology  and 
management  of  the  Mourning  Dove  (T.  S.  Baskett, 


M.  W.  Sayre,  R.  E.  Tomlinson,  and  R.  E.  Mirarchi, 
Eds.).  Stackpole  Books,  Harrisburg,  Pennsylvania. 

Mirarchi,  R.  E.  and  T.  S.  Baskett.  1994.  Mourning 
Dove  ( Zenaida  macroura).  The  Birds  of  North 
America,  no.  1 17. 

Myton,  B.  A.  and  R.  W.  Ficken.  1967.  Seed-size  pref- 
erence in  chickadees  and  titmice  in  relation  to  am- 
bient temperature.  Wilson  Bulletin  79:3 1 9—32 1 . 

Poling,  T.  D.  and  S.  E.  Hayslette.  2006.  Dietary 
overlap  and  foraging  competition  between  Mourn- 
ing Doves  and  Eurasian  Collared-Doves.  Journal 
of  Wildlife  Management  70:  In  press. 

Ramos,  J.  A.  1996.  The  influence  of  size,  shape,  and 
phenolic  content  on  the  selection  of  winter  foods 
by  the  Azores  Bullfinch  ( Pyrrhula  murina).  Jour- 
nal of  Zoology  238:415-433. 

Ricklefs,  R.  E.  2001.  The  economy  of  nature,  5th  ed. 
W.  H.  Freeman  and  Company,  New  York. 

Romagosa,  C.  M.  2002.  Eurasian  Collared-Dove 
( Streptopelia  decaocto ).  The  Birds  of  North 
America,  no.  629. 

Romagosa,  C.  M.  and  R.  F.  Labisky.  2000.  Establish- 
ment and  dispersal  of  the  Eurasian  Collared-Dove 
in  Florida.  Journal  of  Field  Ornithology  71:159- 
165. 

Romagosa,  C.  M.  and  T.  McEneaney.  1999.  Eurasian 
Collared-Dove  in  North  America  and  the  Carib- 
bean. North  American  Birds  53:348-353. 

SAS  Institute,  Inc.  1990.  SAS/STAT  guide  for  per- 
sonal computers,  ver.  6.  SAS  Institute,  Inc.,  Cary, 
North  Carolina. 

Stephens,  D.  W.  and  J.  R.  Krebs.  1986.  Foraging  the- 
ory. Princeton  University  Press,  Princeton,  New 
Jersey. 

van  der  Meij,  M.  A.  A.  and  R.  G.  Bout.  2000.  Seed 
selection  in  the  Java  Sparrow  ( Padda  oryzivora ): 
preference  and  mechanical  constraint.  Canadian 
Journal  of  Zoology  78:1668-1673. 

Willson,  M.  F.  1971.  Seed  selection  in  some  North 
American  finches.  Condor  73:415-429. 

Willson,  M.  F.  1972.  Seed  size  preference  in  finches. 
Wilson  Bulletin  84:449-455. 

Willson,  M.  F.  and  J.  C.  Harmeson.  1973.  Seed  pref- 
erences and  digestive  efficiency  of  Cardinals  and 
Song  Sparrows.  Condor  75:225-234. 


The  Wilson  Journal  of  Ornithology  1 18(1):70— 74,  2006 


LOW  NESTING  SUCCESS  OF  LOGGERHEAD  SHRIKES  IN  AN 
AGRICULTURAL  LANDSCAPE 

JEFFERY  W.  WALK,124  ERIC  L.  KERSHNER,13  AND  RICHARD  E.  WARNER1 


ABSTRACT. — Southeastern  Illinois  is  dominated  by  cropland,  and  the  remaining  pastures  or  grasslands  are 
marginally  suitable  for  breeding  Loggerhead  Shrikes  ( Lanius  ludovicianus),  owing,  in  part,  to  limited  nest  sites. 
From  1998  through  2000,  we  recorded  poor  nest  success  (26%)  among  shrikes,  although  results  of  earlier  studies 
(1967-1972)  in  this  region  indicated  that  nest  success  was  72  to  80%.  Clutch  size  (5.7  eggs)  and  fledglings/ 
successful  nest  (4.4  young/successful  nest)  were  similar  to  those  reported  in  previous  studies.  During  our  study, 
generalist  mammalian  predators  were  abundant,  and  most  nest  failures  (88%)  were  caused  by  predation.  We 
suggest  that  the  loss  of  grassland  habitat  and  agricultural  intensification  has  resulted  in  reduced  nest  success, 
and  this  may  be  true  in  other  areas  of  the  species’  range  as  well.  Received  20  August  2003,  accepted  23  November 
2005. 


The  Loggerhead  Shrike  {Lanius  ludovici- 
anus) is  of  conservation  interest  throughout  its 
range,  and  has  been  designated  a “Bird  of 
Conservation  Concern”  (U.S.  Fish  and  Wild- 
life Service  2002).  The  range  of  the  species 
has  contracted  greatly  over  the  past  half-cen- 
tury (Cade  and  Woods  1997),  and  Christmas 
Bird  Count  and  Breeding  Bird  Survey  data  in- 
dicate a continent-wide  decrease  in  abun- 
dance. The  sharpest  declines  have  occurred  in 
the  core  of  the  shrike’s  range  in  southern  and 
Gulf  Coast  areas  (Yosef  1996). 

Most  studies  of  Loggerhead  Shrikes  have 
revealed  high  nest  success  (mean  of  56%; 
Yosef  1996,  Esely  and  Bollinger  2001),  sug- 
gesting that  problems  associated  with  winter 
habitat  and  survival  may  be  causes  for  popu- 
lation declines  (Haas  and  Sloane  1989,  Brooks 
and  Temple  1990,  Gawlik  and  Bildstein 
1993).  Based  on  reports  of  high  nest  success 
throughout  the  species’  range,  Maddox  and 
Robinson  (2004)  considered  it  fortuitous  that 
habitat  degradation  had  not  resulted  in  elevat- 
ed rates  of  nest  predation  or  decreased  pro- 
ductivity. Our  observations,  and  the  results  of 
some,  more  recent  studies  (DeGeus  1990, 
Yosef  1994,  Collins  1996,  Esely  and  Bollinger 


1 Dept,  of  Natural  Resources  and  Environmental 
Sciences,  Univ.  of  Illinois,  Urbana,  IL  61801,  USA. 

2 Current  address:  Illinois  Natural  History  Survey, 
One  Natural  Resources  Way,  Springfield,  IL  62702, 
USA. 

3 Current  address:  Inst,  for  Wildlife  Studies,  2515 
Camino  del  Rio  South,  Ste.  339,  San  Diego,  CA 
92108,  USA. 

4 Corresponding  author;  e-mail: 
jwalk@dnrmail.state.il. us 


2001) ,  suggest  there  are  landscapes  and  nest- 
site  contexts  in  which  this  presumption  does 
not  apply.  Our  objectives  were  to  measure 
nest  success  of  Loggerhead  Shrikes  in  a re- 
gion of  intensive  agriculture  with  marginal 
habitat,  and  to  determine  whether  land  use 
near  nests,  or  nest-site  context,  influenced  nest 
fate. 

METHODS 

From  1998  through  2000,  we  monitored 
Loggerhead  Shrike  nests  within  a 125-km2 3 4 
area  of  southern  Jasper  County,  Illinois.  The 
study  was  centered  on  Prairie  Ridge  State  Nat- 
ural Area  (88°  12'  W,  38°  57'  N)  and  included 
most  of  Smallwood  Township  and  adjacent 
portions  of  Wade  and  Fox  townships.  Jasper 
County’s  landscape  is  composed  of  71%  row 
crop  (corn,  Zea  mays ; and  soybeans.  Glycine 
max),  6%  wheat  ( Triticum  aestivum\  most 
double-cropped  to  soybeans  after  harvest),  6% 
rural  grassland  (hay,  pasture,  roadsides,  and 
idle  grass),  13%  woodland,  and  1%  roads,  res- 
idential/urban  areas,  and  small  amounts  of 
open  water  and  other  land  covers  (Illinois  In- 
teragency Landscape  Classification  Project 

2002) .  Our  study  area  differed  from  the  coun- 
ty as  a whole  by  having  greater  row  crop  cov- 
er (>85%)  and  less  woodland  cover  (<5%; 
JWW  unpubl.  data). 

Between  1966  and  2000,  North  American 
Breeding  Bird  Survey  results  suggested  de- 
clining abundance  of  Loggerhead  Shrikes  in 
Illinois  (-4.5%/year)  and  the  Midwest 
(—0.3%/year  in  the  eight  states  of  U.S.  Fish 
and  Wildlife  Service  Region  3;  Sauer  et  al. 
2005).  From  1994  to  1996,  roadside  searches 


70 


Walk  et  al.  • LOW  NESTING  SUCCESS  OF  SHRIKES 


71 


within  49  km2  of  the  center  of  our  1998-2000 
study  area  documented  12  (1996)  to  16  (1994) 
shrike  territories  annually  (roughly  0.25-0.33 
shrike  territories/km2;  JWW  unpubl.  data,  re- 
ported to  the  Illinois  Department  of  Natural 
Resource’s  Natural  Heritage  database).  During 
our  1998-2000  study,  the  densities  of  nesting 
shrikes  were  similar  to,  or  lower  than,  that  re- 
ported during  the  1994  to  1996  roadside  sur- 
veys. 

From  March  through  June,  we  located 
shrike  nests  by  locating  adults  or  food  caches, 
searching  nearby  suitable  nest  sites,  observing 
nest  building,  or  observing  provisioning  of  in- 
cubating females  and  nestlings.  Because  the 
study  area  was  almost  entirely  private  land, 
initial  searches  were  limited  to  roadsides. 
When  shrikes  were  located,  we  often  were 
permitted  to  search  for,  and  monitor,  nests  on 
private  land.  We  checked  nests  every  3-5  days 
until  their  fate  was  determined,  and  we  cal- 
culated nest  and  egg  success  based  on  expo- 
sure days  (Mayfield  1975,  Johnson  1979). 
When  nests  failed,  we  assumed  failure  oc- 
curred at  the  midpoint  between  nest  checks. 

We  recorded  nest  context  and  visually  es- 
timated the  percentage  of  land-use  types  with- 
in 100  m (3.1  ha)  of  nest  sites  (after  Gawlik 
and  Bildstein  1993).  Land-use  categories  were 
(1)  row  crop;  (2)  hay  or  pasture;  (3)  idle  grass- 
land; (4)  woody  vegetation  (forest  and  shrub 
areas  combined);  (5)  small  grains;  (6)  road- 
ways, including  grassy  rights-of-way;  and  (7) 
residential  (yards  and  farmsteads).  We  com- 
pared measurements  at  nest  sites  with  20  ran- 
dom locations,  each  also  100  m in  radius  and 
selected  by  overlaying  a map  of  the  study  area 
with  a numbered  grid.  Because  shrike  nest 
sites  are  limited  to  woody  vegetation,  we  cen- 
tered random  locations  on  the  tree  or  shrub 
closest  to  each  randomly  selected  point.  We 
characterized  the  context  of  nest  sites  and  ran- 
dom trees  as  follows:  fencerows,  farmsteads/ 
yards,  watercourses,  woodland  edges,  or  iso- 
lated trees  (>20  m from  another  tree).  Be- 
cause the  proportions  of  land  uses  were  not 
normally  distributed  across  the  study  area,  we 
used  Mann- Whitney  U- tests  to  compare  land 
use  between  random  sites  and  nest  sites,  and 
between  successful  and  depredated  nests. 
Contexts  of  nest  trees  and  random  trees,  and 
trees  with  successful  and  depredated  nests, 
were  compared  using  a chi-square  test.  Unless 


otherwise  noted,  values  are  reported  as  means 
± SE. 

RESULTS 

We  monitored  34  shrike  nests  from  1998 
through  2000.  Ten  nests  (29%)  fledged  >1 
young,  21  nests  (62%)  were  depredated,  2 
(6%)  were  abandoned  during  egg  laying,  and 
1 (3%)  was  dislodged  from  a tree  during  a 
thunderstorm.  Nest  failures  attributed  to  pre- 
dation included  6 empty  nests  with  no  evi- 
dence of  the  predator,  4 tilted  or  compressed 
nests,  and  1 1 highly  disturbed  nests  (lining  re- 
moved, nest  shredded,  or  completely  dis- 
lodged). At  one  depredated  nest  on  a farm- 
stead, the  cooperators  reported  to  us  that  their 
domestic  cat  ( Felis  cattus ) had  killed  one  of 
the  adults.  Shrike  nests  appeared  to  be  more 
vulnerable  to  predation  near  hatching.  Of  22 
nests  surviving  to  at  least  the  14th  day  of  in- 
cubation, 8 were  eventually  lost  to  predators 
before  the  nestlings  were  4 days  old. 

From  the  beginning  of  egg  laying  to  fledg- 
ing, egg  success  was  20.5%  and  nest  success 
was  25.6%  (95%  Cl  = 19.4-33.8%).  Clutch 
size  was  5.7  ± 0.2  eggs  ( n = 30  nests).  In 
nests  that  survived  until  at  least  the  second 
nest  check  after  hatching,  87.5  ± 3.6%  of 
eggs  hatched  successfully  (n  = 10  nests).  Al- 
though 4.4  ± 0.4  young  fledged  per  successful 
nest,  only  1.3  ± 0.4  young  fledged  per  nest 
attempt. 

Land  use  within  100  m of  shrike  nests  in- 
cluded more  hay  and  pasture  than  randomly 
located  points  (Mann-Whitney  U = 241.0;  P 
= 0.049),  but  no  other  variable  that  we  mea- 
sured differed  between  nest  and  random  lo- 
cations (Table  1).  Land  use  did  not  differ  sig- 
nificantly between  successful  and  depredated 
nest  sites.  The  majority  of  the  nests  we  mon- 
itored (87%)  were  located  in  small  (<3  ha) 
pastures  (including  the  enclosing  fences)  or 
yards/farmsteads,  and  were  within  50  m of 
county  roads.  Although  shrike  nests  were 
placed  in  fencerows  more  frequently  than  ex- 
pected (x2  = 25.69,  df  = 4,  P < 0.001),  nests 
in  fencerows  were  more  likely  to  be  depre- 
dated (x2  = 10.94,  df  = 3,  P = 0.012;  Table 
2).  Daily  nest  survival  was  0.957  ± 0.012  in 
fencerows  and  0.973  ± 0.013  in  yards/farm- 
steads. 


72 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


TABLE  1.  Percent  land  use  (±  SE)  within  100  m (3.1  ha)  of  random  trees  and  Loggerhead  Shrike  nest 
locations  in  Jasper  County,  Illinois,  1998-2000. 

Land  use 

Random  sites 
(n  = 20) 

Nest  sites 

All 

(«  = 33) 

Successful 
(n  = 10) 

Depredated 
C n = 21) 

Row  crops 

58.6  ± 6.9 

51.0  ± 4.9 

51.3  ± 9.2 

54.4  ± 6.1 

Hay  and  pasture 

7.5  ± 4.7 

19.4  ± 4.9a 

18.5  ± 9.5 

19.3  ± 5.9 

Idle  grassland 

10.5  ± 6.0 

5.9  ± 2.6 

6.0  ± 5.1 

6.4  ± 3.5 

Woodland 

7.2  ± 2.9 

2.6  ± 0.9 

3.0  ± 1.9 

1.8  ± 0.5 

Small  grains 

2.8  ± 1.6 

3.2  ± 1.6 

0.0  ± 0.0 

2.9  ± 2.0 

Roadway 

5.8  ± 1.4 

9.5  ± 1.3 

7.5  ± 2.0 

10.2  ± 1.8 

Farmstead,  yard 

7.3  ± 2.2 

8.8  ± 2.4 

13.0  ± 5.4 

7.6  ± 2.8 

a Significantly  more  hay  and  pasture  at  nest  sites  than  random  sites  (Mann- Whitney  U = 241.0,  P = 0.049).  There  were  no  other  significant  (all  P > 
0.08)  differences  in  land  use  between  random  and  nest  sites  or  between  successful  and  depredated  nests. 


DISCUSSION 

Reproductive  success  of  Loggerhead 
Shrikes  in  this  agricultural  landscape  (25.6% 
nest  success,  20.5%  egg  success)  is  among  the 
lowest  reported  for  the  species.  Graber  et  al. 
(1973)  reported  80%  nest  success  in  south- 
eastern Illinois  in  1967  and,  in  Jasper  and 
nearby  counties,  Anderson  and  Duzan  (1978) 
observed  72%  nest  success  in  1971-1972.  By 
1991-1992,  Collins  (1996)  found  that  the  pro- 
portion of  fledglings  to  number  of  eggs  laid 
had  dropped  to  25%  in  southern  Illinois.  Our 
methods  for  measuring  nest  success  (Mayfield 
1975)  differ  from  that  used  by  Collins  (1996), 
but  the  results  are  similar  and  suggest  a sub- 
stantial reduction  in  shrike  nesting  success  in 
the  region. 

Clutch  size  (5.7  eggs)  in  our  study  was  sim- 
ilar to  those  of  recent  Midwestern  studies 
(5. 3-5. 7 eggs;  Burton  1990,  DeGeus  1990, 
Collins  1996,  Esely  and  Bollinger  2001).  Our 
measure  of  4.4  fledglings  per  successful  nest 


TABLE  2.  Proportions  of  Loggerhead  Shrike  nests 
and  random  tree  locations  within  various  land-use  con- 
texts in  southeastern  Illinois,  1998-2000. 


Nest  sites 


Context 

Random 

sites 

(n  = 20) 

All 

(n  = 34) 

Successful 
(n  = 10) 

Depredated 
(n  = 21) 

Yard,  farmstead 

0.35 

0.24 

0.40 

0.14 

Fencerow 

0.20 

0.53a 

0.30 

0.62a 

Watercourse 

0.15 

0.15 

0.20 

0.14 

Woodland  edge 

0.15 

0.00 

0.00 

0.00 

Isolated  tree 

0.15 

0.09 

0.10 

0.10 

a Loggerhead  Shrike  nests  were  placed  in  fencerows  more  frequently  than 
expected  (x2  = 25.69,  df  = 4,  P < 0.001),  and  nests  in  fencerows  were 
more  likely  to  be  depredated  (x2  = 10.94,  df  = 3,  P = 0.012). 


was  the  same  as  the  mean  reported  from  14 
studies  compiled  by  Esely  and  Bollinger 
(2001).  Toxicological  problems  affecting  egg 
viability  are  not  implicated;  an  analysis  of  or- 
ganochlorine  residues  in  shrike  eggs,  includ- 
ing several  eggs  collected  in  our  study  area  in 
1995  and  1996,  indicated  that  DDE  levels  had 
decreased  79%  in  the  region  since  the  early 
1970s  (Anderson  and  Duzan  1978,  Herkert 
2004).  Roughly  88%  of  fully  incubated  eggs 
hatched  successfully  during  our  study. 

Intensified  agricultural  land  use  and  a con- 
comitant increase  in  the  abundance  of  gener- 
alist predators  are  likely  the  causes  for  a de- 
crease in  shrike  nesting  success  in  this  land- 
scape. Predation  was  implicated  in  88%  of  the 
nest  failures  and  in  62%  of  all  nesting  at- 
tempts in  our  study.  From  1970  to  2000,  acre- 
age of  row  crops  increased  by  26%  in  Jasper 
County,  while  hay  acreage  decreased  by  85% 
and  pasture  decreased  by  about  47%  (Illinois 
Department  of  Agriculture  1971,  National  Ag- 
ricultural Statistics  Service  undated).  Potential 
nest  sites  in  this  landscape  are  few  and  most 
occur  in  linear  habitats,  often  limiting  shrikes 
and  other  birds  to  nesting  situations  in  prey- 
rich  corridors  that  are  easily  searched  by  pred- 
ators (Major  et  al.  1999).  Furthermore,  human 
structures  and  agricultural  waste  may  subsi- 
dize populations  of  generalist  predators 
(Warner  1985,  Dijak  and  Thompson  2000). 
Road-kill  surveys  conducted  by  the  Illinois 
Department  of  Natural  Resources  from  1975 
through  1998  documented  increases  in  the 
abundance  of  raccoons  ( Procyon  lotor)  and 
opossums  ( Didelphis  virginiana ) of  more  than 
250%  and  100%,  respectively  (Gehrt  et  al. 
2002;  S.  D.  Gehrt  pers.  comm.). 


Walk  et  al.  • LOW  NESTING  SUCCESS  OF  SHRIKES 


73 


Successful  shrike  nests  were  more  likely  to 
be  in  yards,  whereas  depredated  nests  were 
more  likely  to  be  in  fencerows.  Yards  are  not 
benign  breeding  habitat,  however.  Gawlik  and 
Bildstein  (1990)  recognized  the  potential 
threat  of  predation  by  domestic  cats  on  adult 
shrikes  and  their  young,  and  during  our  study, 
a cat  killed  at  least  one  yard-nesting  adult 
shrike.  Row  crops  were  the  most  common 
land  use  near  nests,  reflecting  the  ubiquitous- 
ness of  cropland  in  the  landscape,  but  shrikes 
preferentially  selected  nest  sites  in  or  near 
pastures  (Table  1).  Most  pastures  in  our  study 
area  were  small  (<3  ha)  horse  pastures  adja- 
cent to  residences,  which  were  located  pri- 
marily along  county  roads.  Shrikes  elsewhere 
also  frequently  nest  near  roadsides,  and  the 
shorter  grasses  and  utility  lines  along  road- 
sides may  be  superior  hunting  areas  (Luuk- 
konen  1987,  DeGeus  1990,  Gawlik  and  Bild- 
stein 1990,  Smith  and  Kruse  1992).  The  vis- 
ibility of  shrikes  on  utility  lines  (i.e.,  along 
roads)  may  have  contributed  to  high  represen- 
tation of  pasture  habitat  near  the  shrike  nests 
that  we  monitored,  but  little  suitable  nesting 
habitat  was  available  to  Loggerhead  Shrikes 
away  from  roadways  in  our  study  area.  Al- 
though widely  used,  roadways  and  other  linear 
habitats  may  be  ecological  sinks  for  nesting 
shrikes,  given  the  low  nesting  success  we 
found  and  that  has  been  reported  elsewhere. 
Yosef  (1994)  found  lower  nest  success  among 
shrikes  nesting  in  fencerows  (36%)  than  for 
those  nesting  within  pastures  (54%)  in  Flori- 
da. Likewise,  roadside-nesting  Loggerhead 
Shrikes  had  39%  nest  success  in  Missouri 
(compared  with  76%  success  for  interior 
nests;  Esely  and  Bollinger  2001),  35%  nest 
success  in  Iowa  (DeGeus  1990),  and  31%  egg 
success  in  Indiana  (Burton  1990). 

Pasture  and  hay  acreage  has  declined  by 
50%  in  the  Midwest  over  the  past  50  years 
(Herkert  et  al.  1996).  At  the  same  time, 
shrikes  have  declined  or  disappeared  from 
much  of  the  region  (Cade  and  Woods  1997, 
Sauer  et  al.  2005).  Our  results — documenting 
a sharp  decrease  in  nesting  success  in  a county 
with  recent  land-use  changes  typical  of  the 
Midwest — suggest  that  nesting  success,  often 
thought  to  be  relatively  high  for  shrikes  (Yos- 
ef 1996,  Maddox  and  Robinson  2004),  will 
need  to  be  re-evaluated  as  land-use  changes 


result  in  a less  optimal  environment  for  Log- 
gerhead Shrikes. 

ACKNOWLEDGMENTS 

We  thank  the  many  private  landowners  of  Jasper 
County,  Illinois,  who  permitted  us  to  access  their  prop- 
erties to  search  for  and  monitor  nests.  This  project  was 
supported  by  an  S.  Charles  Kendeigh  Award  from  the 
Champaign  County  Audubon  Society,  and  the  Univer- 
sity of  Illinois.  R.  Yosef,  C.  P.  Woods,  and  anonymous 
reviewers  suggested  several  improvements  to  the  man- 
uscript. 

LITERATURE  CITED 

Anderson,  W.  L.  and  R.  E.  Duzan.  1978.  DDE  resi- 
dues and  eggshell  thinning  in  Loggerhead  Shrikes. 
Wilson  Bulletin  90:215-220. 

Brooks,  B.  L.  and  S.  A.  Temple.  1990.  Dynamics  of 
a Loggerhead  Shrike  population  in  Minnesota. 
Wilson  Bulletin  102:441-450. 

Burton,  K.  M.  1990.  An  investigation  of  population 
status  and  breeding  biology  of  the  Loggerhead 
Shrike  ( Lanius  ludovicianus)  in  Indiana.  M.A.  the- 
sis, Indiana  University,  Bloomington. 

Cade,  T.  J.  and  C.  P.  Woods.  1997.  Changes  in  dis- 
tribution and  abundance  of  the  Loggerhead 
Shrike.  Conservation  Biology  11:21—31. 

Collins,  J.  A.  1996.  Breeding  and  wintering  ecology 
of  the  Loggerhead  Shrike  in  southern  Illinois. 
M.Sc.  thesis.  Southern  Illinois  University,  Carbon- 
dale. 

DeGeus,  D.  W.  1990.  Productivity  and  habitat  prefer- 
ences of  Loggerhead  Shrikes  inhabiting  roadsides 
in  a Midwestern  agroenvironment.  M.Sc.  thesis, 
Iowa  State  University,  Ames. 

Dijak,  W.  D.  and  F.  R.  Thompson,  III.  2000.  Land- 
scape and  edge  effects  on  the  distribution  of  mam- 
malian predators  in  Missouri.  Journal  of  Wildlife 
Management  64:209-216. 

Esely,  J.  D.  and  E.  K.  Bollinger.  2001.  Habitat  se- 
lection and  reproductive  success  of  Loggerhead 
Shrikes  in  northwest  Missouri:  a hierarchical  ap- 
proach. Wilson  Bulletin  113:290-296. 

Gawlik,  D.  E.  and  K.  L.  Bildstein.  1990.  Reproduc- 
tive success  and  nesting  habitat  of  Loggerhead 
Shrikes  in  north-central  South  Carolina.  Wilson 
Bulletin  102:37-48. 

Gawlik,  D.  E.  and  K.  L.  Bildstein.  1993.  Seasonal 
habitat  use  and  abundance  of  Loggerhead  Shrikes 
in  South  Carolina.  Journal  of  Wildlife  Manage- 
ment 57:352-357. 

Gehrt,  S.  D.,  G.  F.  Hubert,  Jr.,  and  J.  A.  Ellis.  2002. 
Long-term  population  trends  of  raccoons  in  Illi- 
nois. Wildlife  Society  Bulletin  30:457-463. 
Graber,  R.  R.,  J.  W.  Graber,  and  E.  L.  Kirk.  1973. 
Illinois  Birds:  Laniidae.  Illinois  Natural  History 
Survey  Biological  Notes,  no.  83. 

Haas,  C.  A.  and  S.  A.  Sloane.  1989.  Low  return  rates 
of  migratory  Loggerhead  Shrikes:  winter  mortal- 


74 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


ity  or  low  site  fidelity?  Wilson  Bulletin  101:458- 
460. 

Herkert,  J.  R.  2004.  Organochlorine  pesticides  are  not 
implicated  in  the  decline  of  the  Loggerhead 
Shrike.  Condor  106:702-705. 

Herkert,  J.  R.,  D.  W.  Sample,  and  R.  E.  Warner. 
1996.  Management  of  Midwestern  grassland  land- 
scapes for  the  conservation  of  migratory  birds. 
Pages  89-116  in  Management  of  Midwestern 
landscapes  for  the  conservation  of  migratory  birds 
(F.  R.  Thompson,  III,  Ed.).  General  Technical  Re- 
port NC-187,  USDA  Forest  Service,  North  Cen- 
tral Forest  Experiment  Station,  St.  Paul,  Minne- 
sota. 

Illinois  Department  of  Agriculture.  1971.  Illinois 
annual  farm  census.  Illinois  Department  of  Agri- 
culture, Division  of  Agricultural  Statistics,  Spring- 
field. 

Illinois  Interagency  Landscape  Classification  Pro- 
ject. 2002.  Land  cover  of  Illinois  statistical  sum- 
mary, 1999-2000.  USDA  National  Agricultural 
Statistics  Services,  Washington,  D.C.;  Illinois  De- 
partment of  Agriculture,  Springfield;  and  Illinois 
Department  of  Natural  Resources,  Springfield. 
www.agr.state.il.us/gis/stats/landcover/  (accessed 
1 1 February  2005). 

Johnson,  D.  H.  1979.  Estimating  nest  success:  the 
Mayfield  method  and  an  alternative.  Auk  96:651- 
661. 

Luukkonen,  D.  R.  1987.  Status  and  breeding  ecology 
of  the  Loggerhead  Shrike  in  Virginia.  M.Sc.  the- 
sis, Virginia  Polytechnic  Institute  and  State  Uni- 
versity, Blacksburg. 

Maddox,  J.  D.  and  S.  K.  Robinson.  2004.  Conserva- 
tion assessment  for  Loggerhead  Shrike  ( Lanius  lu- 


dovicianus).  USDA  Forest  Service,  Eastern  Re- 
gion, Milwaukee,  Wisconsin. 

Major,  R.  E.,  F.  J.  Christie,  G.  Gowing,  and  T.  J. 
Iverson.  1999.  Elevated  rates  of  predation  on  ar- 
tificial nests  in  linear  strips  of  habitat.  Journal  of 
Field  Ornithology  70:351-364. 

Mayfield,  H.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

National  Agricultural  Statistics  Services.  Undat- 
ed. Quick  stats:  agricultural  statistics  database. 
USDA  National  Agricultural  Statistics  Services, 
Washington,  D.C.  www.nass.usda.gov/QuickStats/ 
(accessed  1 1 February  2005). 

Sauer,  J.  R.,  J.  E.  Hines,  and  J.  Fallon.  2005.  The 
North  American  Breeding  Bird  Survey:  results 
and  analysis  1966-2004,  ver.  2005.2.  USGS  Pa- 
tuxent Wildlife  Research  Center,  Laurel,  Mary- 
land. www.mbr-pwrc.usgs.gov/bbs/bbs.html  (ac- 
cessed 19  August  2005). 

Smith,  E.  L.  and  K.  C.  Kruse.  1992.  The  relationship 
between  land-use  and  the  distribution  and  abun- 
dance of  Loggerhead  Shrikes  in  south-central  Il- 
linois. Journal  of  Field  Ornithology  63:420-427. 

U.S.  Fish  and  Wildlife  Service.  2002.  Birds  of  con- 
servation concern  2002.  U.S.  Fish  and  Wildlife 
Service,  Division  of  Migratory  Bird  Management, 
Arlington,  Virginia. 

Warner,  R.  E.  1985.  Demography  and  movements  of 
free-ranging  domestic  cats  in  rural  Illinois.  Journal 
of  Wildlife  Management  49:340-346. 

Yosef,  R.  1994.  Effect  of  fencelines  on  the  reproduc- 
tive success  of  Loggerhead  Shrikes.  Conservation 
Biology  8:281-285. 

Yosef,  R.  1996.  Loggerhead  Shrike,  Lanius  ludovici- 
anus.  The  Birds  of  North  America,  no.  231. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):75 — 80,  2006 


NEST  INTERFERENCE  BY  FLEDGLING  LOGGERHEAD  SHRIKES 

ERIC  L.  KERSHNER12  AND  ERIC  C.  MRUZ1 


ABSTRACT. — Using  video  cameras,  we  documented  at  least  two  fledgling  Loggerhead  Shrikes  ( Lanius  lu- 
dovicianus)  visiting  their  parent’s  second  active  nest.  We  recorded  70  visits  during  a 10-day  period,  with  visits 
averaging  7 min.  We  observed  the  fledglings  sitting  on  the  nest  contents  on  21  occasions.  We  concluded  that 
these  visits  were  not  indicative  of  cooperative  breeding  behavior,  because  the  fledglings  were  destructive  to  the 
nest  structure  and  contents,  and  the  adult  female  exhibited  aggressive  behavior  toward  the  fledglings.  An  early 
reduction  in  post-fledging  parental  care  by  their  father  (who  was  of  captive-bred  origin)  and  slow  development 
of  the  fledglings’  hunting  skills  might  have  caused  them  to  seek  food  resources  from  their  mother.  However, 
this  is  the  first  time  that  we  have  observed  these  behaviors  in  this  intensively  managed  population.  Received  27 
December  2004,  accepted  24  October  2005. 


The  presence  of  extra  individuals  at  nests 
has  been  observed  in  many  groups  of  birds 
(Skutch  1961,  Stacey  and  Koenig  1990). 
Many  of  these  extra  individuals  have  been 
considered  “helpers”  in  cooperative  breeding 
systems.  Helpers  have  been  documented  in  a 
variety  of  species,  including  members  of  the 
Corvidae  (Woolfenden  and  Fitzpatrick  1984), 
Hirundinidae  (Myers  and  Waller  1977,  Fraga 
1979),  Furnariidae  (Skutch  1969),  and  a few 
raptors  (Faaborg  et  al.  1980,  James  and  Oli- 
phant  1986).  These  extra  individuals  are  con- 
sidered to  place  the  good  of  the  species  over 
the  good  of  the  individual,  contrary  to  the  ba- 
sic tenets  of  natural  selection  (Wynne-Ed- 
wards  1962).  In  these  systems,  the  extra  in- 
dividuals help  a breeding  pair  maximize  an- 
nual productivity  by  assisting  with  nest  build- 
ing, attendance,  and  post-fledgling  care. 

However,  not  all  extra  nest  visitors  can  be 
classified  as  helpers.  Lombardo  (1986)  noted 
that  the  sole  purpose  of  extra  Tree  Swallows 
( Tachycineta  bicolor ) visiting  a breeding 
pair’s  nest  was  to  obtain  food  resources. 
House  Wrens  ( Troglodytes  aedon)  and  Acorn 
Woodpeckers  (Melanerpes  formicivorus)  have 
been  observed  using  their  non-active  natal 
nests  for  night  roosting  (Preble  1961,  Koenig 
et  al.  1995),  and  fledgling  Carolina  Wrens 
(Thryothorus  ludovicianus ) have  used  an  ac- 
tive Northern  Cardinal  ( Cardinalis  cardinalis) 
nest  during  a period  of  inclement  weather  (Ja- 
wor  and  Gray  2003).  Thus,  the  motivation  for 


‘Inst,  for  Wildlife  Studies,  2515  Camino  del  Rio 
South,  Ste.  334,  San  Diego,  CA  92018,  USA. 

2 Corresponding  author;  e-mail:  kershner@iws.org 


visiting  the  active  nests  of  parents  or  unrelated 
adults  likely  varies  by  species. 

As  part  of  a larger  study  assessing  the  nest- 
ing behavior  of  Loggerhead  Shrikes  ( Lanius 
ludovicianus)  on  San  Clemente  Island  (SCI), 
California,  we  documented  at  least  two  fledg- 
lings from  a first  brood  visiting  and  interfering 
at  their  parent’s  second  nest.  We  believe  this 
is  the  first  record  of  fledgling  shrikes  returning 
to  their  parent’s  subsequent  active  nest.  We 
document  the  nest  visitation  by  these  fledg- 
lings, and  explore  the  reasons  for  these  visits. 

METHODS 

We  collected  our  fledgling  interference  data 
during  a larger  study  on  the  nesting  behavior 
of  Loggerhead  Shrikes  on  SCI  (32°  50'  N, 
118°  30'  W),  which  is  located  approximately 
109  km  northwest  of  San  Diego,  California. 
SCI  is  administered  by  the  U.S.  Navy  and  is 
used  for  active  military  training  as  part  of  the 
Southern  California  Offshore  Range;  the  U.S. 
Navy  also  has  an  environmental  program  on 
the  island  for  the  protection  and  conservation 
of  natural  resources  (U.S.  Department  of  the 
Navy  2002). 

From  23  May  to  27  June  2003,  we  video- 
taped a pair  of  shrikes  nesting  in  Norton  Can- 
yon, located  on  the  west  side  of  SCI.  The 
breeding  territory  at  this  site  is  at  the  bottom 
of  a steep  canyon,  where  there  are  several 
small  trees  and  shrubs.  As  part  of  the  recovery 
program  for  this  endangered  population,  two 
captive-born  shrikes  were  released  into  the 
wild  in  1999  (male)  and  2001  (female)  as 
hatch-year  birds.  In  2002,  the  male  bred  suc- 
cessfully in  this  same  territory,  whereas  the 
female  bred  successfully  with  another  male  in 


75 


76 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


a nearby  canyon  in  2002.  As  part  of  the  larger 
recovery  effort  to  improve  reproductive  out- 
put and  survival  of  adults  and  offspring,  we 
provided  supplemental  food  to  these  birds  ev- 
ery other  day  during  the  breeding  season  (i.e., 
1 February— 15  July;  feeding  began  when  birds 
took  up  occupancy  at  this  site).  We  provided 
a diet  of  mealworms  ( Tenebrio  sp.),  crickets 
{Gryllus  sp.),  mice  ( Mus  musculus),  and  liz- 
ards ( Anolis  sp.)  in  plastic  tubs  that  we  low- 
ered by  rope  into  the  canyon  bottom,  where 
they  remained  for  the  adults  to  use  for  a 1-hr 
period.  During  this  1-hr  period,  we  recorded 
data  on  the  identity,  health,  and  behavior  of 
all  shrikes  present;  the  amount  of  supplemen- 
tal food  taken;  and  what  each  individual  did 
with  the  supplemental  food  (e.g.,  cached  the 
food,  fed  themselves,  male  fed  the  fledglings 
or  the  female  on  the  nest). 

To  assess  the  behavior  of  nesting  shrikes, 
we  used  miniature  video  cameras  (model 
MVC2000-WP-LED,  Micro  Video  Products, 
Bobcaygeon,  Ontario,  Canada;  7.5  X 4 cm) 
placed  within  30  cm  of  the  nest.  We  set  up 
cameras  during  the  egg-laying  stage.  Each 
camera  was  equipped  with  infrared  light-emit- 
ting diodes  to  allow  data  collection  during 
night  hours.  We  used  coaxial  cable  to  connect 
each  camera  to  a time  lapse  VCR  located 
—500  m from  the  nest  tree.  We  powered  the 
VCR  and  camera  with  a series  of  12-volt 
deep-cycle  marine  batteries  and  used  solar 
panels  to  maintain  battery  charge.  We  pro- 
grammed each  VCR  to  record  five  frames/sec, 
and  we  changed  the  video  tape  every  24  hr. 
We  reviewed  video  tapes  later  to  record  nest- 
ing activity.  We  also  recorded  any  unusual 
events  at  the  nest,  such  as  the  presence  of 
predators  or  competitors  and  the  interactions 
(aggressive  or  not)  between  the  male  and  fe- 
male. We  considered  behaviors  such  as  bill 
snapping  and  physical  contact  as  aggressive 
behaviors  (Yosef  1996).  We  collected  other 
nesting  behavior  data  from  this  site  by  ob- 
serving the  territory  from  the  canyon  rim  dur- 
ing supplemental  feeding  sessions  (i.e.,  every 
other  day).  All  results  are  presented  as  means 
± SE. 

RESULTS 

Five  young  fledged  from  the  first  nest  on 
~6  May.  All  fledglings  were  color  banded  pri- 
or to  fledging  for  individual  identification. 


Both  adults  provided  care  to  the  fledglings  un- 
til —23  May,  when  the  female  began  incubat- 
ing her  second  clutch  in  a different  tree  within 
the  same  territory.  Data  collected  during  sup- 
plemental feeding  observations  indicated  that 
three  to  four  fledglings  were  present  at  the 
second  nest  site  during  the  period  we  collected 
video  data.  We  also  found  that  the  male  allo- 
cated more  time  to  feeding  the  female  on  the 
second  nest  than  to  caring  for  the  first-nest 
fledglings. 

During  the  second  nest  attempt  (23  May- 
27  June),  from  which  five  young  fledged,  we 
observed  at  least  two  different  first-nest  fledg- 
lings visiting  the  second  nest  on  10  different 
days.  Color  bands  were  indistinguishable  on 
black-and-white  video  footage,  but  we  had 
verified  the  identity  of  first-nest  fledglings  re- 
maining in  the  territory  during  supplemental 
feeding  sessions.  The  first  visit  was  made  on 
24  May  and  the  last  took  place  on  8 June, 
although  we  detected  at  least  three  fledglings 
from  the  first  nest  attempt  in  the  general  area 
until  1 1 July.  We  do  not  know  whether  fledg- 
lings were  regularly  present  in  the  nest  tree 
during  this  period,  as  the  camera  was  focused 
on  the  nest  and  immediate  surroundings.  On 
numerous  occasions,  the  adult  female  was  ob- 
served vocalizing  at  something  in  the  nest 
tree,  probably  fledglings  that  may  have  spent 
considerable  time  in  the  nest  tree  and  out  of 
camera  view. 

We  recorded  70  visits  (mean  visits/day  = 
7.0  ± 2.4,  range  = 1-22)  during  the  10  days 
when  first-nest  fledglings  appeared  at  the  nest 
(Fig.  1).  During  these  visits,  fledglings  spent 
a total  of  6 hr  48  min  at  the  nest  (mean  min/ 
visit  = 7.0  ± 1.7,  range  = 3-21).  On  24  May, 
two  first-nest  fledglings  were  present  at  the 
nest  at  the  same  time  on  three  separate  occa- 
sions. Subsequently,  we  witnessed  only  one 
fledgling  at  the  nest  at  any  one  time. 

On  21  occasions,  we  observed  a first-nest 
fledgling  sitting  on  the  nest  while  the  adult 
female  was  away.  This  occurred  16  times  dur- 
ing the  egg  stage  and  5 times  during  the  nest- 
ling stage.  When  the  female  left  the  nest,  the 
fledgling  would  move  into  the  nest  cup  im- 
mediately. The  total  time  spent  sitting  on  the 
nest  contents  by  first-nest  fledglings  was  1 hr 
12  min,  averaging  3.0  ± 1.0  min  per  occasion. 
On  two  occasions  when  a fledgling  was  sitting 
on  the  nest,  the  female  tried  to  evict  the  fledg- 


Kershner  and  Mruz.  • NEST  INTERFERENCE  BY  SHRIKES 


77 


0:25 

0:23 

0:21 

0:19 

0:17  e 
E 

0:15  its 
0:12  ? 


0:10  § 


0:08 

0:06 

0:04 

0:02 

0:00 


FIG.  1.  Number  of  visits  and  mean  time  per  visit  for  fledgling  Loggerhead  Shrikes  returning  to  their  parents’ 
second  active  nest  on  San  Clemente  Island,  California,  2004.  Asterisks  indicate  dates  when  we  provided  sup- 
plemental food. 


ling  by  pecking  at  it.  During  one  visit,  the 
female  sat  on  top  of  the  fledgling  for  15  min, 
poking  underneath  the  fledgling  as  if  to  check 
on  her  eggs.  On  several  occasions,  the  fledg- 
ling would  act  destructively  while  sitting  on 
the  nest.  These  behaviors  included  pulling  up 
the  nest  lining,  pulling  sticks  out  of  the  nest 
structure,  breaking  open  an  egg  and  eating  the 
eggshell,  pecking  at  newly  hatched  nestlings, 
and  stealing  food  from  the  female.  Other  be- 
haviors exhibited  by  fledglings  included  con- 
stant begging  to  the  female,  sleeping  on  the 
rim  of  the  nest,  and  blocking  the  female  from 
entering  the  nest  cup.  Fledglings  would  block 
the  female  by  getting  in  the  nest  and  moving 
so  that  the  female  could  not  resume  incubation 
or  brooding. 

We  observed  the  adult  female  feeding  first- 
nest  fledglings  on  18  occasions  (Fig.  2),  not 
including  when  fledglings  stole  food  from  the 
female  or  consumed  food  laying  in  the  nest. 
More  often,  however,  the  female  was  aggres- 
sive toward  fledglings  at  the  nest.  We  recorded 


215  aggressive  acts  between  the  female  and 
fledglings,  with  the  number  of  aggressive  acts/ 
min  ranging  from  0.25  to  1.16  (Fig.  2). 

DISCUSSION 

We  believe  this  is  the  first  report  of  fledg- 
ling shrikes  returning  to  an  active  nest  of  their 
parents.  The  actual  number  of  different  fledg- 
lings visiting  the  nest  (we  know  of  at  least 
two)  was  uncertain.  It  does  seem  clear,  how- 
ever, that  the  observations  made  at  this  nest 
are  not  behaviors  associated  with  a coopera- 
tive breeding  system.  In  cooperative  breeding 
systems,  helpers  are  present  to  assist  the  nest- 
ing pair  increase  productivity  (Skutch  1961). 
Assisting  with  nest  building,  nest  defense, 
nest  attendance,  and  post-fledging  care  allows 
the  nesting  pair  to  focus  their  energy  on  pro- 
ducing multiple  clutches.  Some  species  would 
not  be  capable  of  double-clutching  without  a 
cooperative  breeding  system  (Poiani  and  Jer- 
miin  1994).  Although  Loggerhead  Shrikes  on 
SCI  are  frequently  double-brooded,  in  this 


78 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


FIG.  2.  Number  of  times  fledglings  were  fed  by  their  parents  while  visiting  their  second  nest,  and  the  number 
of  aggressive  acts  per  minute  exhibited  by  the  adult  female  toward  visiting  fledglings  while  at  the  nest,  San 
Clemente  Island,  California,  2004. 


case,  the  destructive  nature  of  these  fledglings 
(e.g.,  nest  and  egg  destruction,  attacking  nest- 
lings, stealing  food)  suggests  that  they  were 
not  attempting  to  increase  their  parent’s  an- 
nual productivity,  despite  their  incidental  nest 
attendance. 

We  believe  first-nest  fledglings  visited  the 
nest  and  begged  from  the  female  to  extend 
parental  care.  In  Loggerhead  Shrikes,  post- 
fledging  care  is  generally  the  duty  of  the  male 
(Yosef  1996).  During  this  period,  the  male 
provisions  the  young,  who  often  follow  the 
male  around,  presumably  learning  how  to  hunt 
for  food.  This  period  extends  to  independence, 
which  occurs  40  days  after  hatching  (Scott 
and  Morrison  1990).  The  first-nest  fledglings 
in  our  study  should  have  become  independent 
on  29  May.  However,  it  appears  that  when  the 
female  began  incubating  her  second  clutch  on 
23  May,  the  male  turned  his  attention  to  pro- 
visioning the  female  on  the  nest  and  reduced 


his  feeding  of  the  first-nest  fledglings.  There- 
fore, it  appears  that  the  amount  of  parental 
care  might  have  been  reduced  earlier  than  nor- 
mal, although  how  and  when  parental  care  is 
terminated  remains  unclear.  Trivers  (1974) 
suggests  that  there  is  a conflict  between  adults 
and  offspring  regarding  how  long  the  depen- 
dency period  should  be.  There  should  be  se- 
lective pressure  for  young  to  try  to  receive 
more  parental  care  than  is  optimal  for  the  par- 
ents to  give. 

Why  males  might  reduce  parental  care  early 
is  unknown.  The  male  we  observed  was  cap- 
tive-bred and  may  have  exhibited  some  be- 
havioral abnormalities  associated  with  being 
reared  in  captivity.  Selection  pressures  in  cap- 
tivity are  vastly  different  from  those  in  the 
wild,  and,  as  a result,  changes  in  important 
life-history  behavioral  traits  may  occur  (Curio 
1996,  McPhee  2003).  For  example,  Woolaver 
et  al.  (2000)  found  that  over-dependence  on 


Kershner  and  Mruz.  • NEST  INTERFERENCE  BY  SHRIKES 


79 


food  provisioning  resulted  in  behavioral 
changes  in  captive-bred  and  released  Echo 
Parakeets  ( Psittacula  echo),  and  Harvey  et  al. 
(2002)  found  abnormal  nesting  behavior  in 
captive  Hawaiian  Crows  ( Corvus  hawaiien- 
sis ).  Thus,  since  the  male  we  observed  was 
reared  in  captivity  before  being  released  into 
the  wild,  there  could  have  been  some  behav- 
ioral deficiency  causing  the  male  to  terminate 
parental  care  prematurely.  We  do  not  believe 
this  to  be  the  case,  however,  as  this  male  was 
released  into  the  wild  as  a juvenile,  and  young 
birds  are  better  at  assimilating  into  new  wild 
environments  than  adults  (Swinnerton  et  al. 
2000,  Robert  et  al.  2004,  Turner  et  al.  2004). 
In  addition,  this  male  bred  each  year  after  his 
release  and  successfully  raised  four  fledglings 
to  independence  prior  to  2003.  During  nest 
monitoring  of  his  prior  breeding  attempts,  we 
did  not  detect  any  abnormal  behavior  (Insti- 
tute for  Wildlife  Studies  unpubl.  data). 

A more  plausible  explanation  for  our  un- 
usual observation  could  be  the  presence  of 
supplemental  food  provided  as  part  of  the  re- 
covery program.  Supplemental  food  is  meant 
to  increase  survival  and  productivity;  howev- 
er, it  is  unknown  to  what  extent  released  birds 
rely  on  this  food.  If  the  male  shrike  relied  on 
supplemental  food  for  provisioning  the  female 
and  fledglings,  he  may  have  foraged  less  for 
natural  food  than  birds  that  do  not  receive 
supplemental  food.  It  is  also  possible  that  the 
presence  of  supplemental  food  slowed  the 
first-nest  fledglings’  learning  process  in  ac- 
quiring natural  food.  Wheelwright  and  Tem- 
pleton (2003)  suggest  that  the  speed  at  which 
juveniles  acquire  foraging  skills  might  deter- 
mine the  length  of  parental  care.  By  feeding 
regularly  from  the  food  tubs,  the  fledglings 
may  not  have  observed  many  wild  foraging 
tactics  by  the  male.  Thus,  they  needed  more 
time  to  develop  these  skills  and  continued  to 
beg  for  food  from  both  adults — despite  the  po- 
tential cost  to  the  adults  (Trivers  1974). 

Our  supplemental  feeding  observations  re- 
vealed that  first-nest  fledglings  learned  to  for- 
age from  the  food  tubs  rather  quickly  and  reg- 
ularly took  supplemental  food  when  we  of- 
fered it.  This  may  explain  why  the  fledglings 
did  not  visit  the  second  nest  every  day.  We 
provided  supplemental  food  on  6 days  during 
the  period  when  the  fledglings  were  observed 
at  the  nest  (24  May-8  June).  They  did  not 


visit  the  nest  on  4 of  those  days  (25  May,  27 
May,  2 June,  6 June),  and  visited  only  two  or 
three  times  on  each  of  the  other  2 days  we 
provided  supplemental  food  (31  May  and  4 
June;  Fig.  1).  Fledglings  visited  the  nest  on  8 
of  10  days  when  food  was  not  provided,  sug- 
gesting that  the  fledglings  sought  provisioning 
from  the  female. 

Differences  in  the  number  of  visits  each  day 
could  also  be  explained  by  the  abrupt  reduc- 
tion in  food  provisioning.  The  first  day  the 
female  ceased  parental  care  for  her  first  brood 
was  24  May.  The  fledglings  were  unaccus- 
tomed to  not  being  fed  by  her,  potentially  ex- 
plaining the  18  visits  to  the  nest  on  that  day 
(Fig.  1),  On  7 June,  the  second  clutch  of  eggs 
began  to  hatch,  and  there  was  a spike  in  ac- 
tivity surrounding  the  nest  as  the  female  re- 
moved eggshells  and  began  feeding  newly 
hatched  chicks.  This  increase  in  activity,  es- 
pecially with  food  deliveries  to  the  nest,  may 
have  caused  the  high  number  of  visits  record- 
ed ( n = 22;  Fig.  1)  that  day. 

In  general,  we  believe  that  the  visits  of 
these  fledgling  shrikes  to  their  parent’s  second 
nest  were  motivated  by  hunger,  possibly  due 
to  early  reduction  of  parental  care  or  the  re- 
tardation of  foraging  skills  due  to  the  presence 
of  supplemental  food.  On  SCI,  all  captive- 
reared  pairs  released  into  the  wild  received 
supplemental  food.  Since  2003,  we  have 
placed  small  video  cameras  at  10  shrike  nests 
to  study  nesting  behavior,  and  have  not  ob- 
served nest  visits  by  first-nest  fledglings  at 
any  other  time  (Institute  for  Wildlife  Studies 
unpubl.  data). 

Despite  the  potential  benefit  provided  by 
fledglings  “tending”  the  nest  when  the  adult 
female  was  away  from  the  nest,  the  first-nest 
fledglings  likely  interfered  with  the  success  of 
the  second  nesting  attempt,  as  one  egg  was 
destroyed  by  a visiting  fledgling,  food  was 
stolen  from  the  female,  and  visiting  fledglings 
constantly  pecked  at  newly  hatched  young. 

ACKNOWLEDGMENTS 

We  thank  D.  M.  Cooper,  R.  Dempsey,  and  C.  Camp- 
bell for  assisting  with  the  implementation  of  the  cam- 
era systems  and  initial  review  of  video  tapes.  We  thank 
the  Shrike  Working  Group,  consisting  of  the  U.S. 
Navy,  U.S.  Fish  and  Wildlife  Service,  PRBO  Conser- 
vation Sciences,  and  the  Zoological  Society  of  San  Di- 
ego, for  their  input  regarding  the  initial  use  of  the  vid- 
eo cameras  at  shrike  nests.  We  thank  the  biologists 


80 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  1 , March  2006 


from  PRBO  Conservation  Sciences  for  their  assistance 
in  monitoring  shrike  nests  before  and  after  the  instal- 
lation of  nest  cameras.  We  also  thank  the  shrike  release 
crew  of  the  Institute  for  Wildlife  Studies  for  monitor- 
ing shrikes  at  our  camera  site  during  supplemental 
feeding  sessions.  E.  M.  Russell  and  two  anonymous 
referees  provided  comments  on  an  earlier  draft  of  the 
manuscript. 

LITERATURE  CITED 

Curio,  E.  1996.  Conservation  needs  ethology.  Trends 
in  Ecology  and  Evolution  11:260-263. 

Faaborg,  J.,  T.  DeVries,  C.  B.  Patterson,  and  C.  R. 
Griffon.  1980.  Preliminary  observations  on  the 
occurrence  and  evolution  of  polyandry  in  the  Ga- 
lapagos Hawk.  Auk  97:581-590. 

Fraga,  R.  M.  1979.  Helpers  at  the  nest  in  passerines 
from  Buenos  Aires  Province,  Argentina.  Auk  96: 
606-608. 

Harvey,  N.  C.,  S.  M.  Farabaugh,  and  B.  B.  Druker. 
2002.  Effects  of  early  rearing  experience  on  adult 
behavior  and  nesting  in  captive  Hawaiian  Crows. 
Zoo  Biology  21:59-75. 

James,  P.  C.  and  L.  W.  Oliphant.  1986.  Extra  birds 
and  helpers  at  the  nests  of  Richardson’s  Merlin. 
Condor  88:533-534. 

Jawor,  J.  M.  and  N.  Gray.  2003.  Use  of  Northern 
Cardinal  nest  by  fledgling  Carolina  Wrens.  Wilson 
Bulletin  115:95-96. 

Koenig,  W.  B.,  P.  B.  Stacey,  M.  T.  Stanback,  and  R. 
L.  Mumme.  1995.  Acorn  Woodpecker  ( Melaner - 
pes  formicivorus).  The  Birds  of  North  America, 
no.  194. 

Lombardo,  M.  P.  1986.  Attendants  at  Tree  Swallow 
nests.  I.  Are  attendants  helpers  at  the  nest?  Condor 
88:297-303. 

McPhee,  M.  E.  2003.  Generations  in  captivity  increas- 
es behavior  variance:  considerations  for  captive 
breeding  and  reintroduction  programs.  Biological 
Conservation  115:71-77. 

Myers,  G.  R.  and  D.  W.  Waller.  1977.  Helpers  at 
the  nest  in  Barn  Swallows.  Auk  94:596. 

Poiani,  A.  and  L.  S.  Jermiin.  1994.  A comparative 
analysis  of  some  life-history  traits  between  co- 
operatively and  non-cooperatively  breeding  Aus- 
tralian passerines.  Evolutionary  Ecology  8:471- 
488. 

Preble,  C.  S.  1961 . Unusual  behavior  of  House  Wrens. 
Auk  78:442. 

Robert,  A.,  F.  Sarrazin,  D.  Couvet,  and  S.  Legen- 


dre. 2004.  Releasing  adults  versus  young  in  re- 
introduction:  interactions  between  demography 
and  genetics.  Conservation  Biology  18:1 078 — 
1087. 

Scott,  T.  A.  and  M.  L.  Morrison.  1990.  Natural  his- 
tory and  management  of  the  San  Clemente  Log- 
gerhead Shrike.  Proceedings  of  the  Western  Foun- 
dation of  Vertebrate  Zoology  4:23-57. 

Skutch,  A.  F.  1961.  Helpers  among  birds.  Condor  63: 
198-226. 

Skutch,  A.  F.  1969.  A study  of  the  Rufous-fronted 
Thornbird  and  associated  birds,  part  1 . Wilson 
Bulletin  81:5-43. 

Stacey,  P.  B.  and  W.  D.  Koenig.  1990.  Cooperative 
breeding  in  birds:  long-term  studies  of  ecology 
and  behavior.  Cambridge  University  Press,  New 
York. 

Swinnerton,  K.  J.,  C.  G.  Jones,  R.  Lam,  S.  Paul,  R. 
Chapman,  K.  A.  Murray,  and  K.  Freeman.  2000. 
The  release  of  captive-bred  Echo  Parakeets  to  the 
wild,  Mauritius.  Reintroduction  News  19:10-12. 

Trivers,  R.  L.  1974.  Parent  offspring  conflict.  Amer- 
ican Zoologist  14:249-264. 

Turner,  J.  M.,  C.  L.  Sulzman,  E.  L.  Kershner,  and 

D.  K.  Garcelon.  2004.  San  Clemente  Loggerhead 
Shrike  release  program-2003,  final  report.  U.S. 
Navy,  Natural  Resources  Management  Branch, 
Southwest  Division,  Naval  Facilities  Engineering 
Command,  San  Diego,  California. 

U.S.  Department  of  the  Navy,  Southwest  Division. 
2002.  San  Clemente  integrated  natural  resources 
management  plan  draft  final.  Prepared  by  Tierra 
Data  Systems,  Escondido,  California. 

Wheelwright,  N.  T.  and  J.  J.  Templeton.  2003.  De- 
velopment of  foraging  skills  and  the  transition  to 
independence  in  juvenile  Savannah  Sparrows. 
Condor  105:279-287. 

Woolaver,  L.,  C.  G.  Jones,  K.  J.  Swinnerton,  K. 
Murray,  A.  Lalinde,  D.  Birch,  F.  de  Ravel,  and 

E.  Ridgeway.  2000.  The  release  of  captive-bred 
Echo  Parakeets  to  the  wild,  Mauritius.  Reintro- 
duction News  19:12-15. 

WOOLFENDEN,  G.  E.  AND  J.  W.  FITZPATRICK.  1984.  The 
Florida  Scrub- Jay:  demography  of  a cooperative 
breeding  bird.  Princeton  University  Press,  Prince- 
ton, New  Jersey. 

Wynne-Edwards,  V.  C.  1962.  Animal  dispersion  in 
relation  to  social  behavior.  Oliver  and  Boyd,  Ed- 
inburgh, United  Kingdom. 

Yosef,  R.  1996.  Loggerhead  Shrike  ( Lanius  ludovici- 
anus).  The  Birds  of  North  America,  no.  231. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ): 8 1 — 84,  2006 


FIRST  BREEDING  RECORD  OF  A MOUNTAIN  PLOVER  IN 
NUEVO  LEON,  MEXICO 

JOSE  I.  GONZALEZ  ROJAS,1 3 MIGUEL  A.  CRUZ  NIETO,2 
OSCAR  BALLESTEROS  MEDRANO,1 2 3  AND  IRENE  RUVALCABA  ORTEGA1 


ABSTRACT. — We  document  the  first  breeding  record  of  Mountain  Plovers  ( Charadrius  montanus)  in  the 
state  of  Nuevo  Leon,  Mexico.  On  9 July  2004,  we  located  a nest  with  two  eggs  and  one  chick  in  a colony  of 
Mexican  prairie  dogs  ( Cynomys  mexicanus).  Mean  height  of  vegetation  near  the  nest  was  7.1  cm,  and  bare 
ground  cover  was  41.2%  (30  m2  sampled).  Although  this  record  represents  the  second  nesting  for  this  species 
in  Mexico,  it  is  the  first  to  document  successful  breeding.  Received  21  January  2005,  accepted  5 November 
2005. 


The  Mountain  Plover  ( Charadrius  montan- 
us) is  a species  of  North  America’s  grasslands. 
It  is  classified  as  vulnerable  on  the  IUCN  Red 
List  (Birdlife  International  2004),  endangered 
in  Canada  (Committee  on  the  Status  of  En- 
dangered Wildlife  in  Canada  2004),  and 
threatened  in  Mexico  (Diario  Oficial  de  la 
Federacion  2002).  In  the  United  States,  the 
Mountain  Plover  was  proposed  for  listing  as 
a threatened  species  in  1999  (U.S.  Fish  and 
Wildlife  Service  1999),  but  the  proposal  was 
withdrawn  in  2003  (U.S.  Fish  and  Wildlife 
Service  2003).  The  U.S.  Shorebird  Conser- 
vation Plan  rates  the  species  as  highly  imper- 
iled (Brown  et  al.  2001).  Between  1966  and 
1991,  the  entire  population  of  Mountain  Plo- 
vers declined  by  63%  (Knopf  1994);  current- 
ly, the  population  is  estimated  at  1 1 ,000- 
14,000  individuals  (Plumb  et  al.  2005).  The 
population  decline  has  been  attributed  to  loss 
of  nesting  habitat  due  to  cultivation,  urbani- 
zation, livestock  management,  and  declines  in 
native  herbivores,  mainly  black-tailed  prairie 
dogs  ( Cynomys  ludovicianus)  and  North 
American  bison  ( Bison  bison ; Wiersma  1996, 
BirdLife  International  2004). 

The  Mountain  Plover’s  primary  breeding 
range  includes  eastern  Colorado,  central  Wy- 
oming, eastern  Montana  (Graul  and  Webster 
1976),  northeastern  New  Mexico,  and  the 


1 Lab.  de  Ornitologia,  Fac.  de  Ciencias  Biologicas, 
Univ.  Autonoma  de  Nuevo  Leon,  A.P.  25-F,  Cd.  Univ- 
ersitaria,  San  Nicolas  de  los  Garza,  Nuevo  Leon 
66450,  Mexico. 

2 Pronatura  Noreste,  A.  C.  Loma  Larga  331,  Mon- 
terrey, Nuevo  Leon  25268,  Mexico. 

3 Corresponding  author;  e-mail: 
josgonza@fcb.uanl.mx 


Oklahoma  and  Texas  panhandles  (Knopf 
1996).  An  isolated  breeding  population,  which 
may  be  resident  year-round,  occurs  in  the  Da- 
vis Mountains,  Texas  (Knopf  1996).  In  the 
United  States,  the  plover’s  winter  range  ex- 
tends from  Sacramento,  San  Joaquin,  and  the 
Imperial  Valley  in  California  east  to  the  Lower 
Colorado  River  Valley,  and  from  Yuma  east 
to  Phoenix  and  the  Chandler  area  in  southern 
Arizona  (Rosenberg  et  al.  1991,  Knopf  and 
Rupert  1995).  In  Mexico,  the  winter  distri- 
bution has  not  been  well  studied,  but  it  is  be- 
lieved to  extend  along  the  U.S./Mexico  border 
south  through  Baja  California,  Sonora,  Chi- 
huahua, and  Tamaulipas  into  Zacatecas  and 
San  Luis  Potosf  (Phillips  et  al.  1964,  Wilbur 
1987,  Howell  and  Webb  1995,  Gomez  de  Sil- 
va et  al.  1996).  More  surveys  are  needed  to 
document  wintering  as  well  as  year-round  res- 
ident populations. 

Mountain  Plovers  nest  in  shortgrass  and 
mixed  grass  prairies  (Graul  and  Webster  1976, 
Knowles  et  al.  1982,  Knopf  and  Miller  1994, 
Knopf  and  Rupert  1999b).  They  typically  oc- 
cur in  areas  characterized  by  short  vegetation 
(<8  cm  high;  Graul  1975)  and  >30%  bare 
ground  (Knopf  and  Miller  1994),  and  they  are 
commonly  associated  with  prairie  dog  colo- 
nies ( Cynomys  spp.;  Knowles  et  al.  1982). 
Vegetation  at  nest  sites  varies  throughout  the 
breeding  range,  but  is  usually  dominated  by 
blue  grama  ( Bouteloua  gracilis ),  buffalograss 
( Buchloe  dactyloides),  needle-and-thread  (Sti- 
pa  comata),  and  sagebrush  (Artemisia  sp.; 
Finzel  1964,  Graul  1975,  Knowles  et  al.  1982, 
Knopf  and  Miller  1994).  Plovers  often  nest 
near  cow  manure,  rocks,  or  clumps  of  vege- 
tation (Graul  1975,  Olson  and  Edge  1985, 
Knopf  and  Miller  1994). 


81 


82 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


During  the  breeding  period,  Mountain  Plo- 
vers have  been  observed  in  colonies  of  Mex- 
ican prairie  dogs  ( Cynomys  mexicanus) 
around  San  Juan  del  Prado,  Galeana,  Nuevo 
Leon.  These  observations  have  included  in- 
dividuals in  breeding  plumage  and  birds  ex- 
hibiting reproductive  behavior  (calls,  displays, 
etc.;  Knopf  and  Rupert  1999a).  On  5 July 
1994,  observers  found  six  widely  spaced  pairs 
and  one  single  individual;  on  16  June  1997, 
seven  individuals  (including  a pair)  were  de- 
tected in  three  unspecified  prairie  dog  colonies 
in  the  same  area;  and  on  24  April  1998,  seven 
more  individuals  (including  two  pairs)  were 
observed.  Mountain  Plovers  were  also  detect- 
ed on  25  April  1998  at  Galeana  between  El 
Cristal  and  La  Paz  (two  pairs  and  two  terri- 
torial males)  and  on  26  April  1998  at  La  He- 
diondilla  (one  pair  and  one  single)  and  Llano 
La  Soledad  (two  pairs),  where  unsuccessful 
attempts  were  made  to  document  nesting 
(Knopf  and  Rupert  1999a).  During  5-1 1 May 
1999,  Desmond  and  Chavez-Ramirez  (2002) 
observed  30  Mountain  Plovers,  including 
eight  pairs  and  two  groups  of  three  individuals 
each  at  Rancho  Los  Angeles  and  La  India,  in 
Saltillo,  Coahuila  de  Zaragoza,  and  at  La  Ca- 
sita  in  Galeana,  Nuevo  Leon.  On  9 May  1999, 
Desmond  and  Chavez-Ramirez  (2002)  found 
a nest  with  three  eggs  near  La  India,  the  only 
previous  nest  documented  in  Mexico.  The  La 
India  nest  was  not  monitored  and  therefore, 
its  outcome  is  unknown.  Previous  nest  search- 
es conducted  in  Mexico  during  the  plover’s 
known  reproductive  period  (based  on  obser- 
vations in  the  Great  Plains)  yielded  no  other 
nesting  records;  however,  Desmond  and  Cha- 
vez-Ramirez (2002)  suggested  that  the  breed- 
ing season  in  northeastern  Mexico  might  be 
later  than  it  is  farther  north  so  that  hatching 
coincides  with  the  rainy  season  and  the  period 
of  greatest  insect  availability  (June-July). 

On  9 July  2004,  around  17:00  CST,  we  ob- 
served a Mountain  Plover  pair  at  Llano  La 
Soledad,  a gypsophile  grassland  (7,607  ha) 
within  ejido  San  Rafael,  Galeana,  Nuevo  Leon 
(24°  48'  50"  N,  100°  41 ' 54"  W).  Llano  La  So- 
ledad contains  the  largest  known  colony  of 
Mexican  prairie  dogs  (Trevino-Villarreal  and 
Grant  1998).  As  we  approached  the  pair,  one 
individual  feigned  wing  injury,  while  the  oth- 
er emitted  alarm  calls  and  flew  around  us.  Af- 
ter a few  minutes,  one  individual  “squatted” 


on  the  ground,  placed  its  bill  under  its  body, 
and  remained  motionless.  When  we  ap- 
proached within  2 m of  what  appeared  to  be 
a nest,  the  bird  again  feigned  wing  injury.  We 
subsequently  located  the  nest,  which  con- 
tained two  eggs  and  one  chick.  The  chick  re- 
mained motionless  while  the  adults  called  (as 
described  by  Graul  1974). 

Relative  to  the  plover’s  nest,  the  nearest 
Mexican  prairie  dog  burrow  was  approximate- 
ly 15  m away.  There  was  also  a small  cluster 
of  Atriplex  shrubs  ( n = 28;  estimated  mean 
height  = 60  cm)  40  m from  the  nest.  Live- 
stock (cattle,  goats)  were  nearby,  but  the  area 
was  not  overgrazed,  nor  was  cow  manure 
found  near  the  nest.  We  photographed  and 
video-recorded  the  nest.  This  record  repre- 
sents the  first  Mountain  Plover  nest  in  Nuevo 
Leon,  and  the  first  record  of  successful  nesting 
for  Mountain  Plovers  in  Mexico. 

As  part  of  another  study  at  Llano  La  Sole- 
dad, we  had  characterized  the  vegetation  a few 
days  prior  to  finding  the  plover  nest.  After 
finding  the  nest,  we  selected  three  of  our  1 X 
10-m  quadrats  that  were  closest  to  the  nest — 
200,  1,000,  and  1,500  m away — to  character- 
ize the  vegetation.  We  recorded  height,  cover 
diameter,  and  species  of  each  plant.  We  then 
calculated  mean  height,  relative  density  (RD 
= number  of  individuals  of  a given  species  as 
a proportion  of  the  total  number  of  individuals 
of  all  species),  relative  frequency  (RF  = fre- 
quency of  a given  species  as  a proportion  of 
the  sum  of  the  frequencies  for  all  species),  rel- 
ative coverage  (RC  = coverage  for  each  spe- 
cies expressed  as  a proportion  of  the  total  cov- 
erage for  all  species),  and  importance  value  of 
each  species  (IV  = RD  + RF  + RC,  which 
provides  an  overall  estimate  of  the  influence 
or  importance  of  a plant  species  in  the  com- 
munity; Brower  et  al.  1990;  Table  1).  We 
identified  1 1 plant  species,  with  a mean  height 
of  7.1  cm.  The  most  common  forbs  were  sum- 
mer bluet  ( Hedyotis  purpurea ; n = 725,  RD 
= 37,  RF  = 15.9)  and  McVaugh’s  bladderpod 
(. Leonsquerella  mcvaughiana ; n — 273,  RD  = 
13.9,  RF  = 15.9);  Muhlenbergia  sp.  ( n = 654, 
RD  = 33.3,  RF  = 15.9)  and  Karwinski’s 
grama  ( Bouteloua  karwinskii ; n = 140,  RD  = 
7.1,  RF  = 10.5)  were  the  most  common  grass- 
es. Muhlenbergia  sp.  had  the  greatest  RC 
(43.3%)  and  IV  (92.3),  and  the  IV  of  summer 


Gonzalez  Rojas  et  al.  • MOUNTAIN  PLOVER  NEST  IN  NUEVO  LEON 


83 


TABLE  1.  Vegetation  composition  and  structure  in  three  1 X 10-m  quadrats,  placed  2(X),  1,000,  and  1,500  m 


away  from  a Mountain  Plover  nest,  in  the  grassland  at 
2004. 

Llano  La  Soledad, 

Galeana, 

Nuevo  Leon,  Mexico,  July 

Species 

No. 

Ha  (cm) 

RDb  (%) 

RFC  (%) 

RCd  (%) 

IVe 

Forbs 

Summer  bluet  ( Hedyotis  purpurea) 

725 

2.5 

37.0 

15.9 

12.4 

65.2 

McVaugh’s  bladderpod  ( Lesquerella  mcvaughiana) 

273 

3.3 

13.9 

15.9 

19.0 

48.8 

Desert  zinnia  ( Zinnia  acerosa) 

80 

4.5 

4.0 

5.3 

5.7 

15.0 

Woody  crinklemat  {Tiquilia  canescens) 

Al 

5.8 

2.4 

5.2 

3.7 

11.3 

Houston  machaeranthera  ( Machaeranthera  aurea) 

35 

5.1 

1.7 

5.2 

0.5 

7.4 

Texas  sundrops  ( Calylophus  tubicula) 

5 

8.6 

0.3 

10.5 

0.7 

11.5 

Slimpod  fiddleleaf  ( Nama  stenophyllum ) 

1 

3.0 

0.1 

5.2 

0.1 

5.4 

Grasses 

Muhly  ( Muhlenbergia  sp.) 

654 

5.3 

33.3 

15.9 

43.3 

92.3 

Karwinski’s  grama  ( Bouteloua  karwinskii) 

140 

1.7 

7.1 

10.5 

13.0 

30.6 

Havard’s  threeawn  ( Aristida  havardii ) 

1 

25.0 

0.1 

5.2 

1.2 

6.5 

Buffalograss  ( Buchloe  dactyloides) 

1 

13.0 

0.1 

5.2 

0.4 

5.7 

a H = mean  height. 

b RD  = relative  density  (number  of  individuals  of  a given  species  as  a proportion  of  the  total  number  of  individuals  of  all  species). 
c RF  = relative  frequency  (frequency  of  a given  species  as  a proportion  of  the  sum  of  the  frequencies  for  all  species). 
d RC  = relative  coverage  (coverage  for  each  species  expressed  as  a proportion  of  the  total  coverage  for  all  species). 
e Importance  value  = RD  + RF  + RC  (Brower  et  al.  1990). 


bluet  was  65.2.  The  sampled  area  comprised 
41.2%  bare  ground. 

The  continued  documentation  of  Mountain 
Plover  nests  in  northeastern  Mexico  further 
confirms  that  a breeding  population  of  Moun- 
tain Plovers  exists  in  northeastern  Mexico 
(Knopf  and  Rupert  1 999a,  Desmond  and  Cha- 
vez-Ramirez  2002).  Desmond  and  Chavez- 
Ramirez  (2002)  proposed  that  the  breeding 
season  in  northeastern  Mexico  may  be  later 
than  that  known  for  northern  populations,  but 
a more  accurate  hypothesis  might  be  that  the 
breeding  period  in  northeastern  Mexico  is  pro- 
tracted because  the  earliest  observation  of 
pairing  occurred  in  late  April  (Knopf  and  Ru- 
pert 1999a)  and  the  latest  nest  with  eggs  was 
observed  in  early  July. 

Vegetation  characteristics  near  the  nest  we 
found  corresponded  with  those  reported  by 
Graul  (1975;  height  <8  cm),  Knopf  and  Mill- 
er (1994;  bare  ground  >30%),  and  Desmond 
and  Chavez-Ramirez  (2002;  height  = 2.3  cm, 
bare  ground  = 86.4%).  The  presence  of  a 
shading  element  near  the  nest  is  considered 
important  in  nest-site  selection  (Graul  1975, 
Olson  and  Edge  1985,  Knopf  and  Miller 
1994);  the  nearest  shade  we  found  was  40  m 
from  the  nest  (a  cluster  of  Atriplex  sp.).  Dom- 
inant plant  species  differed  from  those  report- 
ed in  association  with  Mountain  Plover  nest 


sites:  blue  grama,  buffalograss,  needle-and- 
thread,  and  sagebrush  (Finzel  1964,  Graul 
1975,  Knowles  et  al.  1982).  In  Llano  La  So- 
ledad,  however,  Muhlenbergia  sp.  and  Kar- 
winski’s  grama  were  the  dominant  grasses, 
and  summer  bluet  and  McVaugh’s  bladderpod 
were  the  dominant  forbs;  buffalograss  occurs 
in  the  area  but  was  not  common  (RD  = 0.1, 
RF  = 5.2;  Table  1). 

The  presence  of  a disjunct  Mountain  Plover 
breeding  population  in  northeastern  Mexico — 
and  its  association  with  colonies  of  Mexican 
prairie  dogs — has  strong  conservation  impli- 
cations for  grasslands  in  that  region.  However, 
the  last  remnants  of  northeastern  Mexico’s  na- 
tive grasslands  and  Mexican  prairie  dog  hab- 
itats are  being  lost,  which  could  have  negative 
effects  on  the  region’s  population  of  Mountain 
Plovers.  Other  avian  species  that  commonly 
occur  in  association  with  Mexican  prairie  dog 
colonies  include  Long-billed  Curlew  ( Numen - 
ius  americanus ),  Ferruginous  Hawk  ( Buteo 
regalis).  Burrowing  Owl  {Athene  cunicular- 
ia ),  and  an  endemic,  Worthen’s  Sparrow  (Spi- 
zella  wortheni );  they,  too,  could  be  at  risk  of 
declines  due  to  habitat  loss. 

ACKNOWLEDGMENTS 

We  thank  F.  L.  Knopf,  C.  C.  Farquhar,  E.  Inigo- 
Elias,  and  two  anonymous  referees  for  their  comments 
and  suggestions  on  the  manuscript. 


84 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


LITERATURE  CITED 

Birdlife  International.  2004.  Species  factsheet: 
Charadrius  montanus.  BirdLife  International, 
Cambridge,  United  Kingdom,  www.birdlife.org/ 
datazone/  (accessed  13  July  2004). 

Brower,  J.  E.,  J.  H.  Zar,  and  C.  N.  Von  Ende.  1990. 
Field  and  laboratory  methods  for  general  ecology, 
3rd  ed.  W.  C.  Brown,  Dubuque,  Iowa. 

Brown,  S.,  K.  Hickey,  B.  Harrington,  and  R.  Gill 
(Eds.).  2001.  United  States  Shorebird  Conserva- 
tion Plan,  2nd  ed.  Manomet  Center  for  Conser- 
vation Sciences,  Manomet,  Massachusetts. 

Committee  on  the  Status  of  Endangered  Wildlife 
in  Canada.  2004.  Canadian  species  at  risk,  No- 
vember 2004.  Committee  on  the  Status  of  Endan- 
gered Wildlife  in  Canada,  Environment  Canada, 
Ottawa,  Ontario. 

Desmond,  M.  J.  and  F.  Chavez-Ramirez.  2002.  Nest 
documentation  confirms  the  presence  of  a breed- 
ing population  of  Mountain  Plovers  ( Charadrius 
montanus ) in  north-east  Mexico.  Cotinga  17: 17 — 
19. 

Diario  Oficial  de  la  Federacion.  2002.  Norma  Ofi- 
cial  Mexicana  NOM-O59-SEMARNAT-2001 . 
Proteccion  ambiental-Especies  nativas  de  Mexico 
de  flora  y fauna  silvestres-Categorias  de  riesgo  y 
especificaciones  para  su  inclusion,  exclusion  o 
cambio-Lista  de  especies  en  riesgo.  6 March  2002. 
Mexico. 

Finzel,  J.  E.  1964.  Avian  populations  of  four  herba- 
ceous communities  in  southeastern  Wyoming. 
Condor  66:496-510. 

Gomez  de  Silva,  H.,  R.  A.  Medilin  Legorreta,  M. 
A.  Amin,  and  S.  Aguilar.  1996.  A concentration 
of  Mountain  Plovers  Charadrius  montanus  in  San 
Luis  Potosi,  Mexico.  Cotinga  5:74-75. 

Graul,  W.  D.  1974.  Vocalizations  of  the  Mountain 
Plover.  Wilson  Bulletin  86:221-229. 

Graul,  W.  D.  1975.  Breeding  biology  of  the  Mountain 
Plover.  Wilson  Bulletin  87:6-31. 

Graul,  W.  D.  and  L.  E.  Webster.  1976.  Breeding 
status  of  the  Mountain  Plover.  Condor  78:265— 
267. 

Howell,  S.  N.  and  S.  Webb.  1995.  A guide  to  the 
birds  of  Mexico  and  northern  Central  America. 
Oxford  University  Press,  New  York. 

Knopf,  F.  L.  1 994.  Avian  assemblages  on  altered  grass- 
lands. Studies  in  Avian  Biology  15:247-257. 


Knopf,  F.  L.  1996.  Mountain  Plover  ( Charadrius  mon- 
tanus). The  Birds  of  North  America,  no.  211. 

Knopf,  F.  L.  and  B.  J.  Miller.  1994.  Charadrius  mon- 
tanus: montane,  grassland,  or  bare-ground  plover? 
Auk  1 1 1 :504-506. 

Knopf,  F.  L.  and  J.  R.  Rupert.  1995.  Habits  and  hab- 
itats of  Mountain  Plover  in  California.  Condor  97: 
743-751. 

Knopf,  F.  L.  and  J.  R.  Rupert.  1999a.  A resident  pop- 
ulation of  Mountain  Plover  Charadrius  montanus 
in  Mexico?  Cotinga  11:17-19. 

Knopf,  F.  L.  and  J.  R.  Rupert.  1999b.  Use  of  culti- 
vated fields  by  breeding  Mountain  Plovers  in  Col- 
orado. Studies  in  Avian  Biology  19:81-86. 

Knowles,  C.  J.,  C.  J.  Stoner,  and  S.  P.  Gieb.  1982. 
Selective  use  of  black-tailed  prairie  dog  towns  by 
Mountain  Plover.  Condor  84:71-74. 

Olson,  S.  L.  and  D.  Edge.  1985.  Nest  site  selection 
by  Mountain  Plovers  in  northcentral  Montana. 
Journal  of  Range  Management  38:280-282. 

Phillips,  A.  R.,  J.  T.  Marshall,  and  G.  Monson. 
1964.  The  birds  of  Arizona.  University  of  Arizona 
Press,  Tucson. 

Plumb,  R.  E.,  F.  L.  Knopf,  and  S.  H.  Anderson.  2005. 
Minimum  population  size  of  Mountain  Plovers 
breeding  in  Wyoming.  Wilson  Bulletin  117:1 5— 
22. 

Rosenberg,  K.  V.,  R.  D.  Ohmart,  W.  C.  Hunter,  and 
B.  W.  Anderson.  1991.  Birds  of  the  lower  Colo- 
rado River  Valley.  University  of  Arizona  Press, 
Tucson. 

Trevino- Villarreal,  J.  and  W.  E.  Grant.  1998.  Geo- 
graphic range  of  the  endangered  Mexican  prairie 
dog  ( Cynomys  mexicanus).  Journal  of  Mammalo- 
gy 79:  i 273-1 287. 

U.S.  Fish  and  Wildlife  Service.  1999.  Endangered 
and  threatened  wildlife  and  plants:  proposed 
threatened  status  for  the  Mountain  Plover.  Federal 
Register  64(30):7587-7601 . 

U.S.  Fish  and  Wildlife  Service.  2003.  Endangered 
and  threatened  wildlife  and  plants:  withdrawal  of 
the  proposed  rule  to  list  the  Mountain  Plover  as 
threatened.  Federal  Register  68(  1 74):53083- 
53101. 

Wiersma,  P.  1996.  Family  Charadriidae  (plovers).  Pag- 
es 41 1-442  in  Handbook  of  the  birds  of  the  world, 
vol.  3:  Hoatzin  to  auks  (J.  del  Hoyo,  A.  Elliott, 
and  J.  Sargatal,  Eds.).  Lynx  Edicions,  Barcelona, 
Spain. 

Wilbur.  S.  R.  1987.  Birds  of  Baja  California.  Univer- 
sity of  California  Press,  Berkeley. 


The  Wilson  Journal  of  Ornithology  1 18(  I ):85-90,  2006 


BREEDING  BIOLOGY  OF  THE  DOUBLE-COLLARED  SEEDEATER 
(SPOROPHILA  CAERULESCENS) 

MERCIVAL  R.  FRANCISCO1 


ABSTRACT. — The  Double-collared  Seedeater  ( Sporophila  caerulescens ) is  the  most  common  seedeater  in 
southern  South  America.  Because  information  on  its  breeding  biology  is  mostly  limited  to  descriptions  of  nests 
and  eggs,  I studied  the  reproductive  biology  of  the  Double-collared  Seedeater  in  southeastern  Brazil.  I found  41 
active  nests  during  seven  breeding  seasons  (1997-2003).  Nesting  occurred  from  December  to  May.  All  nests 
found  during  incubation  contained  two  eggs,  eggs  were  laid  on  consecutive  days,  and  incubation  started  the 
morning  the  female  laid  the  last  egg.  Incubation  and  nestling  periods  were  12  and  12-15  days,  respectively. 
Only  females  incubated  the  eggs.  Mean  time  spent  incubating/hr  was  52.3  min,  and  incubation  recesses  averaged 
6.6  min.  Nestlings  were  fed  7.6  times/hr,  and  although  both  males  and  females  fed  the  young,  the  participation 
of  females  was  significantly  greater  than  that  of  males.  Predation  was  the  major  cause  of  nest  failure.  Daily 
survival  rates  during  the  incubation  (0.990)  and  nestling  (0.935)  stages  differed.  Overall  nesting  success  was 
36%.  Although  studies  conducted  in  disturbed  areas  can  reveal  greater  rates  of  nest  predation  than  those  found 
in  undisturbed  areas,  some  Sporophila  species  seem  to  benefit  from  habitat  disturbance.  The  conversion  of  native 
habitats  to  agricultural  lands  in  Brazil,  as  well  as  the  spread  of  exotic  grasses,  has  resulted  in  the  expansion  of 
the  Double-collared  Seedeater  to  previously  forested  areas.  Received  14  February  2005,  accepted  16  November 
2005. 


The  genus  Sporophila  (Emberizidae)  com- 
prises a diverse  group  of  small  finches  widely 
distributed  in  the  Neotropics.  The  greatest  di- 
versity is  reached  in  interior  South  America, 
where  most  species  inhabit  grassy  semi-open 
areas  (Ridgely  and  Tudor  1994,  Sick  1997). 
However,  detailed  information  on  breeding  bi- 
ology is  lacking  for  most  of  these  species. 
Furthermore,  the  melodious  songs  of  these 
seedeaters  make  them  vulnerable  to  pursuit 
for  the  illegal  pet  trade.  As  a result,  many  spe- 
cies have  been  locally  extirpated,  and  some 
are  severely  threatened  (Collar  et  al.  1992, 
Willis  and  Oniki  1992,  Ridgely  and  Tudor 
1994,  Sick  1997,  Willis  2003). 

The  Double-collared  Seedeater  ( S . caeru- 
lescens) is  the  most  common  seedeater  in 
southern  South  America.  It  inhabits  grasslands 
and  agricultural  areas  (Ridgely  and  Tudor 
1994),  commonly  near  populated  locations. 
Recently,  it  has  expanded  its  distribution  in 
response  to  the  destruction  of  forested  areas 
and  the  consequent  spread  of  exotic  grass  spe- 
cies (Sick  1997).  Although  not  endangered, 
entire  populations  of  the  Double-collared 
Seedeater  have  been  lost  to  the  illegal  pet 
trade,  being  one  of  the  most  popular  cage 


1 Depto.  de  Genetica  e Evolut^ao,  Univ.  Federal  de 
Sao  Carlos,  Rodovia  Washington  Luis,  km  235,  P.O. 
Box  676,  CEP  13565-905,  Sao  Carlos,  SP,  Brazil; 
e-mail:  mercivalfrancisco@uol.com.br 


birds  in  Brazil.  Information  on  its  breeding 
biology  is  limited  to  descriptions  of  nests  and 
eggs  (Euler  1900,  Ihering  1900,  Pereyra  1956, 
De  La  Pena  1981,  Alabarce  1987)  and  the 
length  of  the  nestling  period — obtained  from 
a single  nest  observed  in  Argentina  (Pereyra 
1956).  More  information  on  the  species’  ecol- 
ogy is  needed  before  meaningful  conservation 
objectives  can  be  developed  for  the  species. 
Herein,  I describe  the  reproductive  biology  of 
the  Double-collared  Seedeater  in  southeastern 
Brazil.  Phenology  and  duration  of  the  breed- 
ing season,  length  of  incubation  and  nestling 
periods,  egg  mass,  nest  success,  and  infor- 
mation on  parental  care  are  reported. 

METHODS 

Study  area. — I conducted  my  study  on  the 
campus  of  Sao  Carlos  Federal  University,  lo- 
cated in  the  central  region  of  Sao  Paulo  state, 
southeastern  Brazil  (21°  58'  S,  47°  52'  W). 
The  campus  is  subdivided  into  a semi-urban- 
ized  portion  and  an  adjacent  non-urbanized, 
disturbed  cerrado  area  (savanna  that  ranges 
from  open  grasslands  to  forested  areas,  such 
as  gallery  forests  that  grow  alongside  water- 
courses; Eiten  1972).  The  semi-urbanized  area 
totals  187  ha,  and  is  composed  of  extensive 
lawns,  orchards,  gardens,  and  Eucalyptus  spp. 
and  Pinus  spp.,  with  regenerating  cerrado  un- 
dergrowth. Buildings  and  streets  are  widely 
spaced  and  compose  only  23  ha  (12%).  The 


85 


86 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


non-urbanized  area  is  a 529.6-ha  mosaic  com- 
posed of  sensu  lato  cerrado  (125  ha),  gallery 
forest  (3.6  ha),  regenerating  cerrado  (84  ha), 
an  abandoned  Eucalyptus  spp.  plantation  with 
regenerating  cerrado  undergrowth  (94  ha), 
and  active  Eucalyptus  spp.  silviculture  (223 
ha)  (Paese  1997).  The  climate  in  this  region 
is  tropical,  with  two  well-marked  seasons:  a 
humid,  hot  season  from  October  through 
March  and  a dry,  cold  season  from  April 
through  September.  In  both  the  semi-urban- 
ized  and  non-urbanized  areas,  grass  seeds  are 
abundant  during  the  wet  season. 

During  seven  breeding  seasons  (1997- 
2003),  I conducted  nest  searches  from  early 
November,  when  males  started  to  sing  and  de- 
fend territories,  to  May,  when  males  stopped 
singing.  All  habitats  were  searched  for  nests. 
Although  I conducted  nest  searches  weekly, 
the  number  of  habitats  covered  and  search  ef- 
fort varied  on  each  field  survey.  Nests  were 
located  by  searching  the  areas  defended  by 
males  and  by  following  females  observed  near 
these  territories.  Using  a metal  caliper  (accu- 
rate to  0.01  mm),  I measured  nests  and  eggs, 
and  I used  a spring  scale  (accurate  to  0.1  g) 
to  weigh  eggs. 

Using  a 7 X 35  binocular,  I observed  nests 
during  60-min  periods  to  calculate  the  fre- 
quency of  feeding  visits  and  to  estimate  the 
proportion  of  time  that  females  spent  incubat- 
ing the  eggs.  These  observations  were  always 
made  early  in  the  morning  (06:00-10:00 
UTC  — 3)  and  while  maintaining  a minimum 
distance  of  20  m from  the  nests.  During  the 
nestling  stage,  only  nests  containing  two 
young  (the  most  frequent  brood  size)  were 
considered  for  observations.  The  nestling 
stage  was  subdivided  into  three  observation 
periods:  early  (1—4  days  after  hatching),  mid- 
dle (5-9  days  after  hatching),  and  late  (10-13 
days  after  hatching;  Roper  and  Goldstein 
1997).  I used  the  Kruskal-Wallis  test  to  com- 
pare the  frequencies  of  feeding  trips  among 
these  periods.  To  compare  the  number  of 
times  that  males  and  females  fed  the  young,  I 
used  the  Mann-Whitney  £/-test. 

1 checked  nests  every  1—3  days.  Predation 
was  assumed  to  have  occurred  when  eggs  or 
nestlings  younger  than  fledging  age  disap- 
peared from  a nest.  Abandonment  was  as- 
sumed if  adults  were  not  seen  on  or  near  the 
nest  and  the  eggs  were  cold  or  the  nestlings 


were  dead  (Pletschet  and  Kelly  1990).  When- 
ever possible,  I checked  nests  from  a distance. 
By  using  binoculars,  I was  able  to  see  eggs 
and  young  through  the  thin  nest  walls,  thus 
avoiding  observer  disturbance  (see  Roper  and 
Goldstein  1997).  I estimated  rates  of  daily 
nest  survival  during  the  incubation  and  nest- 
ling stages  by  using  the  Mayfield  method 
(Mayfield  1961),  and  compared  them  accord- 
ing to  Sauer  and  Williams  (1989)  by  using 
program  CONTRAST  (Hines  and  Sauer 
1989).  One  to  six  nests  of  each  stage  were 
analyzed  per  year  in  order  to  calculate  surviv- 
al rate,  but  because  of  small  sample  sizes, 
years  were  pooled.  Means  of  daily  survival 
rate  are  presented  ± SE;  all  other  means  are 
presented  ± SD.  I calculated  standard  errors 
according  to  Johnson  (1979).  Nesting  success 
(probability  of  survival)  from  incubation 
through  fledging  was  also  estimated  following 
Mayfield  (1961). 

RESULTS 

I found  41  active  nests,  26  in  the  semi-ur- 
ban  area  and  15  in  disturbed  cerrado.  Nests 
were  found  in  all  habitats  except  gallery  forest 
and  active  Eucalyptus  spp.  plantations.  Males 
started  defending  territories  in  early  Novem- 
ber, and  I found  the  earliest  nest  on  18  De- 
cember 1999.  The  nest  contained  two  eggs  in 
the  late  stage  of  incubation,  suggesting  that 
breeding  activities  had  started  in  early  Decem- 
ber. The  latest  nesting  activity  was  recorded 
on  9 May  1997,  when  I observed  the  last 
young  in  a nest. 

Nests  were  cup-shaped  and  built  of  thin 
grass  roots  and  spiderweb  silk.  The  walls  were 
thin,  as  the  eggs  and  young  could  be  seen 
through  them.  The  eggs  were  white  or  slightly 
greenish,  with  dark  and  light  brown  spots, 
sometimes  concentrated  at  the  large  end  of  the 
egg  (Euler  1900,  Ihering  1900,  De  La  Pena 
1981).  The  height  of  nests  above  ground 
ranged  from  0.6  to  6 m (2.4  ± 1.2,  n = 25). 
I also  measured  outside  diameter  (6.7  cm  ± 
0.8,  n — 19),  inside  diameter  (5.2  cm  ± 0.7, 
n = 19),  inside  height  (4.0  cm  ± 0.6,  n — 18), 
and  outside  height  (4.8  cm  ± 0.7,  n — 19)  of 
the  nests.  Egg  measurements  were  length  = 
17.7  mm  ± 0.5,  n — 11;  width  =13  mm  ± 
0.5,  n = 11;  and  weight  = 1.4  g ± 0.5,  n — 
11. 

Double-collared  Seedeaters  did  not  appear 


Francisco  • BREEDING  BIOLOGY  OF  SPOROPH/LA  CAFRULESCENS 


87 


to  select  any  particular  plant  species  for  nest 
construction.  Eighteen  species  belonging  to  1 1 
different  families  were  identified,  including 
the  exotic  Pinus  spp.  (Pinaceae),  Cupressus 
spp.  (Cupressaceae),  Eriobotrya  japonica 
(Rosaceae),  Michelia  champaca  (Magnoli- 
aceae),  Ligustrum  lucidum  (Oleaceae),  Mur- 
raya  exotica , Citrus  sp.  (Rutaceae),  and  Eu- 
patorium  sp.  (Asteraceae).  Native  plant  spe- 
cies included  Piptocarpha  rotundifolia , Ver- 
nonia  sp.  (Asteraceae),  Didymopanax  vinosum 
(Araliaceae),  Miconia  albicans , Tibouchina 
granulosa  (Melastomataceae),  Machaerium 
acutifolium,  Caesalpinia  peltophoroides, 
Sweetia  elegans,  Sibipiruna  sibipiruna  (Fa- 
baceae),  and  Casearia  silvestris  (Flacourti- 
aceae). 

All  nests  observed  during  incubation  con- 
tained two  eggs  ( n = 27).  Eggs  were  laid  on 
consecutive  days  and  incubation  started  the 
morning  the  female  laid  the  last  egg  (first  day 
of  incubation).  Hatching  occurred  on  the 
morning  of  the  13th  day  (n  — 4 nests).  During 
33  hr  of  focal  observations  at  seven  different 
nests,  I observed  only  females  incubating  the 
eggs.  Males  did  not  feed  females  on  the  nests. 
The  mean  time  spent  incubating/hr  was  52.3 
min  ± 5.8  (range  = 41.2-60  min),  and  incu- 
bation recesses  were  6.6  min  ± 4.4  (range  = 
0.3—18.7  min,  n = 23). 

Nestlings  fledged  in  12-15  days  (mean  = 
13.3  ± 1.2,  n = 8),  and  invariably,  nestlings 
from  the  same  nest  fledged  on  the  same  day 
(n  = 4 nests).  They  left  the  nests  with  poorly 
developed  feathers  and  weak  flight  capabili- 
ties. In  34  hr  of  focal  observations  at  1 1 dif- 
ferent nests,  nestlings  were  fed  an  average  of 
7.6  ± 4.3  times/hr.  The  number  of  feeding  vis- 
its/hr increased  throughout  the  nestling  period 
(Fig.  1),  and  although  both  males  and  females 
fed  the  young,  the  participation  of  females 
(4.8  visits/hr  ± 2.4)  was  significantly  greater 
than  that  of  males  (2.7  ± 2.5;  U = 341.0,  P 
= 0.001). 

Females  regularly  brooded  nestlings  after 
feedings  (until  the  young  were  up  to  7 days 
old),  and  both  males  and  females  removed  fe- 
cal sacs.  In  one  territory,  adults  fed  one  fledg- 
ling and  young  nestlings  at  the  same  time, 
suggesting  that  the  nestlings  represented  at 
least  a second  brood  for  that  breeding  season. 
On  several  occasions,  one  or  both  adults  of  a 


Nestling  stage 

FIG.  1 . Average  number  of  feedings/hr  in  the  early 
(n  = 14  hr  at  eight  different  nests),  middle  (n  = 9 hr 
at  five  different  nests),  and  late  (n  - 1 1 hr  at  six  dif- 
ferent nests)  nestling  stages.  Error  bars  are  SDs.  The 
frequency  of  feedings  differed  among  the  stages  (Krus- 
kal-Wallis  H = 16.38,  P < 0.001). 

pair  were  observed  chasing  intruding  Double- 
collared  Seedeaters  that  approached  nests. 

Apart  from  one  nest  that  fell  down  during 
a storm,  predation  was  the  only  cause  of  nest 
failure.  No  nests  were  abandoned  and  no  eggs 
were  infertile.  Daily  survival  during  incuba- 
tion was  0.990  ± 0.010  (one  predation  event 
in  104  nest  days,  n = 12  nests).  Survival  dur- 
ing the  nestling  stage  was  0.935  ± 0.024  (sev- 
en predation  events  in  107  nest  days,  n = 13 
nests).  Nest  survival  was  higher  during  incu- 
bation (11  of  12)  than  during  the  nestling 
stage  (6  of  13;  x2  = 4.5,  df  = 1,  P = 0.033). 
Nesting  success  from  incubation  to  fledging 
was  36%. 

The  mean  number  of  female  arrivals  and 
departures  from  nests  during  the  incubation 
stage  was  1.9  ± 2.0/hr  ( n = 33  hr).  During 
the  nestling  stage,  the  mean  number  of  paren- 
tal arrivals  and  departures  was  15.7  ± 9.2  (n 
= 34  hr).  The  mean  number  of  parental  de- 
partures and  arrivals  per  hr  was  greater  during 
the  nestling  stage  than  it  was  during  incuba- 
tion (U  = 31.5,  P < 0.001). 

DISCUSSION 

The  nesting  season  of  Double-collared 
Seedeaters  began  in  December,  which  is  late 
compared  with  the  onset  of  breeding  season 
for  most  passerine  birds  inhabiting  cerrado 
(i.e.,  they  usually  start  in  September;  Sick 
1997).  Nesting  in  Double-collared  Seedeaters 


88 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


coincided  with  the  fruiting  period  of  exotic 
Gramineae  species,  seeds  of  which  are  fed  to 
nestlings.  Entire  seeds  of  Brachiaria  sp.  were 
observed  in  the  crops  of  nestlings.  Other  seed- 
eaters  found  in  the  study  region,  such  as  S. 
lineola  and  S.  leucoptera,  shared  the  same 
breeding  period  (MRF  pers.  obs.). 

The  nests  were  similar  to  those  described 
for  other  Sporophila  species  (e.g.,  S.  collaris, 
S.  ruficollis,  S.  albogularis,  S.  nigricollis,  and 
S.  lineola;  Alderton  1961;  ffrench  1965;  De 
La  Pena  1981;  Marcondes-Machado  1982, 
1997),  and  the  cup  shape  and  thin  walls  were 
typical  of  those  constructed  by  congeners 
(Sick  1997).  The  use  of  spiderwebs  in  nest 
construction  has  also  been  recorded  for  S.  ruf- 
icollis (De  La  Pena  1981)  and  S.  nigricollis 
(ffrench  1965).  Nests  were  not  reused,  as  pre- 
viously reported  for  S.  nigricollis  (Alderton 
1961),  S.  albogularis , and  S.  lineola  (Marcon- 
des-Machado 1982,  1997),  but  females  re- 
used the  material  of  old  nests  to  build  new 
nests. 

The  incubation  period  of  12  days  was  sim- 
ilar to  that  of  S.  nigricollis  and  S.  americana 
(it  lasts  13  days  for  S.  torqueola).  For  S.  cae- 
rulescens,  Pereyra  (1956)  reported  a nestling 
period  of  13—14  days,  and  I observed  a mean 
of  13.25  days.  The  nestling  period  is  8-9  days 
for  S.  nigricollis,  11—13  days  for  S.  ameri- 
cana, 1 1 days  for  S.  torqueola , and  9 days  for 
S.  lineola  (Skutch  1945,  Gross  1952,  Alderton 
1961,  Marcondes-Machado  1997).  Overall, 
both  the  incubation  and  nestling  periods  re- 
ported for  the  Sporophila  species  are  among 
the  shortest  of  Neotropical,  open-cup  nesting 
Passeriformes.  Although  some  nests  of  S. 
americana,  S.  lineola  (Skutch  1945,  Gross 
1952,  Marcondes-Machado  1997),  and  S.  cae- 
rulescens  (De  La  Pena  1981),  have  been 
found  containing  three  eggs  or  young,  two 
seems  to  be  the  usual  brood  size  for  Sporo- 
phila species. 

Predation  was  by  far  the  major  factor  lim- 
iting nesting  success  in  S.  caerulescens,  sim- 
ilar to  reports  for  many  other  open-cup  nesting 
Neotropical  passerines  (Skutch  1949,  1985; 
Snow  1976;  Oniki  1979;  Roper  and  Goldstein 
1997;  Martin  et  al.  2000;  Mezquida  and  Ma- 
rone  2000).  My  data  support  the  hypothesis 
that  parental  activity  may  increase  the  risk  of 
nest  predation  (Skutch  1949,  1985).  The  num- 
ber of  adult  departures  from,  and  arrivals  to. 


nests  were  much  greater,  and  daily  survival 
was  lower  during  the  nestling  stage.  Skutch’s 
hypothesis  predicts  that  the  primary  predators 
should  be  diurnal  and  visually  oriented.  How- 
ever, in  addition  to  parental  activities,  nests 
containing  nestlings  may  be  more  conspicu- 
ous due  to  the  noise  (Haskell  1994,  1999; 
Dearborn  1999)  and  odor  of  the  young,  which 
would  attract  nocturnal  mammalian  predators 
that  use  olfaction.  Nestlings  vocalized  only 
when  parents  were  feeding  them,  and  the  beg- 
ging calls  were  audible  from  15  m when 
broods  were  7-8  days  old,  and  from  about  20 
m when  young  were  in  the  late  nestling  stage. 

Although  little  is  known  about  nest  preda- 
tors in  the  Neotropics,  preliminary  observa- 
tions and  video  data  have  shown  diurnal  birds 
to  be  the  most  important  predators  in  environ- 
ments other  than  wet  forests  (Martin  et  al. 
2000,  Mezquida  and  Marone  2002).  Potential 
predators  in  the  study  area  included  Burrow- 
ing Owl  {Athene  cunicularia;  Mezquida  and 
Marone  2000),  Guira  Cuckoo  {Guira  guira; 
Mason  1985),  Squirrel  Cuckoo  {Piaya  cay- 
ana),  and  anis  {Crotophaga  spp.;  Telleria  and 
Diaz  1995).  During  my  study,  I observed  a 
Great  Kiskadee  {Pitangus  sulphuratus ) prey- 
ing upon  a Double-collared  Seedeater  nest.  I 
have  also  observed  Plush-crested  Jays  ( Cyan - 
ocorax  chrysops)  feeding  on  Common  Quail 
{Coturnix  coturnix)  eggs  placed  in  artificial 
cup-shaped  nests  (MRF  unpubl.  data),  which 
suggests  their  potential  as  a predator  of  seed- 
eaters,  as  well.  Potential  nocturnal  mammalian 
predators  occurring  in  the  study  area  included 
white-eared  opossum  ( Didelphis  albiventris), 
crab-eating  raccoon  {Procyon  cancrivorus), 
grison  (Galictis  vittata ),  striped  hog-nosed 
skunk  ( Conepatus  semistriatus),  tayra  ( Eira 
barbara),  jaguarundi  ( Herpailurus  yaguaron- 
di ),  and  house  cats  ( Felis  catus ). 

Studies  conducted  in  disturbed  areas  can  re- 
veal greater  rates  of  nest  predation  than  those 
in  undisturbed  areas  due  to  the  increased 
abundance  of  mesopredators  in  disturbed  ar- 
eas (Martin  1996,  Martin  et  al.  2000).  How- 
ever, some  Sporophila  species  seem  to  benefit 
from  habitat  disturbance.  Before  its  expansion 
into  anthropogenic  habitats,  the  niche  occu- 
pied by  the  Double-collared  Seedeater  was 
probably  limited  to  non-forested  areas,  such 
as  forest  borders,  cerrados,  and  wetlands 
where  native  grasses  occurred.  Today,  the  in- 


Francisco  • BREEDING  BIOLOGY  OF  SPOROPHILA  CA FRIJLESCFNS 


89 


creasing  extension  of  agricultural  areas  in 
Brazil,  as  well  as  the  spread  of  exotic  grasses, 
has  resulted  in  the  expansion  of  Double-col- 
lared Seedeaters  to  areas  previously  covered 
by  forests.  Gross  (1952)  and  ffrench  (1965) 
provide  additional  records  of  the  expansion  of 
S.  americana  and  S.  nigricollis  into  anthro- 
pogenic habitats. 

ACKNOWLEDGMENTS 

I am  grateful  to  N.  Arguedas,  H.  L.  Gibbs,  M.  Ro- 
drigues, M.  A.  Pizo,  M.  Galetti,  and  three  anonymous 
referees  for  their  important  suggestions  on  previous 
versions  of  this  manuscript,  and  M.  I.  S.  Lima  for  iden- 
tifying plant  species.  This  study  was  supported  by 
Coordenagao  de  Aperfeigoamento  de  Pessoal  de  Nivel 
Superior  (CAPES),  Conselho  Nacional  de  Desenvol- 
vimento  Cientifico  e Tecnologico  (CNPq)  and  Funda- 
gao  de  Amparo  a Pesquisa  do  Estado  de  Sao  Paulo 
(FAPESP). 

LITERATURE  CITED 

Alabarce,  E.  A.  1987.  Notas  sobre  la  biologia  de  al- 
gunos  Passeriformes  del  noroeste  argentino-I. 
Acta  Zoologica  Lilloana  39:23-27. 

Alderton,  C.  C.  1961.  The  breeding  cycle  of  the  Yel- 
low-bellied Seedeater  in  Panama.  Condor  63:390- 
398. 

Collar,  N.  J.,  L.  P.  Gonzaga,  N.  Krabbe,  A.  N.  Man- 
drono,  L.  G.  Naranjo,  T.  A.  Parker,  III,  and  D. 
C.  Wege.  1992.  Threatened  birds  of  the  Americas: 
the  ICBP/IUCN  red  data  book,  3rd  ed.,  part  2. 
Smithsonian  Institution  Press,  Washington,  D.C., 
and  International  Council  for  Bird  Preservation, 
Cambridge,  United  Kingdom. 

Dearborn,  D.  C.  1999.  Brown-headed  Cowbird  nest- 
ling vocalizations  and  risk  of  nest  predation.  Auk 
116:448-457. 

De  La  Pena,  M.  R.  1981.  Notas  nidobiologicas  sobre 
corbatitas  (Aves,  Emberizidae):  segunda  parte. 
Historia  Natural  6:45-48. 

Eiten,  G.  1972.  The  cerrado  vegetation  of  Brazil.  Bo- 
tanical Review  38:201-341. 

Euler,  C.  1900.  Descrigao  de  ninhos  e ovos  das  aves 
do  Brasil.  Revista  do  Museu  Paulista  4:9-148. 
ffrench,  R.  P.  1965.  The  nesting  behavior  of  the  Yel- 
low-bellied Seedeater.  Caribbean  Journal  of  Sci- 
ence 5:149-156. 

Gross,  A.  O.  1952.  Nesting  of  Hicks’  Seedeater  at 
Barro  Colorado  Island,  canal  zone.  Auk  69:433- 
446. 

Haskell,  D.  G.  1994.  Experimental  evidence  that  nest- 
ling begging  behavior  incurs  a cost  due  to  nest 
predation.  Proceedings  of  the  Royal  Society  of 
London,  Series  B 257:161-164. 

Haskell,  D.  G.  1999.  The  effect  of  predation  on  beg- 
ging-call evolution  in  nestling  wood  warblers.  An- 
imal Behaviour  57:893-901. 

Hines,  J.  E.  and  J.  R.  Sauer.  1989.  Program  CON- 


TRAST: a general  program  for  the  analysis  of  sev- 
eral survival  or  recovery  rate  estimates.  Fish  and 
Wildlife  Technical  Report  24:1-7. 

Ihering,  H.  von.  1900.  Catalogo  critico-comparativo 
dos  ninhos  e ovos  das  aves  do  Brasil.  Revista  do 
Museu  Paulista  4:191-300. 

Johnson,  D.  H.  1979.  Estimating  nest  success:  the  May- 
field  method  and  an  alternative.  Auk  96:651-661. 

Marcondes-Machado,  L.  O.  1982.  Notas  sobre  a re- 
produgao  de  Sporophila  alhogularis  (Spix  1825) 
(Passeriformes,  Emberizidae)  em  cativeiro.  Iher- 
ingia  61:81-89. 

Marcondes-Machado,  L.  O.  1997.  Comportamento 
reprodutivo  de  Sporophila  lineola  (Linnaeus) 
(Passeriformes,  Emberizidae).  Revista  Brasileira 
de  Zoologia  14:517-522. 

Martin,  T.  E.  1996.  Life  history  evolution  in  tropical 
and  south  temperate  birds:  what  do  we  really 
know?  Journal  of  Avian  Biology  27:263-271. 

Martin,  T.  E.,  P.  R.  Martin,  C.  R.  Olson,  B.  J.  Hei- 
dinger,  AND  J.  J.  Fontana.  2000.  Parental  care 
and  clutch  sizes  in  North  and  South  American 
birds.  Science  287:1482-1485. 

Mason,  P.  1985.  The  nesting  biology  of  some  passer- 
ines of  Buenos  Aires,  Argentina.  Ornithological 
Monographs  36:954-972. 

Mayfield,  H.  1961.  Nesting  success  calculated  from 
exposure.  Wilson  Bulletin  73:255-261. 

Mezquida,  E.  T.  and  L.  Marone.  2000.  Breeding  bi- 
ology of  Gray-crowned  Tyrannulet  in  the  Monte 
Desert,  Argentina.  Condor  102:205-210. 

Mezquida,  E.  T.  and  L.  Marone.  2002.  Microhabitat 
structure  and  avian  nest  predation  risk  in  an  open 
Argentinean  woodland:  an  experimental  study. 
Acta  Oecologica  23:313-320. 

Oniki,  Y.  1979.  Is  nesting  success  of  birds  low  in  the 
tropics?  Biotropica  11:60-69. 

Paese,  A.  1997.  Caracterizagao  e analise  ambiental  do 
campus  da  Universidade  Federal  de  Sao  Carlos 
(UFSCar),  Sao  Carlos,  Sao  Paulo.  M.Sc.  thesis, 
UFSCar,  Sao  Carlos,  Sao  Paulo,  Brazil. 

Pereyra,  J.  A.  1956.  Notas  biologicas  sobre  el  cor- 
batita  comun.  El  Hornero  10:140-142. 

Pletschet,  S.  M.  and  J.  F.  Kelly.  1990.  Breeding  bi- 
ology and  nest  success  of  Palila.  Condor  92:1012- 
1021. 

Ridgely,  R.  S.  and  G.  Tudor.  1994.  The  birds  of 
South  America,  vol.  2.  University  of  Texas  Press, 
Austin. 

Roper,  J.  J.  and  R.  R.  Goldstein.  1997.  A test  of  the 
Skutch  hypothesis:  does  activity  at  nests  increase 
nest  predation  risk?  Journal  of  Avian  Biology  28: 
111-116. 

Sauer,  J.  R.  and  B.  K.  Williams.  1989.  Generalized 
procedures  for  testing  hypotheses  about  survival 
or  recovery  rates.  Journal  of  Wildlife  Management 
53:137-142. 

Sick,  H.  1997.  Ornitologia  Brasileira.  Editora  Nova 
Fronteira,  Rio  de  Janeiro,  Brasil. 

Skutch,  A.  F.  1945.  Incubation  and  nestling  periods  of 
Central  American  birds.  Auk  62:8-37. 


90 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


Skutch,  A.  F.  1949.  Do  tropical  birds  rear  as  many 
young  as  they  can  nourish?  Ibis  91:430-455. 

Skutch,  A.  F.  1985.  Clutch  size,  nesting  success,  and 
predation  on  nests  of  Neotropical  birds,  reviewed. 
Ornithological  Monographs  36:575-594. 

Snow,  D.  W.  1976.  The  web  of  adaptation:  bird  studies 
in  the  American  tropics.  Cornell  University  Press, 
Ithaca,  New  York. 


Telleria,  J.  L.  and  M.  Diaz.  1995.  Avian  nest  predation 
in  a large  natural  gap  of  the  Amazonian  rainforest. 
Journal  of  Field  Ornithology  66:343-351. 

Willis,  E.  O.  2003.  Aves  do  Estado  de  Sao  Paulo. 

Divisa,  Rio  Claro,  Sao  Paulo,  Brasil. 

Willis,  E.  O.  and  Y.  Oniki.  1992.  Losses  of  Sao  Paulo 
birds  are  worse  in  the  interior  than  in  Atlantic  For- 
ests. Ciencia  e Cultura  44:326-328. 


The  Wilson  Journal  of  Ornithology  1 18(l):91-98,  2006 


SMALL  MAMMAL  SELECTION  BY  THE  WHITE-TAILED  HAWK  IN 

SOUTHEASTERN  BRAZIL 

MARCO  A.  MONTEIRO  GRANZINOLLI1 2 AND  JOSE  CARLOS  MOTTA-JUNIOR1 2 


ABSTRACT. — We  analyzed  diet  and  prey  selection  of  the  relatively  unknown  albicaudatus  subspecies  of  the 
White-tailed  Hawk  ( Buteo  albicaudatus).  Our  study  was  based  on  an  analysis  of  259  pellets  collected  from 
September  2000  to  September  2001  in  the  municipality  of  Juiz  de  Fora  in  southeastern  Brazil.  We  also  assessed 
the  abundance  of  small  mammals  with  pitfall  traps  (2,160  trap-nights).  Small  mammals  composed  52.5%  of  the 
estimated  biomass  consumed  by  the  hawks,  and  selection  appeared  to  be  mediated  by  abundance.  The  Bonferroni 
confidence  intervals  procedure  revealed  that  when  abundance  of  small  mammals  was  higher,  the  hawks  were 
selective,  preying  on  Calomys  tener  more  than  would  be  expected  by  chance  (P  < 0.05);  other  rodents  were 
consumed  less  than  expected.  Oligoryzomys  nigripes,  Oxymycterus  sp.,  and  Gracilinanus  spp.  were  taken  in  the 
same  proportion  as  they  were  found  in  the  field.  During  reduced  prey  abundance  (October-March),  White-tailed 
Hawks  preyed  opportunistically  on  small  mammals.  Differences  in  habits  and  vulnerability  of  small  mammals 
may  explain  prey  selectivity  in  the  White-tailed  Hawk.  Received  5 October  2004,  accepted  3 October  2005. 


The  White-tailed  Hawk  {Buteo  albicauda- 
tus) is  a poorly  known  species  ranging  from 
southern  Texas  to  northern  Argentinean  Pata- 
gonia (Farquhar  1992,  Thiollay  1994).  Infor- 
mation on  its  ecology  is  scarce  and  largely 
descriptive  or  anecdotal,  with  most  studies 
having  been  conducted  in  North  America 
(Stevenson  and  Meitzen  1946,  Kopeny  1988, 
Farquhar  1992).  Data  on  type  and  number  of 
prey  have  received  some  attention  in  Texas 
(see  Farquhar  1992),  but  prey  selection  rela- 
tive to  prey  abundance  remains  unknown. 
Only  three  studies  report  on  the  diet  of  this 
raptor  in  the  Neotropics.  Schubart  et  al.  (1965) 
examined  contents  of  two  stomachs  contain- 
ing mainly  insects;  Brasileiro  et  al.  (2003)  re- 
ported predation  on  a snake,  and  Motta- Junior 
and  Granzinolli  (2004)  observed  consumption 
of  a Ringed  Kingfisher  {Megaceryle  torqua- 
ta ).  The  species  is  thought  to  be  an  opportu- 
nistic predator  (Stevenson  and  Meitzen  1946, 
Kopeny  1988),  and  in  Texas,  half  of  the  prey 
biomass  comprises  mammals  (Farquhar 
1986). 

Opportunistic  predators  generally  take  prey 
in  accordance  with  their  abundance  in  the 
field,  whereas  selective  predators  consume 
prey  in  proportions  that  differ  from  those 
available  (Jaksic  1989).  This  selectivity  or  op- 
portunism may  be  explained  in  relation  to  the 
energy  costs  and  benefits  involved  in  the  cap- 


1  Depto.  de  Ecologia,  Instituto  de  Biociencias,  Univ. 
de  Sao  Paulo,  05508-900  Sao  Paulo,  SP,  Brazil. 

2 Corresponding  author;  e-mail:  mgranzi@usp.br 


ture  and  handling  of  prey.  Predators  may  con- 
sume the  most  profitable,  but  not  necessarily 
the  most  abundant,  prey  (Schoener  1971,  Kor- 
pimaki  1985,  Stephens  and  Krebs  1986,  Iriarte 
et  al.  1989,  Jaksic  1989).  According  to  opti- 
mal foraging  theory,  predators  behave  to  max- 
imize their  fitness,  which  is  done  by  maxi- 
mizing their  net  rate  of  energy  intake  (Emlen 
1966,  1968;  Schoener  1971;  Stephens  and 
Krebs  1986).  Thus,  prey  selection  by  a pred- 
ator not  only  depends  on  prey  energy  content, 
but  also  on  the  predator’s  success  in  three  ba- 
sic stages:  finding,  handling,  and  consuming 
prey.  Selectivity  can  be  assessed  by  observing 
differences  among  the  prey  species  at  any  of 
these  steps.  Prey  selectivity  may  be  a result  of 
both  prey  and  predator  morphology  and  be- 
havior (Corley  et  al.  1995).  Emlen  (1966, 
1968)  hypothesized  that  predators  will  exhibit 
a greater  degree  of  dietary  selection  when 
their  prey  are  abundant,  but  will  be  more  op- 
portunistic when  food  is  scarce.  Additionally, 
a predator  may  eat  more  abundant  prey  at 
greater  frequencies  than  expected  in  relation 
to  abundance  (Emlen  1966).  Here,  we  analyze 
prey  selection  by  the  White-tailed  Hawk  rel- 
ative to  prey  abundance,  evaluating  previous 
assertions  about  the  opportunistic  feeding  be- 
havior of  this  species  (Stevenson  and  Meitzen 
1946,  Farquhar  1986,  Kopeny  1988). 

METHODS 

Study  site. — We  conducted  fieldwork  on  pri- 
vate farmlands  in  northern  Juiz  de  Fora  (21° 
41'  S,  43°  27'  W),  in  the  state  of  Minas  Gerais 


91 


92 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  1 , A/c/rc/?  2P06 


in  southeastern  Brazil  (Fig.  1).  The  elevation 
of  our  study  area  (17,537  ha)  ranged  from  670 
to  800  m;  the  topography  is  mountainous.  The 
climate  is  Humid  Subtropical,  winters  are  dry, 
and  annual  rainfall  averages  1,536  mm.  The 
wet  season  extends  from  October  to  April 
(192  mm  rainfall,  mean  temperature  = 20.2° 
C),  and  the  dry  season  occurs  from  May  to 
September  (37  mm  rainfall,  mean  temperature 
= 16.8°  C).  Originally,  the  dominant  vegeta- 
tion was  semi-deciduous  forest;  now  the  area 
is  primarily  farmland,  pastures,  patches  of 
second-growth  vegetation,  and  plantations  of 
exotics  (e.g..  Eucalyptus  spp.  and  Pinus  spp.; 
Juiz  de  Fora  1996). 

General  diet. — The  analysis  of  the  White- 
tailed Hawk’s  diet  was  based  on  259  pellets, 
which  we  collected  from  seven  nesting  and  six 
roosting  sites  of  approximately  seven  pairs. 
We  collected  and  identified  (by  size  and 
shape)  all  pellets  from  perches  used  exclu- 
sively by  White-tailed  Hawks.  We  oven-dried 
the  collected  material  and  treated  it  with  a 
10%  NaOH  aqueous  solution  (Marti  1987). 
Prior  to  chemical  treatment,  we  removed  re- 
mains of  scales,  fur,  and  feathers,  and  later 
added  them  to  other  remains,  such  as  mandi- 
bles, teeth,  and  invertebrate  exoskeletons.  We 
identified  remains  by  comparing  them  to  a ref- 
erence collection  from  the  study  area.  Inver- 
tebrates were  generally  identified  to  family 
and  order,  whereas  vertebrates  were  identified 
mostly  to  genus  or  species.  Prey  biomass  was 
estimated  by  counting  the  minimum  number 
of  individuals  in  pellets  and  then  multiplying 
this  number  by  the  mean  body  mass  of  each 
species  at  the  study  site  (Marti  1987). 

Prey  selection. — We  estimated  the  relative 
abundance  of  small  mammals  in  the  field  by 
monitoring  five  sets  of  drift-fence  pitfall  traps 
(Friend  et  al.  1989).  Traps  were  distributed 
systematically  around  most  of  the  hawks’ 
hunting  sites  (Fig.  1),  determined  before  and 
during  the  study  period  through  observations 
of  foraging  individuals.  We  collected  pellets 
during  small  mammal  trapping.  Each  set  of 
pitfall  traps  consisted  of  12  buckets  (36  1 
each),  totaling  60  traps.  From  September  2000 
to  September  200 1 , we  operated  traps  monthly 
for  3 consecutive  days,  totaling  2,160  trap- 
nights.  Captured  mammals  were  identified, 
weighed,  sexed,  earmarked,  and  released.  An 
index  of  small  mammal  abundance  for  each 


month  was  based  on  the  total  number  of  in- 
dividual first  captures  (recaptures  were  not 
counted). 

Indices  of  prey  abundance  are  assumed  to 
reflect  prey  availability,  but  this  may  not  nec- 
essarily be  true  (Jaksic  1989).  Traps  should  be 
efficient,  nonselective,  and  catch  the  entire 
range  of  small  mammal  prey.  Moreover,  traps 
should  be  placed  in  patches  where  and  when 
the  predator  hunts.  Our  procedures  fulfilled 
these  assumptions,  in  terms  of  both  time  and 
place  of  foraging.  Our  traps  were  open  24  hr 
per  day,  so  that  both  diurnal  and  crepuscular 
activities  of  White-tailed  Hawks  were  ac- 
counted for  by  the  trapping  procedures.  Pitfall 
traps  appear  to  be  less  selective  and  more  ef- 
ficient, capturing  larger  numbers  of  species, 
individuals,  and  age  classes  compared  with 
traditional  live  traps  (Williams  and  Braun 
1983;  MAMG  unpubl.  data). 

Analyses. — We  conducted  G-tests  to  test  the 
goodness-of-fit  of  the  frequency  distributions 
of  prey  in  the  diet  and  in  the  field  (Zar  1984). 
We  interpreted  nonsignificant  results  to  mean 
that  White-tailed  Hawks  exploited  prey  in 
proportion  to  their  abundance  in  the  field;  sig- 
nificant differences  suggested  that  the  hawks 
“preferred”  or  “avoided”  some  small  mam- 
mal species,  hence  apparently  selecting  or 
avoiding  prey.  To  confirm  selection  or  avoid- 
ance of  prey,  we  used  the  Bonferroni  confi- 
dence intervals  procedure  for  each  prey  spe- 
cies (Neu  et  al.  1974,  Byers  et  al.  1984, 
Plumpton  and  Lutz  1993,  Martinez  and  Jaksic 
1997,  McLoughlin  et  al.  2002).  If  the  expect- 
ed proportion  of  consumption  was  not  includ- 
ed in  the  confidence  interval,  then  the  ob- 
served and  expected  consumption  differed 
significantly.  If  the  confidence  interval  includ- 
ed the  expected  proportion  of  consumption, 
then  the  hypothesis  that  prey  species  were  pre- 
ferred or  avoided  was  rejected.  All  tests  were 
considered  significant  at  P < 0.05. 

RESULTS  AND  DISCUSSION 

General  diet. — Numerically,  the  main  prey 
were  insects,  followed  by  small  mammals, 
reptiles,  and  birds  (Fig.  2).  Small  mammals 
composed  the  bulk  of  biomass,  followed  by 
insects,  reptiles,  and  birds.  Our  results  are 
similar  to  those  of  Stevenson  and  Meitzen 
(1946),  Farquhar  (1986),  and  Kopeny  (1988). 

Only  5 of  12  genera  of  small  mammals 


Granzinolli  and  Motto-Junior  • PREY  SELECTION  BY  THE  WHITE-TAILED  HAWK 


93 


7617814 


7613814 


7609814 


7605814 


7601814 


7597814 


662845  666845  670845 


662845  666845  670845 


7617814 


7613814 


7609814 


7605814 


7601814 


7597814 


1 g g 1_9 3_9 5_8 7_8  km 

Scale  1:97000 

FIG.  I.  Satellite  image  (LANDS AT  7/ETM,  27  June  2000)  of  study  area  in  Juiz  de  Fora  municipality,  Minas 
Gerais,  southeastern  Brazil.  Coordinate  grid  system  is  UTM  (Zone  22,  Corrego  Alegre).  White  squares  are  sites 
of  pitfall  traps;  white  circles  are  nest  and  perch  sites  of  White-tailed  Hawks. 


94 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


Insecta  Other  Reptilia  and  Aves  Mammalia 

Arthropoda  Amphibia 


FIG.  2.  Number  of  individuals  and  estimated  biomass  of  prey  groups  consumed  by  White-tailed  Hawks 
from  September  2000  to  September  2001,  Juiz  de  Fora  municipality,  Minas  Gerais,  southeastern  Brazil. 


( Calomys , Akodon,  Oligoryzomys,  Oxymycte- 
rus, Gracilinanus)  found  in  the  study  area 
(Appendix)  were  identified  in  White-tailed 
Hawk  pellets.  The  genus  Akodon  was  repre- 
sented mostly  by  A.  lindberghi,  with  some  A. 
cursor ; both  were  found  in  pellets  and  in  pit- 
fall  traps.  The  seven  genera  whose  remains 
were  not  found  in  pellets  were  uncommon: 
only  12  individuals  (4.6%  of  total  captures) 
were  trapped  in  pitfalls  (Appendix).  Prey  be- 
havior or  habitat  choice  may  explain  the  ab- 
sence of  some  genera  in  the  diet  of  White- 
tailed Hawks.  Rhagomys , Oryzomys,  and  Ju- 
liomys  {—Wilfredomys)  have  arboreal  or  scan- 


sorial  habits,  whereas  Thaptomys , Bibimys , 
Bolomys , and  Blarinomys  display  subterra- 
nean or  fossorial  habits,  and  all  but  Bolomys 
and  A.  lindberghi  inhabit  mostly  forests  (Em- 
mons 1990,  Eisenberg  and  Redford  1999,  No- 
wak 1999;  JCM-J  pers.  obs.).  Furthermore,  al- 
though the  genus  Oxymycterus  was  as  uncom- 
mon as  the  seven  genera  not  recorded  in 
White-tailed  Hawk  pellets,  its  habitat  is  most- 
ly open  vegetation  (MAMG  unpubl.  data). 

Prey  selection. — White-tailed  Hawks  exhib- 
ited differential  predation  on  small  mammal 
species  when  both  seasons  were  combined  (G 
= 32.54,  P < 0.001;  Table  1).  The  same  pat- 


TABLE  1.  Small  mammal  prey  selection  by  White-tailed  Hawks  in  Juiz  de  Fora  municipality,  Minas  Gerais, 
southeastern  Brazil,  from  September  2000  to  September  2001.  Observed  values  (Obs)  are  actual  frequencies  in 
the  diet;  expected  values  (Exp)  are  frequencies  calculated  from  proportions  obtained  in  the  field  by  pitfall 
trapping. 


Species 

Dry  season 

Wet  season 

Total  diet 

Obs 

Exp 

Obs 

Exp 

Obs 

Exp 

Akodon  spp. 

1 1 

33.5 

6 

7.5 

17 

40.8 

Calomys  tener 

95 

59.1 

18 

23.7 

113 

83.1 

Oligoryzomys  nigripes 

24 

41.3 

14 

6.8 

38 

47.9 

Oxymycterus  sp.a 

2 

0.7 

1 

— 

3 

0.7 

Gracilinanus  spp.a 

4 

1.4 

1 

2.0 

5 

3.5 

Total 

136 

136.0 

40 

40.0 

176 

176.0 

G" 

52.07 

7.68 

32.54 

P 

<0.001 

0.054 

<0.001 

a Oxymycterus  sp.  and  Gracilinanus  spp.  were  grouped  for  G-tests. 
b G-test,  df  = 3. 


95 


Granzinolli  and  Motta-Junior  • PREY  SELECTION  BY  THE  WHITE-TAILED  HAWK 


tern  was  observed  during  the  dry  season  (G 
= 52.07,  P < 0.001),  but  not  in  the  wet 
months  (G  - 7.68,  P = 0.054;  Table  1). 

The  Bonferroni  confidence  intervals  pro- 
cedure revealed  that  in  the  dry  season,  the 
hawks  preyed  more  on  Calomys  tener  and  less 
on  Akodon  spp.  than  expected  based  on  trap- 
ping data  (Table  2).  Conversely,  in  wet 
months,  there  were  no  differences  in  small 
mammal  predation  compared  with  the  avail- 
ability of  small  mammals  in  the  study  area 
(Table  2).  Oligoryzomys  nigripes,  Oxymycte- 
rus  sp.,  and  Gracilinanus  spp.  were  always 
consumed  in  the  same  proportion  that  they 
were  found  in  the  environment  (Table  2). 
Hence,  our  findings  are  not  entirely  congruent 
with  those  of  Stevenson  and  Meitzen  (1946) 
and  Kopeny  (1988). 

Other  studies  on  small  mammal  populations 
in  southeastern  Brazil  indicate  peaks  of  abun- 
dance during  the  dry  season  (e.g.,  Motta-Ju- 
nior  1996,  Vieira  1997,  Talamoni  and  Dias 
1999).  The  same  pattern  was  observed  in  our 
study  (Fig.  3). 

The  high  frequency  of  C.  tener  (sometimes 
considered  a subspecies  of  C.  laucha\  Eisen- 
berg  and  Redford  1999)  in  the  White-tailed 
Hawk’s  diet  may  be  due  to  its  higher  vulner- 
ability. A similar  suggestion  was  proposed  by 
Corley  et  al.  (1995)  for  other  rodent  and  pred- 
ator species  in  Patagonia.  A less  vulnerable 
species  ( Eligmodontia  typus,  better  escape 
ability)  was  preyed  upon  less  than  expected 
by  the  culpeo  fox  ( Dusicyon  culpaeus ),  while 
the  behaviorally  and  morphologically  vulner- 
able Akodon  spp.  were  consumed  more  fre- 
quently than  expected.  Other  diet  studies  of 
owls  (Motta-Junior  1996,  Motta-Junior  and 
Bueno  2004,  Motta-Junior  et  al.  2004)  in 
southeastern  Brazil  have  revealed  that  C.  tener 
is  one  of  the  main  prey  species,  despite  not 
being  the  most  abundant  in  the  field,  suggest- 
ing higher  vulnerability.  C.  tener  is  apparently 
mainly  terrestrial  and  does  not  dig  burrows 
(Eisenberg  and  Redford  1999,  Nowak  1999); 
thus,  it  is  more  vulnerable  because  it  is  likely 
to  be  more  conspicuous  to  the  hawks.  In  con- 
trast, species  of  Akodon  travel  in  tunnels  un- 
der the  leaf  litter  and  nest  in  burrows  (Em- 
mons 1990);  thus,  Akodon  spp.  may  be  able 
to  escape  White-tailed  Hawk  predation  more 
efficiently  than  C.  tener. 

Our  results  suggest  that  prey  selection  by 


Akodon  spp.  0.081  0.304  0.020  < Pl , < 0.141  (-)  0.150  0.187  0.004  < Pi  < 0.295  (0) 

Calomys  tener  0.699  0.435  0.597  < Pi  < 0.799  ( + ) 0.450  0.593  0.247  < P,  < 0.652  (0) 

Oligoryzomys  nigripes  0.176  0.246  0.092  < Pi  < 0.260  (0)  0.350  0.170  0.155  < Pi  ^ 0.544  (0) 

Oxymycterus  sp.  0.015  0.005  0.000  < p,  < 0.041  (0)  0.025  0.000  0.000  < p,  < 0.088  (0) 

Gracilinanus  spp.  0.029  0.010  0.000  < Pi  < 0.066  (0)  0.025  0.050  0.000  < Pi  < 0.088  (0) 


96 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  J,  March  2006 


FIG.  3.  Small  mammal  abundance  from  September  2000  to  September  2001,  Juiz  de  Fora  municipality, 
Minas  Gerais,  southeastern  Brazil.  Data  were  not  available  for  November  2000. 


White-tailed  Hawks  was  mediated  by  prey 
abundance.  When  the  abundance  of  small 
mammals  was  higher  (dry  season),  the  hawks 
selected  the  more  abundant  prey,  Calomys  te- 
ner  (Table  2).  However,  during  a period  of 
lower  abundance  of  prey  (wet  season),  White- 
tailed Hawks  were  opportunistic  relative  to 
small  mammal  species.  Our  results  support  the 
prediction  of  Emlen  ( 1 966)  that  predators  feed 
selectively  on  very  abundant  prey,  thus  sug- 
gesting that  White-tailed  Hawks  exploit  re- 
sources depending  on  their  availability. 

In  conclusion.  White-tailed  Hawks  seem  to 
prey  selectively  on  a more  vulnerable  small 
mammal  (C.  tener ),  which  has  terrestrial  hab- 
its and  uses  open  habitat.  The  semi-fossorial 
Akodon  spp.  were  apparently  less  vulnerable 
to  the  hawks.  Alternatively,  but  not  exclusive- 
ly, our  results  support  Emlen’s  (1966)  hypoth- 
esis that  predators,  in  times  of  high  prey  abun- 
dance, will  prey  selectively  on  species  that  are 
more  abundant.  Further  studies  of  raptor  diet 
selection  in  the  Neotropics  should  stress  mor- 
phological and  behavioral  traits  of  prey  as  a 
way  to  understand  differential  vulnerability  to 
predators  (e.g.,  Kotler  1985,  Corley  et  al. 
1995). 

ACKNOWLEDGMENTS 

We  thank  C.  V.  Rios  for  assistance  in  the  field.  M. 
V.  Vieira,  R.  P.  Kavanagh,  C.  C.  Farquhar.  A.  A. 


Bueno,  and  two  anonymous  referees  provided  helpful 
comments  on  previous  versions  of  our  manuscript.  Fi- 
nancial support  was  provided  by  the  Conselho  Na- 
cional  de  Desenvolvimento  Cientffico  e Tecnologico 
and  the  Fundagao  de  Amparo  a Pesquisa  do  Estado  de 
Sao  Paulo. 

LITERATURE  CITED 

Brasileiro,  C.  A.,  M.  Martins,  and  M.  C.  Kiefer. 
2003.  Lystrophis  nattereri  (NCN).  Predation.  Her- 
petological  Review  34:70. 

Byers,  C.  R.,  R.  K.  Steinhorst,  and  P.  R.  Krausman. 
1984.  Clarification  of  a technique  for  analyses  of 
utilization-availability  data.  Journal  of  Wildlife 
Management  48:1050-1053. 

Corley,  J.  C.,  G.  J.  Fernandez,  A.  F.  Capurro,  A.  J. 
Novaro,  M.  C.  Funes,  and  A.  Travaini.  1995. 
Selection  of  cricetine  prey  by  the  culpeo  fox  in 
Patagonia:  a differential  prey  vulnerability  hy- 
pothesis. Mammalia  59:315-325. 

Eisenberg,  J.  F.  and  K.  H.  Redford.  1999.  Mammals 
of  the  Neotropics,  vol.  3:  the  central  Neotropics. 
University  of  Chicago  Press,  Chicago,  Illinois. 
Emlen,  J.  M.  1966.  The  role  of  time  and  energy  in 
food  preference.  American  Naturalist  100:611- 
617. 

Emlen,  J.  M.  1968.  Optimal  choice  in  animals.  Amer- 
ican Naturalist  102:385-389. 

Emmons,  L.  H.  1990.  Neotropical  rainforest  mammals: 
a field  guide.  University  of  Chicago  Press,  Chi- 
cago, Illinois. 

Farquhar,  C.  C.  1986.  Ecology  and  breeding  behavior 
of  the  White-tailed  Hawk  in  the  northern  Coastal 
Prairies  of  Texas.  Ph.D.  dissertation,  Texas  A&M 
University,  College  Station. 


Granzinolli  and  Motto-Junior  • PREY  SELECTION  BY  THE  WHITE-TAILED  HAWK 


97 


Farquhar,  C.  C.  1992.  White-tailed  Hawk  ( Buteo  al- 
bicaudatus).  The  Birds  of  North  America,  no.  30. 

Friend,  G.  R.,  G.  T.  Smith,  D.  S.  Mitchell,  and  C. 
R.  Dickman.  1989.  Influence  of  pitfall  and  drift 
fence  on  capture  rates  of  small  vertebrates  in 
semi-arid  habitats  of  Western  Australia.  Australian 
Wildlife  Research  16:1-10. 

Iriarte,  J.  A.,  J.  E.  Jimenez,  L.  C.  Contreras,  and  F. 
Jaksic.  1989.  Small  mammal  availability  and  con- 
sumption by  the  fox,  Dusicyon  culpaeus , in  cen- 
tral Chile  scrublands.  Journal  of  Mammalogy  70: 
641-645. 

Jaksic,  F.  1989.  Opportunism  vs.  selectivity  among 
carnivorous  predators  that  eat  mammalian  prey:  a 
statistical  test  of  hypotheses.  Oikos  56:427-430. 

Juiz  de  Fora.  1996.  Plano  Diretor  de  Juiz  de  Fora. 
Juiz  de  Fora:  Instituto  de  Pesquisa  e Planejamento. 
Juiz  de  Fora,  Minas  Gerais,  Brazil. 

Kopeny,  M.  T.  1988.  White-tailed  Hawk.  Pages  97- 
104  in  Proceedings  of  the  southwest  raptor  man- 
agement symposium  and  workshop  (R.  L.  Glinski, 
B.  G.  Pendleton,  and  M.  B.  Moss,  Eds.).  Scientific 
and  Technical  Series,  no.  1 1 . National  Wildlife 
Federation,  Washington,  D.C. 

Korpimaki,  E.  1985.  Prey  choice  strategies  of  the  Kes- 
trel Falco  tinnunculus  in  relation  to  available 
small  mammals  and  other  Finnish  birds  of  prey. 
Annales  Zoologici  Fennici  22:91-104. 

Kotler,  B.  P.  1985.  Owl  predation  on  desert  rodents 
which  differ  in  morphology  and  behavior.  Journal 
of  Mammalogy  66:824-828. 

Marti,  C.  D.  1987.  Raptor  food  habits  studies.  Pages 
67-69  in  Raptor  management  techniques  manual 
(B.  A.  Giron  Pendleton,  B.  A Millsap,  K.  W. 
Cline,  and  D.  M.  Bird,  Eds.).  Scientific  and  Tech- 
nical Series,  no.  10.  National  Wildlife  Federation, 
Washington,  D.C. 

Martinez,  D.  R.  and  F.  M.  Jaksic.  1997.  Selective 
predation  on  scansorial  and  arboreal  mammals  by 
Rufous-legged  Owls  (Strix  rufipes)  in  southern 
Chilean  rainforest.  Journal  of  Raptor  Research  3 1 : 
370-375. 

McLoughlin,  P.  D.,  H.  D.  Cluff,  and  F.  Messier. 
2002.  Denning  ecology  of  barren-ground  grizzly 
bears  in  the  central  Arctic.  Journal  of  Mammalogy 
83:188-198. 

Motta-Junior,  J.  C.  1996.  Ecologia  alimentar  de  co- 
rujas  (Aves:  Strigiformes)  na  regiao  central  do  Es- 
tado  de  Sao  Paulo:  biomassa,  sazonalidade  e se- 
letividade  de  suas  presas.  Ph.D.  thesis,  Universi- 
dade  Federal  de  Sao  Carlos,  Sao  Carlos,  Brazil. 

Motta-Junior,  J.  C.,  C.  J.  R.  Alho,  and  S.  C.  S.  Be- 
lentani.  2004.  Food  habits  of  the  Striped  Owl 
Asio  clamator  in  southeast  Brazil.  Pages  777-784 
in  Raptors  worldwide:  proceedings  of  the  VI 


world  conference  on  birds  of  prey  and  owls  (R. 
Chancellor  and  B.-U.  Meyburg,  Eds.).  World 
Working  Group  on  Birds  of  Prey  and  Owls,  MME 
BirdLife  Hungary,  Budapest. 

Motta-Junior,  J.  C.  and  A.  A.  Bueno.  2004.  Trophic 
ecology  of  the  Burrowing  Owl  in  southeast  Brazil. 
Pages  763-775  in  Raptors  worldwide:  proceed- 
ings of  the  VI  world  conference  on  birds  of  prey 
and  owls  (R.  Chancellor  and  B.-U.  Meyburg, 
Eds.).  World  Working  Group  on  Birds  of  Prey  and 
Owls,  MME  BirdLife  Hungary,  Budapest. 

Motta-Junior,  J.  C.  and  M.  A.  M.  Granzinolli. 
2004.  Consumption  of  a Ringed  Kingfisher  ( Me - 
gaceryle  torquata ) by  White-tailed  Hawk  ( Buteo 
albicaudatus ) in  southeastern  Brazil.  Journal  of 
Raptor  Research  38:191. 

Neu,  C.  W.,  C.  R.  Byers,  and  J.  M.  Peek.  1974.  A 
technique  for  analysis  of  utilization-availability 
data.  Journal  of  Wildlife  Management  38:541- 
545. 

Nowak,  R.  M.  1999.  Walker’s  mammals  of  the  world, 
vol.  2,  6th  ed.  The  Johns  Hopkins  University 
Press,  Baltimore,  Maryland. 

Plumpton,  D.  L.  and  R.  S.  Lutz.  1993.  Prey  selection 
and  food  habits  of  Burrowing  Owls  in  Colorado. 
Great  Basin  Naturalist  53:299-304. 

Schoener,  T.  W.  1971.  Theory  of  feeding  strategies. 
Annual  Review  of  Ecology  and  Systematics  2: 
369-404. 

Schubart,  O.,  A.  C.  Aguirre,  and  H.  Sick.  1965. 
Contribui^ao  para  o conhecimento  da  alimenta^ao 
das  aves  brasileiras.  Arquivos  de  Zoologia  de  Sao 
Paulo  12:95-249. 

Stephens,  D.  W.  and  J.  R.  Krebs.  1986.  Foraging  the- 
ory. Princeton  University  Press,  Princeton,  New 
Jersey. 

Stevenson,  J.  O.  and  L.  H.  Meitzen.  1946.  Behavior 
and  food  habits  of  Sennett’s  White-tailed-Hawk  in 
Texas.  Wilson  Bulletin  58:198-205. 

Talamoni,  S.  A.  and  M.  M.  Dias.  1999.  Population 
and  community  ecology  of  small  mammals  in 
southeastern  Brazil.  Mammalia  63:167-181. 

Thiollay,  J.  M.  1994.  Family  Accipitridae.  Pages  52- 
205  in  Handbook  of  the  birds  of  the  world,  vol. 
2:  New  World  vultures  to  guineafowl  (J.  del  Hoyo, 
A.  Elliot,  and  J.  Sargatal,  Eds.).  Lynx  Edicions, 
Barcelona,  Spain. 

Vieira,  M.  V.  1997.  Dynamics  of  a rodent  assemblage 
in  a cerrado  of  southeastern  Brazil.  Revista  Bras- 
ileira  de  Biologia  57:99-107. 

Williams,  D.  F.  and  S.  E.  Braun.  1983.  Comparison 
of  pitfall  and  conventional  traps  for  sampling 
small  mammal  populations.  Journal  of  Wildlife 
Management  47:841-845. 

Zar,  J.  H.  1984.  Biostatistical  analysis,  2nd  ed.  Pren- 
tice Hall,  Englewood  Cliffs,  New  Jersey. 


98 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


APPENDIX.  Rodents  and  opossums  ( Gracilinanus  spp.)  captured  in  pitfall  traps  in  Juiz  de  Fora  municipality, 
Minas  Gerais,  southeastern  Brazil,  from  September  2000  to  September  2001.  For  each  month,  we  tallied  only 
first  captures.  Data  were  not  available  for  November  2000. 


Species 

weight  (g) 

Sep 

Oct 

Dec 

Jan 

Feb 

Mar 

Apr 

May 

Jun 

Jul 

Aug 

Sep 

Total 

Akodon  cursor* 

17 

1 

b 

— 

1 

— 

1 

1 

2 

— 

— 

— 

— 

6 

Akodon  lindberghi a 

13 

— 

3 

2 

— 

— 

2 

1 

10 

•7 

14 

4 

9 

52 

Bibimys  labiosus 

19 

— 

— 

1 

— 

— 

— 

— 

1 

— 

— 

— 

— 

2 

Blarinomys  breviceps 

12 

1 

1 

Bolomys  lasiurus 

24 

1 

1 

Calomys  tener* 

12 

22 

6 

3 

5 

5 

5 

1 1 

12 

17 

24 

2 

6 

118 

Gracilinanus  agilis a 

20 

1 

— 

1 

— 

— 

— 

— 

— 

1 

— 

— 

— 

3 

Gracilinanus  spp.a 

19 

— 

1 

— 

1 

— 

— 

— 

— 

— 

— 

— 

— 

2 

Juliomys  sp. 

20 

1 

1 

Oligoryzomys  cf.  flavescens 

18 

1 

1 

Oligoryzomys  nigripes a 

11 

4 

1 

1 

1 

1 

1 

5 

6 

13 

18 

6 

10 

67 

Oryzomys  cf.  kelloggi 

29 

2 

1 

3 

Oxymycterus  sp.a 

73 

1 

1 

Thaptomys  nigrita 

22 

— 

— 

— 

— 

— 

1 

— 

— 

— 

— 

— 

— 

1 

Rhagomys  rufescens 

27 

1 

1 

— 

2 

Total 

30 

11 

9 

9 

6 

10 

18 

31 

38 

57 

15 

27 

261 

a Species  preyed  on  by  White-tailed  Hawks. 
b — represents  no  captures. 


Short  Communications 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):99- 101,  2006 


Provisioning  of  Fledgling  Conspecifics  by  Males  of  the  Brood-parasitic 
Cuckoos  Chrysococcyx  klaas  and  C.  caprius 

Irby  J.  Lovette,14  Dustin  R.  Rubenstein,1 2-23  and  Wilson  Nderitu  Watetu3 4 


ABSTRACT. — Although  post-fledging  care  by  adult 
males  seems  unlikely  in  bird  species  that  are  obligate, 
interspecific  brood  parasites,  there  have  been  numer- 
ous reports  of  adult  male  Chrysococcyx  cuckoos  ap- 
parently feeding  conspecific  young.  Most  researchers 
currently  view  these  observations  with  skepticism,  in 
large  part  because  Chrysococcyx  and  other  cuckoo  spe- 
cies engage  in  courtship  feeding,  and  it  is  possible  that 
field  observers  could  mistake  adult  females  receiving 
food  from  courting  males  for  fledglings,  especially  giv- 
en the  similar  appearances  of  females  and  juveniles. 
Here,  we  report  an  observation  of  an  extended  provi- 
sioning bout  by  an  adult  male  Klaas’s  Cuckoo  (C. 
klaas)  feeding  a conspecific  individual  with  juvenile 
plumage  and  behavior,  and  we  summarize  our  obser- 
vations of  similar  occurrences  in  the  Diederik  Cuckoo 
(C.  caprius ) in  Kenya.  We  suggest  that  the  available 
evidence  indicates  that  male  provisioning,  and  hence 
potential  parental  care,  is  present  in  these  brood-para- 
sitic cuckoos  at  a higher  frequency  than  currently  rec- 
ognized. The  mechanism  that  causes  males  to  associate 
with  fledglings  is  unknown,  but  warrants  further  study. 
Received  20  December  2004,  accepted  19  September 
2005. 


The  genus  Chrysococcyx  comprises  15  spe- 
cies of  small.  Old  World  cuckoos  (Sibley  and 
Monroe  1990),  of  which  all  are  thought  to  be 
obligate  brood  parasites  (Davies  2000). 
Klaas’s  Cuckoo  (C.  klaas)  has  a wide  distri- 
bution in  sub-Saharan  Africa,  where  it  is 
known  to  parasitize  a large  number  of  passer- 
ine host  taxa,  often- — but  not  exclusively — 
species  of  Sylviidae  and  Nectarinidae  (Irwin 
1988).  Similarly,  the  Diederik  Cuckoo  (C.  ca- 
prius) breeds  throughout  much  of  sub-Saharan 
Africa  and  has  a broad  range  of  hosts,  pri- 
marily species  of  Ploceidae  (Irwin  1988). 


1 Cornell  Lab.  of  Ornithology,  159  Sapsucker 
Woods  Rd„  Ithaca,  NY  14850,  USA. 

2 Dept,  of  Neurobiology  and  Behavior,  Cornell 
Univ.,  Seeley  G.  Mudd  Hall,  Ithaca,  NY  14853,  USA. 

3 Mpala  Research  Centre,  Box  555,  Nanyuki,  Ken- 
ya. 

4 Corresponding  author;  e-mail:  IJL2@cornell.edu 


Over  the  past  century,  there  have  been  nu- 
merous observations  of  male  Chrysococcyx 
cuckoos  feeding  conspecifics  that  were 
thought  to  be  fledglings  (Moreau  1944,  Fried- 
mann 1968,  Iversen  and  Hill  1983,  Rowan 
1983).  In  a literature  review  of  provisioning 
behavior  in  brood  parasites,  Lorenzana  and 
Sealy  (1998)  found  5 records  of  nestling  or 
fledgling  provisioning  by  Klaas’s  Cuckoo 
males  and  1 1 such  records  for  Diederik  Cuck- 
oo males;  Friedmann  (1968)  discusses  12  and 
15  such  records,  respectively,  including  some 
anecdotal  reports.  There  is  apparently  only 
one  equivalent  report  of  a female  Chrysococ- 
cyx cuckoo  provisioning  fledglings,  and  in  that 
case,  both  the  female  and  young  were  captive 
birds  (Millar  1926,  Lorenzana  and  Sealy 
1998).  Historically,  a number  of  researchers 
(e.g.,  Moreau  1944,  Friedmann  1968)  consid- 
ered parental  care  to  be  common  in  African 
Chrysococcyx  cuckoos  and  believed  that  the 
behavior  might  be  a primitive  condition  as- 
sociated with  a relatively  recent  evolutionary 
transition  to  brood  parasitism.  As  researchers 
continued  to  document  the  prevalence  of 
courtship  feeding  in  these  and  other  cuckoo 
species,  more  recent  authorities  (e.g..  Rowan 
1983,  Irwin  1988,  Lorenzana  and  Sealy  1998, 
Davies  2000)  have  suggested  that  the  behavior 
is  either  misdirected  courtship  feeding  or  the 
result  of  human  observers  misidentifying 
adult  females  as  fledglings.  In  practice,  these 
and  other  possibilities  are  difficult  to  exclude. 
Although  the  plumages  of  adult  African  Chry- 
sococcyx are  highly  sexually  dimorphic,  it  is 
difficult  to  distinguish  females  from  juveniles 
in  the  field  (Rowan  1983). 

Here,  we  report  an  observation  of  an  ex- 
tended provisioning  bout  by  an  adult  male 
Klaas’s  Cuckoo  feeding  a conspecific  individ- 
ual with  juvenal  plumage  and  behavior,  and 
we  summarize  our  observations  of  similar  oc- 
currences in  the  Diederik  Cuckoo.  These  ob- 


99 


100 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


servations  add  to  the  body  of  evidence  sug- 
gesting that  male  Chrysococcyx  cuckoos  may 
engage  in  intensive  provisioning  of  juveniles. 

KLAAS’S  CUCKOO 

Beginning  at  10:08  UTC  + 3 on  15  August 
2004,  IJL,  DRR,  and  WNW  observed  an  adult 
male  (by  plumage)  Klaas’s  Cuckoo  in  Lake 
Nakuru  National  Park,  Kenya  (00°  22'  S,  36° 
03'  E).  This  bird  was  foraging  at  an  extremely 
rapid  rate  of  movement  in  the  open  canopy  of 
a large  yellow-barked  acacia  tree  ( Acacia  xan- 
thophloea ).  After  watching  the  bird  for  a few 
minutes,  we  observed  it  deliver  food  to  a sec- 
ond, sedentary  cuckoo  in  the  same  tree.  We 
noted  the  time,  and  for  the  next  26  min,  we 
were  able  to  keep  both  cuckoos  under  constant 
focal  observation  with  at  least  one  observer 
following  each  bird.  This  is  apparently  the 
longest-duration  period  of  potential  fledgling 
provisioning  reported  for  Chrysococcyx 
(Friedmann  1968). 

During  our  observation,  the  adult  male 
cuckoo  continued  to  forage  rapidly  within  an 
approximate  40-m  radius  around  the  second 
cuckoo.  The  male  returned  to  the  second 
cuckoo  18  times  while  carrying  food  items, 
all  of  which  appeared  to  be  1-  to  3-cm-long 
lepidopteran  larvae  gleaned  from  the  foliage 
and  bark  of  the  acacia.  On  16  of  the  18  visits, 
the  second,  more  sedentary  bird  accepted  and 
ate  the  caterpillar.  On  each  visit,  the  adult 
male  presented  the  food  with  his  tail  slightly 
cocked,  but  we  observed  no  other  conspicuous 
postures  or  behaviors  potentially  related  to 
courtship.  No  copulations  or  attempted  copu- 
lations occurred. 

During  our  observation,  the  presumed  ju- 
venile moved  among  four  perches,  flying  3—4 
m each  time.  These  flights  were  notably  more 
fluttery  than  those  of  the  adult  male  and  ap- 
peared typical  of  the  weak  flight  exhibited  by 
recently  fledged  birds.  While  perched,  this 
bird  also  assumed  the  “fluffed”  posture  typi- 
cal of  recent  fledglings,  and  it  remained  sta- 
tionary between  most  provisioning  visits.  The 
observation  ended  when  the  presumed  juve- 
nile made  a similar,  but  slightly  longer  flight 
into  denser  foliage  and  disappeared  from  our 
sight.  Although  the  plumages  of  female  and 
immature  Klaas’s  Cuckoos  are  variable  and 
overlap  (Irwin  1988),  we  noted  at  the  time 
that  the  bird  being  provisioned  had  a distinct 


white  patch  behind  the  eye  and  a white  throat 
marked  with  substantial,  dark  barring — plum- 
age characters  more  typical  of  juveniles  (Irwin 
1988). 

DIEDERIK  CUCKOO 

On  28  May  2002  at  08:23,  WNW  observed 
a male  Diederik  Cuckoo  feeding  an  apparent 
fledgling  (based  on  plumage)  at  the  Mpala  Re- 
search Centre,  Laikipia,  Kenya  (00°  17'  N, 
36°  54'  E).  The  fledgling  was  perched  about  3 
m above  ground  in  a Balanites  aegyptica  tree. 
During  15  min  of  observation,  the  adult  fed 
the  fledgling  at  least  four  times  and  continued 
to  do  so  when  the  observer  left  the  area.  On 
19  May  2003  at  10:15,  WNW  noted  similar 
behavior  at  a site  100  m from  that  of  the  first 
observation.  During  this  observation,  an  adult 
male  Diederik  Cuckoo  gleaned  insects  from 
long  grass  and  fed  them  to  a fledgling  (based 
on  plumage)  perched  on  a nearby  acacia.  We 
observed  the  male  make  six  feeding  trips  be- 
fore cattle  flushed  the  birds. 

DISCUSSION 

Based  on  the  posture,  behavior,  and  plum- 
age of  the  Klaas’s  Cuckoo  that  we  observed 
being  fed  by  an  adult  male,  it  seems  highly 
likely  that  it  was  a recently  fledged  bird  rather 
than  an  adult  female  being  courted.  We  also 
noted  that  the  adult  male  engaged  in  intensive 
(and,  presumably,  energetically  costly)  forag- 
ing for  an  extended  period  in  order  to  provi- 
sion this  individual.  Friedmann  (1968)  consid- 
ered provisioning  bouts  as  long  as  15  min  as 
“suggestive  of  the  fact  that  the  catering  adult 
was  not  merely  indulging  in  courtship  feed- 
ing.” Our  observation  of  an  intensive  provi- 
sioning period  of  nearly  twice  that  duration 
further  supports  this  interpretation.  In  contrast, 
courtship  feeding  in  Chrysococcyx  typically 
involves  a series  of  stereotyped  behaviors  that 
we  did  not  observe:  the  male’s  presentation  of 
food  while  simultaneously  cocking  his  head 
and  vibrating  his  wings  and  tail,  postural  bow- 
ing movements  by  the  male,  vocalizations  by 
the  male  or  both  individuals,  or  (in  some  cas- 
es) subsequent  copulation  (Irwin  1988). 

When  considered  in  concert,  our  observa- 
tions and  those  in  dozens  of  previous  reports 
describing  equivalent  behaviors  suggest  that 
males  of  several  African  Chrysococcyx  cuck- 
oos may  provision  fledglings  regularly.  Post- 


SHORT  COMMUNICATIONS 


101 


fledging  associations  of  adults  and  offspring 
also  have  been  documented  in  other  brood- 
parasitic  taxa,  such  as  the  Brown-headed 
Cowbird  ( Molothrus  ater\  Hahn  and  Fleischer 
1995).  Indeed,  previous  reports  have  docu- 
mented male  Klaas’s  and  Diederik  cuckoos 
provisioning  both  pre-volant  young  and  mul- 
tiple fledglings  (Moreau  1944,  Friedmann 
1968,  Lorenzana  and  Sealy  1998),  thus  ex- 
cluding misidentification  of  adult  females  as 
sufficient  explanation  for  this  behavior.  We 
speculate  that  not  only  are  females  sometimes 
misidentified  as  fledglings,  but  perhaps  older 
fledglings  being  provisioned  by  males  are 
sometimes  mistaken  for  females  being  court- 
ed. If  earlier  reports  were  correct  and  provi- 
sioning of  fledglings  by  adult  males  is  rela- 
tively common  in  the  African  Chrysococcyx, 
it  raises  interesting  questions  about  the  genetic 
relatedness  of  the  interacting  individuals  and 
their  underlying  social  system. 

ACKNOWLEDGMENTS 

We  thank  M.  Muchai  and  the  many  additional  staff 
members  of  the  Department  of  Ornithology,  National 
Museums  of  Kenya  who  have  greatly  facilitated  our 
field  research  in  their  country.  We  similarly  acknowl- 
edge the  Kenya  Wildlife  Service  and  the  Mpala  Re- 
search Centre  for  allowing  us  to  conduct  this  research. 


LITERATURE  CITED 

Davies,  N.  B.  2000.  Cuckoos,  cowbirds,  and  other 
cheats.  T.  & A.D.  Poyser,  London,  United  King- 
dom. 

Friedmann,  H.  1968.  The  evolutionary  history  of  the 
genus  Chrysococcyx.  U.S.  National  Museum  Bul- 
letin, no.  265,  Smithsonian  Institution,  Washing- 
ton, D.C. 

Hahn,  D.  C.  and  R.  C.  Fleischer.  1995.  DNA  finger- 
print similarity  between  female  and  juvenile 
Brown-headed  Cowbirds  trapped  together.  Animal 
Behaviour  49:1577-1580. 

Irwin,  M.  P.  S.  1988.  Order  Cuculiformes.  Pages  58- 
104  in  Birds  of  Africa,  vol.  Ill  (C.  H.  Fry,  S. 
Keith,  and  E.  K.  Urban,  Eds.).  Academic  Press, 
London,  United  Kingdom. 

Iverson,  E.  and  B.  Hill.  1983.  Diederik  Cuckoo  feeds 
fledgling  of  same  species.  Bee-eater  34:47. 

Lorenzana,  J.  C.  and  S.  G.  Sealy.  1998.  Adult  brood 
parasites  feeding  nestlings  and  fledglings  of  their 
own  species:  a review.  Journal  of  Field  Ornithol- 
ogy 69:364-375. 

Millar,  H.  M.  1926.  Golden  Cuckoo,  C.  cupilus. 
South  African  Journal  of  Natural  History  6:28-29. 

Moreau,  R.  E.  1944.  Food-bringing  by  African 
Bronze  Cuckoos.  Ibis  86:98-100. 

Rowan,  M.  K.  1983.  The  doves,  parrots,  louries,  and 
cuckoos  of  Southern  Africa.  David  Philip,  Cape 
Town,  South  Africa. 

Sibley,  C.  G.  and  B.  L.  Monroe.  1990.  Distribution 
and  taxonomy  of  birds  of  the  world.  Yale  Univer- 
sity Press,  New  Haven,  Connecticut. 


The  Wilson  Journal  of  Ornithology  1 1 8(  1 ):  101-104,  2006 


Widespread  Cannibalism  by  Fledglings  in  a Nesting  Colony  of 
Black-crowned  Night-Herons 

Christina  Riehl12 


ABSTRACT. — I studied  the  diet  and  foraging  be- 
havior of  fledgling  Black-crowned  Night-Herons  (Nyc- 
ticorax  nycticorax ) in  a mixed-species  nesting  colony 
of  Black-crowned  Night-Herons  and  Snowy  Egrets 
(Egretta  thula ) in  New  Orleans,  Louisiana.  In  1 of  5 
years,  cannibalism  of  nestlings  that  had  fallen  or 
climbed  out  of  nests  was  common,  accounting  for  66 
of  94  (70.2%)  prey  items  taken  by  fledglings.  Juveniles 
took  younger  conspecifics  by  both  predation  and  scav- 


1 5500  Camp  St.,  New  Orleans,  LA  70115,  USA. 

2 Current  address:  Dept,  of  Ecology  and  Evolution- 
ary Biology,  Princeton  Univ.,  Princeton,  NJ  08544, 
USA;  e-mail:  criehl@princeton.edu 


enging.  Isolated  incidents  of  cannibalism  among 
Black-crowned  Night-Herons  have  been  reported  pre- 
viously, but  not  on  a colony-wide  scale.  Received  2 
December  2004,  accepted  19  September  2005. 


Many  researchers  have  studied  the  diets  of 
adult  and  nestling  Black-crowned  Night-Her- 
ons ( Nycticorax  nycticorax ; Bent  1 926,  Palm- 
er 1962,  Wolford  and  Boag  1971),  but  there 
are  few  data  on  the  diet  and  foraging  behavior 
of  juveniles  immediately  after  leaving  the 
nest.  Here,  I provide  the  first  report  of  wide- 


102 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1/8,  No.  1,  March  2006 


spread  cannibalism  and  scavenging  of  conspe- 
cifics  among  fledglings  in  a nesting  colony  of 
Black-crowned  Night-Herons. 

METHODS 

From  1 February  to  18  July  2000,  I moni- 
tored a colony  of  Black-crowned  Night-Her- 
ons on  Ochsner  Island,  Audubon  Park,  New 
Orleans,  Fouisiana  (29°  56'  N,  90°  8'  W)  as 
part  of  a long-term  (1997-2001)  study  on  re- 
productive success.  Ochsner  Island  is  a small 
(600  m2)  island  in  a man-made  lagoon;  the 
distance  between  the  island  and  the  shore  of 
the  mainland  is  approximately  6 m.  The  is- 
land’s vegetation  is  dominated  by  Chinese  tal- 
lowtree  ( Sapium  sebiferum ) and  live  oak 
( Quercus  spp.).  In  2000,  143  pairs  of  Black- 
crowned  Night-Herons  and  10  pairs  of  Snowy 
Egrets  ( Egretta  thula ) nested  on  the  island. 
Nest  height  ranged  from  1 to  7 m above 
ground.  I recorded  the  diet  and  foraging  be- 
havior of  approximately  70  juvenile  night-her- 
ons from  fledging  until  the  end  of  the  breeding 
season,  when  the  members  of  the  nesting  col- 
ony dispersed.  Night-herons  were  considered 
to  have  fledged  when  they  left  the  nest  per- 
manently and  were  no  longer  fed  by  adults,  at 
which  point  most  were  capable  of  clumsy 
flight.  Prey  items  were  identified  by  direct  ob- 
servation of  foraging  night-herons.  Observa- 
tions were  made  from  the  mainland,  from 
which  approximately  half  of  the  nests  in  the 
colony  could  be  observed.  I observed  foraging 
juveniles  for  546  hr. 

RESUFTS 

Juvenile  Black-crowned  Night-Herons  were 
fed  by  parents  until  45  ± 3 (SD)  days  after 
hatching  ( n = 23).  However,  juveniles  were 
able  to  climb  out  of  the  nest  and  onto  sur- 
rounding vegetation  as  early  as  30  days  after 
hatching,  returning  to  the  nest  when  a parent 
approached  with  food.  At  35  days,  juveniles 
readily  left  their  nests,  often  climbing  out  of 
the  nest  to  solicit  food  from  a nearby  parent 
or  unrelated  adult  night-heron. 

Juveniles  remained  on  the  island  for  1—3 
weeks  after  leaving  the  nest  permanently, 
forming  small  groups  of  one  to  four  individ- 
uals from  the  same  nest,  or  neighboring  nests. 
Each  group  or  lone  individual  occupied  a 
small  (7—9  m2)  territory  on  the  ground  and 
defended  the  area  from  passing  adults  and  oth- 


er fledglings  (see  Noble  et  al.  1938  for  a full 
description  of  territoriality  in  juvenile  night- 
herons).  Fledglings  rarely  ventured  into  the 
water  to  hunt;  rather,  they  spent  most  of  their 
time  foraging  on  the  ground  under  active 
nests.  Of  94  prey  items  that  I saw  juvenile 
night-herons  consume,  66  (70.2%)  were  youn- 
ger fledgling  or  nestling  night-herons.  I ob- 
served juveniles  feeding  on  both  chicks  that 
they  killed  ( n = 20)  and  chicks  that  were  al- 
ready dead  when  I began  observations  ( n = 
46).  Other  prey  items  included  fish  (10.6%), 
frogs  (8.5%),  brown  rats  (Rattus  norvegicus\ 
4.3%),  carrion  dropped  from  active  nests 
(3.2%),  Wood  Duck  chicks  ( Aix  sponsa; 
2.1%),  and  a dead  Snowy  Egret  nestling 
(1.1%). 

Fledglings  did  not  prey  on  chicks  in  nests 
or  chicks  perched  in  vegetation;  they  limited 
their  attacks  to  nestlings  on  the  ground  that 
had  fallen  or  climbed  out  of  nests.  Adults  de- 
fended chicks  in  nests,  but  I never  observed 
adults  interfering  with  fledglings  that  were 
preying  on  chicks  on  the  ground.  Since  older 
night-heron  nestlings  often  left  the  nest  to 
perch  on  nearby  vegetation  before  fledging 
permanently,  it  was  not  always  clear  whether 
victims  were  nestlings  that  had  fallen  from 
nests  or  younger  fledglings  that  had  just  left 
the  nest.  It  is  probable,  however,  that  preda- 
tion by  fledgling  night-herons  increased  mor- 
tality rates  of  chicks  that  had  climbed  out  of 
the  nest  and  would  have  otherwise  been  able 
to  climb  to  safety.  Older  nestlings  in  low  nests 
(<1.5  m above  ground)  often  climbed  out  of 
the  nest  onto  the  ground  before  fledging,  and 
were  therefore  more  vulnerable  to  attacks  than 
nestlings  in  high  nests. 

Small,  weak,  and  moribund  chicks  were  at- 
tacked more  frequently  than  healthy-looking 
nestlings  near  the  age  of  fledging.  The  victims 
were  approximately  50-70%  of  the  size  of 
fledglings  and  appeared  difficult  to  kill  and 
consume.  Fledglings  killed  younger  conspe- 
cifics  by  striking  them  with  their  bills  for  up 
to  1 hr  or  more,  and  then  consumed  them  by 
repeatedly  striking  the  carcasses  and  labori- 
ously tugging  small  pieces  of  meat  from  them. 

Older  fledglings  were  particularly  skilled  at 
preying  on  nestlings  and  appeared  to  focus 
their  foraging  attempts  on  nestlings  to  the  ex- 
clusion of  other  prey.  When  a fledgling  found 
an  undefended  nestling  and  began  to  attack  it. 


SHORT  COMMUNICATIONS 


103 


other  fledglings  usually  came  to  fight  over  the 
victim.  In  one  case,  I observed  five  fledglings 
attack  and  consume  a 15-day-old  nestling  that 
had  fallen  from  its  nest. 

DISCUSSION 

Black-crowned  Night-Herons  are  among 
the  most  opportunistic  of  North  American  her- 
ons. They  employ  several  different  foraging 
behaviors  (Kushlan  1976)  and  consume  a 
wide  variety  of  prey,  including  fish,  mollusks, 
insects,  reptiles,  amphibians,  rodents,  birds, 
eggs,  carrion,  refuse,  and  plants  (Hancock  and 
Kushlan  1984,  Davis  1993).  Night-herons  will 
alter  their  foraging  methods  to  concentrate  on 
locally  abundant  resources,  including  mice 
(Allen  and  Mangels  1940),  fish  (Spanier 
1980),  and  amphibians  (Wetmore  1920).  They 
have  also  been  reported  to  systematically  ex- 
ploit rookeries  of  other  colonially  nesting 
birds,  including  Common  Terns  {Sterna  hirun- 
do\  Marshall  1942,  Collins  1970,  Shealer  and 
Kress  1991)  and  Franklin’s  Gulls  {Larus  pi- 
pixcan\  Wolford  and  Boag  1971).  Kale  (1965) 
reported  an  instance  of  adult  night-herons  in 
a colony  preying  opportunistically  on  White 
Ibis  ( Eudocimus  albus ) and  Great  Egret  ( Ar - 
dea  alba)  chicks  from  the  same  mixed-species 
rookery,  noting  that  ibis  and  egret  chicks  from 
neighboring  nests  constituted  a major  food 
source  for  night-heron  chicks.  Published  re- 
ports of  night-herons  feeding  on  conspecifics, 
however,  are  limited  to  Wolford  and  Boag’s 
(1971)  report  of  a night-heron  nestling  that 
was  regurgitated  by  another  nestling.  Williams 
and  Nicholson  (1977)  reported  a suspected  in- 
stance of  brood  reduction  in  the  Black- 
crowned  Night-Heron,  but  did  not  find  evi- 
dence of  cannibalism. 

There  is  virtually  no  information  on  the  for- 
aging behavior  of  night-heron  fledglings  dur- 
ing the  period  immediately  after  they  leave  the 
nest — after  the  adults  have  stopped  feeding 
them  but  before  they  become  adept  at  catching 
their  own  prey.  Lorenz  (1938)  and  Palmer 
(1962)  reported  that  fledglings  move  through 
the  colony  and  are  able  to  beg  food  from  any 
adult;  however,  Finley  (1906)  and  Noble  et  al. 
(1938)  found  that  adults  do  not  feed  juveniles 
on  the  ground.  Data  on  the  composition  of 
fledgling  diet  are  scarce,  possibly  because  re- 
cently fledged  juveniles  may  forage  at  night 
(Rockwell  1910,  Davis  1993).  In  this  study,  I 


found  that  juveniles  sometimes  climbed  back 
to  the  nest  in  the  first  2-3  days  after  fledging, 
and  were  usually  fed  by  the  parents.  After  3 
days  post-fledging,  fledglings  on  the  ground 
often  grabbed  the  bills  of  passing  adults  in  an 
attempt  to  stimulate  them  to  regurgitate  food, 
but  were  almost  always  unsuccessful. 

Fledglings  also  seemed  unable  to  fish  effi- 
ciently in  the  deep  water  surrounding  the  is- 
land, at  least  for  the  first  7 or  8 days  after 
fledging.  I frequently  observed  fledglings  in 
the  water  striking  at  floating  sticks  and  pieces 
of  leaves,  but  they  rarely  captured  live  prey. 
Fledglings  occasionally  picked  up  prey 
dropped  by  nestlings  in  active  nests;  on  one 
occasion,  a fledgling  climbed  into  a low  nest 
and  pulled  a fish  from  the  bill  of  the  fledgling 
to  which  it  had  just  been  delivered.  Adults,  by 
contrast,  were  never  observed  feeding  on  dead 
nestlings  or  other  carrion,  suggesting  that  they 
were  more  skilled  at  catching  higher-quality, 
live  prey. 

Although  I spent  similar  amounts  of  time 
observing  the  same  rookery  each  year  ( 1 997- 
2001),  cannibalism  among  Black-crowned 
Night-Heron  fledglings  was  prevalent  only  in 
2000.  I observed  night-heron  fledglings  feed- 
ing on  dead  night-heron  and  egret  chicks  only 
twice  in  1998  and  once  in  2001.  The  species 
composition  of  the  nesting  colony  was  fairly 
constant  across  years,  comprising  120-150 
pairs  of  Black-crowned  Night-Herons  and  5— 
10  pairs  of  Great  Egrets  and  Snowy  Egrets; 
thus,  the  level  of  competition  for  food  among 
fledglings  on  the  island  should  not  have  been 
elevated  in  2000.  In  other  years,  fledgling  di- 
ets were  dominated  by  fish  and  frogs.  How- 
ever, it  is  difficult  to  compare  prey  composi- 
tion across  years  because  I observed  far  fewer 
prey  captures  in  other  years,  possibly  because 
juvenile  Black-crowned  Night-Herons  may 
forage  mostly  at  night. 

It  is  possible  that  cannibalism  rates  were 
exceptionally  high  in  2000  because  local 
shortages  of  fish  or  other  live  prey  forced 
fledglings  to  seek  alternate  food  resources,  but 
I was  unable  to  document  such  a shortage.  A 
food  shortage  would  have  affected  the  diet 
and  foraging  patterns  of  fledglings  more  than 
adults  and  nestlings,  since  adults  often  left  the 
nesting  colony  to  forage  while  fledglings  re- 
mained on  the  island. 


104 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


ACKNOWLEDGMENTS 

I  am  grateful  to  W.  E.  Davis,  Jr.,  J.  A.  Kushlan,  and 

an  anonymous  reviewer  for  their  helpful  comments  on 

the  manuscript. 

LITERATURE  CITED 

Allen,  R.  P.  and  F.  P.  Mangels.  1940.  Studies  of  the 
nesting  behavior  of  the  Black-crowned  Night  Her- 
on. Proceedings  of  the  Linnean  Society  of  New 
York  50-51:1-28. 

Bent,  A.  C.  1926.  Life  histories  of  North  American 
marsh  birds.  U.S.  National  Museum  Bulletin,  no. 
135,  Smithsonian  Institution,  Washington,  D.C. 

Collins,  C.  T.  1970.  The  Black-crowned  Night  Heron 
as  a predator  of  tern  chicks.  Auk  87:584-586. 

Davis,  W.  E.,  Jr.  1993.  Black-crowned  Night-Heron 
( Nycticorax  nycticorax).  The  Birds  of  North 
America,  no.  74. 

Finley,  W.  L.  1906.  Herons  at  home.  Condor  8:35-40. 

Hancock,  J.  and  J.  A.  Kushlan.  1984.  The  herons 
handbook.  Harper  and  Row,  New  York. 

Kale,  H.  W.,  II.  1965.  Nestling  predation  by  herons  in 
a Georgia  heronry.  Oriole  30(March):69-70. 

Kushlan,  J.  A.  1976.  Feeding  behavior  of  North 
American  herons.  Auk  93:86-94. 

Lorenz,  K.  1938.  A contribution  to  the  comparative 
sociology  of  colonial-nesting  birds.  Proceedings 


of  the  International  Ornithological  Congress  8: 
207-218. 

Marshall,  N.  1942.  Night  desertion  by  nesting  Com- 
mon Terns.  Wilson  Bulletin  54:27-31. 

Noble,  G.  K.,  M.  Wurm,  and  A.  Schmidt.  1938.  So- 
cial behavior  of  the  Black-crowned  Night  Heron. 
Auk  55:7-40. 

Palmer,  R.  S.  1962.  Handbook  of  North  American 
birds,  vol.  I:  loons  through  flamingos.  Yale  Uni- 
versity Press,  New  Haven,  Connecticut. 

Rockwell,  R.  B.  1910.  Some  Colorado  night  heron 
notes.  Condor  12:113-121. 

Shealer,  D.  A.  and  S.  W.  Kress.  1991.  Nocturnal 
abandonment  response  to  Black-crowned  Night- 
Heron  disturbance  in  a Common  Tern  colony.  Co- 
lonial Waterbirds  14:51-56. 

Spanier,  E.  1980.  The  use  of  distress  calls  to  repel 
night  herons  ( Nycticorax  nycticorax ) from  fish 
ponds.  Journal  of  Applied  Ecology  17:287-294. 

Wetmore,  A.  1920.  Observations  on  the  habits  of 
birds  at  Lake  Burford,  New  Mexico.  Auk  37:393- 
412. 

Williams,  M.  D.  and  C.  P.  Nicholson.  1977.  Obser- 
vations at  a Black-crowned  Night-Heron  nesting 
colony.  Migrant  48:1-7. 

Wolford,  J.  W.  and  D.  A.  Boag.  1971.  Food  habits 
of  Black-crowned  Night  Herons  in  southern  Al- 
berta. Auk  88:435-437. 


The  Wilson  Journal  of  Ornithology  1 18(1):  104- 106,  2006 


First  Report  of  Black  Terns  Breeding  on  a Coastal  Barrier  Island 


Shawn  R.  Craik,1 3 Rodger  D.  Titman,1  Amelie  Rousseau,1  and  Michael  J.  Richardson2 3 


ABSTRACT. — Black  Terns  ( Chlidonias  niger  suri- 
namensis)  breed  locally  in  freshwater  wetlands  across 
the  northern  United  States  and  central  Canada,  often 
building  their  nests  over  shallow  water  on  a floating 
substrate  of  matted  marsh  vegetation.  Here,  we  report 
the  first  nesting  record  of  this  species  on  a coastal  bar- 
rier island.  The  nest,  which  consisted  of  two  eggs  laid 
in  a slight  scrape  of  sand,  was  located  on  6 July  2004 
in  a large  breeding  colony  of  Common  Terns  ( Sterna 
hirundo)  on  Kelly’s  Island  at  Kouchibouguac  National 
Park.  New  Brunswick,  Canada.  The  observation  also 
represents  the  current  northeastern  breeding  limit  for 
this  species  in  North  America.  Both  eggs  hatched,  but 


' Dept,  of  Natural  Resource  Sciences,  McGill  Univ., 
Ste-Anne-de-Bellevue,  QC  H9X  3V9,  Canada. 

2 Dept,  of  Biological  Sciences,  Bishop's  Univ.,  Len- 
noxville,  QC  JIM  1Z7,  Canada. 

3 Corresponding  author;  e-mail: 
shawn.craik@mail.mcgill.ca 


neither  chick  survived  beyond  4 days.  Received  15  De- 
cember 2004,  accepted  5 October  2005. 


The  North  American  subspecies  of  Black 
Tern  ( Chlidonias  niger  surinamensis ) breeds 
locally  across  the  northern  United  States  and 
central  Canada.  Black  Terns  are  semicolonial, 
typically  nesting  in  productive,  shallow  fresh- 
water marshes,  semipermanent  ponds,  prairie 
sloughs,  and  along  margins  of  lakes  and  rivers 
(Stewart  and  Kantrud  1984,  Dunn  and  Agro 
1995,  Schummer  and  Eddleman  2003).  Nests 
are  generally  placed  in  areas  of  calm  water 
within  stands  of  emergent  bulrush  ( Scirpus 
spp.),  cattail  ( Typha  spp.),  bur-reed  ( Spargan - 
ium  spp.),  or  pickerel  weed  ( Pontederia  corda- 
ta\  Cuthbert  1954,  Dunn  1979,  Mazzocchi  et 


SHORT  COMMUNICATIONS 


105 


al.  1997).  Nests  are  usually  built  over  shallow 
water  (0.5- 1.2  m deep)  on  a floating  substrate 
of  matted,  dead  marsh  vegetation,  floating  root- 
stalks  and  discarded  pieces  of  wood,  or  musk- 
rat feeding  platforms;  occasionally,  nests  are 
built  on  non-floating  substrates,  including 
muskrat  lodges,  flattened  vegetation,  and  mud 
(Cuthbert  1954,  Bergman  et  al.  1970,  Dunn 
1979).  Nests  often  consist  of  dead  vegetation 
arranged  in  a compressed  pile  with  a shallow 
depression  at  the  top  (Dunn  and  Agro  1995). 

Black  Terns  use  coastal  habitats  during  mi- 
gration, winter,  and  in  summer  when  non- 
breeding birds  aggregate  in  large  flocks  (100+ 
birds)  on  saltpans,  marshes,  estuaries,  and 
brackish  wetlands  (Dunn  and  Agro  1995).  Re- 
ports of  Black  Terns  breeding  in  marine  areas 
are  extremely  rare  (Sirois  and  Fournier  1993). 
In  the  mid-1990s,  a single  nest  was  found  at 
Seal  Island  National  Wildlife  Refuge  (NWR), 
Rockland,  Maine  (C.  S.  Hall  pers.  comm.), 
and  in  both  2003  and  2004,  two  nests  were 
located  at  Machias  Seal  Island,  New  Bruns- 
wick (C.  M.  Develin  pers.  comm.).  The  nests 
at  these  marine  sites  consisted  of  a small 
amount  of  dead  vegetation  in  sparse  common 
sheep  sorrel  ( Rumex  acetosella)  and  grasses, 
or  they  were  placed  on  a granite  rock  surface. 
Nests  were  located  in  large,  mixed  colonies  of 
Common  ( Sterna  hirundo)  or  Arctic  ( S . par- 
adisaea ) terns.  The  nest  at  Seal  Island  NWR 
was  ~30  m from  the  high-tide  line,  whereas 
the  nests  at  Machias  Seal  Island  were  —100 
m from  water.  All  five  Black  Tern  nests  in 
marine  areas  failed  to  fledge  young. 

The  Canadian  Maritime  breeding  popula- 
tion of  Black  Terns  was  estimated  to  be  150 
pairs  (Erskine  1992),  with  southern  New 
Brunswick  representing  the  species’  north- 
eastern breeding  limit  in  North  America 
(Dunn  and  Agro  1995).  Since  2000,  however. 
Black  Terns  (<4  birds  annually)  have  been 
observed  in  mid-  to  late  June  with  breeding 
Common  Terns  on  four  coastal  barrier  islands 
of  Kouchibouguac  National  Park,  New  Bruns- 
wick. Surveys  conducted  from  2000  to  2003, 
however,  did  not  confirm  breeding  (Christie  et 
al.  2004;  E.  Tremblay  pers.  comm.). 

Here,  we  report  the  first  evidence  of  Black 
Terns  breeding  on  a coastal  barrier  island.  Kel- 
ly’s Island  (46°  50'  N,  64°  55'  W),  2 ha  in  size, 
is  part  of  a 26-km  crescent  of  barrier  spits  and 
islands  that  separate  Kouchibouguac  Bay  of  the 


Northumberland  Strait  from  the  shallow  estu- 
ary-lagoon system  of  Kouchibouguac  National 
Park  (Beach  1988).  The  island  is  composed  of 
sand  and  is  vegetated  by  extensive  stands  of 
marram  grass  ( Ammophila  brevdigulata );  the 
island’s  outer  edge  consists  of  a gently  sloping 
intertidal  beach  zone.  The  island  supports  a 
large  breeding  colony  of  Common  Terns, 
which  included  1,041  nests  counted  in  2004 
(Parks  Canada  Tern  Survey  2004). 

On  6 July  2004  at  approximately  17:00 
AST,  after  the  entire  tern  colony  at  Kelly’s  Is- 
land had  flushed  and  taken  flight,  we  identi- 
fied a pair  of  adult  Black  Terns  flying  above 
the  center  of  the  island.  One  of  the  Black 
Terns  descended  and  landed,  and  we  subse- 
quently identified  a Black  Tern  nest  with  two 
eggs  laid  in  a slight  scrape  of  sand.  The  long, 
oval  eggs  were  noticeably  smaller  (—34  X 24 
mm)  than  the  subelliptical  eggs  in  nearby 
Common  Tern  nests  (—42  X 31  mm;  SRC 
pers.  obs.).  The  Black  Tern  eggs  were  dark 
olive  and  marked  with  dark  brown  dots  and 
blotches,  the  density  of  which  was  greater 
near  the  large  end.  Nearby  Common  Tern  eggs 
were  generally  cream  colored  and  finely 
marked  with  brown  and  black  dots.  The  Black 
Tern  nest  and  many  of  the  Common  Tern  nests 
consisted  of  a small  amount  of  dead  vegeta- 
tion loosely  lining  a scrape  made  in  the  sand. 
Both  species  nested  in  areas  of  the  island 
where  cover  was  sparse  (5-15%  marram 
grass).  Whereas  Common  Tern  nests  were 
0.5-30  m from  the  high-tide  line,  the  Black 
Tern  nest  was  26.5  m from  the  water.  Two 
Common  Tern  nests  were  within  3 m of  the 
Black  Tern  nest. 

On  20  July  at  17:20,  we  returned  to  the  nest 
and  found  a newly  hatched  chick  and  a pip- 
ping egg.  The  hatchling’s  down  was  predom- 
inantly cinnamon  and  black,  except  for  a 
white  belly  and  a white  mask  over  the  eye  and 
cheek.  A single  adult  Black  Tern  was  ob- 
served flying  5-10  m directly  above  the  nest. 
On  24  July,  we  checked  the  nest  again  and 
found  both  chicks  dead  at  the  nest;  one  adult 
Black  Tern  was  flying  10-15  m above  the  is- 
land. The  young  were  necropsied,  but  the 
cause  of  death  was  undetermined  (S.  McBur- 
ney  pers.  comm.). 

Adult  Common  Terns  at  Kelly’s  Island 
readily  exhibited  aggressive  displays  toward 
the  smaller  Black  Tern  adults.  Overt  aggres- 


106 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


sion  typically  involved  brief  aerial  chases  and 
attack  by  Common  Terns  as  a Black  Tern 
adult  approached  and  descended  toward  its 
nest.  Common  and  Black  terns  occasionally 
form  mixed-breeding  colonies  elsewhere 
(Snow  and  Perrins  1998),  and  Common  Terns 
have  been  known  to  defend  nesting  territories 
against  other  tern  species,  including  Roseate 
Terns  ( Sterna  dougallii ; Burger  and  Gochfeld 
1991,  Nisbet  2002).  Aggressive  displays  by 
Common  Terns,  and  the  close  proximity  of 
tern  nests  at  Kelly’s  Island,  may  have  com- 
promised the  survival  of  the  Black  Tern 
chicks  by  preventing  the  adults  from  provid- 
ing sufficient  food  resources  to  their  young, 
resulting  in  dehydration  or  malnutrition  (S. 
McBurney  pers.  comm.).  Nevertheless,  our 
observations  represent  the  first  confirmed 
breeding  of  Black  Terns  on  the  barrier  islands 
of  Kouchibouguac  National  Park  and  repre- 
sent the  northeastern  breeding  limit  for  this 
species  in  North  America. 

ACKNOWLEDGMENTS 

We  are  grateful  to  Kouchibouguac  National  Park, 
especially  E.  Tremblay,  for  providing  logistical  and 
field  support.  We  appreciate  information  on  Black 
Terns  provided  by  the  National  Audubon  Society’s 
Seabird  Restoration  Program,  the  University  of  New 
Brunswick,  and  I.  C.  T.  Nisbet.  S.  McBurney  willingly 
performed  necropsies.  Observations  and  preparation  of 
this  manuscript  were  conducted  during  SRC’s  graduate 
studies  of  Red-breasted  Mergansers  ( Mergus  serrator) 
supported  by  Bishop’s  University,  the  Canadian  Wild- 
life Federation,  the  New  Brunswick  Wildlife  Trust 
Fund,  and  the  Province  of  Quebec  Society  for  the  Pro- 
tection of  Birds.  We  thank  M.-A.  Hudson  and  three 
anonymous  reviewers  for  helpful  comments  on  this 
manuscript. 

LITERATURE  CITED 

Beach,  H.  (Ed.).  1988.  The  resources  of  Kouchiboug- 
uac National  Park:  resource  description  and  anal- 


ysis. Kouchibouguac  National  Park,  Environment 
Canada-Parks,  New  Brunswick,  Canada. 

Bergman,  R.  D.,  P.  Swain,  and  M.  W.  Weller.  1970. 
A comparative  study  of  nesting  Forster’s  and 
Black  terns.  Wilson  Bulletin  82:435-444. 

Burger,  J.  and  M.  Gochfeld.  1991.  The  Common 
Tern:  its  breeding  biology  and  social  behaviour. 
Columbia  University  Press,  New  York. 

Christie,  D.  S.,  B.  E.  Dalzell,  M.  David,  R.  Doiron, 
D.  G.  Gibson,  M.  H.  Lushington,  P.  A.  Pearce, 
S.  I.  Tingley,  and  J.  G.  Wilson.  2004.  Birds  of 
New  Brunswick:  an  annotated  list.  New  Bruns- 
wick Museum  monographic  series  (natural  sci- 
ence), no.  10.  Saint  John,  New  Brunswick,  Can- 
ada. 

Cuthbert,  N.  L.  1954.  A nesting  study  of  the  Black 
Tern  in  Michigan.  Auk  71:36-63. 

Dunn,  E.  H.  1979.  Nesting  biology  and  development 
of  young  in  Ontario  Black  Terns.  Canadian  Field- 
Naturalist  93:276-281. 

Dunn,  E.  H.  and  D.  J.  Agro.  1995.  Black  Tern  (Chli- 
donias  niger).  The  Birds  of  North  America,  no. 
147. 

Erskine,  A.  J.  1992.  Atlas  of  breeding  birds  of  the 
Maritime  Provinces.  Nimbus  Publishing  and  Nova 
Scotia  Museum,  Halifax,  Nova  Scotia,  Canada. 

Mazzocchi,  I.  M.,  J.  M.  Hickey,  and  R.  L.  Miller. 
1997.  Productivity  and  nesting  habitat  character- 
istics of  the  Black  Tern  in  northern  New  York. 
Colonial  Waterbirds  20:596-603. 

Nisbet,  I.  C.  T.  2002.  Common  Tern  ( Sterna  hirundo). 
The  Birds  of  North  America,  no.  618. 

Schummer,  M.  L.  and  W.  R.  Eddleman.  2003.  Effects 
of  disturbance  on  activity  and  energy  budgets  of 
migrating  waterbirds  in  south-central  Oklahoma. 
Journal  of  Wildlife  Management  67:789-795. 

Sirois,  J.  and  M.  A.  Fournier.  1993.  Clarification  of 
the  status  of  the  Black  Tern  ( Chlidonias  niger)  in 
the  Northwest  Territories,  Canada.  Colonial  Wa- 
terbirds 16:208-212. 

Snow,  D.  W.  and  C.  M.  Perrins.  1998.  Black  Tern 
( Chlidonias  niger).  Pages  796-799  in  The  birds  of 
the  Western  Palearctic,  vol.  1.  Oxford  University 
Press,  Oxford,  United  Kingdom. 

Stewart,  R.  E.  and  H.  A.  Kantrud.  1984.  Ecological 
distribution  and  crude  density  of  breeding  birds 
on  prairie  wetlands.  Journal  of  Wildlife  Manage- 
ment 48:426-437. 


SHORT  COMMUNICATIONS 


107 


The  Wilson  Journal  of  Ornithology  1 18(1):  107— 108,  2006 

First  Observation  of  Cavity  Nesting  by  a Female  Blue  Grosbeak 

Thomas  S.  Rischu  and  Thomas  J.  Robinson12 


ABSTRACT. — On  21  May  2003,  we  discovered  a 
completed  Blue  Grosbeak  ( Passerina  caerulea ) nest  in 
an  Eastern  Bluebird  ( Sialia  sialis ) nest  box.  On  28 
May,  the  nest  contained  four  whitish-tan  eggs  with 
light-brown,  streaky  and  spotty  markings,  an  unusual 
color  pattern  for  Blue  Grosbeak  eggs.  Species’  iden- 
tification was  confirmed  by  capturing  the  breeding  fe- 
male in  the  nest  box,  and  confirmed  again  later  with 
identification  of  the  chicks  as  Blue  Grosbeaks.  To  our 
knowledge,  this  is  the  first  published  account  of  cavity 
nesting,  artificial  or  otherwise,  for  this  species.  Re- 
ceived 27  September  2004,  accepted  31  May  2005. 


The  Blue  Grosbeak  ( Passerina  caerulea ) is 
a large  bunting  in  the  family  Cardinalidae  and 
is  relatively  common  in  the  southeastern  Unit- 
ed States.  However,  little  is  known  of  the 
breeding  ecology  of  this  species  (Ingold 
1993).  The  nest  is  typically  cup-shaped  and 
composed  of  twigs,  rootlets,  and  bark,  is  often 
lined  with  grass  and/or  fine  hair,  and  some- 
times contains  artificial  debris,  such  as  card- 
board, cellophane,  or  newspaper  (Stabler 
1959,  Bent  1968,  Ingold  1993).  Blue  Gros- 
beaks commonly  build  their  nests  in  riparian 
thickets,  fallow  fields,  open  woodlands,  and 
hedgerows,  usually  from  1 to  4 m above  the 
ground  (Stabler  1959,  Bent  1968,  Ehrlich  et 
al.  1988). 

Here,  we  detail  an  observation  of  cavity 
nesting  by  a pair  of  Blue  Grosbeaks.  We  dis- 
covered the  nest  during  an  ongoing  study  of 
Eastern  Bluebirds  ( Sialia  sialis)  in  Craighead 
County,  Arkansas.  During  the  winter  of  2002, 
we  erected  approximately  200  Eastern  Blue- 
bird nest  boxes  at  2 m above  ground,  with 
each  box  being  at  least  100  m from  adjacent 
boxes.  The  site  is  composed  mostly  of  pas- 


1 Dept,  of  Biological  Sciences,  Arkansas  State 
Univ.,  P.O.  Box  599,  State  University,  AR  72467, 
USA. 

2 Current  address:  Dept,  of  Biological  Sciences,  331 
Funchess  Hall,  Auburn  Univ.,  Auburn,  AL  36849, 
USA. 

3 Corresponding  author;  e-mail:  trisch@astate.edu 


tures  and  fallow  fields,  with  some  nest  boxes 
located  along  mixed-hardwood  forest  edge. 

We  checked  all  nest  boxes  at  least  once  per 
week  to  monitor  nesting  activity.  On  21  May 
2003,  we  discovered  an  unidentified,  but  com- 
plete, nest  without  eggs  in  a nest  box  in  an 
area  of  open  woodland  dominated  by  northern 
red  oak  ( Quercus  rubra ) and  bordered  on  one 
side  by  a thin  stand  of  privet  ( Ligustrum  spp.). 
The  nest  was  an  open  cup  composed  of  grass, 
fine  sticks,  and  several  interwoven  pieces  of 
cellophane.  Cellophane  is  commonly  incor- 
porated within  nests  of  Blue  Grosbeaks  (In- 
gold 1993),  possibly  as  a substitute  for  shed 
snakeskin,  a common  item  in  grosbeak  nests 
(Strecker  1926).  It  is  unclear  why  snakeskins 
are  incorporated  into  grosbeak  nests  (Ingold 
1993),  but  their  addition  to  nest  boxes  with 
artificial  nests  may  decrease  predation  (E.  C. 
Medlin  and  TSR  unpubl.  data).  This  behavior 
is  common  in  some  obligate  cavity-nesting 
species,  including  Tufted  Titmouse  ( Baeolo - 
phus  bicolor ) and  Great  Crested  Flycatcher 
( Myiarchus  crinitus ).  We  did  not  measure  the 
nest,  but  the  nesting  material  entirely  covered 
the  floor  of  the  nest  box  (10  cm  wide  X 15 
cm  deep),  and  the  nest  cup  covered  the  rear 
70%  of  the  nest-box  floor.  We  estimated  the 
inside  diameter  of  the  nest  cup  to  be  ~6— 7 
cm,  which  is  similar  to  grosbeak  nest-cup  di- 
ameters reported  by  others  (Ingold  1993). 

On  28  May,  we  checked  the  nest  again  and 
it  contained  four  oval  eggs  with  light-brown, 
streaky  and  spotty  markings,  and  a light,  whit- 
ish-tan background  color.  Although  Blue 
Grosbeak  eggs  are  typically  light  blue  to  white 
and  unmarked  (Ingold  1993),  some  are  lightly 
spotted  with  brown  (Ingold  1993)  or  “dis- 
tinctly marked  with  dots  and  spots  of  chestnut 
and  subdued  lilac”  (Davie  1898:404).  The 
size,  color,  and  markings  of  the  eggs  we  ob- 
served were  similar  to  those  of  Brown-headed 
Cowbirds  ( Molothrus  ater),  so  much  so  that 
we  could  not  distinguish  them  from  cowbird 
eggs.  Although  Blue  Grosbeaks  are  frequent 


108 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol  118,  No.  1,  March  2006 


hosts  of  Brown-headed  Cowbirds,  and  cow- 
birds  are  known  to  parasitize  hosts  nesting  in 
nest  boxes  (Whitehead  et  al.  2000,  2002),  we 
did  not  observe  nest  parasitism  in  any  of  our 
nest  boxes  during  our  2-year  study. 

Prior  to  the  discovery  of  the  nest,  we  had 
observed  a pair  of  Blue  Grosbeaks  near  the 
nest  box  several  times  over  a 2-week  period. 
We  suspected  that  the  pair  was  nesting  nearby, 
but  not  in  the  nest  box.  On  8 June,  however, 
we  captured  a female  Blue  Grosbeak  in  the 
nest  box  by  using  a nest-box  trap  (Robinson 
et  al.  2004);  she  was  incubating  the  four  eggs 
described  above,  which  appeared  to  be  pip- 
ping. When  we  revisited  the  nest  again  on  13 
June,  we  found  four  nestlings  approximately 
5 days  old  and  apparently  in  good  condition. 
We  identified  the  nestlings  as  Blue  Grosbeaks 
(and  not  cowbirds)  by  virtue  of  their  large 
conical  bills  and  yellow  rictal  flanges.  Al- 
though Brown-headed  Cowbirds  also  have 
conical  bills,  grosbeaks’  bills  are  obviously 
larger.  In  addition,  Blue  Grosbeak  chicks  have 
yellow  rictal  flanges  (Baicich  and  Harrison 
1997),  whereas  those  of  Brown-headed  Cow- 
bird  chicks  are  cream-colored  in  the  eastern 
subspecies  (Baicich  and  Harrison  1997). 

On  27  June,  the  nestlings  were  no  longer  in 
the  nest.  We  assumed  they  fledged  success- 
fully because  there  were  no  obvious  signs  of 
nest  predation,  and  predation  at  our  field  site 
is  generally  low  (13%  Eastern  Bluebird  nest 
predation;  TJR  and  TSR  unpubl.  data). 

Our  observation  of  Blue  Grosbeaks  nesting 
in  a nest  box  is  unique  for  two  reasons:  (1)  to 
our  knowledge,  this  is  the  first  record  of  cavity 
nesting  by  Blue  Grosbeaks,  and  (2)  the  color 
pattern  of  the  eggs  was  unusual.  We  know  of 
few  previously  published  reports  of  female 
Blue  Grosbeaks  laying  eggs  with  brown  spot- 
ty markings — a rare  color  pattern  for  Blue 
Grosbeak  eggs  (Davie  1898,  Ingold  1993). 
Avian  ecologists  should  be  aware  that  cavity 


nesting  occasionally  occurs  in  this  species;  the 
behavior  may  merit  closer  examination. 

ACKNOWLEDGMENTS 

We  extend  our  thanks  to  J.  C.  Bednarz,  H.  B.  Fok- 
idis,  J.  L.  Ingold,  M.  A.  Whitehead,  and  an  anonymous 
reviewer  for  comments  on  previous  versions  of  this 
manuscript.  Additionally,  we  appreciate  J.  Cassady  and 
C.  Robertson  for  their  invaluable  assistance  in  the 
field.  This  research  was  funded  by  an  ASU  Faculty 
Improvement  Grant  to  TSR  in  2003  and  2004,  an  Ar- 
kansas Audubon  Society  grant  to  TJR  in  2003  and 
2004;  TSR  further  benefited  from  reassignment  time 
from  the  Environmental  Sciences  Program  at  ASU. 

LITERATURE  CITED 

Baicich,  P.  J.  and  C.  J.  O.  Harrison.  1997.  A guide 
to  the  nests,  eggs,  and  nestlings  of  North  Ameri- 
can birds,  2nd  ed.  Academic  Press,  San  Diego, 
California. 

Bent,  A.  C.  1968.  Eastern  Blue  Grosbeak.  Pages  67- 
75  in  Life  histories  of  North  American  cardinals, 
grosbeaks,  towhees,  finches,  sparrows,  and  allies 
(part  1)  (O.  L.  Austin,  Jr.,  Ed.).  U.S.  National  Mu- 
seum Bulletin,  no.  237,  part  1.  [Reprinted  1968, 
Dover  Publications,  New  York.] 

Davie,  O.  1898.  Nests  and  eggs  of  North  American 
birds.  Landon  Press,  Columbus,  Ohio. 

Ehrlich,  P.  R.,  D.  S.  Dobkin,  and  D.  Wheye.  1988. 
Blue  Grosbeak.  Page  556  in  The  birder’s  hand- 
book: a field  guide  to  the  natural  history  of  North 
American  birds.  Simon  and  Schuster,  New  York. 
Ingold,  J.  L.  1993.  Blue  Grosbeak  ( Guiraca  caerulea). 

The  Birds  of  North  America,  no.  79. 

Robinson,  T.  J.,  L.  M.  Siefferman,  and  T.  S.  Risch. 
2004.  A quick,  inexpensive  trap  for  use  with  nest 
boxes.  North  American  Bird  Bander  29:115-116. 
Stabler,  R.  M.  1959.  Nesting  of  the  Blue  Grosbeak 
in  Colorado.  Condor  61:46—48. 

Strecker,  J.  K.  1926.  On  the  use,  by  birds,  of  snakes’ 
sloughs  as  nesting  material.  Auk  43:501-507. 
Whitehead,  M.  A.,  S.  H.  Schweitzer,  and  W.  Post. 
2000.  Impact  of  brood  parasitism  on  nest  survival 
parameters  and  seasonal  fecundity  of  six  songbird 
species  in  southeastern  old-field  habitat.  Condor 
102:946-950. 

Whitehead,  M.  A.,  S.  H.  Schweitzer,  and  W.  Post. 
2002.  Cowbird  host  interactions  in  a southeastern 
old-field:  a recent  contact  area?  Journal  of  Field 
Ornithology  73:379-386. 


SHORT  COMMUNICATIONS 


109 


The  Wilson  Journal  of  Ornithology  1 18(1):  109-1  12,  2006 

A New  Record  of  the  Endangered  White-winged  Nightjar  ( Eleothreptus 

candicans)  from  Beni,  Bolivia 

Tomas  Grim1 3 and  Radim  Sumbera1 2 3 


ABSTRACT. — The  ecology  of  the  White-winged 
Nightjar  ( Eleothreptus  candicans)  is  poorly  known. 
Only  three  breeding  populations  (one  from  Brazil  and 
two  from  Paraguay)  are  known,  and  populations  are 
decreasing  due  to  continuing  destruction  of  cerrado 
habitat.  On  14  September  2003,  we  took  several  pho- 
tos of  an  unidentified  nightjar  in  Beni  Biosphere  Re- 
serve, Departmento  Beni,  Bolivia.  The  bird  was  later 
determined  to  be  an  adult  male  White-winged  Nightjar. 
Interestingly,  the  only  previous  record  for  Bolivia  was 
a male  collected  in  1987  at  the  same  locality  and  time 
of  year.  Because  the  White-winged  Nightjar  is  non- 
migratory  and  secretive,  we  hypothesize  that  there  may 
be  a sustainable  population  of  White-winged  Nightjars 
in  Bolivia,  and  the  paucity  of  sightings  may  be  due  to 
the  species’  low  detectability.  Received  16  December 
2004,  accepted  1 1 October  2005. 


The  White- winged  Nightjar  ( Eleothreptus 
candicans ),  a member  of  the  Caprimulgidae 
(Cleere  1999,  Pople  2004),  was  recently  re- 
classified from  the  genus  Caprimulgus  to  the 
genus  Eleothreptus  (Cleere  2002).  Its  known 
range  and  population  size  are  very  small,  and 
its  ecology  has  received  attention  only  re- 
cently (Pople  2003).  Parker  et  al.  (1996)  as- 
signed the  species  High  Conservation  Priority 
and  the  IUCN  lists  the  species  as  Endangered 
(IUCN  Red  List;  Pople  2004).  E . candicans  is 
threatened  by  ongoing  loss  of  its  cerrado  hab- 
itat (heavy  grazing,  trampling,  invasive  grass- 
es, habitat  conversion  to  plantations,  and 
large-scale,  uncontrolled  grass  fires;  Cleere 
1999,  Pople  2004). 

Until  the  1980s,  White-winged  Nightjars 
were  known  only  from  two  museum  speci- 
mens collected  at  the  beginning  of  the  19th 
century  in  Or^anga,  Sao  Paulo  state,  and  Cu- 
iaba,  Mato  Grosso  state,  Brazil  (Sclater  1866). 
Only  three  populations  have  been  found,  all 


1 Dept,  of  Zoology,  Palacky  Univ.,  Tr.  Svobody  26, 
771  46  Olomouc,  Czech  Republic. 

2 Dept,  of  Zoology,  Univ.  of  South  Bohemia,  Bran- 
isovska  31,  37005  Ceske  Budejovice,  Czech  Republic. 

3 Corresponding  author;  e-mail:  grim@prfnw.upol.cz 


in  southern  Brazil  and  eastern  Paraguay:  Emas 
National  Park,  Brazil  (Rodrigues  et  al.  1999); 
Aguara  Nu,  Mbaracayu  Forest  Nature  Re- 
serve, Paraguay  (Lowen  et  al.  1996,  Clay  et 
al.  1998);  and  a recently  discovered  popula- 
tion at  Laguna  Blanca,  Departmento  San  Pe- 
dro, central  Paraguay  (Anonymous  2002).  Ad- 
ditionally, in  1987  a single  male  was  captured 
and  collected  at  the  Beni  Biological  Station, 
Departmento  Beni,  Bolivia  (Davis  and  Flores 
1994).  Despite  specific  searches  for  the  spe- 
cies in  subsequent  years,  however,  it  has  not 
been  relocated  at  Beni  (Brace  et  al.  1997, 
Brace  2000,  Pople  2004;  R.  Brace  and  J. 
Hornbuckle  in  lift.). 

Surveys  in  Aguara  Nu  have  resulted  in  a 
population  estimate  of  40—150  individuals 
(Clay  et  al.  1998,  Pople  2003)  at  that  location. 
The  number  of  birds  observed  in  Emas  Na- 
tional Park  was  12  in  September  1985  and 
only  1 in  October  1990  and  in  November 
1997  (Rodrigues  et  al.  1999).  Although  there 
are  no  other  recently  published  records  from 
Emas,  the  national  park  probably  supports  a 
sizeable  population  of  E.  candicans  (Pople 
2004)  because  Emas  encompasses  a large  ex- 
tent of  apparently  suitable  habitat.  The  re- 
cently discovered  population  at  Laguna  Blan- 
ca in  Paraguay  is  estimated  to  include  a min- 
imum of  30  birds  (R.  P.  Clay  in  litt.). 

On  14  September  2003  at  22:00  EDT,  we 
photographed  an  unidentified  nightjar  on  a ter- 
mite mound  between  the  Beni  Biological  Sta- 
tion (Estacion  Biologica  del  Beni;  14°  50'  S, 
66°  17'  W)  and  Laguna  Normandia  (—1.5  km 
northwest  of  the  station;  see  Fig.  3 in  Brace 
et  al.  1997),  Departmento  Beni  in  northern 
Bolivia.  Later  the  bird  was  unambiguously 
identified  as  a male  E.  candicans  (Fig.  1).  Be- 
cause it  lacked  visible  wear  on  the  remiges 
and  pale  flecking  in  the  contour  plumage,  it  is 
probable  that  the  individual  had  recently  com- 
pleted a molt.  If  the  species  undergoes  the 
same  pattern  of  molt  in  both  Beni  Biosphere 


110 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


LIG.  1.  Adult  male  White-winged  Nightjar  ( Eleothreptus  candicans ) photographed  on  14  September  2003 
in  Beni  Biosphere  Reserve,  Departmento  Beni,  Bolivia.  Photo  by  R.  Sumbera. 


Reserve  and  Paraguay  (i.e.,  replacement  of 
flight  feathers  in  a single  post-nuptial  molt),  it 
suggests  that  the  species  may  breed  consid- 
erably earlier  in  Bolivia  than  in  Paraguay 
(where  it  breeds  mainly  between  September 
and  December). 

Beni  Biological  Station  is  180  km  west  of 
Trinidad  and  50  km  east  of  San  Borja  on  El 
Porvenir  Estancia.  El  Porvenir  Estancia  lies  in 
the  Llanos  de  Mojos,  which  is  a lowland  plain 
(—200  m elevation)  characterized  as  savanna 
with  forest  islands.  The  habitat  where  we  ob- 
served the  White-winged  Nightjar  is  a season- 
ally inundated  savanna  with  a high  density  of 
termite  mounds  (Fig.  2). 

Ours  is  only  the  second  record  of  White- 
winged Nightjar  in  Bolivia,  the  first  having 
been  made  in  September  1987  (Davis  and  Flo- 
res 1994).  Interestingly,  both  observations 
were  made  near  Beni  Biological  Station  at  the 
same  time  of  year  (1 1 September  1987  and  14 
September  2003).  Despite  a number  of  re- 
search programs  that  have  been  conducted  at 
the  station  (A.  B.  Hennessey  in  lift.),  there  had 
been  no  additional  records  of  White-winged 
Nightjar  after  1987.  R.  C.  Brace  and  J.  Horn- 


buckle  (in  lift .),  for  example,  searched  for 
White-winged  Nightjars  and  conducted  mist- 
netting  from  mid-July  through  the  end  of  Au- 
gust every  year  from  1992  to  1999,  but  re- 
corded no  White-winged  Nightjars.  Although 
the  White-winged  Nightjar  is  considerably 
less  conspicuous  than  many  other  sympatric 
nightjar  species  common  in  Bolivia  (R.  G.  Po- 
ple  in  litt.),  it  seems  unlikely  that  there  would 
be  so  few  observations  of  the  species  if  the 
area  supported  a small  resident  population. 
Rather,  the  two  individuals  recorded  during 
the  last  2 decades  may  have  come  from  an 
undiscovered  population  elsewhere  in  the 
northern  Bolivian  lowlands.  However,  E.  can- 
dicans is  presumed  to  be  a resident  species. 
Indeed,  radio-tracking  work  in  Paraguay  (Po- 
ple  2003)  revealed  that  White-winged  Night- 
jars are  year-round  residents,  and  a study  of 
captive  birds  revealed  a post-nuptial  molt  pat- 
tern typical  of  a nonmigratory  species.  There- 
fore, the  occurrence  of  the  two  individuals  at 
Beni  Biological  Station  during  the  same  time 
of  year  may  indicate  that  some  birds  make 
local  movements,  possibly  in  response  to  fires 
(Pople  2004). 


SHORT  COMMUNICATIONS 


1 1 1 


FIG.  2.  Typical  habitat  of  the  White-winged  Nightjar — wet  savanna  with  termite  mounds  providing  perches 
above  the  surrounding  young  vegetation.  The  forest  in  the  background  is  Florida  Fragment  south  of  Laguna 
Normandia,  1.5  km  northwest  of  Beni  Biological  Station,  Departmento  Beni,  Bolivia.  The  photo  in  Figure  1 
was  taken  within  this  area.  Photo  by  T.  Grim. 


Neotropical  savannas  are  under  increasing 
human  pressure  due  to  large-scale  conversion 
of  grassland  habitats  to  pastures  (Marris 
2005).  Although  the  White-winged  Nightjar  is 
a typical  savanna  dweller  and  is  adapted  to 
irregular  and  small-scale  fires,  it  likely  has 
been  negatively  affected  by  regular  and  large- 
scale  burning  in  recent  years  (Brace  et  al. 
1997,  Pople  2004).  Conservation  of  savanna 
habitats — including  cerrado,  the  primary  hab- 
itat for  E.  candicans — has  been  neglected  thus 
far.  Because  savanna  habitats  are  facing  great- 
er threats  than  Amazonian  rainforests,  the 
conservation  of  cerrado  habitat  should  be- 
come a top  priority  in  the  Neotropics  (Marris 
2005). 

Our  observation  highlights  the  importance 
of  Beni  Biosphere  Reserve  for  threatened  ( n 
~ 4)  and  near-threatened  {n  = 15)  bird  species 
in  Bolivia  (Brace  et  al.  1997).  Among  these 
19  species  are  1 1 that  rely  wholly  or  partially 
on  savanna  habitat.  So  far,  500  bird  species 
have  been  reported  from  Beni  Biosphere  Re- 


serve (Brace  et  al.  1997,  Brace  2000).  We  add 
to  this  list  one  more  species:  on  the  same  day 
(14  September  2003)  that  we  observed  the 
White-winged  Nightjar,  we  also  recorded  one 
Black-throated  Saltator  ( Saltator  atricollis). 

We  hypothesize  that  Departmento  Beni  in 
northern  Bolivia  holds  a resident  population 
of  E.  candicans , and  that  the  paucity  of  re- 
cords from  Bolivia  reflects  the  lack  of  inten- 
sive searches  during  the  correct  season  and  the 
low  detectability  of  this  species.  We  concur 
with  Brace  et  al.  (1997)  that  more  information 
on  the  White-winged  Nightjar’s  status  is  re- 
quired, and  we  hope  that  our  observation  pro- 
vides an  impetus  for  further  research  on  this 
elusive  species. 

ACKNOWLEDGMENTS 

We  are  grateful  for  the  detailed  and  helpful  com- 
ments and  suggestions  by  R.  R Clay  and  two  anony- 
mous referees.  We  thank  R.  C.  Brace  and  A.  B.  Hen- 
nessey for  their  comments  on  the  manuscript  and  G. 
Dryden  for  reviewing  our  translation  to  English. 


112 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  1,  March  2006 


LITERATURE  CITED 

Anonymous.  2002.  A new  population  of  the  White- 
winged Nightjar.  World  Birdwatch  24:5. 

Brace,  R.  C.  2000.  The  avifauna  of  the  Beni  Biolog- 
ical Station:  records  to  1999.  Estacion  Biologica 
del  Beni,  Bolivia. 

Brace,  R.  C.,  J.  Hornbuckle,  and  J.  W.  Pearce-Hig- 
gins.  1997.  The  avifauna  of  the  Beni  Biological 
Station,  Bolivia.  Bird  Conservation  International 
7:117-159. 

Clay,  R.  R,  D.  R.  Capper,  J.  Mazar  Barnett,  I.  J. 
Burfield,  E.  Z.  Esquivel,  R.  Farina,  C.  P.  Ken- 
nedy, M.  Perrens,  and  R.  G.  Pople.  1998.  White- 
winged Nightjars  Caprimulgus  candicans  and  cer- 
rado  conservation:  the  key  findings  of  Project 
Aguara  Nu  1997.  Cotinga  9:52-56. 

Cleere,  N.  1999.  Family  Caprimulgidae  (nightjars). 
Pages  302-386  in  Handbook  of  the  birds  of  the 
world,  vol.  5:  barn-owls  to  hummingbirds  (J.  del 
Hoyo,  A.  Elliott,  and  J.  Sargatal,  Eds.).  Lynx  Ed- 
icions,  Barcelona,  Spain. 

Cleere,  N.  2002.  A review  of  the  taxonomy  and  sys- 
tematics  of  the  Sickle-winged  and  White-winged 
nightjars  (Caprimulgidae).  The  Bulletin  of  the 
British  Ornithologists’  Club  122:168-179. 

Davis,  S.  E.  and  E.  Flores.  1994.  First  record  of 
White-winged  Nightjar  Caprimulgus  candicans 
for  Bolivia.  The  Bulletin  of  the  British  Ornithol- 
ogists’ Club  1 14:127-128. 


Lowen,  J.  C.,  L.  Bartrina,  T.  M.  Brooks,  R.  P.  Clay, 
and  J.  Tobias.  1996.  Project  YACUTINGA  ’95: 
bird  surveys  and  conservation  priorities  in  eastern 
Paraguay.  Cotinga  5:14-19. 

Marris,  E.  2005.  The  forgotten  ecosystem.  Nature 
437:943-944. 

Parker,  T.  A.,  Ill,  D.  F.  Stotz,  and  J.  W.  Fitzpatrick. 
1996.  Ecological  and  distributional  databases  for 
Neotropical  birds.  Pages  1 1 8-407  in  Neotropical 
birds:  ecology  and  conservation  (D.  F.  Stotz,  J.  W. 
Fitzpatrick,  T.  A.  Parker,  III,  and  D.  K.  Moskov- 
its).  University  of  Chicago  Press,  Chicago,  Illi- 
nois. 

Pople,  R.  G.  2003.  The  ecology  and  conservation  of 
the  White-winged  Nightjar  Caprimulgus  candi- 
cans. Ph.D.  dissertation,  University  of  Cambridge, 
United  Kingdom. 

Pople,  R.  G.  2004.  White-winged  Nightjar  Eleothrep- 
tus  candicans.  In  Threatened  birds  of  the  world 
2004.  CD-ROM.  BirdLife  International,  Cam- 
bridge, United  Kingdom. 

Rodrigues,  F.  H.  G.,  A.  Hass,  O.  J.  Marini-Filho,  M. 
M.  GuimarAes,  and  M.  A.  Bagno.  1999.  A new 
record  of  White-winged  Nightjar  Caprimulgus 
candicans  in  Emas  National  Park,  Goias,  Brazil. 
Cotinga  11:83-85. 

Sclater,  P.  L.  1866.  Additional  notes  on  the  Capri- 
mulgidae. Proceedings  of  the  Zoological  Society 
of  London  1866:581-590. 


The  Wilson  Journal  of  Ornithology  1 18(1):  1 12-1 13,  2006 


Predation  of  Eared  Grebe  by  Great  Blue  Heron 

James  W.  Rivers' 2 and  Michael  J.  Kuehn1 


ABSTRACT. — Great  Blue  Herons  ( Ardea  herodias) 
typically  prey  upon  fish  and  other  aquatic  organisms, 
and  they  occasionally  take  small  mammals  and  birds. 
We  observed  a Great  Blue  Heron  attack,  kill,  and  at- 
tempt to  consume  an  Eared  Grebe  ( Podiceps  nigricol- 
lis).  The  heron  was  unable  to  swallow  the  grebe,  and 
it  abandoned  the  carcass  after  approximately  30  min. 
An  examination  of  the  carcass  showed  that  the  grebe 
lacked  obvious  physical  deformities.  Our  observation, 
coupled  with  a similar  one  nearby,  indicates  that  Great 
Blue  Herons  attack  and  kill  birds  larger  than  reported 
previously.  Received  11  January  2005,  accepted  19 
September  2005. 


1 Dept,  of  Ecology,  Evolution,  and  Marine  Biology, 
Univ.  of  California,  Santa  Barbara,  CA  93106.  USA. 

2 Corresponding  author;  e-mail: 
rivers  @ lifesci  .ucsb.edu 


On  the  morning  of  14  November  2004,  we 
witnessed  an  adult  Great  Blue  Heron  {Ardea 
herodias)  attack,  kill,  and  attempt  to  consume 
an  Eared  Grebe  {Podiceps  nigricollis ) at  Oso 
Flaco  Lake  (35°  00' N,  120°  30' W)  in  San 
Luis  Obispo  County,  California.  The  incident 
occurred  shortly  after  the  heron  landed  near 
the  grebe  and  began  foraging  in  shallow  (~30 
cm  deep)  water.  At  approximately  11:25  PST, 
the  heron  caught  the  grebe  with  a stabbing 
motion  as  the  grebe  swam  underwater.  The 
heron  then  proceeded  to  subdue  the  grebe  by 
grasping  its  neck,  shaking  it,  and  submerging 
it  intermittently.  After  approximately  15  min, 
the  grebe  appeared  to  be  dead.  At  this  point, 
the  heron  briefly  released  the  grebe  to  deliver 
several  sharp  blows  to  its  head  and  chest  area. 


SHORT  COMMUNICATIONS 


1 13 


The  heron  attempted  several  times  to  swal- 
low the  grebe,  but  it  had  difficulty  maneuver- 
ing the  grebe  into  its  mouth.  During  one  at- 
tempt, it  was  able  to  maneuver  the  carcass  into 
position,  but  the  grebe’s  diameter,  its  limp 
wings,  or  both  prevented  the  heron  from  swal- 
lowing it.  After  attempting  to  swallow  the 
grebe  for  approximately  15  min,  the  heron 
abandoned  the  carcass,  preened  briefly,  and 
then  flew  off.  The  grebe  weighed  255  g 
(weighed  after  the  grebe  was  frozen  and  then 
thawed),  and  although  that  is  low  body  weight 
for  this  species  (Cullen  et  al.  1999),  it  is  typ- 
ical of  grebes  arriving  on  a wintering  area  af- 
ter a migratory  flight  (Jehl  1997;  J.  R.  Jehl, 
Jr.  pers.  comm.).  When  we  examined  the 
grebe,  we  found  no  deformities  or  obvious  in- 
dications of  poor  condition  (e.g.,  loss  of  pec- 
toral muscle). 

On  the  day  previous  to  our  observation  (13 
November  2004),  H.  R.  Pedersen  (pers. 
comm.)  observed  a Great  Blue  Heron  at  Lake 
Cachuma  in  Santa  Barbara  County,  California 
(—130  km  southeast  of  Lake  Oso  Flaco),  cap- 
ture an  Eared  Grebe.  The  heron  was  foraging 
and  caught  the  grebe  in  shallow  water,  grasped 
it  by  the  neck  in  the  same  manner  we  wit- 
nessed, and  submerged  it  several  times.  After 
a brief  struggle,  the  grebe  escaped  and  ap- 
peared unharmed  (H.  R.  Pedersen  pers. 
comm.). 

We  know  of  no  previous  reports  of  Great 
Blue  Herons  capturing,  killing,  and  attempting 
to  consume  Eared  Grebes,  or  any  other  bird 
species  of  that  size;  however,  McCanch 
(2003)  reported  a Grey  Heron  ( Ardea  cinerea) 
that  had  choked  to  death  while  attempting  to 
ingest  a Little  Grebe  ( Tachybaptus  ruficollis ). 
Great  Blue  Herons  have  a diverse  diet  that 
includes  songbirds  and  mammals  of  various 
sizes  (Peifer  1979.  Butler  1992),  and  they 
have  been  observed  abandoning  large  prey 
items  that  they  were  unable  to  swallow  (R.  W. 


Butler  pers.  comm.).  Thus,  it  is  possible  that 
the  herons  may  have  targeted  the  grebes  as 
potential  prey  items,  but  were  unable  to  suc- 
cessfully consume  them  because  of  their  size. 
Alternative  explanations  are  (1)  that  the  her- 
ons mistook  the  grebes  for  fish  or  (2)  that  the 
herons  were  acting  to  defend  a foraging  area. 
Indeed,  an  observer  at  Lake  Cachuma  report- 
ed seeing  a foraging  heron  attack  and  kill  an 
American  Coot  ( Fulica  americana)  with  no 
attempt  to  eat  it  (L.  R.  Mason  pers.  comm.). 
The  heron  we  observed,  however,  expended  a 
substantial  amount  of  effort  subduing  and  at- 
tempting to  consume  the  grebe,  indicating  a 
deliberate  act  of  predation.  Evidently,  small 
grebes  are  potential  prey  items  for  Great  Blue 
Herons,  and  herons  may  attack  and  kill  large 
birds  more  commonly  than  is  recognized. 

ACKNOWLEDGMENTS 

We  thank  K.  E.  Jirik,  the  students  of  the  Terrestrial 
Vertebrate  Ecology  Laboratory  course  at  the  Univer- 
sity of  California-Santa  Barbara,  and  members  of  the 
Pomona  Valley  and  Golden  Gate  Audubon  Societies 
for  their  assistance  with  observations.  H.  R.  Pedersen 
and  L.  R.  Mason  kindly  shared  their  observations  of 
foraging  herons;  R.  W.  Butler  provided  helpful  discus- 
sion; and  J.  R.  Jehl,  N.  V.  McCanch,  and  an  anony- 
mous reviewer  provided  valuable  comments  on  the 
manuscript. 

LITERATURE  CITED 

Butler,  R.  W.  1992.  Great  Blue  Heron  {Ardea  hero- 
dias ).  The  Birds  of  North  America,  no.  25. 
Cullen,  S.  A.,  J.  R.  Jehl,  Jr.,  and  G.  L.  Nuechter- 
lein.  1999.  Eared  Grebe  {Podiceps  nigricollis). 
The  Birds  of  North  America,  no.  433. 

Jehl,  J.  R.,  Jr.  1997.  Cyclical  changes  in  body  com- 
position in  the  annual  cycle  and  migration  of  the 
Eared  Grebe  Podiceps  nigricollis.  Journal  of  Avi- 
an Biology  28:132-142. 

McCanch,  N.  2003.  Grey  Heron  choking  on  Little 
Grebe.  British  Birds  96:86. 

Peifer,  R.  W.  1979.  Great  Blue  Herons  foraging  for 
small  mammals.  Wilson  Bulletin  91:630-631. 


114 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


The  Wilson  Journal  of  Ornithology  1 18(1):  1 14-1 16,  2006 


Abnormal  Eggs  and  Incubation  Behavior  in  Northern  Bobwhite 

Fidel  Hernandez,1 2 Juan  A.  Arredondo,1  Froylan  Hernandez,1  Fred  C.  Bryant,1  and 

Leonard  A.  Brennan1 


ABSTRACT. — A long-term  (>5  years)  study  of 
Northern  Bobwhite  ( Colinus  virginianus)  provided  the 
first  record  of  runt  eggs  and  two  observations  of  pro- 
longed incubation.  During  2004,  we  located  two 
clutches  (n  = 11  and  9 eggs) — laid  by  the  same  hen — 
consisting  entirely  of  runt  eggs.  Mean  length,  width, 
and  mass  were  18.8  mm,  15.4  mm,  and  2.0  g,  respec- 
tively, 26%  of  the  volume  and  24%  of  the  mass  of 
typical  bobwhite  eggs.  Based  on  our  long-term  data 
set  for  bobwhites  ( n = 3,566  eggs),  runt  eggs  occur  at 
a frequency  of  0.56%,  within  the  range  (0.02-4.32%) 
reported  for  other  avian  species.  The  two  records  of 
prolonged  incubation  behavior  represented  75  days 
(326%)  and  47  days  (204%)  beyond  the  normal  incu- 
bation period  (23  days)  of  bobwhites.  This  prolonged 
incubation  behavior  is  in  excess  of  the  time  frame  re- 
ported for  most  birds  exhibiting  prolonged  incubation 
(50-100%  beyond  normal  incubation).  Received  31 
January  2005,  accepted  3 October  2005. 


Documenting  anomalies  in  avian  behavior 
often  is  an  opportunistic  endeavor  given  the 
rarity  of  such  behavior  and  the  short-term  na- 
ture (<2  years)  of  most  studies.  An  ongoing, 
long-term  (>5  years)  radiotelemetry  project 
(The  South  Texas  Quail  Research  Project; 
STxQRP)  on  Northern  Bobwhite  ( Colinus  vir- 
ginianus) provided  us  with  the  opportunity  to 
monitor  bobwhite  behavior  over  seven  breed- 
ing seasons  (1998-2004)  on  the  Encino  Di- 
vision of  the  King  Ranch,  Inc..  Brooks  Coun- 
ty, Texas.  We  provide  the  first  record  of  runt 
eggs  for  Northern  Bobwhite  and  two  addition- 
al records  of  prolonged  incubation  behavior. 

First  record  of  runt  eggs. — Runt  eggs,  also 
referred  to  as  dwarf,  cock,  wind,  and  witch 
eggs  (Rothstein  1973),  are  those  noticeably 
smaller  than  the  smallest  expected  for  a given 
species  (Mulvihill  1987;  for  suggested  crite- 
ria, see  Koenig  1980a).  Although  runt  eggs 
have  been  reported  for  several  avian  species 


1 Caesar  Kleberg  Wildlife  Research  Inst.,  Texas 
A&M  Univ.,  Kingsville,  TX  78363,  USA. 

2 Corresponding  author;  e-mail: 
fidel.hernandez@tamuk.edu 


(e.g.,  Canada  Goose,  Branta  canadensis 
[Manning  and  Carter  1977];  woodpeckers  [Pi- 
cidae,  Koenig  1980b];  and  Eastern  Bluebird, 
Sialia  sialis  [Mulvihill  1987]),  they  normally 
occur  at  low  frequencies  (~1  per  1,000  to 
2,000  eggs;  Koenig  1980b.  Mallory  et  al. 
2004).  Furthermore,  runt  eggs  usually  repre- 
sent only  a small  proportion  of  a clutch  (Roth- 
stein 1973,  Ricklefs  1975,  Bartel  1986).  Entire 
clutches  consisting  solely  of  runt  eggs  are  ex- 
tremely rare  and  have  been  reported  only  for 
Song  Thrush  ( Turdus  philomelos;  M’Wiliiam 
1927),  Gray  Catbird  ( Dumetella  carolinensis ; 
Rothstein  1973),  and  Eastern  Bluebird  (Ze- 
leny  1983).  We  report  the  first  record  of  runt 
eggs  for  Northern  Bobwhite  and  provide  es- 
timates of  the  frequency  of  such  eggs. 

On  21  June  2004,  we  located  a radiomarked 
hen  on  a nest  at  the  base  of  brownseed  pas- 
palum  ( Paspalum  plicatulum ).  The  clutch 
consisted  entirely  of  runt  eggs  ( n = 11).  We 
monitored  the  hen  for  several  days  thereafter, 
but  never  located  her  at  the  nest  site  again. 
We  concluded  that  she  had  abandoned  the  nest 
and  we  collected  the  eggs.  During  the  follow- 
ing 5 weeks,  the  hen  again  paired  with  a male, 
and  on  30  July,  we  documented  a second 
clutch  of  runt  eggs  (n  = 9)  in  a nest  con- 
structed in  red  lovegrass  ( Eragrostis  secun- 
diflora).  The  hen  also  abandoned  this  nest,  and 
we  collected  the  clutch  on  2 August. 

None  of  the  runt  eggs  was  viable  (i.e.,  none 
contained  yolk).  Mean  length,  width,  and 
mass  of  the  runt  eggs  {n  = 20)  were  18.8  mm, 
15.4  mm,  and  2.0  g,  respectively.  The  smallest 
reported  measurements  for  bobwhite  eggs  are 
26  mm  (length)  and  22.5  mm  (width)  (Bent 
1932),  and  8.2  g (Case  and  Robel  1974).  Koe- 
nig (1980a)  defined  runt  eggs  as  those  with  a 
relative  volume  (length  X width2  X tt/6) 
<75%  of  the  average.  Mean  length,  width, 
and  mass  of  bobwhite  eggs  are  30  mm,  24 
mm,  and  8.3  g,  respectively  (Bent  1932,  Case 
and  Robel  1974).  Thus,  the  volume  and  mass 


SHORT  COMMUNICATIONS 


of  the  runt  eggs  we  found  were  only  26%  and 
24%,  respectively,  of  the  average. 

We  used  data  from  STxQRP  and  Hernandez 
(1999)  to  estimate  the  frequency  of  runt  eggs 
in  Northern  Bobwhite.  During  1999-2004  of 
the  STxQRP,  we  located  392  nests  and  deter- 
mined clutch  size  for  297  nests  ( n = 3,161 
eggs).  Hernandez  (1999)  located  83  bobwhite 
nests  in  Shackelford  County,  Texas  during 
1997-1998  and  determined  clutch  size  for  35 
nests  ( n = 385  eggs).  Based  on  these  com- 
bined data  (3,546  normal-sized  eggs  + 20  runt 
eggs),  runt  eggs  in  bobwhites  occur  at  a fre- 
quency of  0.56%,  which  is  within  the  range 
(0.02-4.32%)  reported  for  other  avian  species 
(Koenig  1980b,  Mallory  et  al.  2004). 

The  mechanisms  underlying  the  production 
of  runt  eggs  are  not  entirely  understood  (Mul- 
vihill  1987).  However,  runt  eggs  often  are  pro- 
duced after  temporary  disturbance  or  damage 
(e.g.,  injury  or  infection)  to  the  reproductive 
organs  (Pearl  and  Curtis  1916,  Romanoff  and 
Romanoff  1949).  Instances  of  entire  clutches 
being  composed  of  runt  eggs  suggest  a con- 
genital defect  or  permanent  injury  to  the  re- 
productive system  (Mulvihill  1987).  We  pre- 
sume the  bobwhite  hen  that  laid  the  runt  eggs 
may  have  suffered  from  some  type  of  per- 
manent injury  to  her  reproductive  organs. 

Prolonged  incubation  behavior. — Pro- 
longed incubation  beyond  the  normal  time  re- 
quired for  hatching  has  been  reported  for 
many  avian  species,  including  Killdeer  ( Cha - 
radrius  vociferus'.  Powers  1978),  Common 
Loon  ( Gavia  immer\  Sutcliffe  1982),  and 
Long-eared  Owl  ( Asio  otus\  Marks  1983). 
Most  birds  that  exhibit  prolonged  incubation 
appear  to  incubate  for  at  least  50-100%  lon- 
ger than  necessary  to  hatch  a clutch  (Skutch 
1962).  Prolonged  incubation  (56  days)  has 
been  reported  only  once  for  Northern  Bob- 
white  (Stoddard  1931),  which  is  33  days 
(143%)  beyond  the  average  incubation  period 
(23  days).  We  report  two  additional  records  of 
prolonged  incubation  for  Northern  Bobwhite. 

During  our  first  observation  of  prolonged 
incubation,  a bobwhite  hen  exhibited  normal 
incubation  behavior  during  a first  nesting,  and 
the  eggs  successfully  hatched  on  7 July  2003. 
However,  the  hen  exhibited  prolonged  incu- 
bation of  a second  clutch.  We  discovered  the 
nest  on  1 1 August,  and  by  8 September,  only 
1 of  10  eggs  had  hatched.  The  female  was  not 


1 15 

observed  on  the  nest  between  9 and  25  Sep- 
tember, but  on  26  September,  the  hen  returned 
to  the  nest  and  resumed  incubation  until  5 De- 
cember. Thus,  the  hen  incubated  the  eggs  for 
28  days,  abandoned  the  nest  for  17  days,  and 
then  resumed  incubation  for  another  70  days. 
The  98  days  of  incubation  was  75  days 
(326%)  beyond  the  normal  incubation  period 
for  bobwhites. 

We  documented  the  second  occurrence  of 
prolonged  incubation  during  the  2004  nesting 
season.  On  18  June,  we  accidentally  flushed 
an  un-radiomarked  hen  from  a nest.  We  re- 
turned to  the  nest  site  on  12  July,  presuming 
the  clutch  had  hatched,  and  found  her  still  in- 
cubating the  clutch.  The  hen  continued  incu- 
bating until  27  August,  when  the  clutch  was 
depredated.  Assuming  the  hen  had  just  begun 
incubation  when  we  first  found  the  nest,  she 
incubated  for  at  least  70  days,  or  47  days 
(204%)  beyond  the  normal  incubation  period 
for  bobwhites. 

Although  only  1 of  10  eggs  hatched  in  our 
first  observation  of  prolonged  incubation, 
Murray  and  Frye  (1957)  suggest  that  the 
hatching  of  even  one  egg  is  sufficient  to  sat- 
isfy the  nesting  instinct.  In  our  observation, 
however,  the  hen  continued  incubation  even 
though  only  one  egg  hatched.  Hurst  (1978) 
observed  a similar  phenomenon,  during  which 
a bobwhite  hen  continued  incubation  of  par- 
tially hatched,  dead  chicks.  The  clutch  con- 
sisted of  10  eggs:  1 infertile,  1 completely 
hatched,  and  8 partially  hatched.  The  eight 
partially  hatched  eggs  contained  fully  devel- 
oped chicks  that  had  pipped  and  partially 
ringed  their  eggshells  but  had  become  “en- 
tombed.” Hurst  (1978)  did  not  report  the 
length  of  time  that  the  hen  remained  on  the 
partially  hatched  eggs. 

Prolonged  incubation  is  thought  to  provide 
a safety  margin  for  eggs  that  take  longer  than 
normal  to  hatch  (Skutch  1962,  Holcomb 
1970).  However,  Holcomb  (1970)  suggested 
that  prolonged  incubation  would  be  maladap- 
tive for  species  capable  of  renesting.  Bob- 
whites  commonly  renest  two  or  three  times 
per  breeding  season,  regardless  of  previous 
nest  fate  (Stoddard  1931).  Given  that  the  two 
records  of  prolonged  incubation  occurred  to- 
ward the  end  (July-August)  of  the  normal 
nesting  season  for  bobwhites  (May-August), 
the  opportunity  for  renesting  was  limited  and 


116 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


may  have  contributed  to  prolonged  incuba- 
tion. 

ACKNOWLEDGMENTS 

This  research  was  supported  by  the  George  and 
Mary  Josephine  Hamman  Foundation;  Amy  Shelton 
McNutt  Charitable  Trust;  William  A.  and  Madeline 
Welder  Smith  Foundation;  Bob  and  Vivian  Smith 
Foundation;  Robert  J.  Kleberg,  Jr.,  and  Helen  C.  Kle- 
berg Foundation;  Texas  State  Council  of  Quail  Unlim- 
ited; the  South  Texas,  Houston,  East  Texas,  and  Alamo 
Chapters  of  Quail  Unlimited;  the  South  Texas  Celeb- 
rity Weekend;  Quail  Associates;  private  contributions; 
King  Ranch,  Inc.;  and  San  Tomas  Hunting  Camp.  We 
thank  B.  M.  Ballard,  D.  G.  Hewitt,  and  three  anony- 
mous reviewers  for  providing  helpful  comments  on  an 
earlier  version  of  this  manuscript.  Data  for  this  re- 
search were  provided  by  the  South  Texas  Quail  Re- 
search Project.  This  manuscript  is  Caesar  Kleberg 
Wildlife  Research  Institute  Publication  Number  OS- 
Ill. 

LITERATURE  CITED 

Bartel,  K.  E.  1986.  Another  record  of  runt  eggs  for 
the  Tree  Swallow.  North  American  Bird  Bander 
1 1:3-4. 

Bent,  A.  C.  1932.  Eastern,  Florida,  and  Texas  Bob- 
white.  Pages  9-36  in  Life  histories  of  North 
American  gallinaceous  birds.  U.S.  National  Mu- 
seum Bulletin,  no.  162.  [Reprinted  1963,  Dover 
Publications,  New  York.] 

Case,  R.  M.  and  R.  J.  Robel.  1974.  Bioenergetics  of 
the  Bobwhite.  Journal  of  Wildlife  Management 
38:638-652. 

Hernandez,  F.  1999.  The  value  of  prickly  pear  cactus 
as  nesting  cover  for  Northern  Bobwhite.  Ph.D. 
dissertation,  Texas  A&M  University,  Kingsville 
and  College  Station. 

Holcomb,  L.  C.  1970.  Prolonged  incubation  behaviour 
of  Red-winged  Blackbird  incubating  several  egg 
sizes.  Behaviour  36:74-83. 

Hurst,  G.  A.  1978.  Unusual  incubation  behavior  in 
Bobwhite.  Wilson  Bulletin  90:290-291. 


Koenig,  W.  D.  1980a.  The  determination  of  runt  eggs 
in  birds.  Wilson  Bulletin  92:103-107. 

Koenig,  W.  D.  1980b.  The  incidence  of  runt  eggs  in 
woodpeckers.  Wilson  Bulletin  92:169-176. 

Mallory,  M.  L„  L.  Kiff,  R.  G.  Clark,  T.  Bowman, 
P.  Blums,  A.  Mednis,  and  R.  T.  Alisauskas. 
2004.  The  occurrence  of  runt  eggs  in  waterfowl 
clutches.  Journal  of  Field  Ornithology  75:209- 
217. 

Manning,  T.  H.  and  B.  Carter.  1977.  Incidence  of 
runt  eggs  in  the  Canada  Goose  and  Semipalmated 
Sandpiper.  Wilson  Bulletin  89:469. 

Marks,  J.  S.  1983.  Prolonged  incubation  by  a Long- 
eared Owl.  Journal  of  Field  Ornithology  54:199- 
200. 

Mulvihill,  R.  S.  1987.  Runt  eggs:  a discovery,  a syn- 
opsis, and  a proposal  for  future  study.  North 
American  Bird  Bander  12:94-96. 

Murray,  R.  W.  and  O.  E.  Frye,  Jr.  1957.  The  Bob- 
white  quail  and  its  management  in  Florida.  Game 
Publication  2,  Florida  Game  and  Fresh  Water  Fish 
Commission,  Tallahassee,  Florida. 

M’William,  J.  M.  1927.  Some  abnormal  eggs  of  wild 
birds.  Scottish  Naturalist  66:108-110. 

Pearl,  R.  and  M.  R.  Curtis.  1916.  Studies  on  the 
physiology  of  reproduction  in  the  domestic 
fowl — XV.  Dwarf  eggs.  Journal  of  Agricultural 
Research  6:977-1042. 

Powers,  L.  R.  1978.  Record  of  prolonged  incubation 
by  a Killdeer.  Auk  95:428. 

Ricklefs,  R.  E.  1975.  Dwarf  eggs  laid  by  a starling. 
Bird  Banding  46:169. 

Romanoff,  A.  L.  and  A.  J.  Romanoff.  1949.  The  avi- 
an egg.  John  Wiley  and  Sons,  New  York. 

Rothstein,  S.  I.  1973.  The  occurrence  of  unusually 
small  eggs  in  three  species  of  songbirds.  Wilson 
Bulletin  85:340-342. 

Skutch,  A.  F.  1962.  The  constancy  of  incubation.  Wil- 
son Bulletin  74:115-152. 

Stoddard,  H.  L.  1931.  The  Bobwhite  quail:  its  habits, 
preservation  and  increase.  Charles  Scribner’s 
Sons,  New  York. 

Sutcliffe,  S.  1982.  Prolonged  incubation  behavior  in 
Common  Loons.  Wilson  Bulletin  94:361-362. 

Zeleny,  L.  1983.  Miniature  bluebird  eggs.  Sialia  5: 
127-128. 


The  Wilson  Journal  of  Ornithology  1 18(1):  1 17-1 19,  2006 


Once  Upon  a 1 lime  in  [American  Ornithology 


George  Bird  Grinnell,  the  “father  of  Amer- 
ican conservation,”  was  born  in  1849.  He  ul- 
timately would  spearhead  a movement  for  the 
preservation  of  North  American  waterfowl, 
lay  the  foundation  for  the  national  park  sys- 
tem, lead  the  way  in  ending  the  commercial 
taking  of  wildlife,  and  help  found  the  Amer- 
ican Ornithologists’  Union,  Boone  and  Crock- 
ett Club,  and  Audubon  Society.  Schooled  for 
a time  by  Lucy  Bakewell  Audubon,  John 
James  Audubon’s  widow,  Grinnell  grew  up  in 
Audubon  Park,  the  former  12-ha  Audubon  es- 
tate on  Manhattan  Island  in  New  York  City. 
“Grandma”  Audubon’s  tutelage,  hunting  ex- 
periences with  Audubon’s  grandson.  Jack,  and 
frequent  visits  to  the  homes  of  Audubon’s 
sons,  Victor  and  John  Woodhouse  Audubon — 
where  rifles  and  shotguns,  powder  horns  and 
shot,  animal  trophies,  bird  paintings,  and  box- 
es of  bird  skins  were  always  about — were  for- 
mative, and  predisposed  Grinnell’s  future  as  a 
naturalist  and  conservationist.  At  the  age  of 
25,  four  years  after  receiving  a B.A.  from  Yale 
University  in  1870,  Grinnell  was  asked  by  pa- 
leontologist O.  C.  Marsh,  head  of  the  Peabody 
Museum  in  New  Haven,  Connecticut,  to  ac- 
company him  on  an  army-sponsored  expedi- 
tion of  the  Black  Hills,  South  Dakota.  Com- 
manded by  Col.  George  Armstrong  Custer,  the 


60-day  expedition  set  out  on  2 July  1874  from 
Fort  Abraham  Lincoln,  just  across  the  Mis- 
souri River  from  Bismarck,  North  Dakota.  Be- 
cause trouble  was  expected  from  hostile  In- 
dians, the  military  command  consisted  of  10 
companies  of  the  7th  Calvary,  2 companies  of 
Infantry,  and  a battery  of  3 Gatling  guns.  In 
all,  there  were  1,200  men  and  their  horses, 
wagons,  a beef  herd,  and  Indian  scouts  (Fig. 
1).  Military  goals  were  to  explore  unmapped 
Indian  Territory  and  investigate  rumors  of 
gold;  the  scientists,  or  “bug  hunters”  as  the 
military  called  them,  were  along  to  collect 
specimens  and  fossils. 

The  following  ornithological  event  was  re- 
corded by  Grinnell  as  he  accompanied  Cus- 
ter’s exploration  of  the  Black  Hills,  about  22 
months  before  the  battle  of  Little  Bighorn 
(Greasy  Grass).  The  incident  took  place  in  late 
August  as  the  troops  were  on  their  return  trip 
to  Fort  Abraham  Lincoln  from  the  west.  Grin- 
nell was  subsequently  invited  to  accompany 
the  inglorious  1876  expedition  as  naturalist, 
but  he  had  a professional  conflict  that  kept 
him  home.  The  original  reference  is  Grinnell, 
G.  B.  1875.  Zoological  Report.  Pages  79-102 
in  Report  of  a reconnaissance  of  the  Black 
Hills  of  Dakota  made  in  the  summer  of  1874. 
(W.  Ludlow).  U.S.  Army  Department  of  En- 
gineers.—FRITZ  L.  KNOPF 


August  28. — About  6.30  a.m.,  while  we  were  halting  for  a short  time  on  a little 
knoll,  a most  interesting  and  exciting  chase  came  under  my  observation.  The  ground 
was  wet  from  the  rain  that  had  but  just  ceased  to  fall,  and  the  men  were,  most  of 
them,  standing  by  their  horses,  instead  of  lying  asleep  on  the  ground,  as  is  usually 
the  case  when  a halt  is  made.  I was  looking  out  over  the  plain,  when  I observed 
two  birds  in  rapid  flight,  approaching  the  hill  where  we  were  standing.  They  flew 
with  astonishing  velocity,  and  it  was  but  a short  time  before  they  were  quite  near 
us.  From  the  manner  of  their  flight,  I at  first  thought  they  were  two  falcons  engaged 
in  play,  but  a nearer  view  showed  me  that  the  foremost  bird  was  much  the  smallest, 
and  that  it  was  making  most  strenuous  efforts  to  escape  from  its  pursuer  by  darting 
and  twisting  from  one  side  to  the  other,  up  or  down,  or  by  straightforward  flight. 
In  one  of  its  turnings  it  came  quite  close  to  the  column,  and,  forgetting  in  its  intense 
fear  its  natural  shyness,  it  darted  in  among  the  men  and  horses.  The  larger  bird,  a 
peregrine  falcon,  as  I could  now  see,  hesitated  not  an  instant,  but  dashed  after, 
following  the  object  of  its  pursuit  in  every  cut  and  twist  that  it  made,  now  passing 
under  the  horses,  now  low  over  their  backs  or  close  to  the  men’s  heads.  After, 
perhaps,  a minute  of  rapid  pursuit,  the  smaller  bird  by  a quick  double  put  a group 
of  men  and  horses  between  itself  and  the  falcon,  and  then  darted  swiftly  along  the 
ground  to  where  I was  standing,  an  interested  observer.  Here,  almost  exhausted,  it 


117 


118 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


FIG.  1 . The  1 874  Custer  Expedition  returning  from  the  Black  Hills  of  South  Dakota,  photographed  by 
William  H.  Illingworth.  The  expedition  included  1,200  men  and  1 10  wagons,  here  seen  between  the  Black  Hills 
and  Fort  Abraham  Lincoln,  Dakota  Territory,  near  the  modern  day  South  Dakota  and  North  Dakota  border. 
Photograph  taken  in  the  vicinity  of  George  Bird  Grinnell’s  account  of  a Peregrine  Falcon  ( Falco  peregrinus ) 
pursuing  a Passenger  Pigeon  ( Ectopistes  migratorius).  Custer  is  on  horseback  in  the  foreground  with  his  wagon 
behind  him;  in  the  distance,  the  expedition  is  aligned  in  four  columns.  Grinnell  is  believed  to  be  one  of  three 
individuals  (in  the  middle  on  a mule)  mounted  and  slightly  forward  of  the  expedition.  Photo  courtesy  of  the 
South  Dakota  State  Historical  Society-State  Archives. 


alighted  on  the  saddle  of  a horse  standing  within  arm’s  length  of  me,  and  I was 
able  to  distinguish  that  it  was  a passenger  pigeon,  ( Ectopistes  migratoria).  Mean- 
while, the  falcon,  baffled  for  a moment,  had  risen  30  feet  in  the  air,  and  was 
hovering  over  the  group,  looking  for  his  prey.  Hardly  ten  seconds  had  elapsed  since 
the  pigeon  alighted,  when  he  saw  his  pursuer  above  him,  and,  terror-stricken  by 
the  sight,  the  luckless  bird  darted  away  again  over  the  open  prairie.  The  falcon 
followed,  and  the  doubling  and  twisting  recommenced  before  they  had  gone  a 
quarter  of  a mile.  The  pigeon  once  tried  to  regain  the  shelter  of  the  command,  but 
his  relentless  pursuer  cut  him  off  and  drove  him  toward  the  plain,  and,  in  a few 
seconds,  by  a tremendous  burst  of  speed,  caught  up  to  his  victim,  and,  throwing 
out  his  powerful  feet,  seized  him,  and,  without  checking  his  flight,  bore  him  off  to 


ONCE  UPON  A TIME  IN  AMERICAN  ORNITHOLOGY 


1 19 


a neighboring  butte,  there  to  devour  him.  It  was  a splendid  sight,  and  I can  compare 
it  to  nothing  unless  it  be  a scene  of  ancient  falconry,  the  only  difference  being  that 
the  birds  were  so  much  more  evenly  matched  than  in  the  old-time  sports.  It  would, 
I think,  be  difficult  to  name  a harder  bird  to  catch  than  the  pigeon,  and,  perhaps, 
the  only  bird  that  can  do  it  in  a straight-away  chase  is  the  peregrine  falcon.  I should 
mention  that  the  soldiers  made  efforts  to  frighten  the  hawk  away  by  shouting  and 
throwing  their  hats  at  it,  but  it  paid  no  attention  to  their  demonstrations,  except 
once  to  stretch  out  its  feet  as  if  to  grasp  a hat  that  sailed  close  by  it. 


The  Wilson  Journal  of  Ornithology  1 18(1):  120-127,  2006 


Ornithological  Literature 

Compiled  by  Mary  Gustafson 


THE  NORTH  AMERICAN  BANDERS’ 
MANUAL  FOR  BANDING  SHOREBIRDS 
(CH ARADRHFORMES , SUBORDER  CHA- 
RADRII).  By  Cheri  L.  Gratto-Trevor.  North 
American  Banding  Council,  Point  Reyes  Sta- 
tion, California.  2004:  45  pp.,  4 color  plates,  15 
figures,  1 table,  8 appendices.  Available  at  no 
charge  from  www.nabanding.net/nabanding/ 
pubs.html. — This  manual,  intended  to  be  an  in- 
tegral part  of  the  North  American  Banding 
Council  Study  Guide,  should  be  required  read- 
ing for  anyone  capturing  shorebirds  (waders). 
After  introductory  sections  on  the  ethics  of 
banding  and  some  of  the  factors  to  consider 
when  devising  a study  program,  this  publica- 
tion offers  a synthesis  of  the  various  methods 
used  to  capture  shorebirds  in  their  breeding, 
passage,  and  wintering  habitats.  This  group  of 
species  has  tested  human  ingenuity;  thus, 
many  of  the  1 80  references  from  the  published 
literature  included  in  this  manual  are  about 
trapping  techniques.  There  is  also  useful  ma- 
terial on  marking  techniques,  including  bands, 
color  bands,  dye-marking,  radio  tracking,  and 
so  on.  Although  much  of  this  material  is  avail- 
able elsewhere,  it  was  scattered  in  many 
sources,  and  it  is  well  worthwhile  having  it 
compiled  in  one  publication.  The  many  per- 
sonal communications  add  to  the  book’s  val- 
ue, including  many  of  the  little  tricks  that  are 
often  passed  on  by  word  of  mouth. 

The  manual  also  includes  a useful  table  that 
summarizes  all  that  a bander  needs  to  know 
for  each  species:  American  Ornithologists’ 
Union  code.  Birds  of  North  America  refer- 
ence, band  size,  methods  for  determining  age 
and  sex,  and  any  problems  often  encountered 
when  trapping,  handling,  and  banding  the  spe- 
cies. 

As  an  English  ringer  of  wading  birds  (albeit 
with  experience  in  banding  shorebirds  on  four 
continents),  I was  struck  by  the  different  ap- 
proach taken  in  this  publication.  In  many  oth- 
er countries,  extensive  long-term  studies  of 
waders  carried  out  primarily  by  volunteers 
have  provided  ample  opportunities  to  develop 
methods  for  safely  handling  hundreds,  and  oc- 


casionally thousands,  of  birds  at  a time.  The 
target  audience  of  this  manual,  however,  is 
North  American  banders,  who  often  are  pro- 
fessional ornithologists — but  inexperienced  in 
studying  shorebirds — usually  undertaking 
short-term  studies,  often  of  small  numbers  of 
birds.  Capturing  shorebirds  can  indeed  be  a 
specialized  art,  at  times  potentially  dangerous 
for  birds  and  for  banders,  and  should  not  be 
undertaken  lightly.  The  exceptionally  detailed 
and  thorough  treatment  here,  of  all  aspects  of 
the  process,  should  help  ornithologists  maxi- 
mize the  scientific  value  of  their  work  on 
shorebirds,  and  minimize  the  danger  to  them- 
selves or  their  subjects.  The  emphasis 
throughout  is  on  safe  methods  of  capturing 
and  handling.  Given  the  international  knowl- 
edge base  on  these  birds — many  of  which  are 
themselves  great  international  travelers — the 
author  has  succeeded  in  pulling  together  in- 
formation from  around  the  world  to  develop 
this  manual,  and  all  banders  can  probably 
learn  something  from  reading  it. 

There  is,  unfortunately,  one  significant  fail- 
ing in  the  publication — an  Appendix  on  age- 
ing calidrid  shorebirds  in  which  the  photo- 
graphs are  the  worst  that  I have  ever  seen  pub- 
lished. The  birds’  feathers  are  so  disheveled 
that,  not  only  do  they  reflect  poorly  on  band- 
ing, they  make  it  very  difficult  to  discern  the 
plumage  characters  that  the  photographs  are 
intended  to  illustrate.  These  days,  with  pho- 
tographic equipment  so  easy  to  use,  and,  in- 
deed, with  so  many  high-quality  images  ap- 
pearing on  Web  sites  and  elsewhere,  there  is 
no  excuse  for  publishing  such  poor  photo- 
graphs. The  flawed  appendix  should  not  de- 
tract from  the  value  of  this  publication,  but 
users  of  the  manual  should  obtain  other  ref- 
erence materials  for  ageing  shorebirds.  This 
useful  manual  should  surely  be  obligatory 
reading  for  all  who  capture  shorebirds. 

— DAVID  NORMAN,  Merseyside  Ringing 
Group,  England,  and  Carnegie  Museum  of 
Natural  History,  Powdermill  Avian  Research 
Center,  Pennsylvania;  e-mail:  david.norman@ 
physics.org 


120 


ORNITHOLOGICAL  LITERATURE 


121 


A PASSION  FOR  WILDLIFE:  THE  HIS- 
TORY OF  THE  CANADIAN  WILDLIFE 
SERVICE.  By  J.  Alexander  Burnett.  UBC 
Press,  Vancouver,  British  Columbia.  2003:  331 
pp.,  numerous  photos.  ISBN:  0774809604, 
C$85.00  (cloth).  ISBN:  0774809612,  C$27.95 
(paper). — In  1947,  in  what  appears  to  be  an 
endless  series  of  reorganizations,  the  Canadian 
government  reorganized  the  Department  of 
Resources  and  Government  and  gave  birth  to 
the  Dominion  Wildlife  Service,  which  carried 
much  of  the  responsibility  for  wildlife  in  Can- 
ada. Initially,  the  agency  was  staffed  by  fewer 
than  30  people,  but  it  included  several  sea- 
soned ornithologists,  including  George  Boyer 
and  Oliver  Hewitt.  The  early  years  were  chal- 
lenging— in  1949,  Newfoundland  and  Labra- 
dor joined  the  Canadian  Confederation,  bring- 
ing with  them  segments  of  their  population 
that  traditionally  harvested  vast  numbers  of 
seabirds  and  their  eggs.  In  1950,  the  Wildlife 
Service  became  a division  of  the  National 
Parks  Branch,  and  chief  Harrison  secured  per- 
mission to  rename  the  division  the  Canadian 
Wildlife  Service  (CWS) — the  name  it  still 
holds  today.  This  book  recounts  the  nearly 
half-century  history  of  the  CWS. 

The  book  is  divided  into  10  chapters  and 
an  epilogue.  The  first  chapter  covers  the  gen- 
esis of  the  CWS  and  provides  an  historical 
context  through  a synopsis  of  Canadian  wild- 
life policy  up  to  the  1940s.  The  remaining 
chapters  are  topical,  each  focusing  on  an  as- 
pect of  the  CWS’s  diverse  agenda.  Chapter  2 
describes  the  CWS  involvement  in  enforcing 
the  Migratory  Birds  Convention  Act  of  1917. 
This  included  the  difficult  and  sensitive  task, 
conducted  by  Leslie  Tuck  and  others,  of 
bringing  some  level  of  enforcement  to  the  ru- 
ral populations  of  Newfoundland  and  Labra- 
dor who  depended  on  seabird  harvest  for  sub- 
sistence. Managers  required  information,  and 
surveys  and  other  scientific  research  became 
an  integral  part  of  the  CWS.  Chapter  3 em- 
phasizes working  with  birds — during  the  first 
50  years  of  the  CWS,  ornithology  was  the  pre- 
eminent scientific  concern.  In  the  early  years, 
waterfowl  research  tended  to  dominate  the 
agenda,  but  seabird  research  became,  and  re- 
mains, important,  and  research  has  been  di- 
rected at  a broad  spectrum  of  problems  (e.g., 
bird  strikes  at  airports).  The  contributions  of 
prominent  CWS  seabird  biologists  (e.g.,  Les- 


lie Tuck,  David  Nettleship,  Hans  Blokpoel, 
Kees  Vermeer,  Rob  Butler,  Tony  Gaston,  and 
many  others)  are  chronicled  in  the  chapter. 

Chapters  4 and  5 cover  mammals  and  fish, 
and  chapter  6 describes  the  shift  in  conser- 
vation strategy — from  a focus  on  species  to 
habitat  preservation  and  continental-scale 
thinking — that  began  in  the  1970s.  The  chap- 
ter also  traces  changes  in  the  CWS  with  re- 
gionalization of  administrative  control.  Chap- 
ter 7 describes  efforts  to  foster  public  aware- 
ness and  understanding  of  wildlife  values  and, 
hence,  securing  public  support  for  conserva- 
tion initiatives.  This  involved  the  cooperation 
of  CWS  personnel  with  filmmakers  and  the 
establishment  of  wildlife  interpretation  cen- 
ters. This  chapter  also  tells  the  painful  story 
of  consolidation  during  the  late  1970s  and  ear- 
ly 1980s,  when  federal  budget  cuts  caused  a 
thorough  reexamination  of  CWS  priorities,  as 
the  government  cut  off  funds,  for  example,  for 
the  wildlife  interpretation  centers. 

Chapter  8 deals  with  the  growing  field  of 
wildlife  toxicology,  precipitated  by  the  dev- 
astating effects  of  DDT.  It  describes  programs 
designed  to  investigate  avian  ingestion  of 
crude  oil,  as  well  as  problems  with  pesticide 
use  in  agriculture  and  forestry.  Chapter  9 cov- 
ers endangered  species,  including  many  birds. 
Chapter  10,  Defining  the  Rules:  Wildlife  Gov- 
ernance, describes  a series  of  initiatives  that 
had  come  to  fruition  by  the  1990s  (e.g., 
amendments  to  the  Canadian  Wildlife  and  Mi- 
gratory Birds  Convention  acts,  Ramsar  des- 
ignation for  suitable  wetlands  sites  in  Canada, 
and  the  Western  Hemisphere  Shorebird  Re- 
serve Network).  The  Epilogue,  The  Canadian 
Wildlife  Service:  A Work  in  Progress,  high- 
lights important  aspects  of  the  CWS  and 
brings  closure  to  this  historical  account. 

This  book  is  thoroughly  researched  and 
very  well  written.  The  author  does  not  shy 
away  from,  or  gloss  over,  problems  that  have 
been  part  of  the  CWS  (e.g.,  federal  versus  pro- 
vincial authority,  research  versus  management 
mandates,  or  the  disastrous  budget  cuts  of  the 
early  1980s).  He  has  managed  to  provide  an 
even-handed  history  of  the  CWS.  I am  sure 
that  many  people  who  have  had  careers  in  the 
CWS  would  disagree  on  details  and,  perhaps, 
emphasis,  but  I find  this  a well-balanced  his- 
tory of  an  important  North  American  conser- 


122 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  1,  March  2006 


vation  institution.  It  should  be  of  interest  to 
any  historically  oriented  ornithologist. 
—WILLIAM  E.  DAVIS,  JR.,  Boston  Univer- 
sity, Boston,  Massachusetts;  e-mail:  wedavis@ 
bu.edu 


WHALES  & DOLPHINS  OF  THE 
WORLD.  By  Mark  P.  Simmonds,  photogra- 
phy by  seapics.com.  The  MIT  Press,  Cam- 
bridge, Massachusetts.  2004:  160  pp.,  180 
full-color  photographs,  1 color  map.  ISBN 
0262195194.  $29.95  (cloth).— This  book 
takes  on  the  challenging  task  of  introducing 
80-plus  species  of  whales  and  dolphins  to  the 
general  public,  while  at  the  same  time  provid- 
ing many  spectacular  photographs  to  get 
across  the  author’s  conservation  message. 
This  large-format  book  (9.5'  X 12.5')  takes 
advantage  of  its  size  by  including  lots  of  pho- 
tos. The  photos  come  from  seapics.com,  an 
image  library  containing  the  works  of  over 
200  marine  and  underwater  photographers, 
many  of  whom  are  internationally  known. 
This  is  the  marine  wildlife  equivalent  of  the 
bird  photo  archives  provided  by  VIREO  (Vi- 
sual Resources  for  Ornithology),  and  enables 
the  author  to  illustrate  his  book  with  some  tru- 
ly stunning  photos,  including  some  of  very 
rare  species  like  that  of  a breaching  Blain- 
ville’s  beaked  whale  {Mesoplodon  densiros- 
tris).  The  text  is  easy  to  read  and  clearly  writ- 
ten to  introduce  whales  and  dolphins  to  the 
general  public  so  that  people  will  become 
more  informed  about  the  conservation  issues 
affecting  cetaceans.  At  the  same  time,  the 
book  includes  some  of  the  latest  scientific 
findings  and  taxonomic  changes. 

The  book  is  divided  into  5 chapters.  The 
first  2 chapters  cover  whale  biology,  behavior, 
and  some  general  information  for  each  of  the 
14  cetacean  families.  Because  some  families 
are  poorly  known  or  include  fewer  species, 
the  general  accounts  can  vary  in  length  from 
just  2 pages  (porpoises  and  beaked  whales)  to 
12  pages  (marine  dolphins).  The  family  ac- 
counts include  both  general  descriptive  infor- 
mation and  detailed  information,  such  as 
breeding  biology,  habitat,  prey,  and  feeding 
strategies  if  known.  Each  account  also  in- 
cludes information  on  current  and  past  con- 
servation threats,  such  as  whaling  and  habitat 


disturbances.  There  are  also  short  sections  on 
cetacean  physiological  adaptations,  migration, 
intelligence,  and  echolocation.  The  first  2 
chapters  compose  two-thirds  of  the  book 
(—100  pp.),  while  the  final  third  (60  pp.)  in- 
cludes a chapter  on  interactions  between  man 
and  whales  and  2 chapters  on  conservation 
threats  and  current  measures  being  taken  to 
protect  cetaceans.  At  the  end  of  the  book  is  a 
rather  uninteresting  2-page  color  map  of  the 
world  showing  cetacean  habitats;  however,  it 
does  not  include  much  detail  other  than  basic 
ocean  temperature  zones  and  river  dolphin 
ranges.  There  is  also  a nice  comprehensive  list 
of  all  the  cetacean  species,  subdivided  by  fam- 
ilies, that  includes  Latin  names,  a bibliogra- 
phy of  Web  sites  and  book  titles,  a very  gen- 
eral and  basic  1-page  glossary,  and  a page  of 
interesting  facts  and  figures  on  cetaceans  (e.g., 
the  longest-lived  mammal — the  bowhead 
whale,  Balaena  mysticetus — can  live  more 
than  200  years). 

The  180  photographs  are  what  really  make 
this  book  interesting,  especially  since  it  is  just 
160  pages  long.  Included  are  some  incredible 
action  shots  like  an  orca  ( Orcinus  orca ) just 
about  to  make  a meal  of  a mako  shark  ( Isurus 
oxyrinchus ) and  another  of  copulating  Atlantic 
spotted  dolphins  ( Stenella  frontalis ).  My  fa- 
vorite was  a photo  of  a snorkler  alongside  a 
sperm  whale  ( Physeter  macrocephalus ) that 
fills  two  full  pages,  although  it  is  somewhat 
ruined  by  a large  chapter  heading  on  one  of 
the  pages.  By  using  photos  from  seapics.com, 
the  author  is  able  to  draw  from  an  almost  lim- 
itless collection  of  quality  images.  Many  of 
the  photos  appear  to  be  published  for  the  first 
time  in  this  book,  although  some  have  previ- 
ously appeared  in  other  publications.  None- 
theless, the  quality  of  the  reproductions  is 
good. 

This  is  an  easy  book  to  recommend  to  any- 
one with  an  even  slight  interest  in  marine 
mammals.  As  stated  in  the  introduction,  the 
book  “is  intended  as  both  a celebration  of  the 
whales  and  dolphins  of  the  world  and  an  in- 
troduction to  their  diversity,  biology  and  con- 
servation.” It  certainly  meets  that  goal.  The 
author  is  the  Director  of  Science  at  the  Whale 
and  Dolphin  Conservation  Society,  and  the 
book  is  written  to  promote  their  conservation 
views;  in  fact,  the  royalties  from  the  book  are 
donated  to  the  society.  Whereas  several  other 


ORNITHOLOGICAL  LITERATURE 


123 


recent  books  serve  as  excellent  marine  mam- 
mal field  guides,  this  book  was  intended  for  a 
wider  audience  and  would  make  a nice  addi- 
tion to  any  library.  You  will  probably  find 
yourself  looking  through  and  marveling  at  the 
photos  again  and  again,  as  I did. — MICHAEL 
FRITZ,  See  Life  Paulagics,  Seaville,  New  Jer- 
sey; e-mail:  mike@paulagics.com 


PARTNERS  IN  FLIGHT:  NORTH  AMER- 
ICAN LANDBIRD  CONSERVATION 
PLAN.  By  Terrell  D.  Rich,  Carol  J.  Beard- 
more,  Humberto  Berlanga,  Peter  J.  Blancher, 
Michael  S.  W.  Bradstreet,  Greg  S.  Butcher, 
Dean  W.  Demarest,  Erica  H.  Dunn,  W.  Chuck 
Hunter,  Eduardo  E.  Inigo-Elias,  Judith  A. 
Kennedy,  Arthur  M.  Martell,  Arvind  O.  Pan- 
jabi, David  N.  Pashley,  Kenneth  V.  Rosen- 
berg, Christopher  M.  Rustay,  J.  Steven  Wendt, 
and  Tom  C.  Will.  Cornell  Lab  of  Ornithology, 
Ithaca,  New  York.  2004:  84  pp.  Available  at 
no  charge  from  www.partnersinflight.org. — 
The  long-awaited  Partners  in  Flight  [PIF] 
Landbird  Conservation  Plan  arrived  with 
much  fanfare,  and  deservedly  so.  This  broad 
plan  will  serve  as  the  starting  point  for  bird 
conservation  planning  throughout  the  U.S. 
and  Canada.  A future  planned  revision  will 
incorporate  Mexican  species,  expanding  the 
utility  of  the  plan  to  the  continental  scale. 

The  plan  starts  with  a description  of  how  it 
was  created  and  how  it  should  be  implement- 
ed, in  addition  to  definitions  of  terms  and  var- 
ious ranking  factors.  A total  of  448  species 
that  nest  in  North  America  are  included. 
Landbirds  are  defined  to  include  species  in  45 
families.  These  families  include  Cathartidae 
plus  those  within  the  following  orders:  Galli- 
formes,  Falconiformes,  Columbiformes,  Psit- 
taciformes,  Cuculiformes,  Strigiformes,  Ca- 
primulgiformes,  Apodiformes,  Trogoniformes, 
Coraciiformes,  Piciformes,  and  Passeriformes; 
13  more  families  (including  Tinamidae)  will 
be  added  when  the  plan  is  revised  to  include 
Mexico.  The  plan  also  provides  guidance  on 
Conservation  Issues  and  Recommendations 
for  seven  Avifaunal  Biomes:  Arctic,  Northern 
Forest,  Pacific,  Intermountain  West,  South- 
west, Prairie,  and  Eastern. 

At  the  core  of  the  plan  are  the  PIF  Species 
of  Continental  Importance,  composed  of  100 


Watch  List  Species  and  91  Stewardship  Spe- 
cies. The  Watch  List  Species  were  determined 
through  Assessment  Scores  (from  1 to  5)  of 
the  Population  Size,  Breeding  Distribution, 
Non-breeding  Distribution,  Threats  to  Breed- 
ing Population,  Threats  to  Non-breeding  Pop- 
ulation, and  Population  Trend  for  each  indi- 
vidual species.  The  Combined  Score  is  deter- 
mined by  summing  Population  Score,  the 
highest  of  the  Distribution  and  Threats  scores, 
and  the  Population  Trend  score,  for  a maxi- 
mum of  20. 

Species  with  Combined  Scores  of  14  and 
up  comprise  the  Watch  List;  species  with  a 
Combined  Score  of  13  and  a Population  Trend 
of  5 were  also  added  to  the  Watch  List.  Six 
species  had  Combined  Scores  of  12  and  Trend 
Scores  of  5,  including  Northern  Bobwhite 
( Colinus  virginianus ),  Loggerhead  Shrike 
( Lanius  ludovicianus ),  Field  Sparrow  ( Spizella 
pusilla),  Lark  Sparrow  ( Chondestes  gramma- 
cus ),  Black-throated  Sparrow  ( Amphispiza  bil- 
ineata ),  and  Grasshopper  Sparrow  (Ammodra- 
mus  savannarum ).  One  species,  the  Eastern 
Meadowlark  ( Sturnella  magna),  had  a Com- 
bined Score  of  11  and  a Trend  Score  of  5,  but 
no  species  had  a lower  Combined  Score  and 
a Trend  Score  of  5.  A whopping  43  species 
that  had  Combined  Scores  of  13  and  Trend 
Scores  of  less  than  5 did  not  make  the  Watch 
List. 

Several  species  rated  the  maximum  score, 
including  Gunnison  Sage-Grouse  ( Centrocer - 
cus  minimus ),  Lesser  Prairie-Chicken  ( Tym - 
panuchus  pallidicinctus ),  California  Condor 
( Gymnogyps  californianus ),  Thick-billed 
( Rhynchopsitta  pachyrhyncha)  and  Red- 
crowned  parrots  ( Amazona  viridigenalis ),  Ivo- 
ry-billed Woodpecker  ( Campephilus  princi- 
palis), Black-capped  Vireo  ( Vireo  atricapilla ), 
Florida  Scrub-Jay  ( Aphelocoma  coerules- 
cens),  and  Bachman’s  ( Vermivora  bachmanii) 
and  Kirtland’s  warblers  ( Dendroica  kirtlan- 
dii ).  This  varied  group  includes  species  absent 
from  the  USFWS  endangered  species  list 
(Gunnison  Sage-Grouse,  Lesser  Prairie  Chick- 
en, Thick-billed  and  Red-crowned  parrots), 
two  species  that  were  previously  all  but  writ- 
ten off  as  extinct  but  present  on  the  endan- 
gered species  list  (Ivory-billed  Woodpecker 
and  Bachman’s  Warbler),  and  species  that  are 
heavily  managed  endangered  species  (Califor- 


124 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


nia  Condor,  Black-capped  Vireo,  Florida 
Scrub-Jay,  and  Kirtland’s  Warbler). 

The  Population  Size  Ranking  Factor  in- 
cludes a Global  Population  Estimate,  a number 
difficult  to  determine  for  most  bird  species.  I 
find  these  estimates  to  be  interesting  and 
thought  provoking,  though  I continue  to  be 
puzzled  by  the  disparity  in  population  esti- 
mates among  species.  The  percentage  of  the 
population  residing  in  the  U.S.  and  Canada  is 
also  estimated,  and,  for  many  species  included 
in  the  plan,  < 1 % of  the  global  population  nests 
in  the  U.S.  or  Canada.  Expansion  of  the  plan 
to  Mexico  will  be  critical  to  future  conserva- 
tion efforts.  Although  Population  Trend  infor- 
mation for  each  species  is  used  as  a part  of  the 
Combined  Score,  the  information  in  the  Trend 
Score  is  qualified  by  using  the  Monitoring 
Needs  information.  The  Monitoring  Needs 
identifies  species  for  which  trend  data  are  lack- 
ing or  imprecise,  as  well  as  species  affected  by 
poor  survey  coverage  (e.g.,  those  in  boreal  for- 
ests and  far  northern  areas).  The  remainder  of 
the  species  that  lack  an  identified  Monitoring 
Need  have  a qualifier,  that  while  monitoring  is 
considered  adequate  “some  issues,  such  as 
bias,  may  not  have  been  accounted  for.” 

While  it  is  easy  to  find  fault  with  individual 
data  points  or  certain  aspects  of  the  plan,  the 
utility  of  the  ranking  process  is  evident  in  the 
results.  Without  debating  which  species  are 
facing  threats,  what  effect  those  threats  might 
have  on  a population,  or  whether  a Ranking 
Factor  should  be  increased  or  decreased,  the 
plan  will  be  useful  for  achieving  bird  conser- 
vation at  the  biome,  BCR,  state,  or  habitat  lev- 
el. The  plan  is  a starting  point  for  all  future 
bird  conservation  efforts.  Partners  in  Flight 
has  recently  released  revised  ranking  data  for 
landbirds  covered  in  this  plan  on  the  PIF  Web 
site  (www.partnersinflight.org).  The  Landbird 
Conservation  Plan  should  be  required  reading 
for  biologists  and  land  managers  as  well  as 
those  interested  in  bird  conservation. — 
MARY  GUSTAFSON,  Texas  Parks  and  Wild- 
life Department,  Mission,  Texas;  e-mail: 
Mary.Gustafson@tpwd.tx. state. us 


FLIGHT  IDENTIFICATION  OF  EURO- 
PEAN SEABIRDS.  By  Anders  Blomdahl, 
Bertil  Breife,  and  Niklas  Holmstrom.  Chris- 


topher Helm,  London,  United  Kingdom.  2003: 
374  pp.,  over  690  color  photos.  ISBN: 
0713660201.  £35.00  (paper). — Field  guides  to 
bird  identification  are  no  longer  restricted  to 
general  guides  on  the  birds  of  a particular  re- 
gion. Although  this  guide’s  coverage  is  re- 
stricted to  the  European  region,  it  covers  the 
specialized  topic  of  flight  identification  of  sea- 
birds, a group  defined  here  as  including  loons, 
grebes,  tubenoses,  cormorants,  waterfowl, 
skuas,  jaegers,  gulls,  terns,  alcids,  etc.  The  au- 
thors state  that  they  were  inspired  by  their 
study  of  large  numbers  of  migrating  seabirds 
along  the  Baltic  coast  of  Sweden,  but  much 
of  the  information  pertains  to  almost  any  non- 
tropical  coast  along  the  North  Atlantic. 

The  guide  opens  with  a solid  Basics  of 
Field  Identification  section.  It  is  a good  over- 
view of  the  challenges  inherent  to  watching 
fast-flying  birds  in  oftentimes  difficult  condi- 
tions, and  contains  many  cautions  for  the  less 
experienced  birder.  The  book  stresses  the 
shape,  size,  and  flight  style  of  birds  in  flight. 
The  discussion  of  weather,  wind,  and  light  is 
helpful  for  those  not  used  to  scanning  vast 
stretches  of  ocean.  Although  the  next  section 
listing  87  seabird  watching  sites  in  Western 
Europe  is  not  very  useful  on  the  U.S.  side  of 
the  “pond,”  it  is  a good  guide  for  traveling 
North  American  birders. 

Species  are  organized  by  functional  groups: 
some  by  family,  such  as  those  in  the  section 
entitled  “ Divers  Gaviidae others  more  in- 
formally, such  as  those  in  the  section  entitled 
“ Diving  Ducks  and  Sawbills .”  An  overview 
of  identification  points  is  provided  in  each 
section,  including  marks  that  separate  species 
from  other  groups  or  from  other  species  with- 
in groups,  and  marks  related  to  age  and  molt. 
A blue  box  on  the  overview  page  contains  a 
bulleted  list  of  field  marks  to  note  when  at- 
tempting to  separate  species  within  the  group. 
It  stands  out  well  for  easy  reference  in  the 
field  as  that  fast-flying  seabird  goes  whizzing 
past. 

The  individual  species  accounts  are  unique 
among  field  guides  in  that  they  stress  identi- 
fication in  flight.  A short  opening  paragraph 
describes  the  species’  range  and  includes  oth- 
er commentary.  The  accounts  contain  the 
more-expected  information  under  the  head- 
ings Size  and  Plumage  and  Bare  Parts.  Size 
information  is  often  presented  with  a compar- 


ORNITHOLOGICAL  LITERATURE 


125 


ison  to  other  species  covered  by  the  guide. 
The  accounts  also  contain  the  headings  Sil- 
houette and  Flight  and  Flocking.  These  key 
features  make  this  guide  particularly  suited  for 
seabird  watching.  Again,  comparatives  are 
used  liberally  throughout  these  sections. 

A nice  touch  is  that  the  authors  apparently 
were  not  enslaved  by  format.  A Note,  Voice, 
Subspecies,  and/or  Geographical  Variation 
section  appears  at  the  end  of  each  species  ac- 
count, as  warranted.  For  example,  it  would  not 
have  been  very  useful  to  include  a description 
of  Fea’s  Petrel  ( Pterodroma  feae ) vocaliza- 
tions, but  it  is  very  appropriate  that  one  is  in- 
cluded for  Canada  Goose  ( Branta  canaden- 
sis). Notes  include  information  such  as  addi- 
tional identification  points,  the  possibility  of 
hybrids,  the  possibility  of  escapees,  and  com- 
parisons with  other  species  that,  while  very 
rare  to  the  region  and  not  covered  in  the  book, 
are  still  possible. 

Multiple  photographs,  all  of  birds  in  flight, 
of  course,  accompany  almost  every  species 
account.  For  those  who  have  become  used  to 
the  stellar  bird  photos  that  have  cropped  up 
everywhere  these  days,  some  of  the  photos 
might  seem  to  be  of  substandard  quality. 
Many  are  quite  good,  but  even  the  more-dis- 
tant photos  do  an  excellent  job  of  illustrating 
how  the  birds  actually  appear  when  seabird 
watching.  Photos  also  include  images  of  birds 
in  various  plumages. 

This  book  will  be  particularly  useful  as 
more  birders  become  aware  of  the  massive 
bird  migrations  that  can  be  witnessed  in  many 
places  along  the  Atlantic  coastline.  Its  empha- 
sis on  flight  identification  complements  the 
more  standard  field  guides  available.  Use  of 
this  guide  will  speed  birders’  abilities  and 
confidence  as  they  spend  time  in  the  field 
watching  seabirds. 

Because  this  book  was  written  by  Europe- 
ans for  the  purpose  of  identifying  European 
seabirds,  North  American  birders  should  be 
aware  that  some  of  the  book’s  approaches  may 
be  a bit  confusing,  or  less  helpful,  to  them. 
For  example,  the  common  names  used  in  Eu- 
rope do  not  always  match  the  names  used  in 
North  America  (e.g.,  Slavonian  Grebe  [ Podi - 
ceps  auritus]  rather  than  Homed  Grebe,  Arctic 
Skua  [ Stercorarius  parasiticus ] rather  than 
Parasitic  Jaeger).  In  addition,  comparisons  are 
often  made  to  European  species.  For  example, 


“Red-necked  Grebe  lacks  the  abnormally 
elongated  appearance  of  Great  Crested  Grebe 
and  is  a more  compact  and  chubbier  bird,” 
but  many  North  American  birders  are  not  fa- 
miliar with  Great  Crested  Grebe  ( Podiceps 
cristatus).  Finally,  some  species  that  are  fairly 
regular  on  the  U.S.  side  of  the  Atlantic  are 
treated  with  minor  descriptions  and  no  photos 
(e.g.,  Canvasback  [Aythya  valisineria ] and 
Redhead  [A.  americana ])  or  descriptions  are 
missing  altogether  (e.g.,  Black  Skimmer 
[Rhynchops  niger]).  Overall,  however,  this 
book  is  worthwhile  to  those  who  spend  time, 
or  would  like  to  spend  time,  watching  the 
spectacle  of  seabird  migration  along  the  At- 
lantic coast. — PAUL  A.  GURIS,  See  Life 
Paulagics,  Green  Lane,  Pennsylvania;  e-mail: 
info@paulagics.com 


THE  SINGING  LIFE  OF  BIRDS:  THE 
ART  AND  SCIENCE  OF  LISTENING  TO 
BIRDSONG.  By  Donald  E.  Kroodsma,  illus- 
trated by  Nancy  Haver.  Houghton  Mifflin 
Company,  Boston,  Massachusetts  and  New 
York,  New  York.  2005:  482  pp.,  68  figures, 
CD  of  recordings.  ISBN:  0618405682,  $28 
(cloth). — “Somewhere,  always,  the  sun  is 
shining,  and  somewhere,  always,  the  birds  are 
singing.”  So  begins  Don  Kroodma’s  celebra- 
tion of  birdsong,  The  Singing  Life  of  Birds. 
On  every  page,  Kroodsma  reveals  his  passion 
for  birds,  his  infatuation  with  birdsong,  and 
his  desire  to  unravel  the  mysteries  of  avian 
singing  behavior.  More  than  a celebration,  the 
book  is  Kroodsma’s  attempt  to  answer  the 
“why”  questions  of  birdsong.  Why  do  some 
species  learn  their  songs?  Why  are  the  songs 
of  other  species  innate?  Why  do  some  species 
have  dialects,  where  birds  match  the  songs  of 
their  neighbors?  Why  would  other  species  be 
unable  to  learn  neighboring  songs?  Why  do 
mockingbirds  mimic?  Why  do  females  of 
some  species  sing?  Kroodsma  attempts  to  an- 
swer such  questions  with  30  different  adven- 
tures— 30  accounts  of  birds  singing  their  sto- 
ries— and  shares  three  decades  of  recording 
and  analyzing  songs.  Traveling  widely  across 
the  Americas — from  the  eastern  to  the  western 
U.S.  and  from  Saskatchewan  to  Central  and 
South  America — often  enlisting  the  aid  of 
countess  colleagues  and  students,  Kroodsma 


126 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  1,  March  2006 


takes  us  along  on  his  exploits  as  he  recounts 
his  recording  experiences. 

The  common  thread  running  throughout  the 
book  is  an  emphasis  on  the  combination  of 
listening  to  (songs  on  the  CD)  and  seeing 
(sonagrams)  bird  songs.  It  is  the  sight  of 
sound  that  excites  Don  Kroodsma,  and  he  in- 
fects the  reader  with  his  enthusiasm  (“.  . . I 
can’t  imagine  a world  without  sonagrams,  as 
I can’t  imagine  listening  without  also  see- 
ing”). Using  sound  spectrograms  and  the  ac- 
companying CD  of  bird  songs,  he  considers 
how  birds  acquire  their  songs,  what  makes 
their  songs  unique,  what  functions  songs 
serve,  and  “how  the  pieces  of  this  singing 
continent  fit  together.” 

Chapter  1 introduces  readers  to  the  ele- 
ments of  sonagrams — how  to  interpret  the 
time-frequency  displays  of  sonagrams;  how  to 
distinguish  noisy,  complex  sounds  from  pure- 
toned,  whistled  sounds;  how  to  recognize  the 
rhythm  and  amplitude  evident  in  sonagrams; 
and  how  to  learn  to  listen  (“How  do  I hear 
with  my  eyes?”).  Kroodsma  also  shares  his 
personal  beginnings  and  interest  in  birdsong 
in  this  chapter,  crediting  the  Bewick’s  Wren 
( Thryomanes  bewickii ) as  the  bird  that  first 
taught  him  how  to  listen.  He  ends  the  chapter 
by  outlining  the  kinds  of  questions  he  asks, 
and  attempts  to  answer,  throughout  the  book: 
How,  where,  when,  and  from  whom  do  birds 
acquire  their  singing  vocabulary?  What  are 
the  functions  of  different  bird  sounds?  How 
do  a bird’s  life  history  features  and  its  evo- 
lutionary background  influence  song?  How  do 
the  brain,  syrinx,  and  hormones  control  and 
influence  birdsong? 

As  Kroodsma  takes  readers  on  his  pre-dawn 
vigils,  he  reflects  on  the  music  of  nature  and 
the  journeys  on  which  birds  have  taken  him. 
He  bikes  across  Martha’s  Vineyard,  aston- 
ished to  hear  and  record  improbable  sweetie- 
heys  from  Black-capped  Chickadees  ( Poecile 
atricapillus ) (across  the  continent,  nearly  all 
other  chickadees  sing  hey-sweetie).  He  traips- 
es across,  canoes  through,  flies  to,  and  criss- 
crosses, visits,  and  revisits  Illinois,  South  Da- 
kota, New  York,  North  Carolina,  Michigan, 
California,  Colorado,  Saskatchewan,  Iowa, 
and  Nebraska — all  to  identify  “The  Great 
Marsh  Wren  Divide”  that  distinguishes  what 
are  almost  certainly  two  different  species  of 
Marsh  Wren  ( Cistothorus  palustris).  Kroods- 


ma spends  an  entire  early-May  night  (20:10- 
05:04),  following  one  male  Whip-poor-will 
(Caprimulgus  vociferus),  and  counts  20,898 
tuck-wHip-poor-WILLs — 2,300  songs/hr  and 
40  songs/min  in  just  under  9 hr.  And  then  he 
asks  “Why  so  much  song?”  (Because  the 
moon  was  full?  Because  the  weather  was 
warm?  Because  Whip-poor-wills  had  just  re- 
turned from  migration?  Do  high  song  rates  re- 
flect genetic  superiority  or  good  territories?). 
Relentlessly  curious,  always  intrigued, 
Kroodsma  is  continually  searching  for  an- 
swers. 

Kroodsma’s  enthusiasm  is  one  of  the  most 
notable  and  enjoyable  features  of  his  book.  I 
offer  only  a few  examples:  (1)  “Hear  the 
DNA  of  this  flycatcher  speak.  . . ”;  (2)  “I  love 
the  way  song  ‘G’  begins.  . . ”;  (3)  “There’s 
something  universal  in  the  quality  of  these 
sounds  [of  Sooty  Shearwaters,  Pujfinus  gri- 
seus],  and  it  seems  fitting  that  the  birds  them- 
selves have  the  final  comment  about  the  sheer 
wonder  and  joy  of  birdsong”;  (4)  “.  . . I can’t 
help  but.  . . admir[e]  how  the  black  images  of 
songs  against  the  white  paper  reveal  the  magic 
in  the  singing  bird”;  and  (5)  “.  . . songs  of 
some  [Fox  Sparrows,  Passerella  iliaca,  are] 
so  beautiful  that  they  can  bring  tears  to  the 
eyes.” 

Kroodsma  shares  many  of  his  discoveries 
about  birdsong  with  readers.  For  example, 
there  are  two  birdsong  vocabularies  and  two 
species  (eastern  and  western)  of  Marsh  Wrens, 
not  just  one.  The  songs  of  Eastern  Phoebes 
{Sayornis  phoebe)  and  Willow  ( Empidonax 
traillii)  and  Alder  ( E . alnorum)  flycatchers  are 
innate,  not  learned.  Sedge  Wrens  ( Cistothorus 
platensis ) improvise  (make  up  their  songs) 
and  they  do  not  imitate  (learn  songs  from) 
their  neighbors  as  other  wrens  do — because 
Sedge  Wrens  are  nomadic  due  to  the  unpre- 
dictability of  their  sedge-meadow  breeding 
habitats.  Song  Sparrows  ( Melospiza  melodia) 
that  match  and  share  songs  with  their  neigh- 
bors keep  their  territories  longer — and  may 
live  longer.  A young  Bewick’s  Wren  learns  his 
father’s  songs  early  in  life,  but  in  the  follow- 
ing years,  after  occupying  a territory  of  his 
own,  he  replaces  his  father’s  songs  by  match- 
ing those  of  neighboring  males.  Kroodsma 
also  lets  us  in  on  the  fact  that  the  meetcha 
song  “switch”  of  a male  Chestnut-sided  War- 
bler ( Dendroica  pensylvanica ) is  “off”  if  he 


ORNITHOLOGICAL  LITERATURE 


127 


has  a female,  but  it  is  “on”  if  he  is  without  a 
female  (males  sing  several  meetcha  songs, 
e.g.,  wheedle  wheedle  wheedle  wheedle  sweet 
sweet  MEETCHA). 

There  are  68  figures,  nearly  all  of  which  are 
sonagrams;  these  are  flawless  and  impeccably 
prepared  and  presented.  Some  sonagrams  are 
presented  at  an  expanded  time  scale  to  show 
greater  detail,  and  songs  of  these  sonagrams 
can  also  be  heard  on  the  CD,  but  are  played 
at  a correspondingly  slower  pace.  Figure  cap- 
tions offer  straightforward  explanations  about 
how  to  interpret  the  notes  and  “read”  the  son- 
agrams; Kroodsma  points  out  the  intricate  de- 
tails and  encourages  readers  to  follow  along 
on  the  CD — to  hear,  and  see,  birdsong  at  the 
same  time.  The  CD  (98  tracks,  —73  min)  con- 
tains superlative  recordings  of  more  than  50 
species — to  aid  readers  in  the  interpretation  of 
the  sonagrams  or  for  sheer  listening  enjoy- 
ment. 

Appendix  I {Bird  Sounds  on  the  Compact 
Disc)  provides  detailed,  colorful  descriptions 
of  the  bird  sounds  on  the  accompanying  CD. 
Appendix  II  {Techniques)  offers  useful  advice 
on  how  to  listen  to  and  record  birdsong,  on 
the  recording  equipment  needed  to  do  so,  and 
on  the  software  for  making  sonagrams.  At  the 


end  of  Appendix  II,  Kroodsma  notes  that 
“There’s  no  longer  any  mystique  to  what  I 
have  done  all  these  years.  Anyone  can  do  this 
kind  of  stuff.  And  anyone  should.”  The  Notes 
and  Bibliography  chapter  provides  a short 
section  on  recommended  readings,  an  anno- 
tated list  of  readings  for  the  key  topics  dis- 
cussed in  text,  and  a formal,  extensive  bibli- 
ography. A well-organized,  all-inclusive  in- 
dex— referencing  key  topics,  CD  tracks,  the 
locations  of  sonagrams  in  text,  and  the  most 
important  information  for  the  key  species  dis- 
cussed— completes  the  volume. 

Cautious,  meticulous,  thoroughly  prepared, 
objective,  and  determined  to  know,  Kroodsma 
takes  the  reader,  with  lively,  often  stirring 
prose,  on  30  fascinating  journeys.  No  matter 
what  your  level  of  ornithological  expertise,  af- 
ter reading  this  book  you  will  have  learned  to 
listen  to,  and  to  look  at,  birdsong  in  a different 
way,  and  you  will  have  broadened  your  un- 
derstanding of  avian  singing  behavior.  As 
Kroodsma  reminds  us  (quoting  Shakespeare), 
“The  earth  has  music  for  those  who  listen.” 
I highly  recommend  this  book. — JAMES  A. 
SEDGWICK,  USGS  Fort  Collins  Science 
Center,  Fort  Collins,  Colorado;  e-mail: 
jim_sedgwick@usgs.gov 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 


Editor  JAMES  A.  SEDGWICK 
U.S.  Geological  Survey 
Fort  Collins  Science  Center 
2150  Centre  Ave.,  Bldg.  C. 

Fort  Collins,  CO  80256-8118,  USA 
E-mail:  wjo@usgs.gov 


Editorial  Board  KATHY  G.  BEAL 
CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 
KARL  E.  MILLER 


Review  Editor 


Managing  Editor 
Copy  Editor 
Editorial  Assistant 


M.  BETH  DILLON 
CYNTHIA  P.  MELCHER 
ALISON  R.  GOFFREDI 


MARY  GUSTAFSON 
Texas  Parks  and  Wildlife  Dept. 

2800  S.  Bentsen  Palm  Dr. 

Mission,  TX  78572,  USA 
E-mail:  WilsonBookReview@aol.com 


GUIDELINES  FOR  AUTHORS 

Consult  the  detailed  “Guidelines  for  Authors”  found  on  the  Wilson  Ornithological  Society  Web  site  (http:// 
www.ummz.lsa.umich.edu/birds/wilsonbull.html). 


NOTICE  OF  CHANGE  OF  ADDRESS 

If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornithological 
Societies  of  Norht  America,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  %The  Museum  of  Zoology,  The 
Univ.  of  Michigan,  Ann  Arbor,  MI  48109.  Persons  having  business  with  any  of  the  officers  may  address  them 
at  their  various  addresses  given  on  the  inside  of  the  front  cover,  and  all  matters  pertaining  to  the  journal  should 
be  sent  directly  to  the  Editor. 


MEMBERSHIP  INQUIRIES 

Membership  inquiries  should  be  sent  to  James  L.  Ingold,  Dept,  of  Biological  Sciences,  Louisiana  State  Univ., 
Shreveport,  LA  71115;  e-mail:  jingold@pilot.lsus.edu 


THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 

The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  in  the  Univ.  of 
Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  Univ.  of  Michigan  in  1930.  Until  1947 
the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  ornithological  magazines  from 
members  and  friends  of  the  Society.  Two  members  have  generously  established  a fund  for  the  purchase  of  new 
books;  members  and  friends  are  invited  to  maintain  the  fund  by  regular  contribution.  The  fund  will  be  admin- 
istered by  the  Library  Committee.  Terry  L.  Root,  Univ.  of  Michigan,  is  Chairman  of  the  Committee.  The  Library 
currently  receives  over  200  periodicals  as  gifts  and  in  exchange  for  The  Wilson  Journal  of  Ornithology.  For 
information  on  the  Library  and  our  holdings,  see  the  Society’s  web  page  at  http://www.ummz.lsa.umich.edu/ 
birds/wos.html.  With  the  usual  exception  of  rare  books,  any  item  in  the  Library  may  be  borrowed  by  members 
of  the  Society  and  will  be  sent  prepaid  (by  the  Univ.  of  Michigan)  to  any  address  in  the  United  States,  its 
possessions,  or  Canada.  Return  postage  is  paid  by  the  borrower.  Inquiries  and  requests  by  borrowers,  as  well  as 
gifts  of  books,  pamphlets,  reprints,  and  magazines,  should  be  addressed  to:  Josselyn  Van  Tyne  Memorial  Library, 
Museum  of  Zoology,  The  Univ.  of  Michigan,  1109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079,  USA.  Contri- 
butions to  the  New  Book  Fund  should  be  sent  to  the  Treasurer. 


This  issue  of  The  Wilson  Journal  of  Ornithology  was  published  on  6 March  2006. 


130 


Continued from  outside  back  cover 


104  First  report  of  Black  Terns  breeding  on  a coastal  barrier  island 

Shawn  R.  Craik,  Rodger  D.  Titman,  Amelie  Rousseau,  and  Michael  J.  Richardson 

107  First  observation  of  cavity  nesting  by  a female  Blue  Grosbeak 
Thomas  S.  Risch  and  Thomas  J.  Robinson 

109  A new  record  of  the  endangered  White-winged  Nightjar  ( Eleothreptus  candicans ) from  Beni,  Bolivia 
Tomas  Grim  and  Radim  Sumbera 


112  Predation  of  Eared  Grebe  by  Great  Blue  Heron 
James  W.  Rivers  and  Michael  J.  Kuehn 

114  Abnormal  eggs  and  incubation  behavior  in  Northern  Bobwhite 

Fidel  Hernandez,  Juan  A.  Arredondo,  Froylan  Hernandez,  Fred  C.  Bryant,  and  Leonard  A.  Brennan 

1 17  Once  Upon  a Time  in  American  Ornithology 

120  Ornithological  Literature 


The  Wilson  Journal  of  Ornithology 

(formerly  The  Wilson  Bulletin) 


Volume  118,  Number  1 CONTENTS  March  2006 

1 Message  from  the  Editor 
Major  Articles 

3 Variation  in  mass  of  female  Prothonotary  Warblers  during  nesting 
Charles  R.  Blem  and  Leann  B.  Blem 

13  The  rediscovery  and  natural  history  of  the  White-masked  An  third  {Pithys  castaneus) 

Daniel  F.  Lane,  Thomas  Valqui  H.,  Jose  Alvarez  A.,  Jessica  Armenta,  and  Karen  Eckhardt 

23  Nesting  ecology  of  Lesser  Prairie-Chickens  in  sand  sagebrush  prairie  of  southwestern  Kansas 

James  C.  Pitman,  Christian  A.  Hagen,  Brent  E.  Jamison,  Robert  J.  Robel,  Thomas  M.  Loughin,  and  Roger 
D.  Applegate 

36  A comparative  behavioral  study  of  three  Greater  Sage-Grouse  populations 
Sonja  E.  Taylor  and  Jessica  R.  Young 

42  First  known  specimen  of  a hybrid  Buteo : Swainson’s  Hawk  ( Buteo  swainsoni)  x Rough-legged  Hawk 
(B.  lagopus)  from  Louisiana 
William  S.  Clark  and  Christopher  C.  Witt 

53  Nocturnal  hunting  by  Peregrine  Falcons  at  the  Empire  State  Building,  New  York  City 
Robert  DeCandido  and  Deborah  Allen 

59  Field  experiments  on  eggshell  removal  by  Mountain  Plovers 
Tex  A.  Sordahl 

64  Seed-size  selection  in  Mourning  Doves  and  Eurasian  Collared- Doves 
Steven  E.  Hayslette 

70  Low  nesting  success  of  Loggerhead  Shrikes  in  an  agricultural  landscape 
JeJfery  W.  Walk,  Eric  L.  Kershner,  and  Richard  E.  Warner 

75  Nest  interference  by  fledgling  Loggerhead  Shrikes 
Eric  L.  Kershner  and  Eric  C.  Mruz 

81  First  breeding  record  of  a Mountain  Plover  in  Nuevo  Leon,  Mexico 

Jose  I.  Gonzalez  Rojas,  Miguel  A.  Cruz  Nieto,  Oscar  Ballesteros  Medrano,  and  Irene  Ruvalcaba  Ortega 

85  Breeding  biology  of  the  Double-collared  Seedeater  ( Sporophila  caerulescens ) 

Mercival  R.  Francisco 

91  Small  mammal  selection  by  the  White-tailed  Hawk  in  southeastern  Brazil 
Marco  A.  Monteiro  Granzinolli  and  Jose  Carlos  Motta-Junior 

Short  Communications 

99  Provisioning  of  fledgling  conspecifics  by  males  of  the  brood-parasitic  cuckoos  Chrysococcyx  klaas  and 
C.  caprius 

Irby  J.  Lovette,  Dustin  R.  Rubenstein,  and  Wilson  Nderitu  Watetu 

101  Widespread  cannibalism  by  fledglings  in  a nesting  colony  of  Black-crowned  Night-Herons 
Christina  Riehl 


Continued  on  inside  back  cover 


Wilson  Journal 

ofO  rnithology 


Volume  118,  Number  2,  June  2006 

\ ■ ' 


Published  by  the 
Wilson  Ornithological  Society 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  ornithologist. 


President — Doris  J.  Watt,  Dept,  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  IN  46556,  USA;  e-mail: 
dwatt@saintmary  s . edu 

First  Vice-President — James  D.  Rising,  Dept,  of  Zoology,  Univ.  of  Toronto,  Toronto,  ON  M5S  3G5, 
Canada;  e-mail:  rising@zoo.utoronto.ca 

Second  Vice-President — E.  Dale  Kennedy,  Biology  Dept.,  Albion  College,  Albion,  MI  49224,  USA; 
e-mail:  dkennedy@albion.edu 

Editor — James  A.  Sedgwick,  U.S.  Geological  Survey,  Fort  Collins  Science  Center,  2150  Centre  Ave., 
Bldg.  C,  Fort  Collins,  CO  80526,  USA;  e-mail:  wjo@usgs.gov 

Secretary — Sara  R.  Morris,  Dept,  of  Biology,  Canisius  College,  Buffalo,  NY  14208,  USA;  e-mail: 
morriss@canisius.edu 

Treasurer — Melinda  M.  Clark,  52684  Highland  Dr.,  South  Bend,  IN  46635,  USA;  e-mail:  MClark@tcservices.biz 

Elected  Council  Members — Robert  C.  Beason,  Mary  Gustafson,  and  Timothy  O’Connell  (terms  expire 
2006);  Mary  Bomberger  Brown,  Robert  L.  Curry,  and  James  R.  Hill,  III  (terms  expire  2007);  Kathy  G. 
Beal,  Daniel  Klem,  Jr.,  and  Douglas  W.  White  (terms  expire  2008). 

Membership  dues  per  calendar  year  are:  Active,  $21.00;  Student,  $15.00;  Family,  $25.00;  Sustaining, 
$30.00;  Life  memberships,  $500  (payable  in  four  installments). 

The  Wilson  Journal  of  Ornithology  is  sent  to  all  members  not  in  arrears  for  dues. 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 
(formerly  The  Wilson  Bulletin ) 

THE  WILSON  JOURNAL  OF  ORNITHOLOGY  (ISSN  1559-4491)  is  published  quarterly  in  March,  June, 
September,  and  December  by  the  Wilson  Ornithological  Society,  810  East  10th  St.,  Lawrence,  KS  66044-8897.  The 
subscription  price,  both  in  the  United  States  and  elsewhere,  is  $40.00  per  year.  Periodicals  postage  paid  at  Lawrence,  KS. 
POSTMASTER:  Send  address  changes  to  OSNA,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

All  articles  and  communications  for  publications  should  be  addressed  to  the  Editor.  Exchanges  should  be  addressed 
to  The  Josselyn  Van  Tyne  Memorial  Library,  Museum  of  Zoology,  Ann  Arbor,  Michigan  48109. 

Subscriptions,  changes  of  address,  and  claims  for  undelivered  copies  should  be  sent  to  OSNA,  5400  Bosque  Blvd., 
Ste.  680,  Waco,  TX  76710.  Phone:  (254)  399-9636;  e-mail:  business@osnabirds.org.  Back  issues  or  single  copies  are 
available  for  $12.00  each.  Most  back  issues  of  the  journal  are  available  and  may  be  ordered  from  OSNA.  Special  prices 
will  be  quoted  for  quantity  orders.  All  issues  of  the  journal  published  before  2000  are  accessible  on  a free  Web  site  at  the 
Univ.  of  New  Mexico  library  (http://elibrary.unm.edu/sora/).  The  site  is  fully  searchable,  and  full-text  reproductions  of  all 
papers  (including  illustrations)  are  available  as  either  PDF  or  DjVu  files. 

© Copyright  2006  by  the  Wilson  Ornithological  Society 
Printed  by  Allen  Press,  Inc.,  Lawrence,  Kansas  66044,  U.S. A. 


COVER:  Wilson’s  Storm-Petrel  ( Oceanites  oceanicus).  Illustration  by  Don  Radovich. 


® This  paper  meets  the  requirements  of  ANSI/NISO  Z39.48-1992  (Permanence  of  Paper). 


FRONTISPIECE.  Bachman’s  Sparrows  ( Aimophila  aestivalis)  occupy  fire-dependent,  longleaf  pine  ( Pinus  pa- 
lustris)  ecosystems  of  the  southeastern  United  States.  Tucker  et  al.  (p.  131)  found  that  both  densities  and 
reproductive  indices  were  greater  during  the  first  3 years  after  burning  than  in  older  burns;  they  recommend  a 
2-3  year  burn  regime  to  maintain  healthy  populations.  Similarly,  Stober  and  Krementz  (p.  138)  report  that  home- 
range  size  increases  with  habitat  succession:  home  ranges  in  mature  habitats  often  were  twice  the  size  of  those 
in  regeneration  habitats.  Original  painting  (watercolor)  by  Don  Radovich. 


rft>e  Wilson  Journal 

of  Ornithology 


Published  by  the  Wilson  Ornithological  Society 


VOL.  118,  NO.  2 June  2006 PAGES  131-280 

The  Wilson  Journal  of  Ornithology  118(2):  131-137,  2006 

BREEDING  PRODUCTIVITY  OF  BACHMAN’S  SPARROWS  IN 
FIRE-MANAGED  LONGLEAF  PINE  FORESTS 

JAMES  W.  TUCKER,  JR.,1-3 4’5  W.  DOUGLAS  ROBINSON,14  AND  JAMES  B.  GRAND2 


ABSTRACT. — Bachman’s  Sparrows  ( Aimophila  aestivalis)  occupy  fire-dependent,  longleaf  pine  ( Pinus palustris) 
ecosystems  of  the  southeastern  United  States.  Their  populations  have  declined,  due,  in  part,  to  fire  suppression 
and  degradation  of  longleaf  pine  forests.  Populations  decline  when  longleaf  stands  go  more  than  3 years  without 
fire.  The  influence  of  fire  on  breeding  productivity,  however,  is  poorly  understood  because  territories  are  large 
and  it  is  difficult  to  find  the  well-hidden  nests  of  this  ground-nesting  sparrow.  In  an  earlier  study,  densities  of 
Bachman’s  Sparrows  were  similar  across  pine  stands  burned  1 to  3 years  previously,  but  declined  significantly  by 
the  4th  year  since  burning.  To  assess  whether  the  decline  in  density  might  be  associated  with  a decline  in  breeding 
success,  in  2001  we  used  a reproductive  index  to  estimate  breeding  productivity  of  70  territorial  males,  and  from 
1999  to  2001  we  monitored  28  nests.  We  examined  the  influence  of  (1)  season  (growing  versus  dormant)  when 
last  burned  and  (2)  years  since  burning  on  breeding  productivity  of  Bachman’s  Sparrows  in  longleaf  pine  stands 
in  the  Conecuh  National  Forest,  Alabama.  Reproductive  indices  were  greater  (Z  = 1.99,  P = 0.047)  during  the 
first  3 years  after  burning  (mean  = 3-8,  SE  = 0.4,  n = 10)  than  they  were  4 years  after  burning  (mean  = 2.0,  SE 
= 0.5,  n = 3),  similar  to  the  pattern  of  change  in  Bachman’s  Sparrow  density.  We  found  no  effect  of  burn  season 
on  the  breeding  productivity  index  (Z  = 0.075,  P — 0.94).  The  parallel  patterns  of  declining  density  and  lower 
breeding  success  suggest  that  Bachman’s  Sparrow  density  may  be  positively  correlated  with  habitat  quality.  We 
conclude  that  burning  longleaf  pine  forests  on  a 2-3  year  rotation  will  best  maintain  populations  of  Bachman’s 
Sparrows.  Received  8 February  2005,  accepted  25  November  2005. 


Bachman’s  Sparrow  (. Aimophila  aestivalis ) is 
one  of  the  bird  species  most  characteristic  of 


1 Dept,  of  Biological  Sciences,  331  Funchess  Hall, 
Auburn  Univ.,  AL  36849,  USA. 

2 USGS,  Alabama  Coop.  Fish  and  Wildlife  Research 
Unit,  School  of  Forestry  and  Wildlife  Sciences,  Au- 
burn Univ.,  AL  36849,  USA. 

3 Current  address:  Archbold  Biological  Station,  475 
Easy  St.,  Avon  Park,  FL  33825,  USA. 

4 Current  address:  Dept,  of  Fisheries  and  Wildlife, 
104  Nash  Hall,  and  Oak  Creek  Lab.  of  Biology, 
Oregon  State  Univ.,  Corvallis,  OR  97331,  USA. 

5 Corresponding  author;  e-mail: 
jtucker@archbold-station.org 


longleaf  pine  ( Pinus  palustris)  forests  and  it 
ranks  high  among  species  of  management  con- 
cern in  the  southeastern  United  States  (Hunter 
et  al.  1994).  It  is  classified  as  threatened  or  en- 
dangered in  several  states  (Dunning  1993)  and 
in  2002  it  was  red-listed  (i.e.,  one  of  most  at- 
risk  species)  by  the  National  Audubon  Society 
on  its  WatchList  (see  http://audubon2.org/ 
webapp/watchlist/viewSpecies.jsp?id  = 18). 
Loss  and  degradation  of  habitat  are  the  most 
probable  causes  for  the  species’  population  de- 
cline (Haggerty  1988).  Prescribed  fire  has  been 
identified  as  a key  tool  for  managing  Bach- 
man’s Sparrow  habitat  (Plentovich  et  al.  1998, 


131 


132 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Tucker  et  al.  1998).  Until  recently,  however, 
prescribed  fire  has  been  used  mainly  during 
the  winter  (dormant  season)  to  minimize  its 
negative  effects  on  sparrow  reproductive  suc- 
cess, despite  evidence  that  historically,  natural 
fires  occurred  most  often  during  late  spring 
and  summer  (growing  season;  Robbins  and 
Myers  1992).  Growing-season  fires  are  most 
beneficial  to  native  plant  communities  (e.g., 
Platt  et  al.  1988,  Waldrop  et  al.  1992,  Streng  et 
al.  1993),  but  the  way  in  which  fire  timing  in- 
fluences sparrow  breeding  success  is  un- 
known. 

Similarly,  evidence  from  botanical  studies  in- 
dicates that  frequent  fires  are  needed  to  main- 
tain dense,  herbaceous  ground  cover  preferred 
by  Bachman’s  Sparrows  (e.g.,  Platt  et  al.  1988, 
Dunning  and  Watts  1990,  Waldrop  et  al.  1992, 
Streng  et  al.  1993,  Plentovich  et  al.  1998,  Tuck- 
er et  al.  1998).  Engstrom  et  al.  (1984)  followed 
changes  in  bird  species  composition  through 
15  years  of  fire  exclusion  in  a stand  of  “oldfield 
pines”  (mostly  loblolly,  P.  taeda\  and  shortleaf, 
P.  echinata,  pines)  in  northwestern  Florida  that 
had  previously  been  burned  annually  during 
the  dormant  season;  Bachman’s  Sparrows  dis- 
appeared from  the  stand  after  5 years  of  fire 
exclusion.  In  studies  on  Florida  dry  prairies, 
Bachman’s  Sparrow  densities  increased  on 
sites  burned  in  mid-June  relative  to  those  on 
control  sites  (>2.5  years  since  burning;  Shriver 
et  al.  1999),  but  there  were  no  differences  in 
density  or  reproductive  success  during  the  first 
three  breeding  seasons  following  winter  fires 
(Shriver  and  Vickery  2001).  Yet,  no  data  are 
available  to  evaluate  directly  the  influence  of 
time  since  burning  and  season  of  burning  on 
breeding  productivity  of  this  elusive  sparrow 
species. 

In  a previous  study,  we  examined  the  influ- 
ence of  burn  season  and  fire  frequency  on  the 
density  of  Bachman’s  Sparrows  in  longleaf 
pine  forests  in  southern  Alabama  and  north- 
western Florida  (Tucker  et  al.  2004)  and  found 
that  density  was  unaffected  by  burn  season. 
Furthermore,  density  was  similar  within  the 
first  3 years  after  burning,  but  declined  precip- 
itously in  stands  4 or  more  years  after  a fire 
(Tucker  et  al.  2004).  We  hypothesized  that  re- 
duced breeding  success  in  stands  unburned  for 
4 or  more  years  might  explain  this  decline  in 
density.  To  test  this  hypothesis,  we  compared 
the  breeding  productivity  of  Bachman’s  Spar- 


rows across  burned  units  of  longleaf  pine  hab- 
itat that  differed  in  time  since  burning.  We  also 
evaluated  the  potential  influence  of  fire  timing 
within  the  growing  season  on  nesting  success 
by  comparing  daily  survival  rates  between 
nests  initiated  early  and  late  in  the  growing 
season. 

METHODS 

We  estimated  breeding  success  by  monitor- 
ing nests  and  using  a reproductive  index  based 
on  behavioral  observations  (Vickery  et  al. 
1992b).  The  reproductive  ecology  of  Bach- 
man’s Sparrows  is  poorly  known  because  nests 
are  hidden  on  the  ground,  usually  below  tufts 
of  overhanging  grasses,  and  are  therefore  ex- 
ceptionally challenging  to  locate  (Weston  1968, 
Harrison  1975,  Haggerty  1986).  In  response  to 
the  difficulties  of  finding  ground-nesting  spar- 
row nests,  Vickery  et  al.  (1992b)  developed  a 
reproductive  index  based  on  readily  observ- 
able behaviors  that  reduces  the  necessity  of  lo- 
cating nests  to  measure  breeding  success 
(Vickery  et  al.  1992a,  Dale  et  al.  1997).  During 
the  breeding  season  of  2001,  we  monitored  the 
territories  of  70  male  Bachman’s  Sparrows  in 
longleaf  pine  stands  of  the  Conecuh  National 
Forest,  Alabama.  To  complement  this  intensive 
study  of  focal  individuals,  we  monitored  nests 
found  in  the  same  habitat  units  from  1999 
through  2001. 

Between  22  April  and  12  May  2001,  we  lo- 
cated territories  within  13  habitat  compart- 
ments (a  group  of  adjacent  stands  managed  as 
a prescribed  burn  unit),  which  varied  from  387 
to  700  ha  in  size  and  comprised  four  treatment 
combinations  of  burn  season  (dormant  [1  Oc- 
tober-31 March]  or  growing  [1  April-30  Sep- 
tember]) and  time  since  burning  (1—3  years  or 
4 years).  We  sampled  two  compartments  for 
each  treatment  but  one:  there  was  only  one 
compartment  that  had  been  burned  during  the 
growing  season  4 years  earlier  (i.e. , 1997).  No 
stands  were  burned  during  the  2000  growing 
season,  so  territories  within  stands  the  1st  year 
after  growing-season  burning  could  not  be  in- 
cluded. Because  the  number  of  compartments 
was  small  and  the  study  design  was  unbal- 
anced, we  grouped  compartments  burned  ^3 
years  earlier  to  test  our  hypotheses  that  repro- 
ductive success  would  parallel  trends  in  den- 
sity (Tucker  et  al.  2004)  and  be  greater  during 


Tucker  et  al.  • FIRE  AND  BACHMAN’S  SPARROWS 


133 


the  first  3 years  (n  = 10)  than  4 years  (n  = 3) 
post-burning. 

Female  and  juvenile  Bachman’s  Sparrows 
are  very  secretive  and  difficult  to  observe,  so 
we  concentrated  our  efforts  on  searching  in- 
dividual territories,  rather  than  mapping  terri- 
tories within  habitat  compartments,  to  increase 
our  chances  of  observing  evidence  of  repro- 
duction. Furthermore,  Bachman’s  Sparrow  ter- 
ritories are  relatively  large  (see  Dunning  1993) 
and  densities  are  relatively  low,  especially  in 
stands  not  burned  for  >4  years  (Tucker  et  al. 
2004);  thus,  monitoring  individual  territories 
also  allowed  us  to  sample  a sufficient  number 
of  territories  to  characterize  breeding  produc- 
tivity within  each  burn  treatment  (i.e.,  each 
combination  of  burn  season  and  years  since 
burning).  Within  each  compartment,  we  se- 
lected territories  for  monitoring  by  visiting  an 
area  known  to  contain  several  Bachman’s 
Sparrows  and  selecting  the  first  four  or  six 
singing  males  encountered  within  each  com- 
partment. Although  unmated  males  of  many 
species  sing  more  frequently  than  mated  males 
(Best  1981),  the  territories  that  we  monitored 
within  habitat  compartments  were  adjacent  to 
each  other  (although  often  separated  by  ^100 
m)  and  we  did  not  observe  evidence  (e.g.,  ap- 
pearance of  additional  territories)  that  would 
suggest  that  we  overlooked  mated  birds  during 
selection  of  the  territories.  We  monitored  10 
territories  within  each  burn  treatment,  but  we 
divided  territories  unequally  between  the  two 
habitat  compartments  within  treatments  to  al- 
low a team  of  two  observers  traveling  together 
to  efficiently  monitor  two  habitat  compart- 
ments (i.e.,  5 territories  per  observer)  each  day. 
We  marked  singing  perches  for  each  male  with 
plastic  flagging  and  noted  the  territorial  bound- 
aries and  location  of  adjacent  territories  not  se- 
lected for  study.  We  also  used  mist  nets  to  cap- 
ture most  of  the  males  (53  of  70)  and  marked 
them  with  unique  combinations  of  colored  leg 
bands.  All  70  territories  were  monitored  once 
per  week  from  21  May  to  12  July  2001,  span- 
ning the  peak  of  breeding  activity  at  our  study 
site. 

Behavioral  evidence  of  reproductive  activity 
was  monitored  during  45-min  visits  to  each  ter- 
ritory once  per  week.  A visit  began  when  we 
arrived  on  a territory,  and  entailed  recording 
all  evidence  of  reproductive  activity — the  pri- 
mary objective  being  the  discovery  of  an  active 


nest.  In  addition,  we  marked  new  song  perch- 
es to  delineate  more  accurately  territory 
boundaries.  Each  territory  was  assigned  a cu- 
mulative score  indicating  increasing  evidence 
of  breeding  success,  slightly  modified  from  the 
method  of  Vickery  et  al.  (1992b).  The  scores 
for  evidence  of  reproductive  success  were  as- 
signed as  follows:  1 = presence  of  the  territo- 
rial male,  2 = presence  of  a mated  pair,  3 = 
evidence  of  an  active  nest,  4 = adults  carrying 
food  to  presumed  nestlings,  5 = direct  obser- 
vation or  evidence  of  fledglings,  6 = evidence 
of  an  active  nest  after  successful  fledging  of  a 
first  brood,  7 = evidence  of  successful  fledging 
for  two  broods,  8 = evidence  of  an  active  nest 
after  successful  fledging  of  two  broods,  and  9 
= evidence  of  successful  fledgling  for  three 
broods.  Bachman’s  Sparrows  are  not  known  to 
attempt  more  than  three  broods  within  one 
breeding  season  (Stober  and  Krementz  2000). 

A cumulative  reproductive  score  corre- 
sponding to  the  maximum  evidence  of  breed- 
ing success  was  assigned  to  each  of  the  70  ter- 
ritories. Because  individual  territories  within 
habitat  compartments  were  not  independent 
sampling  units,  we  calculated  median  repro- 
ductive scores  for  each  compartment  and  treat- 
ed individual  compartments  as  our  sampling 
units.  The  reproductive  scores  were  ranked 
(i.e.,  ordinal)  data,  so  we  used  a nonparametric 
normal  approximation  to  the  Mann- Whitney  U- 
test  (Zar  1984)  to  compare  median  reproduc- 
tive scores  in  compartments  burned  <3  versus 
4 years  previously  and  compartments  burned 
in  the  growing  versus  dormant  season. 

All  Bachman’s  Sparrow  nests  found  from 
1999  through  2001  were  monitored  according 
to  standard  methods  (Martin  and  Geupel  1993) 
on  a 2-3  day  schedule  until  the  nests  failed  or 
the  offspring  fledged.  We  calculated  daily  sur- 
vival rates  (DSR)  of  nests  (Mayfield  1961,  1975) 
to  evaluate  the  influence  of  burn  season,  years 
since  burning  (<3  versus  4),  and  timing  of  fire 
within  the  growing  season  on  nesting  success. 
For  Mayfield  calculations,  we  used  our  obser- 
vations for  length  of  the  incubation  period  (13 
days)  and  nestling  period  (9  days),  which  were 
similar  to  those  reported  by  Haggerty  (1986, 
1988).  Overall  nest  success  (i.e.,  the  probability 
of  a completed  clutch  producing  ^1  fledgling) 
was  calculated  by  raising  daily  nest  survival  to 
the  22nd  power.  We  calculated  variances  in 
DSR  and  evaluated  effects  by  examining  95% 


134 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  1 18,  No.  2,  June  2006 


5 


<3  4 Growing  Dormant 

Years  since  burned  Season  last  burned 

LIG.  1 . Mean  (±  SE)  reproductive  scores  of  Bach- 
man’s Sparrows  calculated  using  median  scores  from 
individual  habitat  compartments  at  the  Conecuh  Na- 
tional Lorest,  Alabama,  during  2001  were  greater  in 
the  first  3 years  after  burning  (n  = 10)  than  4 years 
after  burning  (n  = 3)  but  did  not  differ  between  sea- 
sons when  last  burned  (n  = 5 and  n = 8 for  growing 
and  dormant  seasons,  respectively).  Reproductive 
scores  were  collected  using  methods  modified  from 
Vickery  et  al.  (1992b). 


confidence  intervals  (±2  SE)  around  the  DSR 
(Johnson  1979). 

RESULTS 

Breeding  productivity. — Of  the  70  Bach- 
man’s Sparrow  territories  monitored,  we  found 
evidence  of  successful  reproduction  (i.e., 
fledglings  observed)  within  30  and  evidence  of 
two  successful  broods  within  4 territories. 
Overall,  28%  (14/50)  of  territorial  males  in 


compartments  burned  ^3  years  earlier  re- 
mained unpaired,  and  50%  (10/20)  of  territorial 
males  in  compartments  burned  4 years  earlier 
remained  unpaired  (x2  = 3.07,  P = 0.080).  Fur- 
thermore, 52%  (26/50)  of  territories  in  com- 
partments burned  ^3  years  earlier  successfully 
produced  young,  but  only  20%  (4/20)  of  ter- 
ritories burned  4 years  earlier  successfully  pro- 
duced young  (x2  = 5.97,  P — 0.015).  Repro- 
ductive scores  of  Bachman’s  Sparrows  were 
greater  (Z  = 1.99,  P = 0.047)  in  the  first  3 years 
after  burning  (mean  = 3.8,  SE  = 0.4,  n = 10) 
than  4 years  after  burning  (mean  = 2.0,  SE  = 
0.5,  n = 3)  but  did  not  differ  (Z  = 0.075,  P = 
0.94)  between  stands  burned  in  the  growing 
season  (mean  = 3.3,  SE  = 0.7,  n — 5)  versus 
those  burned  in  the  dormant  season  (mean  = 
3.4,  SE  = 0.5,  n = 8;  Fig.  1). 

Nesting  success. — We  found  34  nests  during 
the  study:  2,  12,  and  20  in  1999,  2000,  and 
2001,  respectively.  Two  nests  were  found  the 
day  of  fledging,  two  were  destroyed  during 
construction,  and  two  were  burned  during  egg 
laying,  leaving  28  nests  for  calculating  DSR. 
Overall,  13  of  the  28  (46%)  nests  fledged 
young.  All  nest  failures  resulted  from  depre- 
dation; no  parasitism  by  Brown-headed  Cow- 
birds  ( Molotbrus  ater)  was  observed. 

DSR  of  early-season  nests  (found  in  April 
and  May)  were  slightly  greater  than  those  of 
late-season  nests  (found  June  and  July),  al- 
though the  95%  confidence  intervals  over- 
lapped (Table  1).  In  addition,  DSR  of  nests  dur- 


TABLE  1.  Exposure  days  (number  of  nests),  number  of  nest  failures,  daily  survival  rates  (DSR),  and  95% 
confidence  intervals  (95%  Cl)  by  nesting  stage  and  time  within  the  breeding  season  (nest  cycle)  for  28  Bachman’s 
Sparrow  nests  in  the  Conecuh  National  Forest,  Alabama,  from  1999  through  2001. 


95%  CIa 


Stage 

Nest  cycleh 

Exposure  days 

Failures 

DSR 

Lower 

Upper 

IncubatioiT 

Early 

66.0  (8) 

2 

0.970 

0.928 

1.012 

Late 

52.5  (8) 

1 

0.981 

0.943 

1.019 

Total 

1 18.5  (16) 

3 

0.975 

0.946 

1.004 

Nestlingd 

Early 

64.5  (13) 

5 

0.923 

0.856 

0.989 

Late 

46.5  (12) 

7 

0.850 

0.745 

0.954 

Total 

1 11.0  (25) 

12 

0.892 

0.833 

0.951 

Combined1' 

Early 

130.5  (15) 

7 

0.946 

0.907 

0.986 

Late 

99.0  (13) 

8 

0.919 

0.864 

0.974 

Total 

229.5  (28) 

15 

0.935 

0.902 

0.967 

a Calculated  as  mean  ± 2 SE  (Johnson  1979). 

b Early  nest  cycle  included  nests  found  in  April  and  May;  late  nest  cycle  included  nests  found  in  June  and  July. 
c Incubation  stage  included  a 13-day  period  from  laying  of  the  penultimate  egg  until  the  first  egg  hatched. 
d Nestling  stage  included  a 9-day  period  from  1st  day  of  hatching  until  fledging. 
e Includes  the  sum  of  incubation  and  nestling  periods. 


Tucker  et  al.  • FIRE  AND  BACHMAN’S  SPARROWS 


135 


ing  the  incubation  period  tended  to  be  greater 
than  during  the  nestling  period,  but  again  the 
95%  confidence  intervals  overlapped  (Table  1). 
DSR  of  all  nests  from  the  beginning  of  incu- 
bation through  fledging  was  0.935  (Table  1), 
and  the  probability  of  a completed  clutch  pro- 
ducing at  least  one  fledgling  was  0.226.  DSR 
was  similar  between  the  first  3 years  (DSR  = 
0.94,  95%  Cl  = 0.90-0.97,  n = 22  nests)  and 
the  4th  year  (DSR  = 0.93,  95%  Cl  = 0.86-1.00, 
n — 6 nests)  after  burning  and  between  sites 
burned  in  the  growing  (DSR  = 0.89,  95%  Cl  = 
0.81-0.97,  n = 7 nests)  and  dormant  (DSR  = 
0.95,  95%  Cl  = 0.92-0.99,  n = 21  nests)  sea- 
sons. 

DISCUSSION 

Nesting  success  averaged  across  all  our  hab- 
itat compartments  was  23%,  which  falls  within 
the  range  previously  reported  for  Bachman’s 
Sparrows  in  Arkansas  pine  forests  (25%;  Hag- 
gerty 1988),  South  Carolina  dear-cuts  (8-34%; 
Stober  and  Krementz  2000),  and  Florida  dry 
prairies  (7-38%;  Perkins  1999).  Neither  burn 
season  nor  time  since  burning  had  a large  ef- 
fect on  nest  survival  rates  at  our  study  sites. 
Although  our  sample  size  of  nests  was  one  of 
the  largest  yet  obtained  in  a Bachman’s  Spar- 
row study,  the  sample  was  nevertheless  rela- 
tively small,  indicating  that  only  large  effects 
could  be  detected  (Johnson  1979).  In  contrast, 
our  results  from  the  reproductive  scores  (i.e., 
70  territories;  Fig.  1)  suggested  that  breeding 
productivity  was  greater  the  first  3 years  after 
burning  than  in  older  burns.  The  latter  result  is 
consistent  with  our  hypothesis  that  reduced 
breeding  success  in  older  burns  may  help  ex- 
plain the  lower  densities  of  Bachman’s  Spar- 
rows in  those  burns  (Tucker  et  al.  2004). 

Although  logistic  constraints  prevented  us 
from  simultaneously  measuring  density  and 
breeding  productivity  of  Bachman’s  Sparrows, 
our  results  suggest  a positive  correlation  be- 
tween the  two  measures  in  our  study  area.  We 
acknowledge  that  these  results  only  are  sug- 
gestive of  a positive  association  between  den- 
sity and  breeding  productivity,  but  our  consis- 
tent results  among  the  3 years  of  our  studies — 
1999  and  2000  for  density  of  Bachman’s  Spar- 
rows (Tucker  et  al.  2004)  and  2001  for  this 
study  of  breeding  productivity — strongly  sup- 
port the  conclusion  that  a regime  of  burning 
every  2-3  years  will  best  maintain  healthy  pop- 


ulations of  Bachman’s  Sparrows  in  longleaf 
pine  forests.  Bock  and  Jones  (2004)  reviewed 
studies  that  examined  the  association  between 
avian  density  and  reproductive  success  and 
found  that  a preponderance  of  studies  in  rel- 
atively undisturbed  areas  reported  a positive 
association  between  the  two  measures;  most 
studies  that  reported  a decoupling  between  the 
two  measures  were  conducted  in  disturbed 
habitats.  Our  study  area  was  within  the  largest 
remaining  extent  of  longleaf  pine  forest  (Out- 
calt  and  Sheffield  1996),  and  habitats  were  rel- 
atively natural  and  managed  under  a paradigm 
of  ecosystem  management.  Thus,  a positive 
correlation  between  density  and  breeding  pro- 
ductivity of  Bachman’s  Sparrows  in  the  area 
would  be  expected. 

Why  do  density  and  breeding  success  de- 
cline in  older  burns?  Previous  studies  suggest 
that  percent  coverage  by  herbaceous  ground 
cover,  particularly  grass  (Dunning  and  Watts 
1990,  Plentovich  et  al.  1998,  Haggerty  2000, 
Tucker  et  al.  2004),  is  a primary  factor  influ- 
encing habitat  occupancy  by  Bachman’s  Spar- 
rows. Herbaceous  ground  cover  and,  thus, 
habitat  suitability  decreases  with  time  since 
burning  (Engstrom  et  al.  1984,  Tucker  2002). 
Haggerty  (1998)  found  that  territory  sizes  were 
inversely  correlated  with  percent  coverage  of 
herbaceous  ground  cover.  Thus,  higher  spar- 
row densities  are  facilitated  by  smaller  territo- 
ries in  high-quality  habitat. 

It  should  be  noted  that  small  territory  sizes 
could  have  an  effect  on  detectability  of  repro- 
ductive status,  as  well.  The  stealthy  behavior 
of  female  and  juvenile  Bachman’s  Sparrows 
makes  them  difficult  to  detect,  but  they  may  be 
easier  to  detect  in  smaller  territories  (i.e.,  high- 
er quality  habitat)  because  their  activities  are 
confined  to  a smaller  area.  Despite  a potential 
bias  in  detectability  resulting  from  territory 
size,  the  scores  for  reproductive  success  nev- 
ertheless would  be  positively  correlated  with 
habitat  quality. 

Future  studies  should  address  the  effects  of 
timing  of  fires  within  the  breeding  season.  We 
were  unable  to  examine  breeding  productivity 
immediately  before  and  after  growing-season 
fires.  Although  we  did  not  find  differences  in 
sparrow  reproductive  success  between  burn 
seasons,  timing  of  fire  within  the  growing  sea- 
son may  be  an  important  factor  and  needs  ad- 
ditional study.  For  example,  both  our  study 


136 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


(Table  1)  and  one  in  South  Carolina  (Stober 
and  Krementz  2000)  revealed  that  early-season 
nests  tended  to  be  more  successful  than  late- 
season  nests.  Fires  during  late  April  and  early 
May  could  destroy  a large  percentage  of  the 
nestlings  or  young  fledglings  from  the  first 
nesting  cycle  and  result  in  low  annual  recruit- 
ment from  those  nesting  attempts.  Further- 
more, we  do  not  yet  know  whether  territory 
holders  move  to  unburned  sites  and  breed 
elsewhere  or  quit  all  reproduction  efforts  for  a 
given  year  when  their  territories  are  burned 
early  in  the  growing  season  (Seaman  and  Kre- 
mentz 2000).  Alternatively,  productivity  of  food 
resources  (i.e.,  seed  production  and  inverte- 
brates) may  be  enhanced  sufficiently  by  early- 
season  fires  to  compensate  for  the  loss  of  nests 
early  in  the  season;  if  vegetation  re-grows 
quickly  enough,  it  could  provide  cover  for 
nests  that  season.  Although  Seaman  and  Kre- 
mentz (2000)  found  that  Bachman’s  Sparrows 
abandoned  stands  burned  in  the  growing  sea- 
son and  failed  to  return  within  50  days  after 
the  fires,  anecdotal  observations  (JWT  unpubl. 
data,  J.  B.  Dunning  pers.  comm.)  suggest  that 
Bachman’s  Sparrows  often  return  and/or  estab- 
lish territories  in  burned  stands  within  a few 
days  after  fire  and  remain  there  through  the 
remaining  breeding  season.  Shriver  et  al. 
(1996,  1999)  found  that  burning  Florida  dry 
prairies  during  mid-June  resulted  in  an  extend- 
ed breeding  season  for  Florida  Grasshopper 
Sparrows  ( Ammodramus  savannarum  flori- 
danus)  but  fires  in  July  did  not. 

In  conclusion,  results  of  this  study  on  breed- 
ing productivity  and  our  earlier  study  on  den- 
sity of  Bachman’s  Sparrows  (Tucker  et  al. 
2004)  suggest  that  land  managers  interested  in 
providing  habitat  for  Bachman’s  Sparrows  in 
longleaf  pine  forests  should  burn  at  least  every 
3 years,  regardless  of  burn  season.  Sites  left  un- 
burned for  >4  years  host  few  to  no  breeding 
Bachman’s  Sparrows  (Tucker  et  al.  2004)  and 
it  appears  that  breeding  productivity  is  low 
among  birds  that  do  settle  in  those  habitats. 
Thus,  low  breeding  productivity  may  be  a 
plausible  explanation  for  the  low  densities  of 
sparrows  in  pine  stands  unburned  for  more 
than  3 years.  Because  most  natural  fires  histor- 
ically occurred  during  the  growing  season 
(Robbins  and  Myers  1992),  prescribed  burning 
during  the  growing  season  probably  will  be 
most  beneficial  for  longleaf  pine  communities 


overall.  Our  study,  based  on  one  of  the  largest 
sample  sizes  of  reproductive  success  yet  ob- 
tained for  this  elusive  sparrow,  suggest  that 
burn  season  may  be  of  little  consequence  to 
the  reproductive  output  of  Bachman’s  Spar- 
rows; however,  the  effects  of  fire  timing  within 
the  growing  season  still  need  to  be  evaluated. 

ACKNOWLEDGMENTS 

This  research  was  supported  by  funds  from  the  U.S. 
Geological  Survey,  State  Partnership  Program;  a J.  L. 
Landers  Research  Award  of  the  Gopher  Tortoise  Coun- 
cil; grants  from  the  U.S.  Geological  Survey,  Biological 
Resource  Division,  Species  at  Risk  Program;  a F.  M.  Pea- 
cock Scholarship  from  the  Garden  Club  of  America  ad- 
ministered through  the  Cornell  Laboratory  of  Ornithol- 
ogy; and  the  W.  F.  Coxe  Research  Fund  of  the  Birming- 
ham Audubon  Society.  The  research  benefited  from 
partnerships  among  the  following  organizations  and 
agencies:  Department  of  Biological  Sciences,  Auburn 
University;  Solon  Dixon  Forestry  Education  Center, 
School  of  Forestry  and  Wildlife  Sciences,  Auburn  Uni- 
versity; Alabama  Cooperative  Fish  and  Wildlife  Re- 
search Unit;  USDA  Forest  Service,  Conecuh  National 
Forest;  Alabama  Department  of  Conservation  and  Nat- 
ural Resources,  Division  of  Wildlife  and  Freshwater 
Fisheries;  and  The  Longleaf  Alliance.  R.  Johnson,  G. 
Morgan,  R.  Lint,  P.  Brinn,  and  R.  Mullins  provided  lo- 
gistical support  during  fieldwork.  Help  from  the  follow- 
ing field  assistants  is  gratefully  acknowledged:  C.  Kriss- 
man,  C.  Romagosa,  J.  Stratford,  R.  Peters,  A.  Sorenson, 
and  M.  Hershdorfer.  Reviews  by  J.  B.  Dunning,  Jr.,  C. 
A.  Haas,  L.  A.  Powell,  P.  C.  Stouffer,  P.  D.  Vickery,  and 
an  anonymous  reviewer  greatly  improved  this  manu- 
script. 

LITERATURE  CITED 

Best,  L.  B.  1981.  Seasonal  changes  in  detection  of  in- 
dividual bird  species.  Studies  in  Avian  Biology  6: 
252-261. 

Bock,  C.  E.  and  Z.  F.  Jones.  2004.  Avian  habitat  evalu- 
ation: should  counting  birds  count?  Frontiers  in 
Ecology  and  the  Environment  2:403^410. 

Dale,  B.  C.,  P.  A.  Martin,  and  P.  S.  Taylor.  1997.  Effects 
of  hay  management  on  grassland  songbirds  in  Sas- 
katchewan. Wildlife  Society  Bulletin  25:616-626. 
Dunning,  J.  B.  1993.  Bachman’s  Sparrow  ( Aimophila 
aestivalis ).  The  Birds  of  North  America,  no.  38. 
Dunning,  J.  B.,  Jr.,  and  B.  D.  Watts.  1990.  Regional 
differences  in  habitat  occupancy  by  Bachman’s 
Sparrow.  Auk  107:463-472. 

Engstrom,  R.  T.,  R.  L.  Crawford,  and  W.  W.  Baker. 
1984.  Breeding  bird  populations  in  relation  to 
changing  forest  structure  following  fire  exclusion: 
a 15-year  study.  Wilson  Bulletin  96:437-450. 
Haggerty,  T.  M.  1986.  Reproductive  ecology  of  Bach- 
man’s Sparrow  ( Aimophila  aestivalis ) in  central  Ar- 
kansas. Ph.D.  dissertation,  University  of  Arkansas, 
Fayetteville. 


Tucker  et  al.  • FIRE  AND  BACHMAN’S  SPARROWS 


137 


Haggerty,  T.  M.  1988.  Aspects  of  the  breeding  biology 
and  productivity  of  Bachman’s  Sparrow  in  central 
Arkansas.  Wilson  Bulletin  100:247—255. 

Haggerty,  T.  M.  1998.  Vegetation  structure  of  Bach- 
man’s Sparrow  breeding  habitat  and  its  relationship 
to  home  range.  Journal  of  Field  Ornithology  69:45- 
50. 

Haggerty,  T.  M.  2000.  A geographic  study  of  the  veg- 
etation structure  of  Bachman’s  Sparrow  C Aimophila 
aestivalis)  breeding  habitat.  Journal  of  the  Alabama 
Academy  of  Science  71:120-129. 

Harrison,  H.  H.  1975.  A field  guide  to  the  birds’  nests: 
United  States  east  of  the  Mississippi  River.  Hough- 
ton Mifflin,  Boston,  Massachusetts. 

Hunter,  W.  C.,  A.  J.  Mueller,  and  C.  L.  Hardy.  1994. 
Managing  for  Red-cockaded  Woodpeckers  and 
Neotropical  migrants — is  there  a conflict?  Proceed- 
ings of  the  Annual  Conference  of  Southeastern  As- 
sociation of  Fish  and  Wildlife  Agencies  48:383-394. 

Johnson,  D.  H.  1979-  Estimating  nest  success:  the  May- 
field  method  and  an  alternative.  Auk  96:651-661. 

Martin,  T.  E.  and  G.  R.  Geupel.  1993.  Nest-monitoring 
plots:  methods  for  locating  nests  and  monitoring 
success.  Journal  of  Field  Ornithology  64:507-519. 

Mayfield,  H.  1961.  Nesting  success  calculated  from  ex- 
posure. Wilson  Bulletin  73:255-261. 

Mayfield,  H.  F.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

Outcalt,  K.  W.  and  R.  M.  Sheffield.  1996.  The  longleaf 
pine  forest:  trends  and  current  conditions.  Research 
Bulletin  SRS-9,  USDA  Forest  Service,  Southern  Re- 
search Station,  Asheville,  North  Carolina. 

Perkins,  D.  W.  1999.  Breeding  ecology  of  Florida  Grass- 
hopper and  Bachman’s  sparrows  of  central  Florida. 
M.Sc.  thesis,  University  of  Massachusetts,  Amherst. 

Platt,  W.  J.,  G.  W.  Evans,  and  M.  M.  Davis.  1988.  Effects 
of  fire  season  on  flowering  of  forbs  and  shrubs  in 
longleaf  pine  forests.  Oecologia  76:353-363. 

Plentovich,  S.,  J.  W.  Tucker,  Jr.,  N.  R.  Holler,  and  G. 
E.  Hill.  1998.  Enhancing  Bachman’s  Sparrow  hab- 
itat via  management  of  Red-cockaded  Woodpeck- 
ers. Journal  of  Wildlife  Management  62:347-354. 

Robbins,  L.  E.  and  R.  L.  Myers.  1992.  Seasonal  effects  of 
prescribed  burning  in  Florida:  a review.  Miscella- 
neous Publication,  no.  8.  Tall  Timbers  Research 
Station,  Tallahassee,  Florida. 

Seaman,  B.  D.  and  D.  G.  Krementz.  2000.  Movements 
and  survival  of  Bachman’s  Sparrows  in  response  to 
prescribed  summer  burns  in  South  Carolina.  Pro- 
ceedings of  the  Annual  Conference  of  Southeastern 
Association  of  Fish  and  Wildlife  Agencies  54:227- 
240. 

Shriver,  W.  G.  and  P.  D.  Vickery.  2001.  Response  of 


breeding  Florida  Grasshopper  and  Bachman’s  spar- 
rows to  winter  prescribed  burning.  Journal  of  Wild- 
life Management  65:470-475. 

Shriver,  W.  G.,  P.  D.  Vickery,  and  S.  A.  Hedges.  1996. 
Effects  of  summer  burns  on  Florida  Grasshopper 
Sparrows.  Florida  Field  Naturalist  24:68-73. 

Shriver,  W.  G.,  P.  D.  Vickery,  and  D.  W.  Perkins.  1999. 
The  effects  of  summer  burns  on  breeding  Florida 
Grasshopper  and  Bachman’s  sparrows.  Studies  in 
Avian  Biology  19:144-148. 

Stober,  J.  M.  and  D.  G.  Krementz.  2000.  Survival  and 
reproductive  biology  of  the  Bachman's  Sparrow. 
Proceedings  of  the  Annual  Conference  of  South- 
eastern Association  of  Fish  and  Wildlife  Agencies 
54:383-390. 

Streng,  D.  R.,  J.  S.  Glitzenstein,  and  W.  J.  Platt.  1993. 
Evaluating  effects  of  season  of  burn  in  longleaf 
pine  forests:  a critical  literature  review  and  some 
results  from  an  ongoing  long-term  study.  Proceed- 
ings of  the  Tall  Timbers  Fire  Ecology  Conference 
18:227-263. 

Tucker,  J.  W.,  Jr.  2002.  Influence  of  season  and  fre- 
quency of  fire  on  Bachman’s  and  Henslow’s  spar- 
rows in  longleaf  pine  forests  of  the  Gulf  Coastal 
Plain.  Ph.D.  dissertation,  Auburn  University,  Au- 
burn, Alabama. 

Tucker,  J.  W.,  Jr.,  G.  E.  Hill,  and  N.  R.  Holler.  1998. 
Managing  mid-rotation  pine  plantations  to  enhance 
Bachman’s  Sparrow^  habitat.  Wildlife  Society  Bul- 
letin 26:342-348. 

Tucker,  J.  W.,  Jr.,  W.  D.  Robinson,  and  J.  B.  Grand. 
2004.  Influence  of  fire  on  Bachman’s  Sparrow:  an 
endemic  North  American  songbird.  Journal  of 
Wildlife  Management  68:1114-1123. 

Vickery,  P.  D.,  M.  L.  Hunter,  Jr.,  andJ.  V.  Wells.  1992a. 
Is  density  an  indicator  of  breeding  success?  Auk 
109:706-710. 

Vickery,  P.  D.,  M.  L.  Hunter,  Jr.,  andJ.  V.  Wells.  1992b. 
Use  of  a new  reproductive  index  to  evaluate  rela- 
tionship between  habitat  quality  and  breeding  suc- 
cess. Auk  109:697-705. 

Waldrop,  T.  A.,  D.  L.  White,  and  S.  M.  Jones.  1992.  Fire 
regimes  for  pine-grassland  communities  in  the 
southeastern  United  States.  Forest  Ecology  and 
Management  47:195-210. 

Weston,  F.  M.  1968.  Aimophila  aestivalis  bachmani 
(Audubon)  Bachman’s  Sparrow.  Pages  956-970  in 
Life  histories  of  North  American  cardinals,  gros- 
beaks, buntings,  towhees,  finches,  sparrows,  and 
allies  (O.  L.  Austin,  Jr.,  Ed.).  U.S.  National  Museum 
Bulletin,  no.  237.  [Reprinted  1968,  Dover  Publica- 
tions, New  York] 

Zar,  J.  H.  1984.  Biostatistical  analysis,  2nd  ed.  Prentice 
Hall,  Englewood  Cliffs,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  1 1 8(2):  138— 144,  2006 


VARIATION  IN  BACHMAN’S  SPARROW  HOME-RANGE  SIZE  AT 
THE  SAVANNAH  RIVER  SITE,  SOUTH  CAROLINA 

JONATHAN  M.  STOBER1 35  AND  DAVID  G.  KREMENTZ24 


ABSTRACT. — Using  radiotelemetry,  we  studied  variation  in  home-range  size  of  the  Bachman’s  Sparrow 
( Aimophila  aestivalis ) at  the  Savannah  River  Site  (SRS),  South  Carolina,  during  the  1995  breeding  season.  At 
SRS,  sparrows  occurred  primarily  in  two  habitats:  mature  pine  habitats  managed  for  Red-cockaded  Woodpecker 
{Picoides  borealis)  and  pine  plantations  1 to  6 years  of  age.  The  mean  95%  minimum  convex  polygon  home- 
range  size  for  males  and  females  combined  {n  = 14)  was  2.95  ha  ± 0.57  SE,  across  all  habitats.  Mean  home- 
range  size  for  males  in  mature  pine  stands  (4.79  ha  ± 0.27,  n = 4)  was  significantly  larger  than  that  in  4-year- 
old  (3.00  ha  ± 0.31,  n = 3)  and  2-year-old  stands  (1.46  ha  ± 0.31,  n = 3).  Home-range  sizes  of  paired  males 
and  females  (n  = 4 pairs)  were  similar  within  habitat  type;  mean  distances  between  consecutive  locations  differed 
by  habitat  type  and  sex.  We  hypothesize  that  a gradient  in  food  resources  drives  home-range  dynamics.  Received 
16  December  2004,  accepted  28  November  2005. 


The  Bachman’s  Sparrow  {Aimophila  aesti- 
valis) is  a species  of  concern  due  to  its  pop- 
ulation decline  (Sauer  et  al.  2004)  and  large 
reductions  in  range  (Dunning  1993).  The  im- 
pact of  prescribed  fire  and  timber  management 
on  Bachman’s  Sparrow  abundance  (Dunning 
and  Watts  1990;  Gobris  1992;  Plentovich  et 
al.  1998;  Tucker  et  al.  1998,  2004)  and  habitat 
occupancy  (Wan  A.  Kadir  1987;  Haggerty 
1998,  2000)  have  been  well  documented.  The 
sparrow’s  secretive  nature,  however,  makes  it 
difficult  to  obtain  basic  information  on  its  re- 
production, survival,  movement,  and  home- 
range  dynamics  (Dunning  1993). 

Bachman’s  Sparrow  home-range  sizes  have 
been  estimated  using  spot  mapping  of  un- 
marked (McKitrick  1979,  Meanley  1990)  and 
color  banded  (Haggerty  1998)  males,  but  this 
approach  is  problematic  in  some  habitats  be- 
cause detecting  Bachman’s  Sparrows  is  diffi- 
cult in  dense,  early  successional  stands  (Bibby 
et  al.  1992).  Bachman’s  Sparrows  are  ex- 
tremely cryptic  in  dense  vegetation,  particu- 
larly after  3-4  years  of  vegetative  succession 
in  rapidly  growing  pine  plantations.  Males  are 


1 Warnell  School  of  Forest  Resources,  Univ.  of 
Georgia,  Athens,  GA  30602,  USA. 

2 USGS  Patuxent  Wildlife  Research  Center,  Warnell 
School  of  Forest  Resources,  Univ.  of  Georgia,  Athens, 
GA  30602,  USA. 

3 Current  address:  J.  W.  Jones  Ecological  Research 
Center,  Rte.  2,  Box  2324,  Newton,  GA  39870,  USA. 

4 Current  address:  USGS  Arkansas  Coop.  Fish  and 
Wildlife  Research  Unit,  Dept,  of  Biological  Sciences, 
Univ.  of  Arkansas,  Fayetteville.  AR  72701,  USA. 

5 Corresponding  author;  e-mail: 
jonathan.stober@jonesctr.org 


often  only  seen  while  perched  on  singing 
posts;  such  observations  do  not  accurately  re- 
flect their  entire  home  range.  Because  females 
do  not  sing,  it  is  impossible  to  consistently 
follow  or  locate  their  movements.  Using  spot 
mapping,  mean  estimates  of  home-range  size 
ranged  from  5.1  ha  ± 1.2  SD  (range  = 4-6.7, 
n = 6)  in  mature  Florida  pine  flatwoods 
(McKitrick  1979)  to  2.5  ha  ± 0.2  SE  (range 
= 0.7— 4.5,  n — 25)  in  several  Arkansas  clear- 
cuts  during  the  initial  3 years  of  succession 
(Haggerty  1998).  How  home-range  sizes  vary 
across  the  species’  range  or  habitat  types  is 
unknown  (Dunning  1993).  Because  of  wide- 
spread conservation  concern  for  Bachman’s 
Sparrows,  wildlife  managers  require  a better 
understanding  of  the  species’  natural  history. 
We  estimated  home-range  size  using  radiote- 
lemetry in  early  and  late  successional  longleaf 
pine  ( Pinus  palustris ) stands,  examined  how 
home-range  size  varied  with  habitat  type,  and 
monitored  movements  within  territories  by 
habitat  type  and  sex. 

METHODS 

During  the  1995  breeding  season,  we  stud- 
ied Bachman’s  Sparrows  at  the  Savannah  Riv- 
er Site  (SRS)  (33°  14'  N,  81°  31'  W),  an  800- 
km1 2 3 4 5  National  Environmental  Research  Park 
managed  by  the  U.S.  Department  of  Energy. 
The  SRS  is  located  in  western  South  Carolina 
along  the  Savannah  River  in  Aiken,  Barnwell, 
and  Allendale  counties  and  lies  in  the  Upper 
Coastal  Plain  physiographic  province.  At  the 
SRS,  Bachman’s  Sparrows  inhabit  understory 
grass  and  grasslands  found  in  mature  loblolly 


138 


Stober  and  Krementz  • BACHMAN’S  SPARROW  HOME-RANGE  SIZE 


139 


( Pinus  taeda)  and  longleaf  pine  stands  (40- 
98  years  old)  managed  for  Red-cockaded 
Woodpeckers  ( Picoides  borealis );  they  also 
occur  in  regenerating  pine  stands  during  the 
initial  6-10  years  after  planting  (Dunning  and 
Watts  1990,  Gaines  et  al.  1995,  Kilgo  and 
Bryan  2005).  Mature  pine  stands  were  man- 
aged with  periodic  prescribed  fires  on  a 3-  to 
5-year  rotation  during  both  the  growing  and 
dormant  seasons.  All  mature  stands  in  which 
we  monitored  sparrows  had  been  burned  1-2 
years  previously  and  were  on  a 3-year  burn 
rotation.  Both  the  mature  and  regenerating 
stands  were  characterized  by  understories 
dominated  by  Andropogon  spp.  and  Panicum 
spp.  grasses,  rather  than  native  wiregrass  (Ar- 
istida  spp.;  Stober  1996).  Regeneration  stands 
consisted  of  areas  recently  clear-cut  and  ma- 
chine planted  with  bare-root  longleaf  pines  at 
densities  of  1,400-1,700  trees/ha;  site  prepa- 
ration generally  included  a prescribed  burn 
before  planting.  Patches  of  shrubs  within  un- 
derstories of  grasses  and  forbs  occurred  in 
both  regeneration  and  mature  stands.  We  ran- 
domly selected  five  stands  from  groups  with 
similar  management  histories:  one  2-year-old 
stand  (19.2  ha),  one  4-year-old  stand  (15.0 
ha),  and  three  mature  stands  (17.6,  16.7,  and 
5.2  ha).  Selected  stands  were  >1  km  apart. 

To  capture  Bachman’s  Sparrows,  in  each 
stand  we  placed  25  12-m-long  (30-mm  mesh) 
mist  nets  in  a 5 X 5 grid  with  nets  50  m apart 
(Krementz  and  Christie  1999).  Captured  birds 
were  weighed,  sexed,  aged,  and  banded  with 
a federal  leg  band.  We  categorized  sparrows 
as  either  hatch-year  or  after-hatch-year  and 
determined  sex  by  the  presence  or  absence  of 
a brood  patch  (Pyle  et  al.  1987).  Using  the 
Rappole  and  Tipton  (1991)  thigh-harness 
method,  we  attached  radio  transmitters  to  20 
sparrows  in  five  stands:  6 sparrows  (4M:2F) 
in  2-year-old  longleaf  pine  habitat,  6 sparrows 
(4M:2F)  in  4-year-old  longleaf  habitat,  and  8 
sparrows  (6M:2F)  in  three  mature  pine  habi- 
tats. The  radio  with  harness  weighed  1.1  — 1.2 
g (Advanced  Telemetry  Systems,  Isanti,  Min- 
nesota), about  6%  of  body  mass  relative  to  all 
captured  birds  (females:  18.6  g ± 0.24  SE,  n 
= 36;  males:  18.2  g ± 0.31  SE,  n = 69;  Sto- 
ber 1996,  Krementz  and  Christie  1999).  With- 
in a few  hours  after  release,  all  radio-tagged 
sparrows  resumed  normal  activities,  and  we 


observed  no  unusual  behaviors  associated 
with  the  radio-attachment  method. 

We  located  radio-marked  sparrows  daily, 
and  made  observations  on  each  sparrow 
throughout  the  day,  from  sunrise  to  twilight, 
throughout  the  breeding  season.  We  recorded 
status  (live,  dead,  or  lost  radio),  location,  and 
any  reproductive,  foraging,  or  other  behavior. 
Occasionally,  we  monitored  individuals  twice 
a day,  with  a minimum  of  2 hr  between  ob- 
servations. Sparrows  readily  traversed  their 
home  ranges  within  this  time  period;  there- 
fore, consecutive  observations  likely  did  not 
result  in  autocorrelation  problems  (Swihart 
and  Slade  1985)  that  would  have  yielded  un- 
derestimates of  home-range  size  (Cresswell 
and  Smith  1992). 

To  provide  an  index  of  Bachman’s  Sparrow 
density,  we  also  conducted  spot  mapping  three 
times  in  each  stand  by  using  playback  tapes 
of  the  Bachman’s  Sparrow’s  primary  song  and 
counting  all  males  (Bibby  et  al.  1992,  Dun- 
ning et  al.  1995,  Stober  1996).  While  record- 
ing daily  locations  of  marked  sparrows,  we 
also  mapped  the  locations  of  unmarked  spar- 
rows within  each  stand.  Counter-singing  ex- 
changes between  unmarked  and  marked  indi- 
viduals were  recorded  as  well. 

We  marked  sparrow  locations  with  flagging, 
and  we  used  a Trimble  Pathfinder  Pro  GPS  (3- 
D mode)  unit  to  establish  benchmark  Univer- 
sal Transverse  Mercator  (UTM)  coordinates 
within  each  territory.  All  GPS  locations  were 
differentially  corrected  and  were  accurate  to 
<5  m.  Individual  locations  were  then  refer- 
enced to  an  established  UTM  location  using  a 
survey  laser.  The  survey  laser  was  used  to  cal- 
culate distance  (±0.10  m)  and  azimuth  (±0.01 
degrees)  between  locations,  which  were  then 
converted  into  UTM  coordinates.  Once  an  in- 
dividual’s locations  were  mapped,  we  used 
program  HOME  RANGE  (Ackerman  et  al. 
1990)  to  estimate  the  95%  minimum  convex 
polygon  (MCP)  for  home  range  (Mohr  1947). 
We  attempted  to  collect  35  observations  per 
bird  (Ackerman  et  al.  1990).  We  recognize 
that  the  95%  MCP  has  certain  limitations,  but 
all  other  breeding  season  home-range  sizes  for 
Bachman’s  Sparrows  described  in  the  litera- 
ture were  estimated  using  this  metric  (Dun- 
ning 1993).  Distances  moved  between  loca- 
tions were  calculated  for  each  individual,  as 
were  distances  from  each  location  to  the  arith- 


140 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  1.  Home-range  size  estimates  and  densities  of  male  Bachman’s  Sparrows  in  pine  habitats,  by  stand 
age,  during  the  1995  breeding  season  at  the  Savannah  River  Site,  South  Carolina. 


Stand  age  (years) 

Stand  size  (ha) 

No.  marked 
sparrows 

95%  MCP3  (SE) 

Range 

Male  density/ 
10  hab 

2 

19.2 

3 

1.46  (0.31) 

0.99-2.04 

2.59 

4 

15.0 

3 

3.00  (0.31) 

2.80-3.37 

4.65 

Mature 

17.6 

3 

— 

— 

3.41 

Mature 

Mature  (both  stands)0 

16.7 

1 

4 

4.79  (0.27) 

4.23-5.69 

1.79 

a Mean  95%  minimum  convex  polygon  estimates. 

b Includes  both  radio-marked  and  unmarked  singing  males;  densities  determined  using  spot  mapping  technique  within  each  stand. 
c Home-range  estimate  for  mature  stands  is  pooled  across  the  two  stands. 


metic  center  of  each  home  range,  defined  as 
the  mean  distance  from  the  estimated  central 
coordinate  to  each  observation  within  the 
home  range. 

Only  14  of  the  20  radio-marked  sparrows 
were  included  in  the  analysis  of  home-range 
size.  We  obtained  <35  locations  for  two  fe- 
males and  one  male,  and  three  males  were 
treated  as  outliers  and  excluded  from  analyses. 
The  outliers  included  ( 1 ) a bird  with  two  dis- 
tinct home  ranges  (a  combined  total  of  20.9 
ha)  in  a mature  stand,  (2)  one  whose  home 
range  (1.63  ha)  ended  up  outside  the  study 
stand  in  an  adjacent  33-year-old  stand  of 
planted  pine,  and  (3)  one  (in  the  2-year-old 
habitat)  that  behaved  like  a floater  and  used 
part  of  an  adjacent  43-year-old  stand  of  pine 
(5.46  ha).  Due  to  these  exclusions,  we  used 
only  two  (the  17.6-  and  16.7-ha  stands)  of  the 
three  mature  pine  stands  in  our  analyses. 


FIG.  1 . Home-range  size,  by  habitat  age  (2  years, 
4 years,  mature),  using  95%  minimum  convex  polygon 
estimates  for  Bachman’s  Sparrows  («  = 4 females,  10 
males)  during  the  1995  breeding  season.  Savannah 
River  Site,  South  Carolina.  Pairs  designated  by  similar 
alpha  characters  (A-D);  F = female.  M = male. 


We  used  a general  linear  model  procedure 
(PROC  GLM;  SAS  Institute,  Inc.  1987)  to 
conduct  three  pre-planned  tests  comparing 
home-range  sizes  of  males  by  habitat  type.  We 
included  only  males  because  all  marked  fe- 
males were  paired  to  marked  males.  Once  we 
determined  differences  in  home-range  sizes  by 
habitat  type  (F-test),  we  used  Tukey  tests  to 
compare  the  least-squares  means.  Due  to  in- 
sufficient sample  sizes,  we  did  not  test  for  the 
effect  of  sex  (female  n = 4)  or  conspecific 
density  (/?  = 4)  on  home-range  size.  We  also 
used  the  GLM  procedure  and  Tukey  tests  to 
compare  the  least-squares  means  for  mean 
distance  moved  and  distance  from  home-range 
arithmetic  center  by  habitat  type  and  sex.  All 
statistical  tests  were  one-tailed.  The  level  of 
statistical  significance  was  set  at  0.05  and 
means  are  reported  ± SE. 

RESULTS 

For  the  14  radio-marked  individuals  in  our 
analyses,  we  recorded  an  average  of  63  loca- 
tions (range  = 45—81)  per  individual  over  an 
average  of  50  days  of  observation  (range  = 
38—62  days).  Ten  birds  were  monitored  in  the 
2-year-old  (3M:2F)  and  the  4-year-old  (3M: 
2F)  stands,  and  four  males  were  monitored  in 
the  two  mature  stands  (3M:0F  and  1M:0F). 
The  mean  95%  minimum  convex  polygon 
home-range  size  for  males  and  females  com- 
bined ( n — 14)  across  all  habitats  was  2.95  ha 
± 0.57.  Mean  95%  MCP  home-range  size 
across  all  habitats  was  3.26  ha  ± 0.49  for 
males  ( n = 10;  Table  1)  and  2.20  ha  ± 0.48 
(n  = 4)  for  females.  For  males,  home-range 
size  increased  with  habitat  age  (F27  = 33.9,  P 
< 0.001;  Fig.  1).  Home-range  sizes  differed 
between  2-year-old  (mean  = 1.46  ha  ± 0.31, 
n = 3)  and  4-year-old  (mean  = 3.00  ha  ± 


Stober  and  Krementz  • BACHMAN’S  SPARROW  HOME-RANGE  SIZE 


141 


0.31,  n — 3)  regeneration  habitat  (r  = 3.54,  P 
= 0.009),  and  were  significantly  larger  in  ma- 
ture pine  habitats  than  in  the  2-  and  4-year- 
old  habitats  (t  = 8.18,  P < 0.001;  t = 4.40, 
P = 0.003,  respectively). 

Home  ranges  were  always  adjacent  to  a 
stand  edge.  Conspecific  density  was  highest 
(4.65  males/ 10  ha)  in  the  4-year-old  stand  (Ta- 
ble 1)  and  lower  in  the  2-year-old  stand  and 
one  mature  stand,  both  of  which  were  isolated 
with  no  suitable  adjacent  sparrow  habitat.  Of 
the  four  sparrow  pairs  in  which  both  the  male 
and  female  were  marked,  two  inhabited  the  2- 
year-old  stand  and  two  inhabited  the  4-year- 
old  stand.  In  one  pair,  the  female  had  a larger 
home-range  size  than  the  male  (pair  B;  Fig. 
1);  otherwise,  male  and  female  home  ranges 
were  roughly  similar. 

Mean  distance  moved  between  consecutive 
observations  was  83.9  m ± 12.78  ( n = 14); 
distance  moved  differed  among  habitat  types 
(F211  ~ 14.66,  P < 0.001)  and  was  marginally 
different  between  sexes  (FU2  = 3.73,  P = 
0.077).  Mean  distance  moved  in  mature  stands 
(106.6  m ± 6.4)  was  not  different  from  that 
in  the  4-year-old  stand  (88.8  m ± 5.7),  but 
differed  from  the  distance  moved  in  the  2- 
year-old  stand  (61.0  m ± 5.7).  The  mean  dis- 
tance from  the  arithmetic  center  of  an  individ- 
ual’s home  range  to  each  location  differed  by 
habitat  type  (F211  = 12.69,  P — 0.001),  but 
not  by  sex  (F,  12  = 0.78,  P = 0.40).  Mean 
distance  from  arithmetic  center  in  mature 
stands  (81.8  m ± 4.7)  was  not  different  from 
that  in  the  4-year  old  stand  (73.6  m ± 4.2,  t 
= 1.32,  P = 0.21)  but  differed  from  that  in 
the  2-year-old  stand  (51.9  m ± 4.2,  t - 4.77, 
P < 0.001).  The  longest  movement  between 
daily  observations  was  824  m by  a male,  and 
most  long-distance  movements  were  about 
200  m.  In  one  case,  a male  crossed  a riparian 
area  200  m wide  to  an  adjacent  regeneration 
stand,  remained  there  for  2 days,  and  then  re- 
turned to  the  original  stand. 

DISCUSSION 

Because  we  located  birds  through  radiote- 
lemetry rather  than  by  visual  documentation 
at  singing  posts,  our  estimates  of  home-range 
size  were  slightly  larger  and  more  precise  than 
those  reported  by  Haggerty  (1998).  Home- 
range  estimates  of  McKitrick  (1979)  and  Hag- 
gerty (1998)  were  biased  by  their  dependence 


on  visual  records  of  males  (color  banded  or 
unmarked)  perched  in  conspicuous  locations. 
Nonetheless,  home-range  sizes  of  male  spar- 
rows in  our  study  were  similar  to  those  re- 
ported by  Haggerty  (1998),  with  the  smallest 
territories  found  in  the  1-  and  2-year-old  pine 
regeneration  habitat  and  home  range  increas- 
ing with  succession  of  habitat.  Radiotelemetry 
also  allowed  us  to  obtain  the  first  estimates  of 
female  home-range  size,  which  were  similar 
to  male  home-range  size  {n  = 4 pairs).  Home- 
range  size  between  paired  birds  is  probably 
influenced  by  the  mate-guarding  behavior  that 
males  exhibit  during  the  breeding  season 
(Haggerty  1986).  Some  locations  for  the  fe- 
male whose  home  range  was  larger  than  the 
male’s  (2-year-old  stand)  were  recorded  after 
her  brood  had  fledged,  which  may  explain  the 
larger  size  of  her  home  range. 

We  recorded  few  instances  of  direct  conflict 
between  adjacent  sparrows  defending  home 
ranges.  The  persistent  use  of  primary  song  and 
counter-singing  (Meanley  1990,  Dunning 
1993)  apparently  mediated  the  need  for  direct 
conflict  in  establishing  and  maintaining  home 
ranges.  Spot  mapping  revealed  the  highest 
density  of  sparrows  in  the  4-year-old  stand. 
Similarly,  spot  mapping  conducted  by  Stober 
(1996)  in  regeneration  habitats  1-6  years  of 
age  revealed  that  Bachman’s  Sparrow  densi- 
ties were  greatest  in  3-  to  4-year-old  habitats. 
Overlap  of  sparrow  home  ranges  was  limited 
to  three  instances  and  occurred  in  grassy 
patches  in  mature  pine  stands  or  in  regenera- 
tion habitats  where  trees  and  shrubs  were  sup- 
pressed and  grasses  dominated  the  vegetation. 

Although  Haggerty  (1986)  reported  that 
sparrow  density  was  inversely  related  to 
home-range  size,  we  were  unable  to  corrobo- 
rate this.  Stober  (1996)  found  more  sparrows 
in  stands  with  suitable  adjacent  habitat  than  in 
isolated,  disjunct  stands.  Dunning  et  al.  (1995) 
also  found  that  areas  connected  by  corridors 
of  suitable  habitat  had  a greater  probability  of 
sparrow  occupancy  than  isolated  patches  of 
suitable  habitat.  Greater  conspecific  density 
may  constrict  the  size  of  home  ranges  in 
breeding  season,  but  this  hypothesis  needs  to 
be  tested  by  removing  territorial  individuals 
and  monitoring  the  behavior  of  adjacent  in- 
dividuals. Vegetation  succession  and  arthro- 
pod food  resources  also  may  play  important 
roles  in  determining  home-range  size. 


142 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  2,  June  2006 


We  found  that  home-range  size  increased 
with  habitat  succession:  home  ranges  in  ma- 
ture habitats  often  were  twice  the  size  of  those 
in  regeneration  habitats.  We  hypothesize  that 
the  distribution  of  resources  within  home 
ranges  may  explain  this  pattern.  Bachman’s 
Sparrows  are  omnivorous,  foraging  exclusive- 
ly from  the  ground  for  insects  (orthopterans, 
arachnids,  lepidopteran  larvae,  coleopterans, 
hemipterans)  and  grass  seeds,  especially  those 
of  Panicum  spp.  (Allaire  and  Fisher  1975, 
Haggerty  1992,  Dunning  1993).  Early  succes- 
sional  habitats  have  greater  arthropod  produc- 
tivity than  mature  pine  stands  in  the  Southeast 
(Menhinick  1963,  Landers  and  Mueller  1986, 
Hurst  1992).  Cross  (1956)  surveyed  a range 
of  upland  habitats  at  the  SRS  for  Orthoptera 
and  found  >40  species  in  early  successional 
habitats  compared  with  only  7 species  in  ma- 
ture loblolly  pine  stands.  In  contrast,  ground- 
level  arthropod  communities  in  mature  pine 
stands  managed  for  Red-cockaded  Woodpeck- 
ers at  the  SRS  include  an  abundance  of  spi- 
ders and  ants,  but  few  grasshoppers  (New  and 
Hanula  1998).  Stober  (1996)  found  that,  as  a 
percentage  of  total  vegetation  cover,  Panicum 
spp.  were  more  abundant  in  regeneration 
stands  (0.8— 1.3%)  than  in  mature  pine  stands 
(0. 1-0.4%);  thus,  differences  in  home-range 
size  between  habitats  may  be  a reflection  of 
greater  seed  resources  and  arthropod  produc- 
tivity in  early  regeneration  habitats  than  in 
mature  pine  habitats  managed  for  Red-cock- 
aded  Woodpeckers.  In  examining  previous 
studies  on  Bachman’s  sparrows  across  their 
range  (Wan  A.  Kadir  1987;  Dunning  and 
Watts  1990;  Gobris  1992;  Haggerty  1998, 
2000;  Plentovich  et  al.  1998;  Tucker  et  al. 
1998,  2004),  we  observed  that,  in  general, 
sparrow  densities  and  arthropod  communities 
were  reduced  with  succession  of  understory 
vegetation. 

Despite  the  differences  we  observed  in 
home-range  sizes  by  habitat  type  and  the  dif- 
ferences in  male  densities  among  stand  ages, 
Stober  and  Krementz  (2000)  detected  no  sig- 
nificant differences  in  survival  rates  between 
sexes  or  habitat  types.  Apparently,  the  larger 
home-range  sizes  of  Bachman’s  Sparrows  in 
mature  pine  stands  do  not  predispose  those 
birds  to  lower  survival  rates,  as  might  be  ex- 
pected from  longer  movements  throughout 
their  territories.  Breeding  season  survival  rates 


were  high  (0.905,  95%  Cl  = 0.794-0.992), 
with  only  2 mortalities  (raptor  and  mammal 
depredations)  out  of  20  individuals  radio- 
tagged. 

We  found  that  Bachman’s  Sparrows  did  not 
move  far  (~100  m/day)  between  consecutive 
observations,  as  was  also  found  for  radio- 
marked  Eastern  Towhees  (. Pipilo  erythroph- 
thalmus ) at  the  Savannah  River  Site  (Kre- 
mentz and  Powell  2000).  Like  towhees,  Bach- 
man’s Sparrows  moved  among  adjacent 
stands,  but  unlike  towhees,  Bachman’s  Spar- 
rows used  middle-aged  (—20-  to  35-year-old) 
stands  infrequently  (Stober  1996).  Not  sur- 
prisingly, we  found  that  daily  movements  re- 
flected home-range  sizes:  smaller  home  ranges 
among  habitat  types  were  associated  with 
shorter  daily  distances  moved. 

Management  for  Bachman’s  Sparrow  pop- 
ulations in  forested  habitats  often  involves 
prescribed  fire  and  reduced  pine  densities.  If 
small  home-range  size  is  a surrogate  for  hab- 
itat suitability,  managers  should  maintain  a 
continuous  matrix  of  herbaceous  understory 
vegetation.  Clear-cuts  should  be  managed  for 
perches  (Dunning  and  Watts  1990),  abundant 
herbaceous  vegetation  (Mills  et  al.  1991,  Dun- 
ning 1993),  and  connectivity  with  nearby  suit- 
able habitat  (Dunning  et  al.  1995).  Although 
it  is  known  that  mature  stands  of  pine  become 
more  suitable  for  sparrows  with  frequent  pre- 
scribed fire  and  moderate  basal  areas  of  pine, 
further  research  should  ascertain  whether 
home-range  size  in  mature  pine  stands  is  de- 
pendent on  the  distribution  of  herbaceous  un- 
derstory, as  arthropod  communities  in  mature 
pine  stands  are  a function  of  primary  produc- 
tivity occurring  on  the  forest  floor  (Cross 
1956).  Additional  information  on  Bachman’s 
Sparrow  reproduction  and  survival  across  the 
range  of  occupied  habitats  is  needed  to  deter- 
mine the  viability  of  populations  inhabiting 
intensively  managed  industrial  forests  versus 
forests  managed  on  longer  logging  rotations 
with  fire  management. 

ACKNOWLEDGMENTS 

This  project  was  funded  by  the  USGS  Biological 
Resources  Division  and  by  the  U.S.  Department  of  En- 
ergy— Savannah  River  Operations  Office  through  the 
U.S.  Forest  Service — Savannah  River  under  Interagen- 
cy Agreement  DE-AI09-00SR221 88  (Cooperative 
Agreement  contract  no.  12-11-008-876).  J.  C.  Kilgo, 
G.  O.  Ware,  J.  B.  Dunning,  Jr.,  J.  Blake,  and  two  anon- 


Stober  and  Krementz  • BACHMAN’S  SPARROW  HOME-RANGE  SIZE 


143 


ymous  reviewers  commented  on  the  manuscript. 

Thanks  to  J.  S.  Christie,  A.  Allen,  H.  McPherson,  C. 

E.  Moorman,  and  J.  B.  Dunning,  Jr.,  for  assistance  in 

the  field. 

LITERATURE  CITED 

Ackerman,  B.  B.,  F.  A.  Leban,  M.  D.  Samuel,  and  E. 
O.  Garton.  1990.  User’s  manual  for  program 
HOME  RANGE,  2nd  ed.  Technical  Report,  no. 
15.  Forestry,  Wildlife  and  Range  Experiment  Sta- 
tion, University  of  Idaho,  Moscow. 

Allaire,  P.  N.  and  C.  D.  Fisher.  1975.  Feeding  ecol- 
ogy of  three  resident  sympatric  sparrows  in  east- 
ern Texas.  Auk  92:260-269. 

Bibby,  C.  J.,  N.  D.  Burgess,  and  D.  A.  Hill.  1992. 
Bird  census  techniques.  Academic  Press,  London, 
United  Kingdom. 

Cresswell,  W.  J.  and  G.  C.  Smith.  1992.  The  effects 
of  temporally  autocorrelated  data  on  methods  of 
home-range  analysis.  Pages  272-284  in  Wildlife 
telemetry:  remote  monitoring  and  tracking  of  an- 
imals (I.  G.  Priede  and  S.  M.  Swift,  Eds.).  Ellis 
Horwood,  London,  United  Kingdom. 

Cross,  W.  H.  1956.  The  arthropod  component  of  old- 
field  ecosystems:  herb  stratum  population  with 
special  emphasis  on  the  Orthoptera.  Ph.D.  disser- 
tation, University  of  Georgia,  Athens. 

Dunning,  J.  B.  1993.  Bachman’s  Sparrow  ( Aimophila 
aestivalis).  The  Birds  of  North  America,  no.  38. 

Dunning,  J.  B.,  Jr.,  R.  Borgella,  Jr.,  K.  Clements, 
and  G.  K.  Meffe.  1995.  Patch  isolation,  corridor 
effects,  and  colonization  by  a resident  sparrow  in 
a managed  pine  woodland.  Conservation  Biology 
9:542-550. 

Dunning,  J.  B.,  Jr.,  and  B.  D.  Watts.  1990.  Regional 
differences  in  habitat  occupancy  by  Bachman’s 
Sparrow.  Auk  107:463-472. 

Gaines,  G.  D.,  K.  E.  Franzreb,  D.  H.  Allen,  K.  S. 
Laves,  and  W.  L.  Jarvis.  1995.  Red-cockaded 
Woodpecker  management  on  the  Savannah  River 
Site:  a management/research  success  story.  Pages 
81-88  in  Red-cockaded  Woodpecker:  recovery, 
ecology  and  management  (D.  L.  Kulavy,  R.  G. 
Hooper,  and  R.  Costa,  Eds.).  Center  for  Applied 
Studies  in  Forestry,  Stephen  F.  Austin  State  Uni- 
versity, Nacogdoches,  Texas. 

Gobris,  N.  1992.  Habitat  occupancy  during  the  breed- 
ing season  by  Bachman’s  Sparrow  at  the  Piedmont 
National  Wildlife  Refuge  in  central  Georgia. 
M.Sc.  thesis.  University  of  Georgia,  Athens. 

Haggerty,  T.  M.  1986.  Reproductive  ecology  of  Bach- 
man’s Sparrow  {Aimophila  aestivalis ) in  central 
Arkansas.  Ph.D.  dissertation.  University  of  Arkan- 
sas, Fayetteville. 

Haggerty,  T.  M.  1992.  Effects  of  nestling  age  and 
brood  size  on  nestling  care  in  the  Bachman’s  Spar- 
row {Aimophila  aestivalis).  American  Midland 
Naturalist  128:115-125. 

Haggerty,  T.  M.  1998.  Vegetation  structure  of  Bach- 
man’s Sparrow  breeding  habitat  and  its  relation- 


ship to  home  range.  Journal  of  Field  Ornithology 
69:45-50. 

Haggerty,  T.  M.  2000.  A geographic  study  of  the  veg- 
etation structure  of  Bachman’s  Sparrow  {Aimo- 
phila aestivalis)  breeding  habitat.  Journal  of  the 
Alabama  Academy  of  Science  71:120-129. 

Hurst,  G.  A.  1992.  Foods  and  feeding.  Pages  75-95 
in  The  Wild  Turkey:  biology  and  management  (J. 
G.  Dickson,  Ed.).  Stackpole  Books.  Harrisburg, 
Pennsylvania. 

Kilgo,  J.  C.  and  A.  L.  Bryan,  Jr.  2005.  Nongame 
birds.  Pages  223-253  in  Ecology  and  management 
of  a forested  landscape:  fifty  years  on  the  Savan- 
nah River  Site  (J.  C.  Kilgo  and  J.  I.  Blake,  Eds.). 
Island  Press,  Washington,  D.C. 

Krementz,  D.  G.  and  J.  S.  Christie.  1999.  Scrub-suc- 
cessional  bird  community  dynamics  in  young  and 
mature  longleaf  pine-wiregrass  savannahs.  Jour- 
nal of  Wildlife  Management  63:803-814. 

Krementz,  D.  G.  and  L.  A.  Powell.  2000.  Breeding 
season  demography  and  movements  of  Eastern 
Towhees  at  the  Savannah  River  Site,  South  Car- 
olina. Wilson  Bulletin  112:243-248. 

Landers,  J.  L.  and  B.  S.  Mueller.  1986.  Bobwhite 
quail  management:  a habitat  approach.  Miscella- 
neous Publication,  no.  6.  Tall  Timbers  Research 
Station,  Tallahassee,  Florida. 

McKitrick,  M.  C.  1979.  Territory  size  and  density  of 
Bachman’s  Sparrow  in  south  central  Florida.  Flor- 
ida Field  Naturalist  7:33-34. 

Meanley,  B.  1990.  Some  observations  of  the  singing 
behavior  of  Bachman’s  Sparrow.  Chat  54:63. 

Menhinick,  E.  F.  1963.  Density,  diversity,  and  energy 
flow  of  arthropods  in  the  herb  stratum  of  a Sericea 
lespedeza  stand.  Ph.D.  dissertation.  University  of 
Georgia,  Athens. 

Mills,  G.  S.,  J.  B.  Dunning,  Jr.,  and  J.  M.  Bates. 
1991.  The  relationship  between  breeding  bird  den- 
sity and  vegetation  volume.  Wilson  Bulletin  103: 
468-479. 

Mohr,  C.  O.  1947.  Table  of  equivalent  populations  of 
North  American  small  mammals.  American  Mid- 
land Naturalist  37:223-249. 

New,  K.  C.  and  J.  L.  Hanula.  1998.  Effect  of  time 
elapsed  after  prescribed  burning  in  longleaf  pine 
stands  on  potential  prey  of  the  Red-cockaded 
Woodpecker.  Southern  Journal  of  Applied  Forest- 
ry 22:175-183. 

Plentovich,  S.,  J.  W.  Tucker,  Jr.,  N.  R.  Holler,  and 
G.  E.  Hill.  1998.  Enhancing  Bachman’s  Sparrow 
habitat  via  management  of  Red-cockaded  Wood- 
peckers. Journal  of  Wildlife  Management  62:347- 
354. 

Pyle,  R,  S.  N.  G.  Howell,  R.  P.  Yunick,  and  D.  F. 
DeSante.  1987.  Identification  guide  to  North 
American  passerines.  Slate  Creek  Press,  Bolinas, 
California. 

Rappole,  J.  H.  and  A.  R.  Tipton.  1991.  New  harness 
design  for  attachment  of  radio  transmitters  to 
small  passerines.  Journal  of  Field  Ornithology  62: 
335-337. 


144 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  2,  June  2006 


SAS  Institute,  Inc.  1987.  SAS/STAT  guide  for  PC, 
6th  ed.  SAS  Institute,  Inc.,  Cary,  North  Carolina. 

Sauer,  J.  R.,  J.  E.  Hines,  and  J.  Fallon.  2004.  The 
North  American  Breeding  Bird  Survey:  results 
and  analysis  1966-2003,  ver.  2004.1.  USGS  Pa- 
tuxent Wildlife  Research  Center,  Laurel,  Mary- 
land. www.mbr-pwrc.usgs.gov/bbs/bbs.html  (ac- 
cessed 30  September  2004). 

Stober,  J.  M.  1996.  Territory  dynamics  and  basic  bi- 
ology of  the  Bachman’s  Sparrow  ( Aimophila  aes- 
tivalis) at  the  Savannah  River  Site,  South  Caroli- 
na. M.Sc.  thesis.  University  of  Georgia,  Athens. 

Stober,  J.  M.  and  D.  G.  Krementz.  2000.  Survival 
and  reproductive  biology  of  the  Bachman’s  Spar- 
row. Proceedings  of  the  Southeastern  Association 
of  Fish  and  Wildlife  Agencies  54:383-390. 


Swihart,  R.  K.  and  N.  A.  Slade.  1985.  Testing  for 
independence  of  observations  in  animal  move- 
ments. Ecology  66:1176-1184. 

Tucker,  J.  W.,  Jr.,  G.  E.  Hill,  and  N.  R.  Holler. 
1998.  Managing  mid-rotation  pine  plantations  to 
enhance  Bachman’s  Sparrow  habitat.  Wildlife  So- 
ciety Bulletin  26:342-348. 

Tucker,  J.  W.,  Jr.,  W.  D.  Robinson,  and  J.  B.  Grand. 
2004.  Influence  of  fire  on  Bachman’s  Sparrow,  an 
endemic  North  American  songbird.  Journal  of 
Wildlife  Management  68:1114-1123. 

Wan  A.  Kadir,  W.  R.  1987.  Vegetational  characteris- 
tics of  early  successional  sites  utilized  for  breed- 
ing by  the  Bachman’s  Sparrow  (. Aimophila  aesti- 
valis) in  eastern  Texas.  M.Sc.  thesis,  Stephen  F. 
Austin  State  University.  Nacogdoches,  Texas. 


The  Wilson  Journal  of  Ornithology  1 18(2):  145— 15 1 , 2006 


NESTING  SUCCESS  AND  BREEDING  BIOLOGY  OF  CERULEAN 

WARBLERS  IN  MICHIGAN 

CHRISTOPHER  M.  ROGERS1 


ABSTRACT. — The  Cerulean  Warbler  ( Dendroica  cerulea ) is  a Nearctic-Neotropical  migratory  bird  species 
that  has  declined  significantly  over  the  long-term.  Poor  reproductive  success  may  be  an  important  factor  con- 
tributing to  the  observed  decline,  but  reproductive  output  has  been  measured  for  very  few  breeding  populations. 
From  2003  to  2005,  I intensively  monitored  22-23  breeding  territories/year  in  each  of  two  large  forest  habitats 
in  southwestern  Michigan:  oak-  ( Quercus  spp.)  hickory  ( Carya  spp.)  (2003:  Barry  State  Game  Area)  and  black 
locust-  ( Robinia  pseudoacacia)  black  cherry  ( Prunus  serotina ) (2004-2005:  Fort  Custer  U.S.  Army  Michigan 
National  Guard  Reservation).  I also  gathered  descriptive  data  on  nonsong  vocalizations  and  age  of  territorial 
males.  I describe  four  distinctive  call  notes,  by  sex,  including  the  social  and  environmental  contexts  in  which 
they  were  used.  Using  two  independent  methods  of  aging,  there  was  a strong  preponderance  of  after-second- 
year  males  at  both  study  sites.  Only  9 {n  = 7 nests),  12  (n  = 14),  and  30  (n  = 25)  fledglings  were  produced 
during  the  2003,  2004,  and  2005  breeding  seasons,  respectively.  Nest  heights  were  the  highest  recorded  for  this 
species  (mean  = 19-20  m).  During  the  same  period,  male  reproductive  success  was  0.30,  0.32,  and  0.80  male 
fledglings/breeding  male  and  0.60,  0.63,  and  1.58  fledglings/breeding  pair.  Productivity  estimates,  not  thought 
to  be  self-sustaining,  were  even  lower  than  those  of  a well-studied  Cerulean  Warbler  population  in  southern 
Ontario.  Thus,  reproductive  output  was  low  in  two  geographic  regions — representing  three  different  forest 
types — in  the  northern  portions  of  the  Cerulean  Warbler’s  breeding  range.  The  preponderance  of  after-second- 
year  males  at  the  Michigan  study  sites  and  in  southern  Ontario  suggests  a need  for  regional  models  of  Cerulean 
Warbler  population  dynamics.  Received  21  March  2005,  accepted  22  December  2005. 


The  population  declines  and  conservation 
status  of  Nearctic-Neotropical  migratory  bird 
species  are  the  subjects  of  much  debate  (As- 
kins  1993,  Martin  and  Finch  1995,  Robinson 
et  al.  1995,  James  et  al.  1996,  Faaborg  2002). 
The  Cerulean  Warbler  {Dendroica  cerulea)  is 
a Nearctic-Neotropical  migratory  bird  species 
that  has  declined  significantly  over  the  long- 
term throughout  its  breeding  range,  prompting 
a recent  petition  to  the  U.S.  Fish  and  Wildlife 
Service  to  assign  the  species  threatened  status 
under  the  Endangered  Species  Act  (U.S.  Fish 
and  Wildlife  Service  2002).  Breeding  Bird 
Survey  data  (1966-2000)  indicate  a popula- 
tion decline  of  3. 04 %/year  (Link  and  Sauer 

2002) .  In  Canada,  the  Cerulean  Warbler  is  a 
Species  of  Special  Concern  (Committee  on 
the  Status  of  Endangered  Wildlife  in  Canada 

2003) . 

Populations  of  Cerulean  Warblers  may  be 
negatively  affected  by  numerous  alterations  to 
their  breeding  habitats,  including  the  loss  of 
large  tracts  of  mature  deciduous  forest,  forest 
fragmentation  and  associated  negative  factors 
(e.g.,  increased  brood  parasitism  and  nest  pre- 


1  Dept,  of  Biological  Sciences,  Wichita  State  Univ., 
Wichita,  KS  67260,  USA;  e-mail: 
chris.rogers@wichita.edu 


dation),  increasing  forest  immaturity  via  ac- 
celerated harvest  cycles,  and  loss  of  key  tree 
species  (Robbins  et  al.  1992,  Hamel  2000). 
Despite  a recent  increase  in  studies  of  breed- 
ing Cerulean  Warblers  (Oliamyk  and  Robert- 
son 1996;  Jones  et  al.  2000,  2001,  2004;  Gab- 
be  et  al.  2002),  critical  information  concerning 
nest  success  and  breeding  biology  in  different 
parts  of  the  breeding  range  remains  scarce. 
Because  the  Cerulean  Warbler  is  apparently 
expanding  its  range  along  a northeastern  front 
(Hamel  2004),  northern  populations  may  be 
important  to  its  continued  persistence.  My  ob- 
jectives were  to  (1)  gather  data  on  nesting  suc- 
cess in  two  distinct  forest  habitats  (oak-hick- 
ory, locust-cherry)  in  southwestern  Michigan, 
and  (2)  describe  the  age  structure  of  breeding 
populations  and  the  social  context  of  nonsong 
vocalizations,  both  poorly  known  for  this  pa- 
rulid  species.  Because  certain  vocalizations 
may  be  given  near  the  nest  by  breeding  adults, 
the  study  of  nonsong  vocalizations  is  impor- 
tant in  evaluating  nest  productivity  (Barg 
2002). 

METHODS 

Study  sites  and  periods. — I studied  nest  suc- 
cess and  breeding  biology  in  two  large  forest 
tracts  in  southwestern  Michigan  known  to  har- 


145 


146 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


bor  breeding  populations  of  the  Cerulean  War- 
bler (Barrows  1912,  Brewer  et  al.  1991,  Ro- 
senberg et  al.  2000).  Site  1 is  a 1,813-ha  tract 
of  largely  unfragmented  oak-  ( Quercus  spp.) 
hickory  (Cary a spp.)  and  black  walnut  (Jug- 
lans  nigra ) forest  in  the  Barry  State  Game 
Area  (BSGA),  Barry  County,  Michigan  (42° 
35' N,  85°  26'  W).  BSGA  is  currently  man- 
aged for  multiple  wildlife  conservation  pur- 
poses, including  the  production  of  game  and 
nongame  species.  The  site  was  originally 
abandoned  farmland  purchased  piecemeal  by 
the  state  of  Michigan  in  the  1940s;  because  of 
natural  succession,  the  forest  at  BGSA  is  now 
relatively  mature.  The  topography  is  low  roll- 
ing hills  and  depressions.  Site  2 is  a 2,849-ha 
black  locust  ( Robinia  pseudoacacia ) and 
black  cherry  ( Prunus  serotina ) forest  in  the 
Fort  Custer  U.S.  Army  Michigan  National 
Guard  Reservation  (FTCU),  Kalamazoo 
County,  Michigan  (42°  18'  N,  85°  19'  W).  The 
site  was  obtained  as  farmland  and  scattered 
homesteads  by  the  U.S.  government  in  1917. 
The  topography  is  gentle  hills  with  little  relief. 
BSGA  and  FTCU  are  35  km  apart,  separated 
by  a landscape  of  small  towns,  small  wood- 
lots,  marshes,  lakes,  and  farmland.  Both  sites 
are  characterized  by  maturing  forest  with  large 
trees,  occasional  gaps,  and  an  open  understo- 
ry— all  habitat  features  preferred  by  Cerulean 
Warblers.  I studied  Cerulean  Warblers  at 
BSGA  from  2 June  to  27  July  2003,  and  at 
FTCU  from  15  May  to  23  July  2004,  and  from 
19  May  to  25  July  2005. 

Territory  mapping,  nest  monitoring,  and 
nest-site  characteristics . — In  2003  and  2004, 
I mapped  Cerulean  Warbler  territories  accord- 
ing to  Bibby  et  al.  (2000),  whereby  foci  of 
territorial  male  activity  are  the  primary  means 
of  identifying  individual  breeding  territories. 
At  each  site,  males  were  often  recognized  by 
song  type  (several  frequently  gave  a distinctly 
less-buzzy  ending  to  their  typical  song),  plum- 
age variation  (several  had  a distinct  white  su- 
percilium),  and  differences  in  their  stage  of 
the  reproductive  cycle.  I marked  locations  of 
male  song  perches  on  USGS  1:24,000  topo- 
graphic maps  that  were  enlarged  (by  comput- 
er) 3X;  enlarged  maps  showed  topographic 
detail  clearly,  including  specific  recognizable 
ridges,  depressions,  small  wetlands,  and  roads. 
Mapping  revealed  preferred  (tall)  trees  from 
which  individual  males  repeatedly  sang 


throughout  the  breeding  season,  usually  in  the 
center  of  the  territory.  I gained  additional  in- 
formation on  territory  boundaries  by  follow- 
ing males  as  they  patrolled  and  sang  through- 
out their  territories,  and  by  observing  bound- 
ary disputes  (counter-singing  and  direct  fights 
involving  male-male  contact).  I color-banded 
14  territorial  males  in  2005  at  FTCU,  10  of 
which  subsequently  aided  me  in  determining 
territory  boundaries  that  year;  the  remainder 
were  on  territories  that  I did  not  monitor.  All 
males  captured  and  banded  were  used  in  an 
analysis  of  age  structure.  Established  territory 
boundaries  were  evident  by  mid- June  at  both 
sites.  Global  Positioning  System  coordinates 
were  determined  for  estimated  territory  cen- 
ters and  plotted  on  topographical  maps. 

I intensively  searched  territories  every  2—6 
days  (0. 5-2.0  hr/visit)  for  the  presence  of  ac- 
tive nests  or  newly  fledged  young;  I observed 
territories  from  06:00  to  16:00  EST,  with  oc- 
casional evening  visits  from  16:00  to  21:30. 
Each  territory  check  involved  a complete  tra- 
verse through  the  entire  territory,  with  stops 
in  and  near  all  forest  gaps  to  search  for  active 
nests  and  adults  giving  contact  calls.  I moni- 
tored nests  every  2—4  days  early  in  the  nesting 
cycle,  and  every  day  as  fledging  neared.  I de- 
fined successful  nests  as  those  from  which  > 1 
warbler  young  fledged;  failed  nests  were  those 
from  which  no  warblers  or  only  Brown-head- 
ed Cowbirds  (Molothrus  ater ) fledged.  Most 
nests  were  very  high  (19—20  m)  in  the  canopy 
and  nest  contents  could  not  be  observed  di- 
rectly. I used  a spotting  scope  (20-60X)  to 
observe  female  incubation  and  brooding  be- 
havior, and  to  count  warbler  and  cowbird  nest- 
lings as  they  grew  large  enough  to  be  seen 
above  the  nest  rim.  As  fledging  approached, 
large  cowbird  young  were  easily  distinguished 
from  the  much  smaller  warbler  young  by 
plumage  features,  size,  and  vocalizations. 
Fledglings  recently  fledged  from  previously 
undiscovered  nests  (n  — 9)  were  counted  di- 
rectly when  they  emitted  loud  begging  calls. 
To  minimize  underestimating  Cerulean  War- 
bler reproduction,  I identified  to  species  and 
recorded  the  locations  of  all  begging  fledg- 
lings and  alarm-calling  adults  of  all  avian  spe- 
cies in  all  foliage  layers  of  each  Cerulean 
Warbler  territory. 

After  each  nest  had  failed  or  the  young  had 
fledged,  I recorded  the  following  nest-tree  and 


Rogers  • CERULEAN  WARBLER  BREEDING  BIOLOGY 


147 


nest  characteristics:  nest-tree  species;  nest 
height  (distance  from  ground  to  nest  bottom); 
nest-tree  height  (distance  from  ground  at  stem 
base  to  highest  foliage);  trunk  distance  (dis- 
tance from  nest  to  central  tree  axis,  measured 
from  the  ground);  foliage  distance  (distance 
from  nest  to  nearest  foliage  below  it);  nest- 
tree  diameter  at  breast  height  (dbh);  and  gap 
distance  (distance  from  nest  to  any  obvious 
forest  discontinuity  >25  m2).  Height  ratio  was 
calculated  as  nest  height/nest-tree  height.  All 
heights  were  measured  with  a rangefinder  ex- 
cept foliage  distance,  which  was  estimated  by 
eye  (nearest  meter).  1 measured  trunk  distance 
and  gap  distance  with  a transit. 

Estimating  age  of  territorial  males. — Age 
of  breeding  male  Cerulean  Warblers  was  es- 
timated using  two  methods.  Method  1 (2004- 
2005),  which  provides  a general  estimate  of 
male  age-class  frequency,  entailed  using  10  X 
32  binoculars  to  observe  whether  the  bird  had 
a distinct  white  supercilium,  purportedly  pre- 
sent only  in  second-year  (SY)  males  and  ab- 
sent in  after-second-year  (ASY)  males  (Dunn 
and  Garrett  1997);  nearly  all  males  were  ob- 
served from  <15  m.  Method  2 (2005)  was 
conducted  by  an  experienced  bird  bander,  who 
relied  on  a combination  of  molt  limits  and  the 
colors  of  flight  and  body  contour-feathers 
(Pyle  1997)  to  age  the  14  captured  territorial 
males  while  they  were  in  hand.  The  two  in- 
vestigators using  the  two  different  aging 
methods  recorded  ages  independently  of  one 
another. 

Nonsong  vocalizations. — When  monitoring 
nests  and  territories  for  reproductive  output,  I 
described  nonsong  vocalizations  emitted  by 
both  sexes  of  Cerulean  Warblers  and  docu- 
mented the  social  and  environmental  context 
of  those  vocalizations.  The  resulting  set  of  de- 
scriptions likely  represents  the  most  common 
nonsong  vocalizations  that  this  species  makes 
on  the  breeding  grounds.  I did  not  make  audio 
recordings  and  sonograms  of  these  vocaliza- 
tions; rather,  I described  them  with  previously 
established  terminology  used  for  describing 
vocalizations  of  wood  warblers  (Nolan  1978, 
Getty  1993). 

RESULTS 

Territory  defense. — From  mid-  to  late  May 
each  year,  I often  observed  male  Cerulean 
Warblers  at  FTCU  engaged  in  territorial  chas- 


ing and  occasional  physical  fights  (n  = 5 ob- 
served). Observations  of  the  BSGA  popula- 
tion began  in  early  June,  when  males  had  al- 
ready established  territories  and  were  no  lon- 
ger chasing  one  another.  Throughout  June  and 
much  of  July  at  both  sites  (and  in  mid-  to  late 
May  at  FTCU),  males  sang  prodigiously,  often 
within  10-30  m of  one  another  during  bouts 
of  intense  counter-singing  at  their  territorial 
borders;  counter-singing  sometimes  involved 
three  males. 

Nonsong  vocalizations. — Adult  Cerulean 
Warblers  emitted  four  distinct  nonsong  vocal- 
izations over  the  3-year  study.  (1)  Both  sexes 
gave  a metallic,  buzzy  zzee  call  note,  singly 
or  in  series  of  1-6  notes.  I heard  female  zzee 
calls  127  times,  and  knew  the  behavioral  con- 
text of  87:  23  occurred  when  females  were 
foraging  alone  or  with  the  male  or  a fledgling 
nearby;  33  when  near,  leaving,  or  approaching 
an  active  nest;  29  when  at  an  active  nest;  and 
2 (a  series  of  loud  zzees)  occurred  when  my 
presence  near  the  female’s  nest  apparently 
caused  alarm.  Males  gave  only  24  zzee  calls, 
with  the  behavioral  context  known  in  23  in- 
stances: 13  were  given  shortly  after  counter- 
singing near  a territorial  border;  6 while  ap- 
parently foraging  alone  or  with  a fledgling 
nearby;  3 when  near,  leaving,  or  approaching 
an  active  nest;  and  1 when  my  presence 
caused  apparent  alarm.  The  zzee  call  is  prob- 
ably the  metallic  call  note  described  as  a flight 
and  contact  call  (Oliarnyk  and  Robertson 
1996).  (2)  Both  adults  frequently  gave  long 
series  of  sweet,  nonmetallic  chip  notes  when 
I was  near  a nest  containing  nestlings  or  fledg- 
lings; in  three  cases,  a female  with  a nest  un- 
der construction  engaged  in  extensive  chip- 
ping when  I was  in  the  territory.  (3)  A high- 
pitched,  nonmetallic  alarm  tchip  was  heard  six 
times:  twice  from  females,  apparently  alarmed 
by  a nearby  Turkey  Vulture  ( Cathartes  aura) 
or  a female  Brown-headed  Cowbird  (the  latter 
near  an  active  nest);  once  from  a territorial 
male,  apparently  alarmed  by  a nearby  Red- 
tailed Hawk  ( Buteo  jamaicensis);  and  three 
times  from  three  different  females  when  I was 
near  a nest  with  nestlings.  (4)  On  10  occa- 
sions, I heard  territorial  males — always  ac- 
companying a female  with  an  active  nest 
(eggs  or  young)  or  a nest  under  construction — 
give  a series  of  very  soft,  almost  warbled 
notes;  apparently,  this  was  the  Cerulean  War- 


148 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  1.  Nest-tree  and  nest-location  variables  (mean  ± SE)  for  Cerulean  Warblers  at  Barry  State  Game 
Area  (BSGA;  n = 6 nests  in  2003)  and  Fort  Custer  (FTCU;  n = 12  nests  in  2004,  n = 18  nests  in  2005).  All 
values  are  in  meters,  except  dbh  (cm)  and  nest  height/tree  height  (proportion).  See  methods  for  definitions  of 
variables. 


Nest-tree  Nest  height/  Distance  Distance  Distance 


Site 

Nest-tree  dbh 

height 

Nest  height 

tree  height 

to  bole 

to  foliage 

to  gap 

BSGA 

FTCU 

FTCU 

45.5  ± 6.6 
38.1  ± 2.9 
41.9  ± 1.0 

21.8  ± 2.0 
26.0  ±1.1 
26.6  ± 4.0 

18.7  ± 2.1 

19.0  ± 1.4 

20.1  ± 0.2 

0.84  ± 0.04 
0.73  ± 0.04 
0.75  ± 0.37 

3.5  ± 0.6 
3.8  ± 0.5 
4.1  ± 0.2 

4.9  ± 1.9 
5.8  ± 1.1 
7.2  ± 0.4 

17.7  ± 7.2 
1.5  ± 0.8 
2.9  ± 0.3 

bier’s  whisper  “song.”  I describe  this  vocali- 
zation as  nonsong,  as  it  differed  strongly  from 
the  typical  song. 

Nest  placement  and  tree  species. — Nests 
usually  were  placed  on  a horizontal  limb  with 
a bifurcation  immediately  distal  to  the  nest, 
but  occasionally  they  were  placed  on  a lateral 
branch  in  a cluster  of  small,  upright  shoots 
with  leaves.  One  nest  was  inside  a spherical 
mass  of  Virginia  creeper  (P arthenocissus 
quinquefolia)  vines  at  the  end  of  a short  lateral 
branch.  Mean  nest-tree  dbh  exceeded  38  cm, 
and  nest-tree  height  exceeded  21  m in  all  3 
years;  at  both  sites,  nest  height  averaged  19- 
20  m (Table  1).  Trunk  and  foliage  distances 
were  essentially  similar  at  the  two  sites, 
whereas  gap  distance  differed  strongly;  nests 
tended  to  be  found  near  roads  (unpaved  sand 
and  gravel)  at  FTCU,  but  not  at  BSGA,  where 
roads  were  fewer  and  narrower  (Table  1).  At 
BSGA,  the  closest  gap  for  one  nest  was  a dirt 
road,  and  for  five  nests  the  closest  gaps  were 
natural  forest  openings  (one  small  marsh  and 
four  light  gaps  where  trees  had  fallen,  allow- 
ing light  to  reach  the  forest  floor);  at  FTCU 
(years  pooled),  the  closest  gaps  were  roads 
(13  nests)  and  natural  forest  openings  (appar- 
ently all  light  gaps;  17  nests).  Nest  height  as 
a percentage  of  tree  height  was  slightly  greater 
at  BSGA  than  at  FTCU  (Student’s  t = 2.60, 
P = 0.014,  df  = 32;  Table  1).  At  BSGA,  four 
tree  species  were  used  for  nesting:  black  oak 
( Q . velutina;  n = 3),  northern  red  oak  ( Q . rub- 
ra; n = 1),  white  oak  ( Q . alba ; n = 1),  and 
black  walnut  {n  = 1).  At  FTCU  (years 

pooled),  six  tree  species  were  used  for  nest- 
ing: black  locust  ( n = 17),  black  walnut  ( n = 
7),  black  cherry  ( n = 3),  sugar  maple  ( Acer 
saccharum\  n — 1),  and  American  sycamore 
( Platanus  occidentalis\  n = 1). 

Male  age. — Most  males  lacked  a white  su- 
percilium  (visible  through  binoculars)  at 


BSGA  (9  of  10),  FTCU  in  2004  (14  of  16), 
and  FTCU  in  2005  (16  of  20).  The  frequency 
distributions  of  the  two  male  plumage  types 
did  not  differ  between  sites  in  2004  (x2  — 
0.04,  df  = 1,  P = 0.85).  Pooling  all  years, 
84.8%  (39  of  46)  of  males  lacked  a white  su- 
percilium.  In  2005  at  FTCU,  10  of  the  14 
males  captured  and  aged  in  the  hand  were 
identified  as  ASY  males,  and  4 were  identified 
as  SY  males. 

Pairing,  nest  success,  and  brood  parasit- 
ism.— Fifteen  females  were  found  on  23  ter- 
ritories at  BSGA,  and  19  females  were  found 
on  23  (2004)  and  22  (2005)  territories  at 
FTCU.  The  relative  frequency  of  paired  and 
unpaired  males  did  not  differ  between  study 
sites  (2003  versus  2004,  x2  = 1.80,  df  = 1,  P 
= 0.18).  Apparent  nest  success  was  43%  at 
BSGA  and  FTCU  in  2004  and  52%  at  FTCU 
in  2005.  (As  most  nests  were  high  in  the  can- 
opy, exact  hatching  dates  could  not  be  deter- 
mined, and  Mayfield  estimates  of  nest  surviv- 
al could  not  be  calculated.)  At  FTCU,  two 
nests  were  found  for  each  of  three  females  in 
2004,  and  two  nests  were  found  for  each  of 
four  females  in  2005,  indicating  renesting  af- 
ter nest  failure;  no  confirmed  renests  were  re- 
corded at  BSGA.  Despite  an  appreciable  rate 
of  nest  failure,  the  exact  cause  of  nest  failure 
could  be  determined  for  only  12  nests:  3 failed 
due  to  brood  parasitism,  (one  cowbird  young 
fledged  in  each  case),  2 failed  due  to  exposure 
(initial  nest  superstructure  destroyed  by  heavy 
rain  or  nest  branch  broken  off  by  high  winds), 
and  7 failed  due  to  predation.  In  the  last  case, 
the  nest’s  rim  was  torn  and/or  the  entire  nest 
was  tipped.  In  2004,  Cerulean  Warbler  pairs 
at  BSGA  and  FTCU  fledged  0.1  cowbird/ 
breeding  pair  and  0.1  cowbird/nest;  in  2005, 
warbler  pairs  at  FTCU  fledged  0.2  cowbird/ 
breeding  pair  and  0.1  cowbird/nest. 

In  23  territories  intensively  monitored  at 


Rogers  • CERULEAN  WARBLER  BREEDING  BIOLOGY 


149 


BSGA  in  2003,  six  nests  and  one  recently 
fledged  brood  were  found,  and  nine  fledglings 
(3.0/successful  nest)  were  produced.  In  23  ter- 
ritories monitored  at  FTCU  in  2004,  12  nests 
and  2 recently  fledged  broods  were  found,  and 
12  young  fledged  (2.0  per  successful  nest).  In 
22  territories  monitored  at  FTCU  in  2005,  19 
nests  and  6 recently  fledged  broods  were 
found,  and  30  fledglings  (2.3  per  successful 
nest)  were  produced. 

Male  reproductive  success. — At  BSGA, 
male  reproductive  success  was  0.20  (all 
males)  and  0.30  (paired  males  only)  male 
fledglings/male.  Corresponding  values  at 
FTCU  were  0.26  and  0.32  male  fledglings/ 
male  in  2004,  and  0.68  and  0.79  in  2005.  The 
number  of  fledglings/breeding  pair  was  0.60 
at  BSGA,  and  0.63  (2004)  and  1.58  (2005)  at 
FTCU. 

DISCUSSION 

Nest  placement. — As  previous  workers  have 
found  in  other  parts  of  the  species’  breeding 
range  (Oliarnyk  and  Robertson  1996,  Hamel 
2000,  Jones  and  Robertson  2001,  Jones  et  al. 
2001),  Cerulean  Warblers  in  Michigan  chose 
a diversity  of  tree  species  for  nest  placement. 
Nests  at  BSGA  and  FTCU  averaged  19-20  m 
in  height;  typical  (pre-ice  storm)  nest  height 
in  a southern  Ontario  population  was  1 1 .6— 
11.8  m (Oliarnyk  and  Robertson  1996,  Jones 
et  al.  2001)  and  the  range-wide  mean  nest 
height  (excluding  the  present  study)  is  1 1 .4  m 
(Hamel  2000).  Nest  height/nest-tree  height  of 
nests  in  the  Ontario  population  (0.61)  was 
lower  than  in  southern  Michigan  (0.73-0.84). 
Thus,  not  only  did  Michigan  Cerulean  War- 
blers choose  high  nest  sites,  they  also  nested 
relatively  high  within  a given  nest  tree.  Given 
the  intensive  season-long  nest  searches  con- 
ducted throughout  all  territories  studied,  it  is 
unlikely  that  any  low  nests  were  missed. 

Nonsong  vocalizations. — The  zzee  call  note 
appears  multifunctional,  as  it  was  used  by 
both  sexes  in  a variety  of  situations.  It  was 
given  more  often  by  females,  which  appar- 
ently used  the  note  as  a contact  call  when  they 
were  at,  or  close  to,  the  nest;  the  call  may  also 
function  as  an  alarm  note.  Therefore,  I used 
female  zzee  calls  as  cues  for  finding  nests; 
when  heard,  I attempted  to  watch  the  female 
return  to  an  active  nest,  or  to  find  the  nest  if 
the  call  was  thought  to  have  been  given  by  a 


sitting  female.  The  relatively  few  occurrences 
of  male  zzees  were  nearly  all  associated  with 
active  territorial  defense  against  a nearby  rival 
male,  but  several  were  given  near  an  active 
nest.  The  sweet  chip  notes — given  in  response 
to  my  presence  near  nests  containing  older 
nestlings  or  fledglings — were  alarm  notes  gen- 
erally resembling  the  alarm  chips  of  other 
North  American  parulids  (Getty  1993).  The 
higher-pitched  tchip  alarm  note  was  rarely 
heard,  and  only  in  response  to  either  a poten- 
tial predator,  a cowbird  near  the  nest,  or  my 
presence  in  the  territory.  I heard  the  whisper 
“song”  10  times,  all  in  the  context  of  a male 
interacting  with  a female  near  an  active  nest. 
Intersexual  behavior  was  difficult  to  observe, 
as  birds  generally  remained  high  in  the  forest 
canopy,  but  males  may  give  this  vocalization 
as  a “song  cue”  to  nesting  females  (Barg 
2002). 

Male  age. — Cerulean  Warbler  males  with  a 
white  supercilium  (SY  males,  Dunn  and  Gar- 
rett 1997)  composed  10-20%  of  all  territorial 
males  at  my  study  sites.  Only  4 of  the  14 
banded  males  in  2005  were  in  the  SY  age 
class.  Although  aging  by  supercilium  alone  is, 
at  best,  an  approximation,  general  agreement 
between  the  two  aging  methods  used  suggests 
that  a significant  majority  of  the  breeding  ter- 
ritorial males  were  in  their  second  or  a sub- 
sequent breeding  season.  In  southern  Ontario, 
15%  of  males  are  thought  to  be  SY  birds 
(Jones  et  al.  2004).  Thus,  at  least  two  popu- 
lations in  the  northern  part  of  the  breeding 
range  are  biased  toward  older  males.  The  two 
southwestern  Michigan  populations  I studied 
are  approximately  774  km  from  the  southern 
Ontario  study  site.  Within  the  range  of  another 
Dendroica  species,  the  Black- throated  Blue 
Warbler  ( D . caerulescens ),  Graves  (1997) 
found  latitudinal  segregation  among  males,  by 
age  class.  Furthermore,  first-year  American 
Redstart  ( Setophaga  ruticilla ) males  are 
forced  into  marginal  breeding  habitat  by  older 
males  (Ficken  and  Ficken  1967,  Sherry  and 
Holmes  1989).  Further  field  study  is  required 
to  investigate  possible  habitat-specificity  and/ 
or  broader  geographical  extent  of  age-struc- 
tured breeding  populations  of  Cerulean  War- 
blers. 

Population  productivity. — Excluding  un- 
paired males,  male  reproductive  success  was 
0.30-0.32  male  fledglings/breeding  male  at 


150 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


BSGA  (2003)  and  FTCU  (2004)  and  0.79 
male  fledglings/breeding  male  at  FTCU 
(2005).  Reproductive  output  in  southwestern 
Michigan  was,  therefore,  poor  in  two  distinct 
habitat  types  in  2 of  3 years,  and  in  all  3 years 
it  was  lower  than  the  productivity  of  a south- 
ern Ontario  population  (0.94  male  fledglings/ 
breeding  male) — thought  to  be  a sink  popu- 
lation requiring  1.7  male  fledglings/breeding 
male  for  sustainability  (Jones  et  al.  2004). 
Survival  estimates  for  the  Michigan  study 
populations  are  needed,  however,  before  de- 
termining whether  they  are  self-sustaining  or 
not. 

A previous  study  of  passerine  populations 
in  southwestern  Michigan  (Rogers  and  Caro 
1998)  indicated  that  corvid  nest  predators 
(American  Crow,  Corvus  brachyrhynchos; 
Blue  Jay,  Cyanocitta  cristata ) and  brood  par- 
asites (Brown-headed  Cowbird)  are  regular  in- 
habitants of  all  habitat  types,  including  the  in- 
terior of  large  forest  tracts,  as  well  as  subur- 
ban, exurban,  and  agricultural  areas.  Avian 
nest  predators  and  cowbirds  were  frequently 
observed  in  both  forest  edge  and  interior  on 
annual  point  counts  from  1993  to  2004  (CMR 
unpubl.  data).  Corvids,  therefore,  are  candi- 
dates for  causing  nest  failure  in  Cerulean  War- 
blers at  BSGA  and  FTCU.  Nest  predation  also 
may  have  been  caused  by  eastern  fox  squirrels 
(, Sciurus  niger),  which  were  common  at  both 
study  sites,  and  eastern  chipmunks  ( Tamias 
striatus ),  which  were  abundant  at  BSGA. 
Both  sciurid  species  were  observed  in  trees  at 
heights  exceeding  10  m.  Effects  of  cowbird 
parasitism  were  low  in  all  3 years;  however, 
>60%  of  the  Hooded  Warbler  ( Wilsonia  ci- 
trina ) nests  at  FTCU  were  parasitized  by  cow- 
birds  in  2004-2005  (R.  Adams  pers.  comm.). 
The  Cerulean  Warbler’s  status  as  a poor  cow- 
bird host  deserves  further  attention. 

A possible  source  of  error  in  my  study  was 
the  failure  to  detect  fledglings  in  all  territories. 
Although  this  cannot  be  completely  discount- 
ed, any  error  was  probably  negligible,  as  Ce- 
rulean Warbler  adults  feeding  older  nestlings 
or  fledglings  typically  became  excited,  and 
gave  frequent  and  obvious  alarm  calls  (chips); 
in  addition,  all  fledged  broods  and  adults  (of 
all  species)  emitting  alarm-chips  were  identi- 
fied to  species.  It  is  unlikely  that  enough 
fledglings  were  missed  to  bias  my  estimates 
of  (low)  reproductive  output.  Some  territories 


yielded  no  fledglings  for  an  entire  breeding 
season.  This  is  not  necessarily  surprising,  as 
the  breeding  season  is  short:  no  July  nest 
starts  were  found  in  2 years,  and  the  species 
begins  breeding  after  spring  migration  in  mid- 
May.  At  BSGA  in  2003,  females  were  not  de- 
tected on  8 of  23  territories  (35%);  thus,  some 
females  may  have  been  missed.  However, 
Holmes  et  al.  (1996)  found  a similar  percent- 
age (27%)  of  unpaired  males  among  Black- 
throated  Blue  Warblers. 

Low  reproductive  output  among  Cerulean 
Warblers  may  be  a factor  contributing  to  their 
long-term  population  decline.  To  test  this  hy- 
pothesis more  rigorously,  additional  studies  of 
this  species  are  needed.  Specifically,  repro- 
ductive output  should  be  measured  in  more 
regions.  In  addition,  longer-term  studies 
would  be  useful  for  assessing  temporal  vari- 
ation in  reproduction  within  individual  sites. 
Finally,  age  structure  strongly  biased  toward 
older  males  suggests  a need  for  regional,  as 
opposed  to  local,  models  of  the  Cerulean  War- 
bler’s population  dynamics. 

ACKNOWLEDGMENTS 

The  able  field  assistance  of  staff  members  of  the 
Kalamazoo  Nature  Center  is  gratefully  acknowledged, 
particularly  that  of  B.  Nelson,  who  captured  and  aged 
male  Cerulean  Warblers.  The  U.S.  Army  Fort  Custer 
Reserve  unit  and  the  Michigan  Department  of  Natural 
Resources  kindly  permitted  access  to  the  study  sites, 
and  Kellogg  Biological  Station  provided  important  lo- 
gistical support.  J.  Jones  and  two  anonymous  referees 
provided  useful  comments  on  an  earlier  version  of  the 
manuscript.  This  is  contribution  number  1210  of  Kel- 
logg Biological  Station. 

LITERATURE  CITED 

Askins,  R.  A.  1993.  Population  trends  in  grassland, 
shrubland,  and  forest  birds  in  eastern  North  Amer- 
ica. Current  Ornithology  11:1-34. 

Barg,  J.  J.  2002.  Small-scale  biological  phenomena  in 
a migrant  songbird.  M.Sc.  thesis.  Queen’s  Uni- 
versity, Kingston,  Ontario. 

Barrows,  W.  B.  1912.  Michigan  bird  life.  Michigan 
Agricultural  College,  East  Lansing. 

Bibby,  C.  J.,  N.  D.  Burgess,  D.  A.  Hill,  and  S.  Mus- 
toe.  2000.  Bird  census  techniques,  2nd  ed.  Aca- 
demic Press,  London,  United  Kingdom. 

Brewer,  R.,  G.  A.  McPeek,  and  R.  J.  Adams,  Jr. 
1991.  The  atlas  of  breeding  birds  of  Michigan. 
Michigan  State  University  Press,  East  Lansing. 
Committee  on  the  Status  of  Endangered  Wildlife 
in  Canada.  2003.  Assessment  and  update  status 
report  on  the  Cerulean  Warbler  Dendroica  cerulea 
in  Canada.  Ottawa,  Ontario. 


Rogers  • CERULEAN  WARBLER  BREEDING  BIOLOGY 


151 


Dunn,  J.  L.  and  K.  L.  Garrett.  1997.  Warblers. 
Houghton-Mifflin,  New  York. 

Faaborg,  J.  A.  2002.  Saving  migrant  birds:  developing 
strategies  for  the  future.  University  of  Texas  Press, 
Austin. 

Ficken,  M.  S.  and  R.  W.  Ficken.  1967.  Age-specific 
differences  in  the  breeding  behavior  and  ecology 
of  the  American  Redstart.  Wilson  Bulletin  79: 
188-199. 

Gabbe,  A.  R,  S.  K.  Robinson,  and  J.  D.  Brawn.  2002. 
Tree-species  preferences  for  foraging  insectivo- 
rous birds:  implications  for  floodplain  forest  res- 
toration. Conservation  Biology  16:462-470. 

Getty,  S.  R.  1993.  Call-notes  of  North  American 
wood  warblers.  Birding  25:159-168. 

Graves,  G.  R.  1997.  Geographic  dines  of  age  ratios 
of  Black- throated  Blue  Warblers  ( Dendroica  cae- 
rulescens).  Ecology  78:2524-2531. 

Hamel,  P.  B.  2000.  Cerulean  Warbler  ( Dendroica  ce- 
rulea).  The  Birds  of  North  America,  no.  511. 

Hamel,  P.  B.,  D.  K.  Dawson,  and  P.  D.  Keyser.  2004. 
How  we  can  learn  more  about  the  Cerulean  War- 
bler ( Dendroica  cerulea ).  Auk  121:7-14. 

Holmes,  R.  T.,  P.  P.  Marra,  and  T.  W.  Sherry.  1996. 
Habitat-specific  demography  of  breeding  Black- 
throated  Blue  Warblers  ( Dendroica  caerulescens ): 
implications  for  population  dynamics.  Journal  of 
Animal  Ecology  65:183-185. 

James,  F.  C.,  C.  E.  McCullough,  and  D.  E.  Wieden- 
feld.  1996.  New  approaches  to  the  analysis  of 
population  trends  in  landbirds.  Ecology  77:13-27. 

Jones,  J.,  J.  J.  Barg,  T.  S.  Sillett,  M.  L.  Veit,  and 
R.  J.  Robertson.  2004.  Minimum  estimates  of 
survival  and  population  growth  for  Cerulean  War- 
blers ( Dendroica  cerulea)  breeding  in  Ontario, 
Canada.  Auk  121:15-22. 

Jones,  J.,  R.  D.  DeBruyn,  J.  J.  Barg,  and  R.  J.  Rob- 
ertson. 2001.  Assessing  the  effects  of  natural  dis- 
turbance on  a Neotropical  migrant  songbird.  Ecol- 
ogy 82:2628-2635. 

Jones,  J.,  P.  Ramoni-Perlazzi,  E.  H.  Carruthers,  and 
R.  J.  Robertson.  2000.  Sociality  and  foraging  be- 
havior of  the  Cerulean  Warbler  in  Venezuelan 
shade-coffee  plantations.  Condor  102:958-962. 

Jones,  J.  and  R.  J.  Robertson.  2001.  Territory  and 


nest-site  selection  of  Cerulean  Warblers  in  eastern 
Ontario.  Auk  1 18:727-735. 

Link,  W.  A.  and  J.  R.  Sauer.  2002.  A hierarchical 
analysis  of  population  change  with  application  to 
Cerulean  Warblers.  Ecology  83:2832-2840. 

Martin,  T.  E.  and  D.  M.  Finch  (Eds.).  1995.  Ecology 
and  management  of  Neotropical  migratory  birds. 
Oxford  Press,  New  York. 

Nolan,  V.,  Jr.  1978.  Ecology  and  behavior  of  the  Prai- 
rie Warbler  ( Dendroica  discolor).  Ornithological 
Monographs,  no.  26. 

Oliarnyk,  C.  J.  and  R.  J.  Robertson.  1996.  Breeding 
behavior  and  reproductive  success  of  Cerulean 
Warblers  in  southeastern  Ontario.  Wilson  Bulletin 
108:673-684. 

Pyle,  P.  1997.  Identification  guide  to  North  American 
birds,  part  I.  Columbidae  to  Ploceidae.  Slate  Creek 
Press,  Bolinas,  California. 

Robbins,  C.  S.,  J.  W.  Fitzpatrick,  and  P.  B.  Hamel. 
1992.  A warbler  in  trouble:  Dendroica  cerulea. 
Pages  549-562  in  Ecology  and  conservation  of 
Neotropical  migrant  landbirds  (J.  M.  Hagan,  III, 
and  D.  W.  Johnston,  Eds.).  Smithsonian  Institution 
Press,  Washington,  D.C. 

Robinson,  S.  K.,  F.  R.  Thompson,  III,  T.  M.  Donovan, 
D.  R.  Whitehead,  and  J.  A.  Faaborg.  1995.  Re- 
gional forest  fragmentation  and  the  nesting  suc- 
cess of  migratory  birds.  Science  267:1987-1990. 

Rogers,  C.  M.  and  M.  J.  Caro.  1998.  A test  of  the 
mesopredator  release  hypothesis  in  wetland-edge 
Song  Sparrows  ( Melospiza  melodia).  Oecologia 
116:227-233. 

Rosenberg,  K.  V.,  S.  E.  Barker,  and  R.  W.  Rohr- 
baugh.  2000.  An  atlas  of  Cerulean  Warbler  pop- 
ulations. Cornell  Lab  of  Ornithology,  Ithaca,  New 
York. 

Sherry,  T.  W.  and  R.  T.  Holmes.  1989.  Age-specific 
social  dominance  affects  habitat  use  by  breeding 
American  Redstarts  ( Setophaga  ruticilla):  a re- 
moval experiment.  Behavioural  Ecology  and  So- 
ciobiology 25:327-333. 

U.S.  Fish  and  Wildlife  Service.  2002.  Endangered 
and  threatened  wildlife  and  plants:  90-day  finding 
on  a petition  to  list  the  Cerulean  Warbler  as  threat- 
ened with  critical  habitat.  Federal  Register  67: 
65083-65086. 


The  Wilson  Journal  of  Ornithology  1 18(2):  152— 163,  2006 


MIGRANT  SHOREBIRD  PREDATION  ON  BENTHIC 
INVERTEBRATES  ALONG  THE  ILLINOIS  RIVER,  ILLINOIS 

GABRIEL  L.  HAMER,1  246  EDWARD  J.  HESKE,12  JEFFREY  D.  BRAWN,23  AND 

PATRICK  W.  BROWN15 


ABSTRACT. — We  evaluated  the  effect  of  shorebird  predation  on  invertebrates  at  a wetland  complex  along 
the  Illinois  River,  west-central  Illinois,  during  spring  migration.  Using  a new  exclosure  experiment  design  adapted 
to  the  shifting  nature  of  foraging  microhabitat  of  interior  wetlands,  we  found  that  shorebird  predation  did  not 
significantly  deplete  total  invertebrate  density  or  total  biomass  in  open  (no  exclosure)  versus  exclosure  treatments. 
Chironomids  and  oligochaetes  were  the  most  common  invertebrates  occurring  in  substrate  samples.  The  density 
of  oligochaetes  was  lower  in  open  treatments,  though  the  degree  of  difference  varied  both  spatially  and  tem- 
porally. Shorebird  density  was  positively  correlated  with  the  amount  of  invertebrate  biomass  removed  from  the 
substrate  during  the  late-May  sampling  period.  Our  results  suggest  that  shorebirds  use  an  opportunistic  foraging 
strategy  and  consume  the  most  abundant  invertebrate  prey.  The  dynamic  hydrology  at  our  study  site  likely 
played  a role  in  preventing  invertebrate  depletion  by  continually  exposing  new  foraging  areas  and  prey.  Received 
16  February  2005,  accepted  30  December  2005. 


Migrating  shorebirds  (Charadriiformes)  re- 
quire stopover  resources  for  rest  and  rapid  ac- 
cumulation of  energy  to  fuel  their  transconti- 
nental migration  (Myers  et  al.  1987).  As  fresh- 
water wetlands  in  the  United  States  continue 
to  be  converted  to  agriculture  and  develop- 
ment (Dahl  2000),  the  reduction  in  stopover 
areas  is  believed  to  have  negative  effects  on 
shorebird  populations  (Sutherland  and  Goss- 
Custard  1991,  Harrington  et  al.  2002).  Con- 
sequently, many  North  American  shorebirds 
are  listed  as  threatened,  endangered,  or  species 
of  special  concern  (Brown  et  al.  2001,  Mor- 
rison et  al.  2001),  including  Greater  Yellow- 
legs  ( Tringa  melanoleuca ),  Short-billed  Dow- 
itcher  ( Limnodromus  griseus),  and  Buff- 
breasted Sandpiper  ( Tryngites  subruficollis ) in 
the  Mississippi  Alluvial  Valley  and  Great 
Lakes  region. 


1 Center  for  Wildlife  and  Plant  Ecology,  Illinois  Nat- 
ural History  Survey,  607  E.  Peabody  Dr.,  Champaign, 
IL  61820,  USA. 

2 Dept,  of  Natural  Resources  and  Environmental 
Sciences,  Univ.  of  Illinois,  W-503  Turner  Hall,  1102 
S.  Goodwin  Ave.,  Urbana,  IL  61801,  USA. 

3 Dept,  of  Animal  Biology,  Univ.  of  Illinois,  Shel- 
ford  Vivarium,  606  E.  Healey,  Champaign,  IL  61820, 
USA. 

4 Current  address:  Dept,  of  Fisheries  and  Wildlife, 
Michigan  State  Univ.,  13  Natural  Resources,  East  Lan- 
sing, MI  48824,  USA. 

5 Current  address:  Michigan  Natural  Features  Inven- 
tory, Michigan  State  Univ.  Extension.  P.O.  Box  30444, 
Lansing.  MI  48909.  USA. 

6 Corresponding  author;  e-mail:  ghamer@msu.edu 


While  migrating  through  the  interior  United 
States,  shorebirds  are  faced  with  unpredictable 
habitats  that  are  much  different  from  coastal 
systems  (Skagen  and  Knopf  1994a).  The  pre- 
dictability of  tidal  cycles  and  blooms  of  food 
resources  in  the  intertidal  zones  of  coastal  sys- 
tems support  large  concentrations  of  shore- 
birds  and  high  levels  of  site  fidelity  in  loca- 
tions such  as  Delaware  Bay  along  the  north- 
east Atlantic  coast  and  the  Copper  River  Delta 
in  the  Gulf  of  Alaska.  In  contrast,  shorebirds 
using  interior  flyways  are  more  dispersed  and 
occur  at  stopover  habitats  in  smaller  numbers 
than  those  using  coastal  flyways  (Skagen  and 
Knopf  1993).  Some  shorebirds  undertake 
long,  nonstop  flights;  many  other  species  do 
not  depart  with  enough  fuel  to  reach  their  final 
destinations  and  must  make  multiple  stops  to 
refuel  during  migration  (White  and  Mitchell 

1990,  Skagen  and  Knopf  1994b,  Farmer  and 
Wiens  1999) — a less  energetically  challenging 
strategy  (Piersma  1987). 

Shorebirds  are  opportunistic  feeders  and 
readily  shift  their  diet  to  exploit  locally  abun- 
dant invertebrate  resources  (Skagen  and  Oman 
1996).  Studies  of  shorebird  diet  among  inte- 
rior stopover  habitats  indicate  that  chironomid 
larvae  are  the  dominant  prey  items  (Helmers 

1991,  Mihue  et  al.  1997).  Much  less  is  known 
about  the  importance  of  oligochaetes — often 
the  most  abundant  invertebrates  in  freshwater 
mudflats  in  the  Mississippi  Alluvial  Valley 
(Elliott-Smith  2003,  Hamer  2004,  Mitchell 
and  Grubaugh  2005) — as  prey  (Safran  et  al. 


152 


Hamer  et  al.  • SHOREBIRD  PREDATION  ON  BENTHIC  INVERTEBRATES 


153 


1997).  The  importance  of  oligochaetes  may  be 
underestimated  because  they  are  small,  frag- 
ile, sensitive  to  post-mortem  digestion  in 
esophageal,  proventricular,  and  gizzard  con- 
tents, and  are  thus  often  ignored  in  analysis 
(Rundle  1982,  Safran  et  al.  1997).  However, 
oligochaetes  are  comparable  to  chironomids  in 
caloric  value  (5,575  and  5,424  calories/g  dry 
weight,  respectively),  crude  protein,  and  gross 
energy  (Cummins  and  Wuycheck  1971,  An- 
derson and  Smith  1998). 

Observational  studies,  esophageal  analyses, 
and  exclosure  experiments  have  been  used  to 
assess  the  interactions  between  shorebirds  and 
their  prey  (Brooks  1967,  Schneider  1978, 
Evans  et  al.  1979,  Rundle  1982,  Swennen 
1990).  Food  consumption  has  been  measured 
using  indirect  visual  methods  in  many  studies 
of  the  foraging  ecology  of  Palearctic,  coastal 
shorebirds  (Evans  et  al.  1979,  Moreira  1997). 
These  indirect  methods,  however,  are  often 
challenging  to  use  in  inland  systems  where 
prey  are  small  and  successful  and  unsuccess- 
ful foraging  pecks  and  probes  are  not  distin- 
guishable. Collecting  individual  shorebirds  for 
esophageal  analysis  provides  valuable  infor- 
mation on  diet,  but  it  does  not  determine  the 
effect  of  shorebird  predation  on  the  inverte- 
brate community  and  may  produce  bias 
caused  by  missing  soft-bodied  invertebrates 
(Rundle  1982).  A less  invasive  technique  for 
investigating  shorebird-prey  relationships  is  to 
use  exclosure  experiments,  also  termed  caging 
experiments,  which  entail  structures  that  pre- 
vent shorebirds  from  feeding  on  invertebrates 
within  the  enclosed  substrate.  The  invertebrate 
community  within  the  exclosure  can  be  com- 
pared with  that  in  equivalent  substrate  outside 
the  exclosure  for  an  indirect  measure  of  shore- 
bird  predation  on  invertebrates. 

Recently,  researchers  have  implemented  ex- 
closure experiments  at  freshwater  shorebird 
stopover  sites  (Mihue  et  al.  1997,  Ashley 
2000,  Mitchell  and  Grubaugh  2005),  but  pre- 
viously the  majority  had  been  conducted  in 
marine  intertidal  systems  (Wilson  1991,  Mer- 
cier  and  McNeil  1994,  Weber  and  Haig  1997). 
Results  of  these  exclosure  experiments  are 
varied;  some  studies  have  revealed  up  to  90% 
reductions  in  prey  densities  due  to  shorebird 
predation  (Schneider  and  Harrington  1981, 
Szekely  and  Bamberger  1992),  whereas  other 
studies  document  no  measurable  effect  (Raf- 


faelli  and  Milne  1987,  Mitchell  and  Grubaugh 
2005).  During  migration  in  the  interior  fly- 
ways,  the  extent  of  shorebird  predation  on  dif- 
ferent invertebrate  taxa  at  stopover  areas  is  not 
clear. 

We  conducted  an  exclosure  experiment  at  a 
shorebird  stopover  location  in  the  Upper  Mis- 
sissippi Alluvial  Valley.  Our  primary  objec- 
tives were  to  evaluate  (1)  whether  shorebird 
predation  depletes  invertebrate  prey  during 
migration  along  an  interior  flyway,  (2)  which 
invertebrates  and  size  classes  are  removed 
from  the  substrate,  (3)  the  chronology  in 
abundance  and  biomass  of  benthic  inverte- 
brates, and  (4)  a new  exclosure-experiment 
design  adapted  to  the  unpredictable  nature  of 
interior  shorebird  foraging  habitats. 

METHODS 

Study  area. — Our  study  was  conducted  at 
Chautauqua  National  Wildlife  Refuge  (NWR) 
(40°  38'  N,  89°  99'  W)  and  Emiquon  NWR 
(40°  32' N,  90°09'W),  which  are  part  of  a 
large  wetland  complex  along  the  Illinois  River 
in  west-central  Illinois  near  Havana  (Fig.  1 A). 
The  1,816-ha  refuge  at  Chautauqua  NWR  was 
established  in  1936  and  consists  of  large  back- 
water lakes,  and  bottomland  and  upland  for- 
est. Chautauqua  also  has  been  designated  a 
stopover  of  international  importance  by  the 
Western  Hemisphere  Shorebird  Reserve  Net- 
work (Harrington  and  Perry  1995).  The  late 
drawdown  in  July  and  August  at  this  refuge 
creates  extensive,  shallow-water  mudflats  at- 
tracting an  estimated  100,000  to  250,000 
shorebirds  each  fall  (Bailey  2003).  Compara- 
tively little  shorebird  habitat  is  available  at 
Chautauqua  in  the  spring,  when  water  levels 
are  elevated  to  prevent  encroachment  of  ex- 
otic invasives — black  willow  ( Salix  nigra ) and 
cocklebur  ( Xanthium  strumarium ) — that  inter- 
fere with  moist-soil  plant  production. 

Emiquon  NWR  is  an  856-ha  refuge  com- 
posed of  backwater  lakes,  sloughs,  forested 
wetlands,  and  a variety  of  other  terrestrial  hab- 
itats. Because  Emiquon  was  only  just  acquired 
in  1993,  much  of  the  refuge  comprises  newly 
established  wetland,  and  portions  will  remain 
in  agriculture  until  leases  with  private  land- 
owners  expire.  The  refuge  is  divided  into  two 
main  units:  Wilder  Tract  (197  ha)  and  South 
Globe  (288  ha).  The  Wilder  Tract  was  taken 
out  of  agricultural  production  in  1998  and  is 


154 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


A 


FIG.  1 . (A)  Location  of  the  three  study  sites  near 

Havana,  Illinois  (Chautauqua  South  Pool,  Emiquon 
South  Globe,  Emiquon  Wilder  Tract)  where  shorebird 
predation  was  studied  from  February  to  June  2004. 
White  squares  show  approximate  location  of  study 
plots.  (B)  Depiction  of  a plot  (1  ha)  containing  one 
exclosure  and  one  open  (no  exclosure)  treatment  used 
in  this  study.  The  dashed  lines  indicate  approximate 
location  of  the  shoreline  (mud/water  interface  where 
shorebirds  foraged). 


managed  as  a moist-soil  unit.  The  South 
Globe  unit  was  taken  out  of  production  for  the 
first  time  in  2004,  when  the  remaining  corn 
and  bean  stubble  were  flooded  to  create  ex- 
tensive shallow  water  habitat. 

Field  methods. — The  exclosure  experiment 
was  conducted  during  spring  shorebird  migra- 
tion from  March  through  June  2004.  Three 
plots  were  established  at  each  of  the  three  field 
sites  (Chautauqua  South  Pool,  Emiquon  Wil- 
der Tract,  Emiquon  South  Globe)  for  a total 
of  nine  plots  (Fig.  1A).  Each  plot  was  1 ha  in 
size  (100  X 100  m,  designated  by  flags  at  each 
corner)  and  contained  both  an  exclosure  treat- 
ment and  an  open  treatment.  The  exclosure 
consisted  of  a sheet  (16  X 1 m)  of  metal  fenc- 
ing (mesh  = 5X10  cm)  positioned  horizon- 
tally and  supported  10  cm  above  the  substrate 
by  metal  stakes  at  each  corner  and  at  5-m  in- 
tervals along  both  sides  (Fig.  IB).  The  long 
axis  of  the  exclosure  was  placed  perpendicular 


to  the  shoreline  so  that  the  shoreline  always 
remained  within  some  part  of  the  exclosure  as 
water  levels  fluctuated.  Because  the  fence 
sagged  between  the  metal  stakes,  small  sec- 
tions of  black  willow  branches  were  used  to 
prop  up  the  fence  to  maintain  the  entire  unit 
at  a 10-cm  height.  Few  predators  of  benthic 
invertebrates — other  than  shorebirds,  largely 
predatory  invertebrates,  and  crayfish — occur 
in  this  inland  system.  The  lack  of  sides  on  the 
exclosure,  however,  allowed  access  by  other 
predators  and  excluded  only  avian  predators. 
The  open  treatment  lacked  any  fencing  but 
was  marked  by  flags  to  the  same  dimensions 
of  the  exclosure.  The  open  and  exclosure 
treatments  were  placed  40  m apart  and  30  m 
from  the  edges  of  the  plot  (Fig.  IB).  Because 
of  the  changing  hydrology  and  changing  lo- 
cations of  shorebird  habitat,  plots  were  not  es- 
tablished at  the  same  time.  The  first  plot  was 
established  on  27  February  and  the  last  on  29 
April. 

We  determined  shorebird  use  of  the  plots 
by  conducting  censuses  twice  per  week  at 
each  plot  during  the  peak  of  migration  (mid- 
April  to  the  end  of  May)  and  once  per  week 
during  the  remainder  of  spring  migration. 
Means  were  calculated  for  each  2-week  period 
for  each  plot  to  determine  average  shorebird 
density  in  the  2-week  period  before  inverte- 
brate sampling.  The  first  survey  was  on  6 
March  and  the  last  was  on  16  June.  During 
each  census,  we  identified  and  counted  all 
shorebirds  in  the  1-ha  plot  (from  a vehicle  or 
on  foot)  using  8 X 42  binoculars  or  a 15-45X 
spotting  scope.  We  recorded  water  levels  dur- 
ing each  census  using  a PVC  pipe  (vertical 
pole)  marked  at  1-cm  intervals;  a pole  was 
placed  permanently  outside  each  plot  in  water 
that  was  deeper  than  it  was  inside  the  plot.  We 
determined  change  in  water  level  by  compar- 
ing the  water  level  from  each  2-week  sam- 
pling period  at  each  plot.  The  absolute  value 
of  the  change  in  water  level  was  used  in  the 
analysis. 

We  sampled  for  benthic  invertebrates  in 
both  treatments  when  each  plot  was  estab- 
lished and  then  at  2-week  intervals  throughout 
spring  migration.  The  first  samples  were  taken 
on  27  February  and  the  last  on  6 June.  Each 
treatment  was  sampled  at  the  shoreline  (where 
edge  of  surface  water  meets  mudflat),  which 
was  the  primary  shorebird  foraging  zone. 


Hamer  et  al.  • SHOREBIRD  PREDATION  ON  BENTHIC  INVERTEBRATES 


155 


Only  one  core  sample  per  2-week  interval  was 
taken  from  each  treatment  to  avoid  potential 
resampling  of  the  same  area  in  subsequent 
sampling  periods  and  to  avoid  sediment  dis- 
turbance. Ashley  et  al.  (2000)  conducted  a 
study  in  which  two  cores  were  sampled  in 
each  treatment;  they  found  no  difference  be- 
tween the  subsamples  and  recommended  elim- 
inating them  in  future  exclosure  studies.  We 
used  core  samplers,  similar  to  those  developed 
by  Swanson  (1978),  that  were  modified  by  us- 
ing metal  conduit  piping  with  a sharpened 
edge.  We  extracted  core  samples  5 cm  in  di- 
ameter to  a depth  of  5 cm  (Sherfy  et  al.  2000). 
After  inserting  the  core  sampler  into  the  sub- 
strate, we  placed  a plumber’s  stopper  plug  in 
the  end  of  the  core  sampler  to  aid  in  removal 
of  the  core.  Contents  of  the  sampler  were 
placed  in  a resealable  plastic  bag  containing 
95%  ethyl  alcohol,  stained  with  Rose  Bengal, 
and  kept  cool  until  sorted. 

Laboratory  methods. — Invertebrates  were 
removed  from  the  preserved  sample  using  a 
number  30  mesh  sieve  and  identified  to  order 
or  family  according  to  Pennak  (1989)  and 
Merritt  and  Cummins  (1996).  All  samples 
were  sorted  by  one  observer  to  reduce  bias. 
Chironomids  and  gastropods  were  sorted  into 
two  size  classes:  <5  mm  and  >5  mm.  All 
invertebrates,  excluding  gastropods,  were 
dried  at  70°  C for  24  hr  on  pre-dried  and  pre- 
weighed glass  microfiber  filters.  To  determine 
biomass,  we  weighed  samples  to  the  nearest 
0.0001  g using  a Mettler  balance.  Invertebrate 
densities  (no.  individuals)  and  biomasses  (g) 
are  reported  per  m2. 

Statistical  analysis. — To  determine  whether 
differences  existed  between  the  two  treat- 
ments prior  to  the  experiment,  we  used  paired 
r-tests  to  compare  measures  of  invertebrate 
density  and  biomass  before  we  established  the 
plots.  To  analyze  invertebrate  density  and  bio- 
mass, we  used  a repeated  measures  mixed- 
model  analysis  of  variance  using  PROC 
MIXED  (Littell  et  al.  1998,  Sherfy  and  Kirk- 
patrick 2003)  in  SAS  8.0  (SAS  Institute,  Inc. 
2000).  Fixed  factors  in  the  model  included 
sampling  period,  site,  predation,  and  all  two- 
way  and  three-way  interactions.  Predation 
(defined  as  the  number  of  invertebrates  re- 
moved) was  determined  by  subtracting  the 
values  for  invertebrates  in  the  open  treatment 
from  values  for  invertebrates  in  the  exclosure 


treatment,  for  each  pair.  Values  above  zero  in- 
dicate greater  invertebrate  densities  in  the  ex- 
closures, suggesting  that  shorebirds  removed 
invertebrates  from  outside  the  exclosure  treat- 
ment. The  random  factor  of  plot  (site)  was  in- 
cluded as  an  error  term  in  the  model;  site  rep- 
resents the  main  blocking  factor.  To  avoid 
problems  with  different  initiation  dates  for  the 
plots,  we  used  samples  only  from  early  May, 
late  May,  and  early  June  in  the  PROC  MIXED 
analysis,  which  matched  the  timing  of  shore- 
bird  migration.  We  also  included  shorebird 
density  (log10  [X  + 1 ]-transformed)  and 
change  in  water  level  as  covariates  in  the 
model. 

A separate  analysis  was  performed  for  all 
eight  invertebrate  density  (individuals/m2) 
variables  (oligochaete,  total  chironomid,  small 
chironomid,  large  chironomid,  total  gastropod, 
small  gastropod,  large  gastropod,  total  inver- 
tebrate) and  for  invertebrate  biomass  (g/m2). 
Data  on  large  chironomids  included  many 
zero  values  that  resulted  in  an  infinite  likeli- 
hood error;  therefore,  they  are  not  reported. 
To  meet  assumptions  of  normality,  we  trans- 
formed all  invertebrate  data  (log10  [X  + 1]) 
prior  to  analysis. 

PROC  MIXED  allows  specification  of  the 
covariance  structure  of  the  R matrix  (Littell  et 
al.  2000).  We  used  the  compound-symmetry 
structure,  which  has  constant  variance  and  co- 
variance  between  repeated  measures  and  as- 
sumes that  all  repeated  measures  on  a subject 
(i.e.,  plots)  are  equally  correlated  regardless  of 
their  temporal  relationship.  We  used  linear  re- 
gression to  analyze  correlations  between 
shorebird  density  and  invertebrate  density,  and 
between  shorebird  density  and  biomass  re- 
moved, in  the  nine  plots  for  the  early  May  and 
late  May  sampling  periods.  Statistical  signifi- 
cance was  set  at  P < 0.05  and  all  means  are 
presented  ± SE. 

RESULTS 

We  found  no  difference  in  oligochaete  den- 
sity ( t = 0.25,  df  = 15,  P = 0.81)  or  inver- 
tebrate biomass  ( t = 0.02,  df  = 15,  P = 0.98) 
between  the  exclosure  and  open  treatments 
from  the  initial  samples  taken  just  before  the 
plots  were  established.  Differences  in  chiron- 
omid density  (t  = 2.15,  df  = 15,  P — 0.048) 
and  invertebrate  density  ( t = 2.22,  df  = 15,  P 
= 0.043)  between  the  exclosure  and  open 


156 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


treatment  indicated  a heterogeneous  inverte- 
brate community  at  the  onset  of  the  experi- 
ment. 

We  conducted  116  shorebird  surveys  and 
observed  15  shorebird  and  11  waterfowl  spe- 
cies foraging  inside  the  plots.  We  observed 
838  shorebirds,  89%  of  which  consisted  of 
Least  Sandpiper  ( Calidris  minutillcr,  n = 309), 
Pectoral  Sandpiper  ( Calidris  melanotos’,  n = 
268),  Lesser  Yellowlegs  ( Tringa  flavipes',  n = 
118),  and  Killdeer  ( Charadrius  vociferus;  n = 
49).  We  observed  463  waterfowl,  94%  of 
which  were  Green- winged  Teal  {Anas  crecca\ 
n = 145),  Northern  Shoveler  (A.  clypeata\  n 
= 110),  Blue-winged  Teal  (A.  discors\  n = 
105),  and  Mallard  (A.  platyrhynchos\  n = 76). 
During  the  early-May  to  early- June  sampling 
periods  used  in  the  PROC  MIXED  analysis, 
only  22  waterfowl  and  677  shorebirds  were 
observed  in  the  plots.  Mean  shorebird  density 
across  all  sites  from  late  March  to  early  June 
was  6.3/ha  ± 1.5  (n  = 36);  peak  density  oc- 
curred in  early  May  (12.3/ha  ± 2.8,  n = 9; 
Fig.  2).  The  highest  shorebird  density  (39.8/ 
ha)  occurred  at  Chautauqua  on  20  May. 

We  collected  108  benthic  core  samples,  but 
not  all  of  these  were  used  in  the  analysis  due 
to  the  dynamic  hydrology.  Oligochaete  den- 
sity (all  sites  combined)  from  late  March  to 
early  June  was  15,137.5/m2  ± 3,005.1  in  ex- 
closure treatments  ( n — 36;  Fig.  2)  versus 
11,798.8/m2  ± 3,131.4  {n  = 36)  in  open  treat- 
ments. Chironomid  density  was  2,291.9/m2  ± 
461.1  ( n = 36)  in  exclosure  treatments  and 
2,306.0/m2  ± 573.0  ( n = 36)  in  open  treat- 
ments. Oligochaete  density  peaked  in  late 
May  (22,975.1/m2  ± 8,999.8;  n = 36)  and  chi- 
ronomid density  peaked  in  early  May 
(5.715.5/m2  ± 1,548.5;  n = 36).  The  greatest 
oligochaete  density  observed  in  a single  sam- 
ple occurred  on  20  May  in  an  open  treatment 
at  Emiquon  Wilder  Tract  (88,618.2/m2),  and 
the  greatest  chironomid  density  was  recorded 
on  7 May  from  the  same  site  (16,297.6/m2). 

Oligochaete  density  (F126  = 7.20,  P = 
0.013)  and  large  gastropod  density  {Fx  26  = 
0.21,  P = 0.049)  differed  between  treatments, 
indicating  a significant  predation  effect  (Table 
1);  a significant  predation  X period  X site  in- 
teraction for  oligochaetes  indicated  that  the  ef- 
fect varied  both  spatially  and  temporally  (F426 
= 3.19,  P = 0.029).  The  grand  mean  for  oli- 
gochaete density  was  1.2X  greater  in  the  ex- 


closure than  in  the  open  treatments.  Based  on 
the  total  of  mean  invertebrate  densities  for  all 
the  plots,  shorebirds  removed  18.9%  of  the 
total  invertebrates  from  the  substrate.  Density 
of  chironomids,  total  invertebrate  density,  and 
total  invertebrate  biomass  did  not  differ  be- 
tween treatments. 

Mean  change  in  water  level  (all  sites  com- 
bined) was  10.33  ± 2.23  cm  {n  = 36).  The 
change  in  water  level  influenced  only  oligo- 
chaete density  (F126  = 4.45,  P = 0.045); 
shorebird  density  had  no  influence  on  any  re- 
sponse variables  (Table  1).  Shorebird  density 
was  positively  correlated  with  invertebrate 
biomass  removed  (r2  = 0.64,  P = 0.010)  and 
invertebrate  density  removed  (r2  = 0.39,  P = 
0.071)  in  late  May  (Fig.  3).  Chautauqua  con- 
tributed the  most  to  the  positive  correlation 
between  shorebird  density  and  invertebrate 
biomass  removed. 

DISCUSSION 

Exclosure  design. — A concern  with  exclo- 
sure experiments  in  soft  sediments  is  the  pres- 
ence of  artifacts  produced  by  the  exclosure 
structure  (Vimstein  1978).  Many  of  these  ar- 
tifacts, however,  are  associated  with  marine 
intertidal  systems,  where  the  influences  of  ex- 
closure structure  appear  greater  than  in  non- 
intertidal  systems.  Hulberg  and  Oliver  (1980) 
found  that  exclosures  alter  the  level  of  sedi- 
mentation, which  in  turn  influences  popula- 
tions of  polychaetes.  Their  study  was  per- 
formed on  a wave-exposed  coastal  beach  that 
is  a very  different  environment  from  our  sys- 
tem, which  lacked  wave  perturbations  and  a 
diurnal  tide.  Quammen  (1981)  established  an 
exclosure  design  to  separate  the  effects  of 
multiple  predators  within  a system:  a floating 
exclosure  without  sides  prevented  access  by 
shorebirds  while  allowing  fish  to  enter  the  ex- 
closure during  high  tide.  This  design,  how- 
ever, is  not  as  appropriate  for  a system  without 
tides  and  with  fewer  predators  of  benthic  in- 
vertebrates. Although  common  carp  {Cyprinus 
carpio ) were  observed  in  our  impoundments, 
no  fish  were  observed  foraging  at  the  soil/wa- 
ter interface  where  core  samples  were  taken. 
Even  if  other  predators  of  benthic  inverte- 
brates went  unnoticed,  the  lack  of  sides  on  our 
exclosure  should  have  allowed  normal  access. 
We  also  had  no  evidence  that  the  exclosure 


Hamer  et  al.  • SHOREBIRD  PREDATION  ON  BENTHIC  INVERTEBRATES 


157 


C\J 

70,000  n 

b 

0 

c 

60,000  - 

>> 

50,000  - 

CO 

c 

40,000  - 

0 

0 

30,000  - 

0 

CO 

20,000  - 

-C 

o 

10,000  - 

U) 

b 

0 - 

CM 

c= 

70,000  -| 

■0 

c 

60,000  - 

50,000  - 

c/5 

c 

40,000  - 

0 

0 

0 

30,000  - 

0 

CO 

20,000  - 

0 

o 

o 

10,000  - 

0) 

b 

0 - 

CM 

E 

70,000  n 

0 

c 

60,000  - 

O 10,000  H 

D) 

b o 


♦ Exclosure 

Chautauqua  South  Pool 


I i 


Emiquon  Wilder  Tract 


" T 

n 

A o 

< 

_ 

► 

i- 

I l 

Emiquon  South  Globe 


□ Open  a Shorebird  density  (ind/ha) 

Chautauqua  South  Pool 


5 l 


14  Mar  3 Apr  23  Apr  13  May  2 Jun  22  Jun 


T 25 

CsJ 

14,000  n 

- 20 

E 

0 

12,000  - 

c 

>> 

10,000  - 

- 15 

c/5 

c 

8,000  - 

- 10 

0 

0 

0 

6,000  - 

E 

4,000  - 

- 5 

o 

c 

o 

2,000  - 

- 0 
T 25 

1— 

z 

O 

C\l 

0 - 
14,000  -| 

E 

0 

12,000  - 

- 20 

c 

10,000  - 

- 15 

CO 

c 

8,000  - 

- 10 

0 

0 

0 

6,000  - 

E 

4,000  - 

- 5 

w 

c 

o 

2,000  - 

- 0 
T 25 

JZ 

O 

C\l 

0 - 

14,000  -, 

E 

0 

12,000  - 

- 20 

c 

>, 

10,000  - 

- 15 

CO 

c 

8,000 

- 10 

0 

0 

0 

6,000  - 

E 

o 

4,000  - 

--  5 

c 

o 

2,000  - 

— 0 

Z 

O 

0 - 

Emiquon  Wilder  Tract 


Sr 


Emiquon  South  Globe 


25 
- 20 
- 15 
10 
5 
0 


25 
- 20 

- 15 

- 10 

- 5 
0 

T 25 

- 20 

15 
10 


0 

14  Mar  3 Apr  23  Apr  13  May  2 Jun  22  Jun 


FIG.  2.  Mean  density  of  oligochaetes  and  chironomids  (mean  ± SE)  in  exclosure  and  open  (no  exclosure) 
treatments  at  three  study  sites:  Chautauqua  South  Pool  ( n = 15),  Emiquon  Wilder  Tract  ( n = 12),  and  Emiquon 
South  Globe  ( n = 9)  in  Havana,  Illinois,  from  late  March  to  early  June  2004.  Shorebird  density  (filled  triangles; 
individuals/ha;  n = 36)  shown  without  error  bars  for  clarity. 


represented  either  shelter  or  obstruction  for 
larger  predators,  such  as  crayfish. 

A potential  problem  with  exclosure  exper- 
iments is  the  build-up  of  algae  on  the  cage 
structure  (Vimstein  1978).  Algae  grew  on  sev- 
eral of  our  exclosures,  but  only  where  the 
fence  was  immersed  in  deeper  water  (>10 
cm),  and  algae  were  never  present  at  the  sam- 
pling locations.  If  water  levels  had  dropped 


quickly  at  an  exclosure  with  algal  growth,  the 
physical  nature  of  the  soil/water  interface 
could  have  been  influenced;  however,  this  did 
not  occur  during  our  study. 

Exclosure  structures  are  often  used  as  avian 
roosts,  which  could  influence  the  nutrient  lev- 
els in  the  exclosure  through  the  addition  of 
feces.  Weber  and  Haig  (1997)  reduced  tern 
and  gull  roosting  on  wooden  stakes  by  sharp- 


158 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  1.  Results  of  repeated  measures  mixed-model  analysis  of  variance  for  shorebird  predation  effects 
on  invertebrate  density  (individuals/m2)  and  biomass  (g/m2)  in  mudflats  at  Chautauqua  and  Emiquon  NWR  near 
Havana,  Illinois,  during  early  May,  late  May,  and  early  June,  2004. 


Oligochaete  density 

Total  chironomid 
density 

Small  chironomid 
density 

Effect 

df 

F 

p 

F 

p 

F 

p 

Site 

2,6 

0.05 

0.95 

2.44 

0.17 

1.08 

0.40 

Period 

2,11 

0.89 

0.44 

5.69 

0.020 

3.47 

0.068 

Period  X Site 

4,11 

2.40 

0.11 

1.20 

0.37 

0.63 

0.65 

Predation 

1,26 

7.20 

0.013 

0.08 

0.79 

0.00 

0.97 

Predation  X Site 

2,26 

5.20 

0.013 

0.06 

0.95 

0.22 

0.80 

Predation  X Period 

2,26 

4.47 

0.022 

0.15 

0.86 

0.08 

0.92 

Predation  X Period  X Site 

4,26 

3.19 

0.029 

1.09 

0.38 

0.62 

0.65 

Shorebird  density 

1,26 

0.00 

0.98 

1.20 

0.28 

0.61 

0.44 

Change  in  water  level 

1,26 

4.45 

0.045 

1.09 

0.31 

0.42 

0.52 

a Indicates  mixed-model  error  to  an  infinite  likelihood  from  too  many  zero  values  in  the  data. 


ening  their  ends.  Our  metal  stakes  were  oc- 
casionally used  as  roosts  by  Red-Winged 
Blackbirds  (. Agelaius  phoeniceus ),  and  feces 
at  the  base  of  some  stakes  were  present  in 
small  amounts.  Core  samples,  however,  were 
taken  from  the  middle  of  the  exclosure  and 
the  open  treatments,  thus  avoiding  the  base  of 
stakes  by  at  least  0.5  m. 

Interior  freshwater  wetlands  are  challenging 
environments  for  exclosure  experiments  be- 
cause of  their  unpredictable  hydrology.  The 
zone  of  shorebird  foraging  habitat  constantly 
shifts  as  water  levels  fluctuate.  The  exclosure 
design  commonly  used  in  marine  intertidal 
systems  consists  of  1-m2  treatments,  which  is 
not  appropriate  in  an  interior  system  because 
the  exclosure  would  not  be  long  enough  to 
ensure  that  the  fluctuating  shoreline  foraging 
zone  would  always  remain  within  the  exclo- 
sure. Mitchell  and  Grubaugh  (2005)  used  the 
traditional  square  exclosure  design  and  estab- 
lished 113  plots  in  the  Lower  Mississippi  Al- 
luvial Valley.  The  plots  were  repeatedly  sam- 
pled over  the  course  of  two  summer/fall  mi- 
grations, but  only  the  plots  representing  shore- 
bird  foraging  habitat  (wet  substrate  or  water 
depth  <10  cm)  were  sampled.  As  a result, 
many  plots  were  never  sampled  during  their 
study.  Our  new  design  was  implemented  to 
compensate  for  the  dynamic  hydrology  by  es- 
tablishing each  treatment  as  a linear  transect 
perpendicular  to  the  shoreline.  This  allowed 
repeated  sampling  as  water  levels  changed 
throughout  the  migration  period.  However, 
even  with  this  modified  design,  only  9 of  16 
plots  originally  established  were  used  in  our 


study;  the  water  level  changed  so  dramatically 
in  the  other  7 plots  that  the  shoreline  did  not 
remain  within  the  treatments. 

When  the  height  of  the  exclosure  structures 
was  maintained  at  10  cm  above  the  substrate, 
prevention  of  shorebird  predation  was  accom- 
plished. On  two  occasions,  however,  we  found 
evidence  that  shorebirds  had  been  inside  the 
exclosure  (presence  of  tracks  and  feces).  This 
occurred  when  the  fence  sagged  below  5 cm 
(shorebirds  walked  over  the  fence),  or  was 
above  15  cm  (shorebirds  walked  under  fence). 

We  believe  that  the  only  major  factor  ac- 
counting for  differences  in  the  response  vari- 
ables (e.g.,  invertebrate  density)  between  the 
two  treatments  was  the  exclusion  of  avian 
predators.  We  observed  22  waterfowl  and  677 
shorebirds  inside  plots  during  the  sampling 
period  used  in  the  analysis.  Most  of  the  wa- 
terfowl observed  foraged  in  deeper  water  and 
likely  did  not  influence  the  benthic  inverte- 
brates at  the  shoreline.  Therefore,  most  differ- 
ences between  the  treatments  were  likely  at- 
tributed to  shorebird  predation. 

Exclosure  experiments  continue  to  be  valu- 
able tools  for  studying  predator-prey  interac- 
tions. Future  studies  in  non-intertidal,  soft 
sediments  may  benefit  from  implementation 
of  an  experimental  design  similar  to  the  one 
used  in  this  study.  Researchers  are  well  aware 
of  exclosure  artifacts  in  marine  systems,  but 
little  is  known  about  the  influences  of  exclo- 
sure structures  in  interior  wetlands.  A third 
treatment  (in  addition  to  the  exclosure  and 
open  control)  used  in  many  marine  studies  is 
the  use  of  a “cage  control”  that  has  a top 


Hamer  et  al.  • SHOREBIRD  PREDATION  ON  BENTHIC  INVERTEBRATES 


159 


TABLE  1.  Extended. 

Total  gastropod 
density 

Small  gastropod 
density 

Large  gastropod 
density 

Invertebrate 

density 

Invertebrate 

biomass 

F 

p 

F 

p 

F 

p 

F 

p 

F 

p 

1.23 

0.36 

0.84 

0.48 

1.01 

0.42 

0.42 

0.68 

0.43 

0.67 

3.34 

0.073 

2.18 

0.16 

0.14 

0.87 

0.51 

0.61 

2.79 

0.10 

2.63 

0.092 

3.09 

0.062 

0.66 

0.63 

1.47 

0.28 

1.23 

0.35 

0.26 

0.62 

0.02 

0.90 

4.21 

0.049 

0.32 

0.58 

1.20 

0.28 

6.76 

0.014 

3.32 

0.049 

1.20 

0.31 

1.29 

0.29 

0.01 

0.99 

5.65 

0.024 

1.17 

0.29 

1.77 

0.19 

0.31 

0.74 

2.34 

0.12 

— a 

— 

— 

— 

— 

— 

1.18 

0.34 

2.35 

0.081 

0.17 

0.68 

0.40 

0.53 

0.14 

0.71 

0.17 

0.69 

0.86 

0.36 

0.39 

0.54 

0.34 

0.56 

0.11 

0.75 

0.32 

0.58 

0.26 

0.62 

cover  and  two  sides,  which  is  designed  to 
identify  the  effects  of  the  cage  structure  while 
allowing  normal  predation  to  occur  (fish  or 
crabs  could  enter  the  cage  from  the  two  open 
sides).  The  presence  of  the  exclosure  cover, 
however,  is  likely  to  influence  normal  shore- 
bird  foraging.  Weber  (1994)  attempted  to  ac- 
count for  this  effect  by  establishing  a cage 
control  identical  to  the  exclosure  treatment  but 
without  the  cover,  which  evaluated  the  influ- 
ence of  the  stakes  but  not  the  potential  effects 
of  the  exclosure  cover. 

Predator— prey  interactions. — Our  results 
indicate  that  migrating  shorebirds  did  not  lo- 
cally deplete  invertebrate  populations  at  our 
study  sites,  and  only  oligochaete  density  was 
reduced  by  shorebird  foraging.  We  were  sur- 
prised to  find  that  shorebirds  affected  oligo- 
chaete densities,  but  not  chironomid  densities. 
Chironomids  are  known  to  be  important 
shorebird  prey  throughout  interior  stopover  lo- 
cations (Eldridge  1987,  Helmers  1991,  Skagen 
and  Omen  1996,  Mihue  et  al.  1997),  but  our 
results  suggest  that  shorebirds  did  not  select 
chironomids  over  other  prey.  Oligochaetes  are 
often  the  most  abundant  freshwater  inverte- 
brate in  mudflats  in  the  Mississippi  Alluvial 
Valley  (Elliott-Smith  2003,  Mitchell  and  Gru- 
baugh  2005),  and  they  were  the  most  abun- 
dant prey  at  our  study  sites  (Hamer  2004).  Our 
results  support  Skagen  and  Omen’s  (1996)  as- 
sertion that  dietary  flexibility  allows  shore- 
birds  to  exploit  variable  resources.  The  effect 
of  shorebird  predation  varied  spatially,  and  we 
identified  at  least  four  factors  that  could  have 


influenced  shorebird  predation  pressure  on 
benthic  invertebrates. 

First,  the  energy  demands  of  shorebirds  are 
highly  variable.  Different  intensities  of  shore- 
bird  predation  occurring  seasonally  on  the 
coast  of  Venezuela  were  explained  by  the  dif- 
ferent energy  demands  of  molt,  fat  deposition, 
and  foraging  habitat  (Mercier  and  McNeil 
1994).  Wilson  (1991)  compared  episodic 
shorebird  predation  at  the  Bay  of  Fundy,  Nova 
Scotia,  and  at  Grays  Harbor,  Washington,  and 
found  a significant  reduction  of  major  prey  at 
the  Bay  of  Fundy  but  no  effects  of  predator 
exclusion  at  Grays  Harbor.  The  difference  in 
the  intensity  of  predation  was  explained  by 
differing  migration  strategies  at  these  two 
sites.  Shorebirds  using  Grays  Harbor  tend  to 
migrate  in  short  hops  (Iverson  et  al.  1996, 
Warnock  and  Bishop  1998)  and  do  not  need 
to  accumulate  the  massive  fat  reserves  re- 
quired for  a transoceanic  migration  strategy 
like  shorebirds  departing  from  the  Bay  of  Fun- 
dy. The  short  hop  migration  strategy  of  inte- 
rior shorebirds  (Skagen  and  Knopf  1994b, 
Farmer  and  Wiens  1999)  may  explain  why 
other  studies  of  shorebird  predation  in  the  in- 
terior U.S.  also  show  little  effect  of  predator 
exclusion  on  invertebrate  prey  (Mihue  et  al. 
1997,  Ashley  et  al.  2000,  Mitchell  and  Gru- 
baugh  2005).  Multiple  stops  reduce  the  need 
to  accumulate  large  amounts  of  fuel  at  one 
location. 

Second,  shorebird  territoriality  may  influ- 
ence the  degree  of  episodic  predation  on  in- 
vertebrates. As  shorebird  densities  increase. 


160 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


0 Chautauqua  South  Pool  x Emiquon  Wilder  Tract  + Emiquon  South  Globe 

Early  May 


Late  May 


FIG.  3.  Relationship  between  invertebrate  biomass  removed  (g/m2)  and  density  removed  (individuals/m2) 
versus  shorebird  densities  (individuals/ha)  at  Chautauqua  South  Pool,  Emiquon  South  Globe,  and  Emiquon 
Wilder  Tract  near  Havana,  Illinois,  in  early  May  and  late  May  of  2004.  Values  for  biomass  and  density  removed 
were  calculated  by  subtracting  open  from  exclosure  values.  A value  of  zero  (dashed  line)  represents  equal 
biomass  (or  density)  in  the  exclosure  and  open  treatments.  Values  >0  indicate  greater  biomass  (or  density)  in 
the  exclosure.  Note  difference  in  scales. 


interference  (fighting,  kleptoparasitism,  distur- 
bance) between  territorial  birds  limits  the  de- 
pletion of  resources  (Goss-Custard  1980). 
Duffy  et  al.  (1981)  studied  shorebird  compe- 
tition for  prey  resources  at  a wintering  ground 
in  Peru  and  did  not  find  depletion  of  inverte- 
brate prey;  one  factor  reducing  the  importance 
of  competition  may  have  been  territoriality 
among  the  wintering  birds.  Migrant  shorebirds 
at  our  stopover  location  are  mostly  nonterri- 
torial (Hamer  2004);  thus,  territorial  interac- 


tions likely  did  not  play  a role  in  the  shore  - 
bird/prey  dynamics  at  our  study  sites. 

Third,  shorebird  predation  pressure  is  great- 
er in  locations  with  greater  densities  of  for- 
aging birds.  Shorebird  densities  observed  dur- 
ing our  study  averaged  6.3/ha,  peaking  at 
39.8/ha.  Coastal  flyways  receive  much  greater 
concentrations  of  shorebirds  where  densities 
can  approach  100/ha  (in  coastal  South  Caro- 
lina; Weber  and  Haig  1997)  to  4,500/ha  (in 
coastal  Venezuela;  Mercier  and  McNeil  1994). 


Hamer  et  al.  • SHOREBIRD  PREDATION  ON  BENTHIC  INVERTEBRATES 


161 


The  dispersed  migration  through  interior  hab- 
itats results  in  lower  shorebird  densities  and 
possibly  reduces  predation  pressure. 

Finally,  the  dynamic  water  levels  recorded 
during  our  study  may  have  been  an  additional 
factor  that  reduced  the  effect  of  shorebird  pre- 
dation on  benthic  invertebrates.  Water  levels 
fluctuated  an  average  of  8.9  cm  during  2-week 
intervals.  Gradual  drawdown  or  flooding  con- 
tinuously shifts  the  location  of  foraging  hab- 
itat and  exposes  new  invertebrate  prey  (Run- 
dle  and  Fredrickson  1981).  Even  though  man- 
agers at  Chautauqua’s  South  Pool  attempted 
to  maintain  a stable  water  level  over  the 
course  of  the  spring,  the  average  fluctuation 
over  each  2-week  period  was  7.6  cm.  Much 
of  this  variation  can  be  explained  by  wind- 
driven  seiches  (wind  fetch),  which  can  expose 
previously  unexploited  foraging  habitat  in 
large,  shallow  wetlands  (Laubhan  and  Fred- 
rickson 1993).  Without  this  phenomenon, 
shorebird  reduction  of  invertebrates  at  Chau- 
tauqua may  have  been  greater. 

Because  shorebirds  are  size-selective  when 
preying  on  invertebrates,  they  can  influence 
the  invertebrate  community  structure  in  soft 
sediments  (Peterson  1979,  Kent  and  Day 
1983,  Wilson  1989).  Shorebird  predation  on 
marine  polychaetes  often  targets  large  (adult) 
individuals,  which  can  lead  to  increased  re- 
cruitment of  juveniles  and  increased  densities 
of  smaller  invertebrates.  As  a consequence, 
exclosure  experiments  in  which  only  prey 
densities  are  measured  may  fail  to  account  for 
the  interactions  of  size-class  predation  and 
size-dependent  competition.  Our  results,  how- 
ever, do  not  suggest  that  such  episodic  shore- 
bird  predation  influenced  the  invertebrate 
community  structure  in  our  study.  There  was 
no  evidence  of  size-selection  of  chironomids, 
but  the  mean  density  of  large  gastropods  was 
more  than  seven  times  greater  in  the  exclosure 
than  the  open  treatment  (106.1/m2  versus  14.1/ 
m2,  respectively).  Thus,  it  seems  likely  that 
shorebirds  selected  large  gastropods,  which 
has  been  observed  elsewhere  in  the  Mississip- 
pi Alluvial  Valley  (Brooks  1967,  Rundle 
1982). 

Competition  for  prey  resources  at  migration 
stopover  locations  may  result  when  early  mi- 
grants deplete  prey  resources  and  reduce  the 
successful  foraging  rate  of  later-arriving 
shorebirds,  thus  increasing  the  length  of  stay 


for  later  arrivals  (Wilson  1991).  Although  this 
occurs  at  some  locations  (Schneider  and  Har- 
rington 1981),  later  migrants  at  our  study  site 
were  not  likely  disadvantaged  by  reductions 
in  prey  density  by  early  migrants  because  the 
dynamic  hydrology  constantly  exposed  pre- 
viously unexploited  food  resources. 

Our  results  suggest  that  migrating  shore- 
birds  along  the  Illinois  River  may  have  re- 
duced oligochaetes  and  larger  gastropods. 
Flexible  and  opportunistic  foraging  strategies 
are  beneficial  to  shorebirds  facing  the  unpre- 
dictable nature  of  interior  flyways.  The  re- 
moval of  oligochaetes,  the  most  abundant  in- 
vertebrates at  our  study  sites,  suggests  that 
shorebirds  fed  opportunistically  on  the  most 
available  prey.  The  dynamic  hydrology,  and 
the  resulting  continuously  renewing  availabil- 
ity of  invertebrate  prey,  likely  offer  sufficient 
invertebrate  resources  for  migrating  shore- 
birds  in  the  Illinois  River  valley. 

ACKNOWLEDGMENTS 

This  research,  conducted  as  part  of  a master’s  thesis, 
was  funded  by  the  Illinois  Natural  History  Survey,  the 
Illinois  Department  of  Natural  Resources  Wildlife 
Preservation  Fund,  The  Nature  Conservancy,  and  the 
Champaign  County  Audubon  Society.  We  thank  the 
staff  at  the  Forbes  Biological  Field  Station,  D.  J.  Sou- 
cek,  and  J.  M.  Levengood  for  their  assistance  with  the 
project;  M.  J.  Wetzel  and  R.  E.  Dewalt  for  aid  in  iden- 
tification of  invertebrates;  J.  Dassow,  B.  J.  O’Neal,  A. 
Bartlett,  and  B.  T.  Kapusta  for  their  assistance  in  the 
field  and  lab;  and  the  staff  at  the  Illinois  River  National 
Wildlife  and  Fish  Refuge  and  Rice  Lake  State  Fish  and 
Wildlife  Area.  Comments  on  earlier  drafts  by  G.  O. 
Batzli  and  three  anonymous  reviewers  greatly  im- 
proved this  paper. 

LITERATURE  CITED 

Anderson,  J.  T.  and  L.  M.  Smith.  1998.  Protein  and 
energy  production  in  playas:  implications  for  mi- 
gratory bird  management.  Wetlands  18:437-446. 
Ashley,  M.  C.,  J.  A.  Robinson,  L.  W.  Oring,  and  G. 
A.  Vinyard.  2000.  Dipterian  standing  stock  bio- 
mass and  effects  of  aquatic  predation  at  a con- 
structed wetland.  Wetlands  20:84-90. 

Bailey,  S.  D.  2003.  Lake  Chautauqua  and  counting 
shorebirds.  Meadowlark:  a Journal  of  Illinois 
Birds  12:54. 

Brooks,  W.  S.  1967.  Food  and  feeding  habits  of  au- 
tumn migrant  shorebirds  at  a small  midwestern 
pond.  Wilson  Bulletin  79:307-315. 

Brown,  S.,  C.  Hickey,  B.  Harrington,  and  R.  Gill. 
2001.  United  States  Shorebird  Conservation  Plan, 
2nd  ed.  Manomet  Center  for  Conservation  Sci- 
ences, Manomet,  Massachusetts. 


162 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Cummins,  K.  W.  and  J.  C.  Wuycheck.  1971.  Caloric 
equivalents  for  investigations  in  ecological  ener- 
getics. Komitee  Fur  Limnologische  Methoden, 
no.  18.  E.  Schweizerbart,  Stuttgart,  Germany. 

Dahl,  T.  E.  2000.  Status  and  trends  of  wetlands  in  the 
conterminous  United  States  1886  to  1997.  U.S. 
Department  of  the  Interior,  Fish  and  Wildlife  Ser- 
vice, Washington,  D.C. 

Duffy,  D.  C.,  N.  Atkins,  and  D.  C.  Schneider.  1981. 
Do  shorebirds  compete  on  their  wintering 
grounds?  Auk  98:215-229. 

Eldridge,  J.  L.  1987.  Ecology  of  migrant  sandpipers 
in  mixed-species  foraging  flocks.  Ph.D.  disserta- 
tion, University  of  Minnesota,  Minneapolis. 

Elliott-Smith,  E.  S.  2003.  Mudflat  subsidence  in  a 
man  made  reservoir:  the  importance  of  topogra- 
phy to  migrant  shorebirds.  M.Sc.  thesis.  Southern 
Illinois  University,  Carbondale. 

Evans,  P.  R.,  D.  M.  Knights,  and  M.  W.  Pienkowski. 
1979.  Short-term  effects  of  reclamation  of  part  of 
Seal  Sands,  Teesmouth,  on  wintering  waders  and 
Shelduck.  Oecologia  41:183-206. 

Farmer,  A.  H.  and  J.  A.  Wiens.  1999.  Models  and 
reality:  time-energy  trade-offs  in  Pectoral  Sand- 
piper ( Calidris  melanotos)  migration.  Ecology  80: 
2566-2580. 

Goss-Custard,  J.  D.  1980.  Competition  for  food  and 
interference  among  waders.  Ardea  68:31-52. 

Hamer,  G.  L.  2004.  Migrant  shorebird  ecology  in  the 
Illinois  River  valley.  M.Sc.  thesis,  University  of 
Illinois,  Urbana-Champaign. 

Harrington.  B.  A.,  S.  C.  Brown,  J.  Corven,  and  J. 
Bart.  2002.  Collaborative  approaches  to  the  evo- 
lution of  migration  and  the  development  of  sci- 
ence-based conservation  in  shorebirds.  Auk  199: 
914-921. 

Harrington,  B.  and  E.  Perry.  1995.  Important  shore- 
bird  staging  sites  meeting  Western  Hemisphere 
Shorebird  Reserve  Network  criteria  in  the  United 
States.  U.S.  Fish  and  Wildlife  Service,  Washing- 
ton, D.C. 

Helmers,  D.  L.  1991.  Habitat  use  by  migrant  shore- 
birds  and  invertebrate  availability  in  a managed 
wetland  complex.  M.Sc.  thesis.  University  of  Mis- 
souri, Columbia. 

Hulberg,  L.  W.  and  J.  S.  Oliver.  1980.  Caging  ma- 
nipulations in  marine  soft-bottom  communities: 
importance  of  animal  interactions  or  sedimentary 
habitat  modifications.  Canadian  Journal  of  Fish- 
eries and  Aquatic  Sciences  37:1 130-1 139. 

Iverson,  G.  C.,  S.  E.  Warnock,  R.  W.  Butler,  M.  A. 
Bishop,  and  N.  Warnock.  1996.  Spring  migration 
of  Western  Sandpipers  along  the  Pacific  coast  of 
North  America:  a telemetry  study.  Condor  98:10- 
21. 

Kent,  A.  C.  and  R.  W.  Day.  1983.  Population  dynam- 
ics of  an  infaunal  polychaete:  the  effect  of  pred- 
ators and  an  adult-recruit  interaction.  Journal  of 
Experimental  Marine  Biology  and  Ecology  73: 
185-203. 

Laubhan,  M.  K.  and  L.  H.  Fredrickson.  1993.  Inte- 


grated wetland  management:  concepts  and  oppor- 
tunities. Transactions  of  the  58th  North  American 
Wildlife  and  Natural  Resources  Conference  58: 
323-333. 

Littell,  R.  C.,  P.  R.  Henry,  and  C.  B.  Ammerman. 
1998.  Statistical  analysis  of  repeated  measures 
data  using  SAS  procedures.  Journal  of  Animal 
Science  76:1216-1231. 

Littell,  R.  C.,  J.  Pendergast,  and  R.  Natarajan. 
2000.  Modeling  covariance  structure  in  the  anal- 
ysis of  repeated  measures  data.  Statistics  in  Med- 
icine 19:1793-1819. 

Mercier,  F.  and  R.  McNeil.  1994.  Seasonal  variations 
in  intertidal  density  of  invertebrate  prey  in  a trop- 
ical lagoon  and  effects  of  shorebird  predation.  Ca- 
nadian Journal  of  Zoology  72:1755-1763. 

Merritt,  R.  W.  and  K.  W.  Cummins.  1996.  An  intro- 
duction to  the  aquatic  insects  of  North  America, 
3rd  ed.  Kendall-Hunt,  Dubuque,  Iowa. 

Mihue,  J.  R.,  C.  H.  Trost,  and  T.  B.  Mihue.  1997. 
Shorebird  predation  on  benthic  macroinverte- 
brates in  an  irrigation  reservoir.  Great  Basin  Nat- 
uralist 57:245-252. 

Mitchell,  D.  W.  and  J.  W.  Grubaugh.  2005.  Impacts 
of  shorebirds  on  macroinvertebrates  in  the  lower 
Mississippi  Alluvial  Valley.  American  Midland 
Naturalist  154:188-200. 

Moreira,  F.  1997.  The  importance  of  shorebirds  to  en- 
ergy fluxes  in  a food  web  of  a south  European 
estuary.  Estuarine,  Coastal  and  Shelf  Sciences  44: 
67-78. 

Morrison,  R.  G.,  R.  E.  Gill,  B.  A.  Harrington,  S. 
Skagen,  G.  W.  Page,  C.  L.  Gratto-Trevor,  and 
S.  M.  Haig.  2001.  Estimates  of  shorebird  popu- 
lations in  North  America.  Occasional  Paper,  no. 
104.  Canadian  Wildlife  Service,  Ottawa,  Ontario. 

Myers,  J.  P,  I.  G.  Morrison,  P.  Z.  Antas,  B.  A.  Har- 
rington, T.  E.  Lovejoy,  M.  S.  Sallaberry,  S.  E. 
Seener,  and  A.  Tarak.  1987.  Conservation  strat- 
egy for  migratory  species.  American  Scientist  75: 
19-26. 

Pennak,  R.  W.  1989.  Fresh-water  invertebrates  of  the 
United  States:  protozoa  to  mollusca,  3rd  ed.  Wiley 
and  Sons,  New  York. 

Peterson,  C.  H.  1979.  Predation,  competitive  exclu- 
sion, and  diversity  in  the  soft-sediment  benthic 
communities  of  estuaries  and  lagoons.  Pages  233- 
264  in  Ecological  processes  in  coastal  and  marine 
systems  (R.  J.  Livingston,  Ed.).  Plenum  Press, 
New  York. 

Piersma,  T.  1987.  Hop,  skip,  or  jump?  Constraints  on 
migration  of  arctic  waders  by  feeding,  fattening, 
and  flight  speed.  Limosa  60:185-194. 

Quammen,  M.  L.  1981.  Use  of  exclosures  in  studies 
of  predation  by  shorebirds  on  intertidal  mudflats. 
Auk  98:812-817. 

Raffaelli,  D.  and  H.  Milne.  1987.  An  experimental 
investigation  of  effects  of  shorebird  and  flatfish 
predation  on  estuarine  invertebrates.  Coastal  and 
Shelf  Science  24:1-13. 

Rundle,  W.  D.  1982.  A case  for  esophageal  analysis 


Hamer  et  al.  • SHOREBIRD  PREDATION  ON  BENTHIC  INVERTEBRATES 


163 


in  shorebird  food  studies.  Journal  of  Field  Orni- 
thology 53:249-257. 

Rundle,  W.  D.  and  L.  H.  Fredrickson.  1981.  Man- 
aging seasonally  flooded  impoundments  for  mi- 
grant rails  and  shorebirds.  Wildlife  Society  Bul- 
letin 9:80-87. 

Safran,  R.  J.,  C.  R.  Isola,  M.  A.  Colwell,  and  O. 
E.  Williams.  1997.  Benthic  invertebrates  at  for- 
aging locations  of  nine  waterbird  species  in  man- 
aged wetlands  of  the  northern  San  Joaquin  Valley, 
California.  Wetlands  17:407-415. 

SAS  Institute,  Inc.  2000.  SAS/STAT,  ver.  8.0.  SAS 
Institute,  Inc.,  Cary,  North  Carolina. 

Schneider,  D.  C.  1978.  Equalisation  of  prey  numbers 
by  migratory  shorebirds.  Nature  271:353-354. 

Schneider,  D.  C.  and  B.  A.  Harrington.  1981.  Tim- 
ing of  shorebird  migration  in  relation  to  prey  de- 
pletion. Auk  98:801-811. 

Sherfy,  M.  H.  and  R.  L.  Kirkpatrick.  2003.  Inver- 
tebrate response  to  Snow  Goose  herbivory  on 
moist-soil  vegetation.  Wetlands  23:236—249. 

Sherfy,  M.  H.,  R.  L.  Kirkpatrick,  and  K.  D.  Richkus. 
2000.  Benthos  core  sampling  and  chironomid  ver- 
tical distribution:  implications  for  assessing  shore- 
bird  food  availability.  Wildlife  Society  Bulletin 
28:124-130. 

Skagen,  S.  K.  and  F.  L.  Knopf.  1993.  Toward  conser- 
vation of  midcontinental  shorebird  migrations. 
Conservation  Biology  7:533-541. 

Skagen,  S.  K.  and  F.  L.  Knopf.  1994a.  Migrating 
shorebirds  and  habitat  dynamics  at  a prairie  wet- 
land complex.  Wilson  Bulletin  106:91-105. 

Skagen,  S.  K.  and  F.  L.  Knopf.  1994b.  Residency  pat- 
terns of  migrating  sandpipers  at  a midcontinental 
stopover.  Condor  96:949-958. 

Skagen,  S.  K.  and  H.  D.  Oman.  1996.  Dietary  flexi- 
bility of  shorebirds  in  the  Western  Hemisphere. 
Canadian  Field-Naturalist  110:419-444. 

Sutherland,  W.  J.  and  J.  D.  Goss-Custard.  1991. 
Predicting  the  consequence  of  habitat  loss  on 


shorebird  populations.  Acta  XX  Congressus  Inter- 
nationalis  Ornithologici  4:2199-2207. 

Swanson,  G.  A.  1978.  A simple  lightweight  core  sam- 
pler for  quantifying  waterfowl  foods.  Journal 
Wildlife  Management  42:426-428. 

Swennen,  C.  1990.  Oystercatchers  feeding  on  giant 
bloody  cockles  on  the  Banc  D’  Arguin,  Maurita- 
nia. Ardea  78:53-62. 

Szekely,  T.  and  Z.  Bamberger.  1992.  Predation  of 
waders  (Charadii)  on  prey  populations:  an  exclo- 
sure experiment.  Journal  of  Animal  Ecology  61: 
447-456. 

Virnstein,  R.  W.  1978.  Predator  caging  experiments 
in  soft  sediments:  caution  advised.  Pages  261-274 
in  Estuarine  interactions  (M.  L.  Wiley,  Ed.).  Ac- 
ademic Press,  New  York. 

Warnock,  N.  and  M.  A.  Bishop.  1998.  Spring  stop- 
over ecology  of  migrant  Western  Sandpipers. 
Condor  100:456-467. 

Weber,  L.  M.  1994.  Foraging  ecology  and  conserva- 
tion of  shorebirds  in  South  Carolina  coastal  wet- 
lands. Ph.D.  dissertation,  Clemson  University, 
Clemson,  South  Carolina. 

Weber,  L.  M.  and  S.  M.  Haig.  1997.  Shorebird  diet 
and  size  selection  of  nereid  polychaetes  in  South 
Carolina  coastal  diked  wetlands.  Journal  of  Field 
Ornithology  68:358-366. 

White,  D.  H.  and  C.  A.  Mitchell.  1990.  Body  mass 
and  lipid  content  of  shorebirds  overwintering  on 
the  south  Texas  coast.  Journal  of  Field  Ornithol- 
ogy 61:445-452. 

Wilson,  W.  H.,  Jr.  1989.  Predation  and  the  mediation 
of  intraspecific  competition  in  an  infaunal  com- 
munity in  the  Bay  of  Fundy.  Journal  of  Experi- 
mental Marine  Biology  and  Ecology  1 32:22 1 — 
245. 

Wilson,  W.  H.,  Jr.  1991.  The  foraging  ecology  of  mi- 
gratory shorebirds  in  marine  soft-sediment  com- 
munities: the  effects  of  episodic  predation  on  prey 
populations.  American  Zoologist  31:840-848. 


The  Wilson  Journal  of  Ornithology  1 18(2):  164— 172,  2006 


COMPOSITION  AND  TIMING  OF  POSTBREEDING  MULTISPECIES 
FEEDING  FLOCKS  OF  BOREAL  FOREST  PASSERINES  IN 

WESTERN  CANADA 

KEITH  A.  HOBSON1 2 AND  STEVE  VAN  WILGENBURG1 2 


ABSTRACT. — The  aggregation  of  nonbreeding  insectivorous  songbirds  into  multispecies  feeding  flocks  dur- 
ing migration  and  on  their  wintering  grounds  is  a well-known  and  important  aspect  of  their  ecology.  The 
establishment  of  multispecies  feeding  flocks  on  the  temperate  breeding  grounds  of  North  American  Neotropical 
migrants,  however,  remains  poorly  known  or  understood.  To  address  this  gap,  we  investigated  the  composition 
and  timing  of  flocking  behavior  among  several  species  occurring  in  the  southern  boreal  mixed-wood  forest  of 
western  Canada.  Of  67  species  observed  in  216  flocks,  the  most  abundant  were  Tennessee  Warbler  ( Vermivora 
peregrina ) and  several  resident  species:  Black-capped  Chickadee  ( Poecile  atricapillus ),  Red-breasted  Nuthatch 
(Sitta  canadensis ),  and  Boreal  Chickadee  ( Poecile  hudsonica).  Consistent  with  previous  work  on  Eurasian  boreal 
species,  residents  appeared  to  play  a pivotal  role  in  flock  occurrence  and  cohesion.  Flocking  tended  to  begin  in 
late  June,  and  flock  sizes  increased  throughout  the  summer.  This  suggests  that  unsuccessful  breeders,  early 
breeders,  and  early  migrants  are  the  first  to  join  flocks,  whereas  later-nesting  species  may  delay  joining  flocks 
until  after  their  young  fledge.  We  also  investigated  the  propensity  of  several  species  to  display  flocking  behavior 
in  areas  with  and  without  a superabundant  food  source — the  spruce  budworm  ( Choristoneura  fumiferana).  These 
data  provided  some  support  for  the  hypothesis  that  flocking  facilitates  foraging,  as  species  tended  to  flock  in 
areas  where  food  abundance  was  lower.  Received  24  January  2005,  accepted  14  December  2005. 


The  aggregation  of  individual  insectivorous 
songbirds  into  multispecies  feeding  flocks  is  a 
phenomenon  that  has  been  noted  for  some 
time  (e.g.,  Newton  1896,  Sharpe  1905)  and 
has  received  considerable  attention  recently 
(Hutto  1994,  Latta  and  Wunderle  1996, 
Monkkonen  et  al.  1996).  Such  flocking  be- 
havior is  interesting  from  several  perspectives, 
and  a number  of  hypotheses  have  been  put 
forth  to  explain  the  evolution  of  such  inter- 
specific associations,  primarily  focusing  on 
the  avoidance  of  predation  (Pulliam  1973,  El- 
gar 1989)  and  the  facilitation  of  food  finding 
(Morse  1970,  1977).  The  establishment  and 
maintenance  of  hierarchies  within  flocks  and 
the  role  of  interspecific  competition  in  struc- 
turing these  aggregations  also  are  areas  of 
considerable  interest  (Munn  and  Terborgh 
1979,  Powell  1979,  Hutto  1994). 

To  date,  research  on  multispecies  feeding 
flocks  involving  forest  passerines  has  focused 
primarily  on  the  wintering  grounds,  particu- 
larly in  the  Neotropics  (reviewed  by  Monk- 
konen et  al.  1996;  see  also  Buskirk  et  al.  1972; 
Hutto  1987,  1994;  Ewert  and  Askins  1991; 


1 Prairie  and  Northern  Wildlife  Research  Centre,  Ca- 
nadian Wildlife  Service,  1 15  Perimeter  Rd.,  Saskatoon, 
SK  S7N  0X4,  Canada. 

2 Corresponding  author;  e-mail: 

Keith. Hobson@ec.gc.ca 


Latta  and  Wunderle  1996).  This  is  in  spite  of 
the  fact  that  mixed-species  flocks  of  North 
American  songbirds  are  conspicuous  on  their 
breeding  grounds  or  during  the  early  post- 
breeding migration  period.  In  the  continental 
United  States,  Morse  (1970)  was  the  first  to 
conduct  a quantitative  study  on  ecological  as- 
pects of  mixed-species  foraging  flocks  of 
songbirds  during  late  summer  through  winter, 
but  virtually  no  studies  of  foraging  flocks  have 
been  conducted  on  North  American  breeding 
grounds  since  then.  Research  by  Monkkonen 
et  al.  (1996)  on  mixed-species  foraging  aggre- 
gations and  heterospecific  attraction  in  boreal 
bird  communities  in  Finland  represents  an  im- 
portant advance  in  the  study  of  flocking  be- 
havior among  temperate-breeding  songbirds. 
These  authors  determined  that  feeding  asso- 
ciations occurred  during  the  breeding  season 
and  that  titmice  ( Parus  spp.)  seemed  to  play 
a focal  role  in  the  occurrence  of  these  flocks. 
They  also  suggested  that  flocking  might  pro- 
duce variation  in  species  numbers,  local  abun- 
dances, and  spatial  patterns,  both  within  and 
between  communities  in  boreal  forests.  To  ad- 
dress the  paucity  of  information  on  multispe- 
cies aggregations  of  boreal  forest  songbirds  on 
their  breeding  grounds  in  North  America,  we 
investigated  the  composition  and  timing  of 
flocking  behavior  among  several  species. 


164 


Hobson  and  Van  Wilgenburg  • BOREAL  MULTISPECIES  FEEDING  FLOCKS 


165 


Within  the  boreal  forest  of  North  America, 
the  southern  boreal  mixed-wood  ecozone  sup- 
ports one  of  the  most  diverse  breeding  bird 
assemblages  of  any  forest  type  in  the  conti- 
nent (Robbins  et  al.  1986,  Price  et  al.  1995). 
Most  of  the  breeding  birds  are  Neotropical  or 
short-distance  migrants.  In  addition,  much  of 
the  forest  occurring  in  this  region  is  contigu- 
ous primary  forest  that  has  not  yet  been  al- 
tered by  logging  (but  see  Stelfox  1995).  This 
is  in  contrast  to  Scandinavian  boreal  forest, 
which  has  less  complex  avian  communities 
with  fewer  migrants  (see  Schmiegelow  and 
Monkkonen  2002).  The  first  objective  of  our 
study  was  to  establish  the  timing  and  impor- 
tance of  flocking  throughout  the  breeding  and 
immediate  postbreeding  periods.  To  accom- 
plish this,  we  aimed  to  document  occurrences 
of  flocking  in  relation  to  overall  breeding  phe- 
nology of  the  avian  community.  Second,  we 
sought  to  document  flock  composition  and  ev- 
idence of  associations  among  flock  members; 
specifically,  we  were  interested  in  identifying 
species  integral  to  flock  formation  and  wheth- 
er species  associations  were  random  or  based 
on  foraging  guilds  or  taxonomic  affinities.  Our 
third  objective  was  to  determine  whether 
flocking  was  associated  with  areas  where  food 
resources  were  superabundant — areas  with  in- 
festations of  spruce  bud  worm  ( Choristoneura 
fumiferana).  If  flocking  was  a response  to  in- 
creased foraging  efficiency,  we  expected  that 
species  occurring  in  areas  where  food  is  su- 
perabundant might  be  less  likely  to  participate 
in  foraging  flocks. 

METHODS 

Study  area  and  field  observations. — The 
study  was  conducted  from  mid-May  to  mid- 
September  1992-1996,  in  the  southern  boreal 
mixed-wood  forest  of  west-central  Saskatch- 
ewan, Canada,  primarily  in  the  vicinity  of 
Prince  Albert  National  Park  (53°  35'  N,  106° 
00'  W),  a 387,500-ha  block  of  contiguous  pri- 
mary forest.  The  dominant  tree  species  in  this 
region  were  trembling  aspen  ( Populus  tremu- 
loides ),  white  spruce  ( Picea  glauca ),  jack  pine 
( Pinus  banksiana ),  black  spruce  ( Picea  mari- 
ana ),  and  paper  birch  ( Betula  papyrifera).  The 
majority  of  the  study  area  burned  in  1919; 
therefore,  it  is  largely  uniform  in  age  structure 
(Weir  and  Johnson  1998).  Forest  west  of  the 
park  had  undergone  an  outbreak  of  spruce 


budworm,  providing  us  with  an  opportunity  to 
sample  similar  forest  habitats  in  budworm-in- 
fested  (hereafter  “infested”)  and  uninfested 
(i.e.,  no  budworm  infestation)  areas.  By  1993, 
approximately  30,000  ha  were  infested  and 
classified  as  moderately  to  severely  defoliated 
(i.e.,  >50%  defoliation;  Saskatchewan  Natural 
Resources  Forest  and  Lands  Branch  1993). 
From  mid-May  to  mid-July  in  >30  mature 
forest  stands,  we  opportunistically  recorded 
all  feeding  flocks  encountered.  For  the  pur- 
poses of  this  study,  we  defined  a flock  as  any 
assemblage  of  individuals  composed  of  more 
than  one  species  clearly  moving  together.  We 
did  not  include  single  family  groups;  however, 
we  did  include  amalgamations  of  single-spe- 
cies flocks  composed  of  more  than  one  family 
group.  After  mid-July  and  until  mid-August, 
when  flocks  became  more  common,  one  ob- 
server spent  at  least  6 hr  per  day  searching  for 
flocks  along  forest  trails  and  riparian  edges. 
Thereafter,  observations  were  again  made  op- 
portunistically during  the  course  of  other 
fieldwork. 

In  total,  we  observed  215  flocks,  distributed 
relatively  equally  amongst  infested  ( n = 102) 
and  uninfested  sites  ( n = 113).  Upon  encoun- 
tering a flock,  we  followed  it  for  -20—30  min 
while  counting  or  estimating  the  number  of 
individuals  of  each  species  and  noting  the 
presence  of  family  groups — as  evidenced  by 
begging  or  feeding  of  young.  Flock  height, 
forest  type,  and  location  were  also  recorded. 

As  part  of  another  study  in  the  same  area, 
from  31  May  to  1 July  1992-1996,  we  also 
conducted  395  point  counts  across  spruce-  or 
aspen-dominated  mixed-wood  stand  types 
(Hobson  and  Bayne  2000).  Points  were  rough- 
ly equally  distributed  among  infested  ( n = 
204)  and  uninfested  (n  = 191)  forest.  Six 
highly  skilled  observers  with  at  least  6 years 
experience  conducted  10-min  point  counts 
from  04:00  to  08:30  CST,  during  which  ob- 
servers recorded  all  birds  heard  or  seen  within 
an  unlimited-distance  radius.  Two  visits  were 
made  to  each  station  during  the  survey  period, 
once  prior  to  15  June,  and  once  after  15  June. 
Relative  abundance  estimates  were  based  on 
the  maximum  count  for  a species  during  these 
two  visits.  Because  these  data  were  collected 
prior  to  the  routine  use  of  methods  to  correct 
for  detectability  biases,  we  do  not  have  de- 
tectability functions  to  correct  these  data; 


166 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


therefore,  estimates  of  flocking  propensity 
should  be  interpreted  with  caution.  However, 
this  dataset  allowed  us  to  quantify  relative 
abundances  as  determined  by  point  counts  and 
contrast  them  with  the  relative  occurrences  of 
species  in  mixed-species  flocks  at  the  regional 
scale. 

Statistical  analyses. — Basic  descriptive  sta- 
tistics were  used  to  examine  the  magnitude 
and  frequency  of  species’  occurrences  in 
flocks;  values  are  reported  as  means  ± 1 SD. 
To  examine  the  probability  of  species  co-oc- 
curring in  flocks,  we  conducted  tests  of  in- 
dependence using  G-tests  with  Williams’  cor- 
rection for  continuity.  Using  2X2  contingen- 
cy tables,  we  contrasted  the  number  of  flocks 
(frequency  of  occurrence)  in  which  species 
co-occurred  and  the  number  of  flocks  in  which 
one  species  occurred  but  the  others  did  not. 
We  used  Fisher’s  exact  test  of  independence 
when  expected  frequencies  were  <5  (Zar 
1996).  To  evaluate  flocking  propensity  (the 
occurrence  of  a species  more  or  less  frequent- 
ly than  expected  due  to  chance)  of  the  most 
abundant  species  in  both  infested  and  unin- 
fested forests,  we  used  a 2 X 2 Yate’s-cor- 
rected  chi-square  test  for  independence;  this 
test  contrasted  a given  species’  abundance  in 
flocks  and  on  point  counts  with  the  total  abun- 
dance of  all  species  in  flocks  and  on  point 
counts. 

To  compare  estimated  flock  size  in  infested 
versus  uninfested  forests,  we  used  analysis  of 
covariance  (ANCOVA)  on  rank-transformed 
data  and  included  Julian  date  as  the  covariate. 
Shannon  Evenness,  species  richness,  and 
Simpson’s,  McIntosh,  and  Shannon  diversity 
indices  were  used  to  evaluate  flock  composi- 
tion (Magurran  1988).  Flock-size  estimates 
were  log-transformed  and  we  used  linear  re- 
gression to  analyze  change  in  flock  size 
throughout  the  season.  Finally,  temporal  pat- 
terns of  flock  composition  were  depicted 
graphically  and  Mann-Whitney  G-tests  were 
employed  to  test  for  significance  of  change 
through  time  and  between  infested  and  unin- 
fested areas.  We  set  statistical  significance  at 
a < 0.05;  however,  Bonferroni  adjustments 
were  used  for  multiple  comparisons,  resulting 
in  species  co-occurrence  being  assessed  at  a 
< 0.0004  (0.05/120  pairwise  comparisons) 
and  flocking  propensity  being  assessed  at  a < 


0.001  (0.05/52  tests).  Scientific  names  of  all 
bird  species  are  given  in  Table  1. 

RESULTS 

We  recorded  5,753  individuals  representing 
67  species  in  216  flocks  (Table  1).  The  mean 
number  of  species  per  flock  was  6.6  ± 3.3  and 
the  mean  number  of  individuals  per  flock  was 
41.1  ± 60.4.  The  six  species  occurring  most 
frequently  in  flocks  included  a long-distance 
migrant  (Tennessee  Warbler),  two  short-dis- 
tance migrants  (Yellow-rumped  Warbler, 
Chipping  Sparrow),  and  three  resident  species 
(Black-capped  Chickadee,  Red-breasted  Nut- 
hatch, Boreal  Chickadee). 

We  evaluated  the  probability  of  the  15  most 
commonly  observed  (i.e.,  number  of  flock  oc- 
currences >30)  species  co-occurring  in  flocks. 
Of  the  120  possible  pair-wise  comparisons,  we 
found  only  5 significant  (positive  or  negative) 
associations.  Black-capped  Chickadee  co-oc- 
curred with  Bay-breasted  Warbler  1.6  times 
less  frequently  than  expected  by  chance  (G  = 
15.03,  P < 0.001),  and  there  was  also  a neg- 
ative association  between  Boreal  Chickadee 
and  American  Redstart  (2.6  times;  G = 10.66, 
P < 0.001).  Red-breasted  Nuthatch  associated 
positively  with  Brown  Creeper  1 .4  times  more 
frequently  than  expected  by  chance  (G  = 
15.44,  P < 0.001).  Among  migrants,  Ameri- 
can Redstart  was  positively  associated  with 
Red-eyed  Vireo  1 .9  times  more  frequently 
than  expected  by  chance  (G  = 18.06,  P < 
0.001)  and  with  Bay-breasted  Warbler  7.8 
times  less  frequently  than  expected  by  chance 
(G  = 12.14,  P < 0.001). 

In  both  infested  and  uninfested  stands,  we 
compared  the  abundances  of  species  in  flocks 
with  their  relative  abundances,  as  determined 
by  regional  point  counts  (Table  2).  This  pro- 
vided us  with  another  measure  of  flocking  ten- 
dency and  whether  it  changed  with  resource 
availability.  Controlling  for  Julian  date,  flock 
size  was  larger  in  uninfested  sites  (61.1  ± 
78.0  individuals)  than  in  infested  areas  (20.1 
± 15.6  individuals;  Fu52  = 13.23,  P < 0.001). 
In  the  infested  sites,  seven  species  occurred  in 
flocks  more  than  expected  and  seven  less  than 
expected  on  the  basis  of  their  regional  relative 
abundances;  12  species  showed  no  significant 
association  (Table  2).  Of  the  same  26  species 
considered  above,  only  9 occurred  more  fre- 
quently in  flocks  than  expected  on  the  basis 


Hobson  and  Van  Wilgenburg  • BOREAL  MULTISPECIES  FEEDING  FLOCKS 


167 


of  their  relative  abundances,  all  but  1 of  which 
(Brown  Creeper)  showed  a similar  tendency 
in  uninfested  sites  (Black-capped  Chickadee, 
Yellow-rumped  Warbler,  Red-breasted  Nut- 
hatch, Boreal  Chickadee.  Dark-eyed  Junco, 
Yellow  Warbler).  Nine  species  avoided  flocks 
in  uninfested  areas  and,  of  these,  five  species 
also  avoided  flocking  in  the  infested  sites. 
Pine  Siskin  showed  less  tendency  to  flock  in 
the  infested  than  in  uninfested  sites.  Six  spe- 
cies showed  a significant  tendency  to  either 
avoid  or  join  flocks  in  one  of  the  two  habitats, 
with  no  significant  trend  in  the  other  habitat 
(Chipping  Sparrow,  Bay-breasted  Warbler, 
Magnolia  Warbler,  Black-and-white  Warbler, 
Solitary  Vireo,  and  Dark-eyed  Junco). 

Flock  size  and  the  number  of  species  in 
flocks  generally  increased  through  the  season; 
however,  the  trend  was  only  significant  for 
flock  size  (F{  152  = 40.305,  P < 0.001;  Fig. 
1).  For  all  years  combined,  we  compared  flock 
attributes  before  and  after  29  July — the  mid- 
point of  our  observation  period  and  the  date 
by  which  most,  if  not  all,  nests  were  expected 
to  have  fledged.  The  number  of  individuals 
detected  in  flocks  after  29  July  (61.8  ± 77.8, 
n = 79  flocks)  was  greater  than  that  detected 
before  (19.4  ± 14.8,  n — 75  flocks;  Mann- 
Whitney  U = 1,221.0,  two-tailed  P < 0.001). 
The  number  of  species  per  flock  was  similar 
in  the  first  (6.0  ± 2.7,  n = 111  flocks)  and 
second  periods  (7.4  ± 3.8,  n — 103  flocks; 
Mann-Whitney  U = 4,892.5,  two-tailed  P = 
0.067).  We  also  compared  indices  of  flock  di- 
versity by  infested  versus  uninfested  areas  and 
time  period  (before  and  after  29  July).  The 
McIntosh  diversity  index  (Magurran  1988) 
was  higher  in  uninfested  (Me U = 15.5  ± 
14.6,  n = 114)  than  in  the  infested  sites  (Met/ 
= 9.1  ± 7.3,  n = 100;  Mann-Whitney  U = 
3,568.5,  two-tailed  P < 0.001),  but  no  signif- 
icant difference  was  found  for  Simpson  or 
Shannon  diversity  measures.  Shannon  Even- 
ness, however,  was  greater  in  the  infested  {J' 
= 0.90  ± 0.09)  than  in  uninfested  sites  ( J ' = 
0.87  ± 0.12;  Mann-Whitney  U = 4,522.0, 
two-tailed  P = 0.021).  The  McIntosh  index 
was  also  higher  for  flocks  observed  after  29 
July  (Met/  = 15.9  ± 15.2)  compared  with 
those  observed  earlier  (Mcf/  = 9.3  ± 7.0,  n 
= 214;  Mann-Whitney  U = 3,842.5,  two- 
tailed  P < 0.001);  again,  however,  we  detect- 
ed no  difference  in  the  other  measures  of  di- 


versity. Shannon  Evenness  was  greater  in 
flocks  observed  before  29  July  ( J ' = 0.90  ± 
0.09)  compared  with  those  observed  later  ( J ' 
= 0.87  ± 0.12;  Mann-Whitney  U = 4,244.0, 
two-tailed  P = 0.003). 

DISCUSSION 

The  tendency  for  species  to  flock  in  our 
study  area  was  widespread  among  migrants 
and  residents.  Tennessee  Warbler  was  one  of 
the  migrants  most  frequently  observed  flock- 
ing, a phenomenon  that  may  be  related  to  its 
relatively  earlier  breeding  and  dispersal  in  the 
boreal  forest,  as  well  as  to  its  high  abundance 
(Rimmer  and  McFarland  1998).  This  species 
is  one  of  the  earliest  fall  migrants  to  be  re- 
corded at  Delta  Marsh  Bird  Observatory 
(DMBO),  a constant-effort  mist-netting  sta- 
tion just  south  of  our  study  area  (DMBO  un- 
publ.  data).  Among  residents,  Black-capped 
Chickadee,  Red-breasted  Nuthatch,  and  Bo- 
real Chickadee  were  among  the  most  fre- 
quently observed  flocking  species.  Similarly, 
other  studies  in  temperate  North  America  and 
Europe  have  revealed  that  parids  and  nut- 
hatches occur  frequently  in  multispecies  for- 
aging flocks;  parids,  in  particular,  have  been 
classified  as  nuclear  species  in  these  aggre- 
gations (Morse  1970,  Berner  and  Grubb  1985, 
Monkkonen  et  al.  1996).  In  our  study  area, 
resident  boreal  species  typically  breed  earlier 
than  migrants  and  are  observed  moving  in 
family  groups  during  June  when  most  mi- 
grants are  still  incubating.  This  phenology 
may  predispose  them  to  serving  as  catalysts 
for  flocking,  similar  to  their  roles  in  forming 
fall  and  winter  flocks. 

Despite  many  thousands  of  hours  of  field- 
work in  the  southern  boreal  mixed-woods 
from  May  through  September,  we  observed  no 
mixed-species  foraging  aggregations  until  late 
June.  Thereafter,  the  probability  of  encounter- 
ing flocks  increased  as  birds  dispersed  beyond 
their  territory  boundaries.  Additionally,  flocks 
tended  to  be  larger  later  in  the  season;  thus, 
even  though  flocks  of  resident  species  would 
move  through  the  territories  of  migrant  spe- 
cies, the  migrants  apparently  did  not  tempo- 
rarily join  the  residents  as  they  passed  through 
(Monkkonen  et  al.  1996).  Rather,  failed  breed- 
ers or  birds  in  dispersing  family  groups  likely 
constituted  the  earliest  migrants  joining 
flocks. 


168 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  1.  Summary  of  flocking  data  for  avian  species  recorded  in  the  southern  boreal  mixed- wood  forest 
of  Saskatchewan,  Canada,  1992-1996. 


Species 

No.  individuals 
(%) 

No.  flocks 
in  which 
present  (%) 

Mean  no.  per 
flock  (SD) 

Tennessee  Warbler  ( Vermivora  peregrina) 

952 

(16.55) 

146 

(67.59) 

6.5 

(9.1) 

Black-capped  Chickadee  ( Poecile  atricapillus ) 

854 

(14.84) 

142 

(65.74) 

6.0 

(7.8) 

Yellow-rumped  Warbler  ( Dendroica  coronato ) 

635 

(11.04) 

132 

(61.11) 

4.8 

(6.3) 

Red-breasted  Nuthatch  (Sitta  canadensis ) 

383 

(6.66) 

122 

(56.48) 

3.1 

(2.7) 

Boreal  Chickadee  ( Poecile  hudsonica ) 

378 

(6.57) 

81 

(37.50) 

4.7 

(5.1) 

Chipping  Sparrow  ( Spizella  passerina) 

355 

(6.17) 

89 

(41.20) 

4.0 

(3.0) 

American  Redstart  ( Setophaga  ruticilla ) 

238 

(4.14) 

35 

(16.20) 

6.8 

(9.9) 

Red-eyed  Vireo  ( Vireo  olivaceus ) 

207 

(3.60) 

78 

(36.11) 

2.7 

(2.2) 

Pine  Siskin  ( Carduelis  pinus) 

178 

(3.09) 

26 

(12.04) 

6.8 

(6.2) 

Ruby-crowned  Kinglet  ( Regulus  calendula ) 

171 

(2.97) 

57 

(26.39) 

3.0 

(2.3) 

Blackburnian  Warbler  {Dendroica  fused) 

161 

(2.80) 

45 

(20.83) 

3.6 

(2.6) 

Bay-breasted  Warbler  {Dendroica  castanea ) 

150 

(2.61) 

48 

(22.22) 

3.1 

(2.5) 

White-throated  Sparrow  {Zonotrichia  albicollis ) 

127 

(2.21) 

33 

(15.28) 

3.8 

(4.6) 

Chestnut-sided  Warbler  {Dendroica  pensylvanica) 

111 

(1.93) 

21 

(9.72) 

5.3 

(5.9) 

Brown  Creeper  {Certhia  americana) 

104 

(1.81) 

47 

(21.76) 

2.2 

(1.2) 

Magnolia  Warbler  {Dendroica  magnolia ) 

73 

(1.27) 

40 

(18.52) 

1.8 

(1.3) 

Black- throated  Green  Warbler  {Dendroica  virens) 

55 

(0.96) 

22 

(10.19) 

2.5 

(1.5) 

Black-and-white  Warbler  {Mniotilta  varia ) 

52 

(0.90) 

16 

(7.41) 

3.3 

(3.5) 

Cape  May  Warbler  {Dendroica  tigrina ) 

51 

(0.89) 

26 

(12.04) 

2.0 

(1.2) 

Blue-headed  Vireo  {Vireo  solitarius) 

50 

(0.87) 

16 

(7.41) 

3.1 

(3.4) 

Ovenbird  {Seiurus  aurocapilla ) 

38 

(0.66) 

27 

(12.50) 

1.4 

(0.8) 

Dark-eyed  Junco  {Junco  hyemalis ) 

36 

(0.63) 

14 

(6.48) 

2.6 

(2.0) 

Cedar  Waxwing  {Bombycilla  cedrorum ) 

32 

(0.56) 

4 

(1.85) 

8.0 

(8.1) 

Yellow  Warbler  {Dendroica  petechia) 

30 

(0.52) 

11 

(5.09) 

2.7 

(2.1) 

Golden-crowned  Kinglet  {Regulus  satrapa) 

30 

(0.52) 

6 

(2.78) 

5.0 

(5.5) 

Rose-breasted  Grosbeak  {Pheucticus  melanocephalus) 

29 

(0.50) 

9 

(4.17) 

3.2 

(3.0) 

Canada  Warbler  {Wilsonia  canadensis ) 

27 

(0.47) 

15 

(6.94) 

1.8 

(1.1) 

American  Robin  {Turdus  migratorius ) 

25 

(0.43) 

5 

(2.31) 

5.0 

(8.4) 

Yellow-bellied  Sapsucker  {Sphyrapicus  varius) 

23 

(0.40) 

15 

(6.94) 

1.5 

(0.6) 

White-winged  Crossbill  {Loxia  leucoptera) 

17 

(0.30) 

3 

(1.39) 

5.7 

Swainson’s  Thrush  {Catharus  ustulatus ) 

14 

(0.24) 

6 

(2.78) 

2.3 

(2.0) 

Flycatcher  spp.  {Empidonax  spp.) 

14 

(0.24) 

6 

(2.78) 

2.3 

(0.4) 

Hairy  Woodpecker  {Picoides  villosus ) 

12 

(0.21) 

11 

(5.09) 

1.1 

(0.3) 

Mourning  Warbler  {Oporornis  Philadelphia) 

12 

(0.21) 

9 

(4.17) 

1.3 

(0.7) 

Philadelphia  Vireo  {Vireo  philadelphicus) 

11 

(0.19) 

7 

(3.24) 

1.6 

(0.5) 

Palm  Warbler  {Dendroica  palmarum) 

9 

(0.16) 

4 

(1.85) 

2.3 

(1.9) 

Purple  Finch  {Carpodacus  purpureus) 

9 

(0.16) 

3 

(1.39) 

3.0 

Least  Flycatcher  {Empidonax  minimus) 

8 

(0.14) 

5 

(2.31) 

1.6 

(0.9) 

Northern  Flicker  {Colaptes  auratus) 

8 

(0.14) 

5 

(2.31) 

1.6 

(0.9) 

Downy  Woodpecker  {Picoides  pubescens) 

7 

(0.12) 

6 

(2.78) 

1.2 

(0.4) 

Gray  Jay  {Perisoreus  canadensis) 

6 

(0.10) 

3 

(1.39) 

2.0 

Western  Tanager  {Piranga  ludoviciana) 

6 

(0.10) 

3 

(1.39) 

2.0 

Alder  Flycatcher  {Empidonax  alnorum) 

5 

(0.09) 

1 

(0.46) 

5.0 

Connecticut  Warbler  {Oporornis  agilis) 

4 

(0.07) 

3 

(1.39) 

1.3 

Wilson’s  Warbler  {Wilsonia  pusilla) 

4 

(0.07) 

3 

(1.39) 

1.3 

Evening  Grosbeak  {Coccothraustes  vespertinus) 

4 

(0.07) 

2 

(0.93) 

2.0 

Blue  Jay  {Cyanocitta  cristata) 

4 

(0.07) 

1 

(0.46) 

4.0 

Kinglet  spp.  {Regulus  spp.) 

4 

(0.07) 

1 

(0.46) 

4.0 

Ruby-throated  Hummingbird  {Archilochus  colubris) 

4 

(0.07) 

1 

(0.46) 

4.0 

Song  Sparrow  {Melospiza  melodia) 

4 

(0.07) 

1 

(0.46) 

4.0 

Common  Yellowthroat  {Geothlypis  trichas) 

3 

(0.05) 

2 

(0.93) 

1.5 

Pileated  Woodpecker  {Dryocopus  pileatus) 

3 

(0.05) 

2 

(0.93) 

1.5 

Swamp  Sparrow  {Melospiza  georgiana) 

3 

(0.05) 

2 

(0.93) 

1.5 

Warbling  Vireo  {Vireo  gilvus) 

3 

(0.05) 

2 

(0.93) 

1.5 

Blackpoll  Warbler  {Dendroica  striata) 

2 

(0.03) 

2 

(0.93) 

1.0 

Northern  Waterthrush  {Seiurus  noveboracensis) 

2 

(0.03) 

2 

(0.93) 

1.0 

Hobson  and  Van  Wilgenburg  • BOREAL  MULTISPECIES  FEEDING  FLOCKS 


169 


TABLE  1 . Continued. 

No.  flocks 

No.  individuals 

in  which 

Mean  no.  per 

Species 

(%) 

present  (%) 

flock  (SD) 

Thrush  spp.  ( Catharus  spp.) 

2 (0.03) 

2 (0.93) 

1.0 

Traill’s  Flycatcher  ( Empidonax  traillii) 

2 (0.03) 

2 (0.93) 

1.0 

American  Three-toed  Woodpecker  ( Picoides  dorsalis ) 

2 (0.03) 

2 (0.93) 

1.0 

American  Goldfinch  ( Carduelis  tristis) 

2 (0.03) 

1 (0.46) 

2.0 

Winter  Wren  ( Troglodytes  troglodytes ) 

2 (0.03) 

1 (0.46) 

2.0 

Common  Grackle  ( Quiscalus  quiscula) 

1 (0.02) 

1 (0.46) 

1.0 

Eastern  Phoebe  (Say o mis  phoebe) 

1 (0.02) 

1 (0.46) 

1.0 

House  Wren  (Troglodytes  aedon ) 

1 (0.02) 

1 (0.46) 

1.0 

Orange-crowned  Warbler  (Vermivora  celata) 

1 (0.02) 

1 (0.46) 

1.0 

Olive-sided  Flycatcher  (Contopus  cooperi) 

1 (0.02) 

1 (0.46) 

1.0 

White-breasted  Nuthatch  (Sitta  carolinensis ) 

1 (0.02) 

1 (0.46) 

1.0 

Total 

5,753  (100.00) 

216 

Overall,  we  found  relatively  few  significant 
(positive  or  negative)  species  co-occurrences 
in  flocks.  Brown  Creeper  was  positively  as- 
sociated with  Red-breasted  Nuthatch;  this 
likely  reflects  common  foraging  habitats.  Sim- 
ilarly, the  strong  negative  association  between 
American  Redstart  and  Bay-breasted  Warbler 
likely  reflects  the  very  different  habitats  that 
these  species  prefer  (i.e.,  deciduous  understo- 
ry versus  coniferous  canopy).  Instead  of 
strong  tendencies  for  species  to  associate  with 
others  during  flocking,  we  observed  random 
associations  of  individuals  and  species  more 
often. 

Our  flocking  propensity  results  suggest  that 
some  species  show  stronger  tendencies  to 
flock  than  others.  In  both  infested  and  unin- 
fested sites.  Black-capped  and  Boreal  chick- 
adees, Yellow-rumped  Warbler,  Red-breasted 
Nuthatch,  and  Yellow  Warbler  consistently 
showed  high  tendencies  to  flock,  whereas 
Red-eyed  Vireo,  White-throated  Sparrow,  Ov- 
enbird.  Cape  May  Warbler,  and  Black-throated 
Green  Warbler  consistently  showed  negative 
tendencies  to  flock,  based  on  their  abundance. 
There  appear  to  be  no  strong  underlying  pat- 
terns other  than  an  increased  propensity  for 
residents  to  flock.  The  flocking  propensity 
trends  we  observed  could  have  been  biased  by 
the  low  detectabilities  of  a few  species  with 
high-pitched  songs  (e.g.,  Brown  Creeper, 
Black-and-white  Warbler,  Cape  May  Warbler, 
and  Bay-breasted  Warbler);  if  that  were  the 
case,  however,  our  estimates  of  flocking  pro- 
pensity should  have  been  greater  instead  of 
lower  because  high-pitched  species  would 


likely  be  more  detectable  in  flock  surveys  (vi- 
sual) than  during  point  counts  (largely  audi- 
tory). 

The  two  primary  hypotheses  explaining  the 
occurrence  of  mixed-species  foraging  flocks 
are  (1)  the  reduction  of  per  capita  predation 
risk  and  (2)  greater  facilitation  of  successful 
foraging  due  to  decreased  need  for  vigilance 
or  insect  flushing  (Morse  1977,  1980;  Krebs 
and  Davies  1981).  At  sites  in  Ohio,  Berner 
and  Grubb  (1985)  sought  evidence  for  each 
hypothesis  by  experimentally  manipulating 
food  abundance  during  winter  and  examining 
the  tendency  for  resident  species  to  flock. 
They  found  less  flocking  in  a food-supple- 
mented site  versus  a control  site  and  so  argued 
that  flocking  was  related  more  to  foraging 
than  to  antipredator  strategies  per  se.  Al- 
though we  did  not  manipulate  food  abundance 
on  our  sites,  we  were  able  to  examine  flock 
composition  in  uninfested  and  infested  forests 
of  similar  composition.  Forests  infested  with 
spruce  budworm  are  known  to  provide  a su- 
perabundant food  source  for  many  forest 
songbirds,  including  budworm  specialists  and 
non-specialists  (Zach  and  Falls  1975,  Morse 
1989).  Flocks  in  budworm-infested  areas  were 
less  diverse  (i.e.,  greater  community  even- 
ness) than  those  outside  of  infested  areas,  like- 
ly because  flocks  in  the  infested  stands  were 
dominated  by  budworm  specialists,  such  as 
Cape  May  Warbler,  Tennessee  Warbler,  and 
Bay-breasted  Warbler.  However,  we  found  no 
general  pattern  of  a greater  propensity  to  flock 
in  uninfested  versus  infested  areas:  nine  spe- 
cies showed  a greater  flocking  propensity  in 


TABLE  2.  Avian  flocking  propensity  (FP)  in  sites  with  and  without  spruce  budworm  infestation.  Relative  abundance,  from  point-count  data  and  within  flocks, 
is  given  for  each  species.  FP  denotes  trend  ( + , 0 = no  trend)  in  flocking  propensity — the  occurrence  of  a species  in  flocks  more  or  less  than  expected  (based 

on  its  relative  abundance)  due  to  chance. 


170 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Hobson  and  Van  Wilgenburg  • BOREAL  MULTISPECIES  FEEDING  FLOCKS 


171 


FIG.  1.  Temporal  patterns  in  avian  flocking,  by 
species  richness  and  flock  size,  for  all  species  observed 
in  the  southern  boreal  mixed-wood  forest  of  Saskatch- 
ewan, Canada,  1992-1996. 

uninfested  sites  and  seven  had  a higher  pro- 
pensity in  the  infested  sites.  Flock  size,  how- 
ever, tended  to  be  much  larger  in  uninfested 
sites  than  in  the  infested  sites.  While  our  re- 
sults are  not  entirely  unequivocal,  they  are 
congruent  with  the  findings  of  Berner  and 
Grubb  (1985)  in  linking  flocking  propensity 
to  relative  food  availability. 

Combined,  the  few  trends  in  species  co-oc- 
currences, the  inconsistent  trends  in  flocking 
propensity  for  most  species  examined,  and  the 
contrasting  diversity  measures  between  infest- 
ed and  uninfested  sites  suggest  that  flocks  are 
largely  representative  of  local  avian  commu- 
nities. Other  than  resident  species,  flock  struc- 
ture appears  little  affected  by  species’  migra- 
tory patterns  or  foraging  and  nesting  guilds. 
This  suggests  that  the  advantages  of  flocking 
extend  to  most  species,  despite  different  life- 
history  strategies. 

Although  we  found  little  structure  in  pat- 
terns of  flocking  on  the  breeding  grounds,  it 
is  well  established  that  flocking  does  occur  in 
boreal  forest  bird  communities  immediately 
after  the  young  fledge  (i.e.,  as  soon  as  birds 


are  no  longer  constrained  by  nesting).  The  ex- 
istence of  mixed-species  flocks  during  south- 
bound migration  suggests  that  this  behavior 
continues  for  migrants,  possibly  until  they 
reach  their  wintering  grounds  (Morse  1970). 
Whether  resident  or  migrant,  species  that  join 
flocks  may  engage  in  non-flocking  behavior 
only  during  the  relatively  short  breeding  pe- 
riod in  their  life  cycle.  Additional  studies  on 
the  breeding  grounds  as  soon  as  multispecies 
feeding  flocks  begin  to  form  are  now  needed 
to  investigate  how  flocking  relates  to  other  de- 
mands, such  as  post-fledging  parental  care  and 
molt. 

ACKNOWLEDGMENTS 

We  thank  the  numerous  field  assistants  who  helped 
collect  field  data  on  mixed-species  flocks  over  the 
course  of  our  investigations,  in  particular,  E.  Cum- 
ming,  who  also  provided  additional  data.  Our  work 
was  funded  by  the  Canadian  Wildlife  Service  and  the 
Prince  Albert  Model  Forest  (Project  No.  211).  We 
thank  M.  Monkkdnen  and  two  anonymous  reviewers 
for  constructive  comments  that  improved  the  manu- 
script. 

LITERATURE  CITED 

Berner,  T.  O.  and  T.  C.  Grubb,  Jr.  1985.  An  experi- 
mental analysis  of  mixed-species  flocking  in  birds 
of  deciduous  woodland.  Ecology  66:1229-1236. 
Buskirk,  W.  H.,  G.  V.  N.  Powell,  J.  F.  Wittenberger, 
R.  E.  Buskirk,  and  T.  U.  Powell.  1972.  Interspe- 
cific bird  flocks  in  tropical  highland  Panama.  Auk 
89:612-624. 

Elgar,  M.  A.  1989.  Predator  vigilance  and  group  size 
in  mammals  and  birds:  a critical  review  of  the 
empirical  evidence.  Biological  Reviews  64:13-33. 
Ewert,  D.  N.  and  R.  A.  Askins.  1991.  Flocking  be- 
havior of  migratory  warblers  in  winter  in  the  Vir- 
gin Islands.  Condor  93:864-868. 

Hobson,  K.  A.  and  E.  M.  Bayne.  2000.  Breeding  bird 
communities  in  boreal  forest  of  western  Canada: 
consequences  of  “unmixing”  the  mixedwoods. 
Condor  102:759-769. 

Hutto,  R.  L.  1987.  A description  of  mixed-species 
insectivorous  bird  flocks  in  western  Mexico.  Con- 
dor 89:282-292. 

Hutto,  R.  L.  1994.  The  composition  and  social  orga- 
nization of  mixed-species  flocks  in  a tropical  de- 
ciduous forest  in  western  Mexico.  Condor  96: 
105-118. 

Krebs,  J.  R.  and  N.  B.  Davies.  1981.  An  introduction 
to  behavioral  ecology.  Sinnauer  Associates,  Sun- 
derland, Massachusetts. 

Latta,  S.  C.  and  J.  M.  Wunderle,  Jr.  1996.  The  com- 
position and  foraging  ecology  of  mixed-species 
flocks  in  pine  forests  of  Hispaniola.  Condor  98: 
595-607. 


172 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Magurran,  A.  E.  1988.  Ecological  diversity  and  its 
measurement.  Princeton  University  Press,  Prince- 
ton, New  Jersey. 

Monkkonen,  M.,  J.  T.  Forsman,  and  P.  Helle.  1996. 
Mixed-species  foraging  aggregations  and  heter- 
ospecific attraction  in  boreal  bird  communities. 
Oikos  77:127-136. 

Morse,  D.  H.  1970.  Ecological  aspects  of  some  mixed- 
species  foraging  flocks  of  birds.  Ecological  Mono- 
graphs 40:1 19-167. 

Morse,  D.  H.  1977.  Feeding  behavior  and  predator 
avoidance  in  heterospecific  groups.  Bioscience  27: 
332-339. 

Morse,  D.  H.  1980.  Behavioral  mechanisms  in  ecol- 
ogy. Harvard  University  Press,  Cambridge,  Mas- 
sachusetts. 

Morse,  D.  H.  1989.  American  warblers:  an  ecological 
and  behavioral  perspective.  Harvard  University 
Press,  Cambridge,  Massachusetts. 

Munn,  C.  A.  and  J.  W.  Terborgh.  1979.  Multi-species 
territoriality  in  Neotropical  foraging  flocks.  Con- 
dor 81:338-347. 

Newton,  A.  1896.  A dictionary  of  birds.  A.  and  C. 
Black,  London,  United  Kingdom. 

Powell,  G.  V.  N.  1979.  Structure  and  dynamics  of 
interspecific  flocks  in  a Neotropical  mid-elevation 
forest.  Auk  96:375-390. 

Price,  J.,  S.  Droege,  and  A.  Price.  1995.  The  summer 
atlas  of  North  American  birds.  Academic  Press, 
London,  United  Kingdom. 

Pulliam,  H.  R.  1973.  On  the  advantages  of  flocking. 
Journal  of  Theoretical  Biology  38:419-422. 

Rimmer,  C.  C.  and  K.  P.  McFarland.  1998.  Tennessee 


Warbler  ( Vermivora  peregrina).  The  Birds  of 
North  America,  no.  350. 

Robbins,  C.,  S.  D.  Bystrak,  and  P.  H.  Geissler.  1986. 
The  Breeding  Bird  Survey:  its  first  fifteen  years, 
1965-1979.  Resource  Publication,  no.  157.  U.S. 
Fish  and  Wildlife  Service,  Washington,  D.C. 

Saskatchewan  Natural  Resources  Forest  and 
Lands  Branch.  1993.  Proposal  for  an  aerial  ap- 
plication of  a biopesticide  to  control  an  infestation 
of  spruce  budworm.  Saskatchewan  Natural  Re- 
sources, Prince  Albert,  Saskatchewan,  Canada. 

SCHMIEGELOW,  F.  K.  A.  AND  M.  MONKKONEN.  2002. 
Habitat  loss  and  fragmentation  in  dynamic  land- 
scapes: avian  perspectives  from  the  boreal  forest. 
Ecological  Applications  12:375-389. 

Sharpe,  L.  B.  1905.  On  further  collections  of  birds 
from  Efulen  district  of  Cameroon,  West  Africa, 
including  field  notes  of  G.  L.  Bates.  Ibis  1905: 
461-476. 

Stelfox,  J.  B.  (Ed.).  1995.  Relationships  between 
stand  age,  stand  structure,  and  biodiversity  in  as- 
pen mixedwood  forests  in  Alberta.  Alberta  Envi- 
ronmental Centre,  Vegreville  (AECV95-R1),  and 
Alberta  Land  and  Forest  Service  and  Canadian 
Forest  Service,  Edmonton,  Alberta,  Canada. 

Weir,  J.  M.  H.  and  E.  A.  Johnson.  1998.  Effects  of 
escaped  settlement  fires  and  logging  on  forest 
composition  in  the  mixedwood  boreal  forest.  Ca- 
nadian Journal  of  Forest  Research  28:459-467. 

Zach,  R.  and  J.  B.  Falls.  1975.  Response  of  the  Ov- 
enbird  (Aves:  Parulidae)  to  an  outbreak  of  the 
spruce  budworm.  Canadian  Journal  of  Zoology 
53:1669-1672. 

Zar,  J.  H.  1996.  Biostatistical  analysis,  3rd  ed.  Pren- 
tice Hall,  Upper  Saddle  River,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  1 18(2):  173— 177,  2006 


VARIATION  IN  SIZE  AND  COMPOSITION  OF  BUFFLEHEAD 
(. BUCEPHALA  ALBEOLA)  AND  BARROW’S  GOLDENEYE 
( BUCEPHALA  ISLANDICA)  EGGS 

JENNIFER  L.  LAVERS,1 35  JONATHAN  E.  THOMPSON,24 
CYNTHIA  A.  PASZKOWSKI,1  AND  C.  DAVISON  ANKNEY2 3 4 5 


ABSTRACT. — We  investigated  the  relationships  between  egg  nutrient  constituents  and  fresh  egg  mass  in 
Bufflehead  ( Bucephala  albeola)  and  Barrow’s  Goldeneye  ( B . islandica ).  We  found  consistently  positive  rela- 
tionships between  egg  mass  and  yolk,  albumen,  lipid,  mineral,  and  water  (absolute  amounts);  however,  the 
proportions  of  nutrient  components  to  fresh  mass  were  highly  variable  in  the  eggs  of  both  species  (allometric 
relationships).  In  Bufflehead  eggs,  all  components  except  mineral  exhibited  negative  allometry  with  fresh  egg 
mass.  In  Barrow’s  Goldeneye  eggs,  only  mineral  exhibited  negative  allometry,  whereas  yolk,  lipid,  and  water 
all  exhibited  positive  allometry  with  fresh  egg  mass.  Overall,  larger  eggs  of  both  species  contained  greater 
absolute  amounts  of  nutrients;  therefore,  larger  eggs  were  of  better  quality  than  smaller  eggs.  Nutrient  content, 
however,  was  more  highly  correlated  with  mass  in  Barrow’s  Goldeneye  eggs  than  in  Bufflehead  eggs.  We  propose 
that  this  may  be  due  to  the  source  of  egg  nutrients:  because  of  their  smaller  body  size,  Buffleheads  typically 
rely  more  on  exogenous  nutrients  than  Barrow’s  Goldeneyes.  Received  5 January  2005,  accepted  16  December 
2005. 


For  many  bird  species,  nutrient  content  is 
positively  correlated  with  egg  size.  Conse- 
quently, egg  size  is  often  used  as  an  indicator 
of  egg  and  hatchling  quality  (Birkhead  1984, 
Sotherland  and  Rahn  1987,  Pelayo  and  Clark 
2002).  There  are  many  potential  benefits  to 
laying  larger,  and  presumably  better  quality, 
eggs,  including  increased  hatchling  size  (Ali- 
sauskas  1986,  Dawson  and  Clark  1996,  An- 
derson and  Alisauskas  2001,  Pelayo  and  Clark 
2002),  increased  growth  rate  of  both  embryos 
and  hatchlings  (Martin  1987,  Badzinski  et  al. 
2002),  and  higher  probability  of  survival  after 
hatching  (Dawson  and  Clark  1996).  Such  ben- 
efits may  lead  to  selective  pressure  for  females 
to  produce  larger  eggs  with  greater  protein 
and  lipid  stores  (Lack  1967).  However,  the  se- 
lective pressure  to  produce  larger  eggs  is  con- 
strained by  a number  of  factors,  including  he- 
redity (Martin  1987),  the  female’s  metabolic 
and  physiological  capabilities  (Rohwer  1988, 


1 Univ.  of  Alberta,  Dept,  of  Biological  Sciences,  Ed- 
monton, AB  T6G  2E9,  Canada. 

2 Univ.  of  Western  Ontario,  Ecology  and  Evolution 
Group,  Dept,  of  Zoology,  London,  ON  N6A  5B7,  Can- 
ada. 

3 Current  address:  Memorial  Univ.  of  Newfound- 
land, Dept,  of  Biology,  St.  John’s,  NL  A1B  3X9,  Can- 
ada. 

4 Current  address:  Ducks  Unlimited  Canada,  #200, 
10720-178  St.,  Edmonton,  AB  T5S  1J3,  Canada. 

5 Corresponding  author;  e-mail:  b06jll@mun.ca 


Thomson  et  al.  1998),  and  nutrient  availability 
(Alisauskas  and  Ankney  1992). 

The  eggs  of  species  with  precocial  young, 
such  as  waterfowl  (Anseriformes),  have  larger 
yolks  than  do  those  of  species  with  altricial 
young  (Ricklefs  1977).  Newly  hatched  Buffle- 
head (. Bucephala  albeola ) and  Barrow’s  Gold- 
eneye (. B . islandica)  ducklings  often  struggle 
to  exit  their  nest  cavity,  and,  once  out,  they 
follow  the  female  to  the  nearest  body  of  water, 
which  may  be  located  immediately  below  the 
nest  or  up  to  2 km  away  (Savard  et  al.  1991, 
Gauthier  1993,  Eadie  et  al.  2000).  Ducklings 
must  rely  on  stored  yolk  reserves  until  they 
reach  the  water  and  begin  to  feed  (Birkhead 
1985).  Barrow’s  Goldeneye  and  Bufflehead 
ducklings  can  experience  high  mortality  rates 
in  the  1st  week  after  hatch  due  to  their  inex- 
perience in  foraging  (Savard  et  al.  1991). 
Thus,  ducklings  with  large  yolk  reserves  like- 
ly survive  for  longer  periods  with  little  or  no 
food  than  do  those  with  relatively  small  yolk 
reserves. 

In  some  species,  nutrient  content  does  not 
depend  on  egg  size,  and  the  benefits  of  laying 
larger  eggs  may  not  exist.  In  the  European 
Starling  ( Sturnus  vulgaris),  for  example,  larg- 
er eggs  contained  proportionately  less  yolk 
and  lipid  than  smaller  eggs  (Ricklefs  1984), 
suggesting  that  in  some  species,  the  chicks 
that  hatch  from  larger  eggs  may  not  experi- 


173 


174 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


ence  the  advantages  of  proportionately  larger 
yolk  reserves. 

The  primary  objective  of  this  study  was  to 
determine  the  relationship  between  size  and 
nutrient  composition  in  the  eggs  of  Bufflehead 
and  Barrow’s  Goldeneye  breeding  in  British 
Columbia.  The  two  species  nest  sympatrically 
and  have  a similar  diet  (Thompson  and  An- 
kney  2002),  but  they  exhibit  significant  dif- 
ferences in  body  and  egg  size. 

METHODS 

Study  area. — The  study  area  included  ap- 
proximately 250  km2  of  the  Cariboo  Parklands 
in  central  British  Columbia,  Canada  (52°  07' 
N,  122°  27'  W,  approximate  center  point). 
Montane  and  boreal  wetlands  used  by  breed- 
ing Bufflehead  and  Barrow’s  Goldeneye  were 
typically  too  alkaline  and/or  too  shallow  to 
support  fish,  and  had  well  developed  and  di- 
verse aquatic  invertebrate  communities  (for  a 
more  detailed  description  of  the  study  area, 
see  Thompson  1996). 

Egg  collection  and  preparation. — Buffle- 
head (n  = 21)  and  Barrow’s  Goldeneye  (n  = 
40)  clutches  were  collected  in  1993  and  1994 
in  conjunction  with  a broader  study  investi- 
gating nutritional  strategies  for  reproduction 
in  these  species  (Thompson  1996).  Digital  cal- 
ipers were  used  to  measure  egg  length  and 
width  (breadth)  to  the  nearest  0.1  mm,  and  a 
Mettler  balance  was  used  to  weigh  fresh  eggs 
to  the  nearest  0.1  g.  Eggs  were  then  boiled 
and  frozen,  pending  analysis.  Later,  the  boiled 
eggs  were  thawed  and  separated  into  their 
component  parts:  yolk,  albumen  (including 
egg  membranes),  and  shell.  Egg  components 
were  dried  to  a constant  mass  at  80°  C and 
measured  to  the  nearest  0.01  g.  Because  egg 
lipid  is  confined  to  the  yolk,  the  dried  yolk 
was  washed  with  petroleum  ether  in  a modi- 
fied Soxhlet  apparatus  to  extract  the  lipid 
component  (Dobush  et  al.  1985). 

Statistical  analyses. — High  rates  of  intra- 
specific  brood  parasitism,  particularly  for  Bar- 
row’s Goldeneye,  precluded  reliable  discrim- 
ination between  eggs  of  the  host  and  parasite; 
therefore,  within-clutch  analyses  of  variation 
in  egg  size  and  composition  were  not  con- 
ducted. For  each  variable,  we  inspected  a scat- 
ter plot  to  identify  eggs  that  were  significantly 
larger  or  smaller  than  average  (outliers).  Us- 
ing Principle  Component  Analyses  (PCA), 


outliers  were  identified  as  points  on  the  scatter 
plot  that  lay  distinctly  apart  from  all  others 
(McGarigal  et  al.  2000).  Outliers  exert  undue 
pull  on  the  direction  of  the  component  axes, 
strongly  affecting  the  ecological  efficacy  of 
the  ordination  (McGarigal  et  al.  2000).  A few 
eggs  that  deviated  noticeably  from  the  norm 
were  removed  from  the  data  set.  Final  sample 
sizes  for  Bufflehead  and  Barrow’s  Goldeneye 
(after  eliminating  outliers)  were  123  and  226 
eggs,  respectively. 

Preliminary  analysis  indicated  that  the  re- 
siduals were  normally  distributed  and  the  data 
did  not  exhibit  any  nonlinear  trends.  We  used 
linear  regression  to  determine  the  relationship 
between  absolute  amounts  of  individual  egg 
components  (dependent  variables:  dry  yolk, 
dry  albumen,  lipid,  mineral,  and  water)  and 
fresh  egg  mass  (independent  variable).  We  ex- 
amined proportional  nutrient  content  by  log10 
— log10  (hereafter  log-log)  regressions  of  egg 
components  versus  fresh  egg  mass  (Alisaus- 
kas  1986).  A regression  slope  of  unity  ( b = 
1.0)  signifies  that  a component  makes  up  a 
constant  fraction  of  the  total  egg  mass.  Slopes 
significantly  <1  or  > 1 imply  that  components 
make  up  a decreasing  or  increasing  fraction  of 
the  total  egg  as  egg  mass  increases.  For  each 
species,  we  tested  both  absolute  and  propor- 
tional variation  in  egg  composition.  We  used 
analysis  of  covariance  (ANCOVA)  to  test 
whether  there  was  differential  partitioning  of 
egg  nutrients  between  the  two  species.  Means 
and  slopes  are  reported  ± SE,  and  significance 
was  set  at  P — 0.05.  All  analyses  were  con- 
ducted using  MINITAB  (Minitab,  Inc.  2003). 

RESULTS 

Dimensions  and  composition  of  Bufflehead 
and  Barrow’s  Goldeneye  eggs  are  presented  in 
Table  1.  Fresh  mass  of  Bufflehead  eggs  con- 
sisted of  42%  wet  yolk,  40%  wet  albumen, 
and  9%  mineral.  Overall,  water  composed  ap- 
proximately 52%  of  fresh  egg  mass.  Similarly, 
the  composition  of  Barrow’s  Goldeneye  eggs 
averaged  40%  wet  yolk,  45%  wet  albumen, 
and  9%  mineral.  Water  composed  approxi- 
mately 57%  of  fresh  egg  mass. 

There  was  a consistently  positive  relation- 
ship between  fresh  egg  mass  and  absolute 
amounts  of  dry  yolk,  dry  albumen,  lipid,  min- 
eral, and  water  in  the  eggs  of  both  species 
(Table  2).  In  Bufflehead  eggs,  all  components 


Lavers  et  al.  • EGG  VARIATION  IN  BUFFLEHEAD  AND  BARROW’S  GOLDENEYE 


175 


TABLE  1.  Dimensions  (mm)  and  composition  (g)  of  Bufflehead  ( n 
226)  eggs  collected  in  central  British  Columbia,  1993-1994. 

= 123)  and  Barrow’s 

Goldeneye  ( n = 

Variable 

Bufflehead 

Barrow’s  Goldeneye 

Mean  ± SE 

CVa  (%) 

Mean  ± SE 

CVa  (%) 

Length 

50.20  ±0.15 

3.33 

61.69  ± 0.13 

3.06 

Breadth 

36.22  ± 0.07 

2.13 

43.76  ± 0.06 

2.10 

Fresh  egg  mass 

36.68  ±0.19 

5.89 

66.41  ± 0.22 

5.04 

Mineral 

3.36  ± 0.03 

9.82 

6.24  ± 0.02 

6.09 

Wet  albumen 

14.71  ± 0.22 

16.93 

30.17  ± 0.23 

1 1.60 

Dry  albumen 

2.66  ± 0.02 

9.77 

4.94  ± 0.02 

7.09 

Wet  yolk 

15.46  ± 0.23 

16.24 

26.00  ± 0.27 

15.65 

Dry  yolk 

7.60  ± 0.06 

8.55 

13.31  ± 0.06 

6.99 

Yolk  lipid 

5.14  ± 0.04 

8.95 

8.94  ± 0.04 

7.27 

Yolk  protein 

2.45  ± 0.02 

8.57 

4.31  ± 0.02 

7.19 

Water 

19.91  ± 0.14 

7.89 

37.92  ± 0.21 

8.47 

a Coefficient  of  variation. 


except  mineral  exhibited  negative  allometry 
with  egg  mass  (Table  3).  The  log-log  regres- 
sion slope  for  mineral  did  not  differ  from  uni- 
ty ( b = 0.96  ± 0.11),  indicating  that  mineral 
mass  made  up  a constant  proportion  of  total 
egg  mass.  In  Barrow’s  Goldeneye,  yolk,  lipid, 
and  water  all  exhibited  positive  allometry, 
whereas  mineral  exhibited  negative  allometry 
and  albumen  exhibited  isometry  with  egg 
mass  (Table  3).  Results  of  the  ANCOVA  in- 
dicated that  the  nutrients  of  Bufflehead  and 
Barrow’s  Goldeneye  eggs  are  partitioned  in 
different  ways;  the  slopes  of  the  regression 
lines  for  each  nutrient  differed  (all  P < 0.001) 
between  species. 

DISCUSSION 

The  percentages  of  wet  yolk  in  Bufflehead 
(42%)  and  Barrow’s  Goldeneye  eggs  (40%) 
were  similar  to  those  reported  by  Lack  (1967) 
for  other  waterfowl,  such  as  Common  Gold- 


eneye ( Bucephala  clangula\  44%)  and  Mus- 
covy Duck  ( Cairina  moschata\  40%),  but 
were  greater  than  those  reported  for  Greater 
Snow  Goose  ( Anser  caerulescens  atlanticus; 
36%)  and  Mute  Swan  ( Cygnus  olor,  34%).  In 
Bufflehead  and  Barrow’s  Goldeneye,  yolk,  al- 
bumen, lipid,  mineral,  and  water  (absolute 
amounts)  all  exhibited  a positive  relationship 
with  egg  size.  Log-log  regression  analysis  of 
component  masses  versus  fresh  egg  mass  in- 
dicated interspecific  differences.  In  Bufflehead 
eggs,  all  components  except  mineral  exhibited 
negative  allometry  with  egg  mass.  In  Barrow’s 
Goldeneye  eggs,  only  mineral  exhibited  neg- 
ative allometry,  whereas  yolk,  lipid,  and  water 
exhibited  positive  allometry  with  egg  mass. 
Thus,  on  average,  large  Bufflehead  eggs  do 
not  contain  proportionately  more  nutrients 
than  small  eggs,  whereas  large  Barrow’s 
Goldeneye  eggs  do  contain  more  nutrients 
than  small  eggs.  The  results  for  Bufflehead  are 


TABLE  2.  Summary  of  linear  regression  analyses  (egg  components  versus  fresh  egg  mass;  absolute  amounts; 
all  P < 0.001)  for  Bufflehead  ( n = 123)  and  Barrow’s  Goldeneye  ( n = 226)  eggs  collected  in  central  British 
Columbia,  1993-1994. 


Component 

Bufflehead 

Barrow’s 

; Goldeneye 

b (SE)a 

Intercept 

r2 

b (SE)a 

Intercept 

r2 

Mineral 

0.09  (0.01) 

-0.04 

0.37 

0.06  (0.01) 

2.45 

0.26 

Dry  albumen 

0.04  (0.01) 

0.95 

0.15 

0.07  (0.00) 

-0.01 

0.51 

Dry  yolk 

0.15  (0.02) 

1.76 

0.28 

0.23  (0.01) 

-1.93 

0.69 

Yolk  lipid 

0.11  (0.02) 

1.12 

0.27 

0.16  (0.01) 

-1.71 

0.67 

Yolk  protein 

0.04  (0.01) 

0.83 

0.21 

0.07  (0.01) 

-0.23 

0.53 

Water 

0.32  (0.06) 

8.03 

0.20 

0.67  (0.05) 

-6.55 

0.49 

a Slope  of  regression  ± SE. 


176 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  3.  Summary  of  allometric  regression  analyses  (egg  components  versus  fresh  egg  mass;  all  P < 
0.001)  for  Bufflehead  (n  = 123)  and  Barrow’s  Goldeneye  (n  = 226)  eggs  collected  in  central  British  Columbia, 
1993-1994. 

Component 

Bufflehead 

Barrow’s  Goldeneye 

b (SE)a 

Intercept 

r2 

b (SE)a 

Intercept 

r2 

Mineral 

0.96  (0.11) 

-0.98 

0.35 

0.60  (0.06) 

-0.29 

0.26 

Dry  albumen 

0.61  (0.13) 

-0.54 

0.14 

1.00  (0.06) 

-1.13 

0.52 

Dry  yolk 

0.81  (0.12) 

-0.38 

0.27 

1.12  (0.05) 

-0.92 

0.68 

Yolk  lipid 

0.80  (0.12) 

-0.54 

0.25 

1.16  (0.05) 

-1.16 

0.67 

Yolk  protein 

0.68  (0.01) 

-0.67 

0.20 

1.03  (0.05) 

-1.25 

0.54 

Water 

0.58  (0.11) 

0.39 

0.18 

1.16  (0.08) 

-0.54 

0.45 

a Slope  of  regression  ± SE;  a regression  slope  of  unity  (b  = 1 .0)  signifies  that  a component  makes  up  a constant  fraction  of  the  total  egg  mass.  Slopes 
significantly  <1  or  >1  indicate  that  components  make  up  a decreasing  or  increasing  fraction  of  the  total  egg  as  egg  mass  increases. 


similar  to  those  of  Jager  et  al.  (2000),  who 
found  that  larger  Eurasian  Oystercatcher 
(Haematopus  ostralegus ) eggs  contained  more 
lean  dry  matter  and  lipid  (absolute  amounts) 
than  smaller  eggs,  but  the  proportion  of  both 
constituents  decreased  with  egg  size. 

In  several  bird  species,  hatchlings  from 
large  eggs  have  a higher  probability  of  sur- 
vival to  fledging  than  do  hatchlings  from 
small  eggs  (Payne  1978).  Bufflehead  and  Bar- 
row’s Goldeneye  hatchlings  were  not  mea- 
sured or  monitored  in  this  study,  therefore  it 
is  not  known  whether  large  eggs  of  these  spe- 
cies do  indeed  produce  larger  ducklings.  How- 
ever, larger  Ruddy  Duck  ( Oxyura  jamaicensis ) 
eggs  produced  larger,  more  mature  ducklings 
that  were  provisioned  with  greater  energy  re- 
serves and  exhibited  greater  survival  rates 
than  ducklings  from  smaller  eggs  (Pelayo  and 
Clark  2002). 

Overall,  larger  eggs  in  both  species  con- 
tained more  nutrients,  although  nutrient  con- 
tent of  Barrow’s  Goldeneye  eggs  was  more 
highly  correlated  with  egg  mass  than  it  was  in 
Bufflehead  eggs  (Table  2).  This  suggests  that 
nutrients  in  Bufflehead  and  Barrow’s  Gold- 
eneye eggs  are  partitioned  differently.  A pos- 
sible mechanism  for  this  difference  is  the 
source  of  egg  nutrients:  because  Buffleheads 
have  a smaller  body  size,  they  rely  more  on 
exogenous  nutrients,  whereas  the  larger  Bar- 
row’s Goldeneyes  can  rely  more  on  endoge- 
nous nutrients  (Thompson  1996,  Hobson  et  al. 
2005).  This  may  explain  the  higher  CVs  for 
the  constituents  of  Bufflehead  eggs,  as  they 
are  less  able  to  buffer  the  effects  of  variable 
food  supplies  by  drawing  on  endogenous  re- 
serves. 


Our  results  show  that  larger  eggs  of  Buffle- 
heads and  Barrow’s  Goldeneyes  contain  more 
nutrients  than  smaller  eggs,  which  may  in- 
crease the  survival  of  their  hatchlings  during 
the  1st  crucial  week  of  life.  This  is  especially 
important  given  that  Buffleheads  and  Bar- 
row’s Goldeneyes  nest  in  boreal  and  montane 
regions  where  food  typically  is  less  available 
than  in,  for  example,  the  prairie  wetlands  of 
North  America,  used  by  many  temperate  nest- 
ing ducks  (Thompson  1996,  Thompson  and 
Ankney  2002).  Further  studies  should  be  con- 
ducted on  these  species  to  examine  variation 
in  egg  composition  within  and  between 
clutches  and  to  determine  whether  hatchlings 
from  larger  eggs  are  larger  and  have  lower 
mortality  than  those  from  smaller  eggs. 

ACKNOWLEDGMENTS 

Funding  for  this  project  was  provided  by  the  Na- 
tional Science  and  Engineering  Research  Council  of 
Canada  (NSERC)  through  an  operating  grant  awarded 
to  C.  D.  Ankney.  Additional  financial  support  was  pro- 
vided through  a joint  Canadian  Wildlife  Service  and 
NSERC  grant  awarded  by  the  Pacific  Wildlife  Re- 
search Network  Program.  We  thank  the  University  of 
Western  Ontario  and  University  of  Alberta  for  further 
financial  support  for  this  project  as  well  as  use  of  com- 
puter and  research  labs  for  data  analysis  and  sample 
preparation.  This  research  was  facilitated  by  generous 
cooperation  and  logistical  support  provided  by  M. 
Clarke,  B.  Arner,  D.  Regier,  and  E.  Hennan  of  Ducks 
Unlimited  Canada,  and  A.  Breault  and  D.  Dockerty  of 
the  Canadian  Wildlife  Service.  Permission  to  work  in 
portions  of  the  study  area  was  kindly  granted  by  the 
Canadian  Department  of  National  Defense  (Chilcotin 
Military  Training  Area)  and  B.  Durrell  of  the  Wine 
Glass  Ranch.  Dedicated  field  and  lab  assistance  for  this 
project  was  provided  by  T.  Matthews  and  S.  A.  Lee. 
Finally,  we  thank  G.  Robertson  and  three  anonymous 


Lavers  et  al.  • EGG  VARIATION  IN  BUFFLEHEAD  AND  BARROW’S  GOLDENEYE 


177 


reviewers  for  their  constructive  feedback  on  previous 

versions  of  this  manuscript. 

LITERATURE  CITED 

Alisauskas,  R.  T.  1986.  Variation  in  the  composition 
of  the  eggs  and  chicks  of  American  Coots.  Condor 
88:84-90. 

Alisauskas,  R.  T.  and  C.  D.  Ankney.  1992.  The  cost 
of  egg  laying  and  its  relationship  to  nutrient  re- 
serves in  waterfowl.  Pages  30-61  in  Ecology  and 
management  of  breeding  waterfowl  (B.  D.  J.  Batt, 
A.  D.  Afton,  M.  G.  Anderson,  C.  D.  Ankney,  D. 
H.  Johnson,  J.  A.  Kadler,  and  G.  L.  Krapu,  Eds.). 
University  of  Minnesota  Press,  Minneapolis. 

Anderson,  V.  R.  and  R.  T.  Alisauskas.  2001.  Egg 
size,  body  size,  locomotion,  and  feeding  perfor- 
mance in  captive  King  Eider  ducklings.  Condor 
103:195-199. 

Badzinski,  S.  S.,  C.  D.  Ankney,  J.  O.  Leafloor,  and 
K.  F.  Abraham.  2002.  Egg  size  as  a predictor  of 
nutrient  composition  of  eggs  and  neonates  of  Can- 
ada Geese  ( Branta  canadensis  interior)  and  Lesser 
Snow  Geese  ( Chen  caerulescens  caerulescens). 
Canadian  Journal  of  Zoology  80:333-341. 

Birkhead,  M.  1984.  Variation  in  the  weight  and  com- 
position of  Mute  Swan  ( Cygnus  olor ) eggs.  Con- 
dor 86:489-490. 

Birkhead,  M.  1985.  Variation  in  egg  quality  and  com- 
position in  the  Mallard,  Anas  platyrhynchos . Ibis 
127:467-475. 

Dawson,  R.  D.  and  R.  G.  Clark.  1996.  Effects  of 
variation  in  egg  size  and  hatching  date  on  survival 
of  Lesser  Scaup  Aythya  affinis  ducklings.  Ibis  138: 
693-699. 

Dobush,  G.  R.,  C.  D.  Ankney,  and  D.  G.  Krementz. 
1985.  The  effect  of  apparatus,  extraction  time,  and 
solvent  type  on  lipid  extractions  of  Snow  Geese. 
Canadian  Journal  of  Zoology  63:1917-1920. 

Eadie,  J.  M.,  J.-P.  L.  Savard,  and  M.  L.  Mallory. 
2000.  Barrow’s  Goldeneye  ( Bucephala  islandica ). 
The  Birds  of  North  America,  no.  548. 

Gauthier,  G.  1993.  Bufflehead  ( Bucephala  albeola). 
The  Birds  of  North  America,  no.  67. 

Hobson,  K.  A.,  J.  E.  Thompson,  M.  R.  Evans,  and  S. 
Boyd.  2005.  Tracing  allocation  to  reproduction  in 
Barrow’s  Goldeneye.  Journal  of  Wildlife  Manage- 
ment 69:1221-1228. 

Jager,  T.  D.,  J.  B.  Hulscher,  and  M.  Kersten.  2000. 


Egg  size,  egg  composition,  and  reproductive  suc- 
cess in  the  Oystercatcher  Haematopus  ostralegus. 
Ibis  142:603-613. 

Lack,  D.  1967.  The  significance  of  clutch  size  in  wa- 
terfowl. Wildfowl  18:125-128. 

Martin,  T.  E.  1987.  Food  as  a limit  on  breeding  birds: 
a life  history  perspective.  Annual  Review  of  Ecol- 
ogy and  Systematics  18:453-487. 

McGarigal,  K.,  S.  Cushman,  and  S.  Stafford.  2000. 
Multivariate  statistics  for  wildlife  and  ecology  re- 
search. Springer- Verlag,  New  York. 

Minitab,  Inc.  2003.  MINITAB  Statistical  Software, 
ver.  14.  Minitab,  Inc.,  State  College,  Pennsylva- 
nia. 

Payne,  R.  B.  1978.  Dependence  of  fledging  success  on 
egg  size,  parental  performance,  and  egg  compo- 
sition among  Common  and  Roseate  terns.  Sterna 
hirundo  and  Sterna  dougallii.  Ibis  120:207-214. 

Pelayo,  J.  T.  and  R.  G.  Clark.  2002.  Variation  in  size, 
composition,  and  quality  of  Ruddy  Duck  eggs. 
Condor  104:457-462. 

Ricklefs,  R.  E.  1977.  Composition  of  eggs  of  several 
bird  species.  Auk  94:350-356. 

Ricklefs,  R.  E.  1984.  Variation  in  the  size  and  com- 
position of  eggs  of  the  European  Starling.  Condor 
86:1-6. 

Rohwer,  F.  C.  1988.  Inter-  and  intraspecific  relation- 
ships between  egg  size  and  clutch  size  in  water- 
fowl.  Auk  105:161-176. 

Savard,  J.  L.,  G.  E.  J.  Smith,  and  J.  N.  M.  Smith. 
1991.  Duckling  mortality  in  Barrow’s  Goldeneye 
and  Bufflehead  broods.  Auk  108:568-577. 

Sotherland,  P.  R.  and  H.  Rahn.  1987.  On  the  com- 
position of  bird  eggs.  Condor  89:48-65. 

Thompson,  J.  E.  1996.  Comparative  reproductive  bi- 
ology of  female  Buffleheads  ( Bucephala  albeola ) 
and  Barrow’s  Goldeneye  ( Bucephala  islandica ) in 
central  British  Columbia.  Ph.D.  dissertation.  Uni- 
versity of  Western  Ontario,  London,  Ontario,  Can- 
ada. 

Thompson,  J.  E.  and  C.  D.  Ankney.  2002.  Role  of 
food  in  territoriality  and  egg  production  of  Buffle- 
heads ( Bucephala  albeola)  and  Barrow’s  Gold- 
eneye ( Bucephala  islandica).  Auk  119:1 075 — 
1090. 

Thomson,  D.  L.,  P.  Monaghan,  and  R.  W.  Furness. 
1998.  The  demands  of  incubation  and  avian  clutch 
size.  Biological  Reviews  73:293-304. 


The  Wilson  Journal  of  Ornithology  1 18(2):  178-186,  2006 


SITE-SPECIFIC  SURVIVAL  OF  BLACK-HEADED  GROSBEAKS 
AND  SPOTTED  TOWHEES  AT  FOUR  SITES  WITHIN  THE 
SACRAMENTO  VALLEY,  CALIFORNIA 

THOMAS  GARDALI1 2 AND  NADAV  NUR1 2 


ABSTRACT. — We  estimated  apparent  annual  survival  and  recapture  probabilities  for  adult  Black-headed 
Grosbeaks  ( Pheucticus  melanocephalus ) and  Spotted  Towhees  ( Pipilo  maculatus)  at  four  sites  along  the  Sacra- 
mento River,  California.  To  calculate  our  estimates,  we  used  capture-recapture  mist-net  data  collected  over  two 
time  periods  at  four  study  sites:  from  1993  to  1995  at  Flynn,  Ohm,  and  Sul  Norte,  and  from  1995  to  2000  at 
Ohm  and  Phelan  Island.  Our  primary  objective  was  to  determine  whether  there  were  site-specific  differences  in 
adult  survival  and  recapture  probabilities  for  each  species.  Such  differences  are  rarely  investigated,  yet,  if  present, 
suggest  site-specific  differences  in  habitat  quality,  with  important  implications  for  source/sink  dynamics.  We 
found  site-specific  variation  in  Black-headed  Grosbeak  survival  within  both  the  1993-1995  dataset  (Flynn  = 
0.797  ± 0.496,  Ohm  = 0.158  ± 0.191,  Sul  Norte  = 0.773  ± 0.131)  and  the  1995-2000  dataset  (Ohm  = 0.088 
± 0.090,  Phelan  Island  = 0.664  ± 0.111).  For  Spotted  Towhees  (1993-1995  data),  the  most  supported  model 
assumed  constant  survival  across  sites  (0.602  ± 0.240),  but  there  was  some  support  for  site  variation  in  survival, 
as  well  (Flynn  = 0.653  ± 0.365,  Ohm  = 0.214  ± 0.253,  Sul  Norte  = 0.632  ± 0.258).  These  results  clearly 
suggest  site  variation  for  Black-headed  Grosbeaks,  and  weak  evidence  of  site  variation  for  Spotted  Towhees. 
For  both  species,  the  general  pattern  was  low  survival  at  Ohm,  suggesting  low-quality  habitat  there  and/or 
reduced  site  fidelity.  The  magnitude  of  site-to-site  variation  in  survival  observed  in  the  Black-headed  Grosbeak, 
and  suggested  for  Spotted  Towhee,  has  strong  implications  for  determining  source  versus  sink  population  status. 
To  determine  source  versus  sink  status,  we  conclude  that  investigators  must  not  only  take  into  account  site 
variation  in  reproductive  success,  but  also  consider  site-specific  estimation  of  adult  survival.  Received  28  March 
2005,  accepted  4 January  2006. 


Measuring  adult  survival — the  probability 
that  an  adult  will  survive  from  one  year  to  the 
next — is  a critical  step  toward  understanding 
population  dynamics,  as  low  survival  rates 
may  be  responsible  for  population  declines  for 
some  species  (Nur  and  Sydeman  1999).  It  has 
been  hypothesized  that  tropical  deforestation 
has  led  to  decreases  in  over-winter  survival 
(e.g.,  Askins  et  al.  1990,  Rappole  and  Mc- 
Donald 1994),  and  several  recent  studies  sug- 
gest that  events  at  migratory  stopover  areas 
also  may  have  significant  consequences  (e.g., 
Moore  et  al.  1995,  Yong  et  al.  1998,  Sillett 
and  Holmes  2002).  Few  researchers,  however, 
have  examined  the  potential  role  of  the  breed- 
ing grounds  in  affecting  annual  survival 
(Chase  et  al.  1997,  Powell  et  al.  2000,  Sillett 
and  Holmes  2002). 

Many  factors  that  are  related  to  a particular 
species’  life-history  characteristics  operate  to 
influence  adult  survival  at  various  periods  in 
the  annual  cycle.  Survival  of  migratory  spe- 
cies, for  example,  may  be  regulated  primarily 


1 PRBO  Conservation  Science,  4990  Shoreline 
Hwy.,  Stinson  Beach,  CA  94970,  USA. 

2 Corresponding  author;  e-mail:  tgardali@prbo.org 


by  events  during  migration  or  on  their  win- 
tering grounds  in  the  tropics  (e.g.,  habitat 
loss).  In  contrast,  all  factors  influencing  the 
survival  of  resident  species  occur  on  their 
year-round  home  ranges.  Survival  also  may  be 
influenced  by  events  during  the  breeding  sea- 
son in  the  temperate  zone.  Reproductive  effort 
can  affect  survival  rates  for  some  species  (Nur 
1988a,  1988b),  and  individuals  that  must 
make  repeated  nesting  attempts  due  to  high 
levels  of  nest  depredation  may  pay  a greater 
cost  in  terms  of  survival.  For  example,  female 
Common  House-Martins  ( Delichon  urbicum ) 
that  double-brooded  experienced  lower  rates 
of  survival  than  single-brooded  females  (Bry- 
ant 1979).  Additionally,  environments  where 
the  predator  community  is  rich  and  abundant 
and  habitat  cover  is  poor  could  negatively  in- 
fluence survival  rates. 

Despite  the  widely  recognized  assumption 
that  survival  plays  a critical  role  in  regulating 
populations,  few  studies  of  passerines  have 
been  designed  to  specifically  look  for  site-  or 
habitat-specific  differences  (Peach  1993), 
though  several  researchers  have  examined  site 
fidelity  in  relation  to  various  indices  of  site 
quality  (e.g.,  Bollinger  and  Gavin  1989,  Sedg- 


178 


Gardali  and  Nur  • SITE-SPECIFIC  SURVIVAL 


179 


wick  2004).  This  is  likely  because  survival  is 
relatively  difficult  to  measure;  it  requires  sev- 
eral years  of  study,  and  often-small  sample 
sizes  from  individual  sites  prohibit  proper 
analyses.  However,  site-specific  estimates  of 
survival  can  provide  insight  into  habitat  qual- 
ity, and  differences  in  survival  could  alert  land 
managers  to  potential  problems. 

Here,  we  present  site-specific  survival  es- 
timates for  two  species  that  differ  in  life  his- 
tory characteristics — the  migratory  Black- 
headed Grosbeak  ( Pheucticus  melanocephal- 
us ) and  the  resident  Spotted  Towhee  ( Pipilo 
maculatus).  Our  estimates  were  based  on  data 
collected  during  a multi-site,  multi-year,  con- 
stant-effort mist-netting  program  (Nur  and 
Geupel  1993)  conducted  along  the  Sacramen- 
to River.  We  also  investigated  differences  in 
recapture  probability — the  probability  that  an 
individual  that  has  survived  from  year  jc  to 
year  x + 1 is  also  recaptured  in  year  x + 1 
(Nur  and  Clobert  1988).  As  is  often  the  case 
in  attempting  to  estimate  survival,  we  could 
not  distinguish  mortality  from  permanent  dis- 
persal (that  is,  we  measured  “local  survival”; 
Lebreton  et  al.  1992);  thus,  our  estimates  are 
conservative  (Lebreton  et  al.  1992).  Site  dif- 
ferences in  the  survival  estimates  we  present 
may  be  explained  by  variation  in  survival 
probability  from  one  year  to  the  next,  varia- 
tion in  permanent  emigration,  or  both.  How- 
ever, local  movements  of  individuals  from 
year  to  year  (e.g.,  in  some  years  individuals 
may  have  nested  closer  to,  or  farther  from  the 
array  of  mist  nets)  should  not  have  biased  our 
survival  estimates;  such  local  dispersal  (af- 
fecting recapture  from  one  year  to  the  next)  is 
incorporated  into  our  recapture  probability 
calculations  (Nur  and  Clobert  1988). 

METHODS 

Study  sites. — Our  four  study  sites  were  in 
the  Sacramento  Valley,  California:  Flynn  (40° 
06'  N,  122°  12'  W),  Sul  Norte  (39°  46'  N, 
121°  99'  W),  Ohm  (40°  09'  N,  122°  12'  W), 
and  Phelan  Island  (39°  69' N,  121°  97' W). 
Ohm  and  Flynn  were  the  northern-most  sites 
(3.4  km  from  each  other),  Phelan  Island  was 
south  of  these  sites  by  —50  km,  and  Sul  Norte, 
located  —100  km  south  of  Ohm  and  Flynn, 
was  the  southern-most  site  (see  map  in  Gar- 
dali et  al.  in  press).  Sites  ranged  in  elevation 
from  39  to  70  m.  Dominant  trees  included 


Fremont  cottonwood  ( Populus  fremontii),  val- 
ley oak  ( Quercus  lobata ),  and  willow  ( Salix 
spp.)  with  varied  understory  communities 
consisting  of  mugwort  ( Artemesia  douglasi- 
ana ),  Santa  Barbara  sedge  ( Carex  barbarae), 
blue  wildrye  ( Elymus  glaucus ),  California 
blackberry  ( Rubus  ursinus),  and  various  ex- 
otic, weedy  species  (e.g.,  Johnson  grass  [Sor- 
ghum halepense],  Bermuda  grass  [ Cynodon 
dactylon ],  Himalayan  blackberry  [Rubus  dis- 
color]).  Flynn  and  Sul  Norte  (in  Tehama  and 
Glenn  counties,  respectively)  were  riparian 
remnants  of  relatively  old  forests.  Ohm  (in  Te- 
hama County)  was  also  a remnant  forest,  but 
differed  in  that  it  was  grazed  by  cattle  for  the 
duration  of  the  study;  thus,  the  density  of  the 
shrub  community  was  diminished  in  compar- 
ison (TG  pers.  obs.).  Ohm  also  had  more 
black  walnut  ( Juglans  californica ) trees  than 
the  other  sites  (TG  pers.  obs.).  Phelan  Island 
(in  Glenn  County)  was  a riparian  restoration 
site  planted  in  1991  and  1992  (see  Alpert  et 
al.  1999  for  details).  Land  use  surrounding  all 
sites  was  primarily  agricultural  (orchards).  Al- 
though all  sites  were  broadly  similar  in  plant 
species  composition,  landscape  context,  and 
climate,  there  were  likely  some  differences  in 
habitat  structure/complexity  and  flooding  fre- 
quency. 

Field  methods. — Black-headed  Grosbeaks 
and  Spotted  Towhees  were  sampled  through 
standardized  effort  mist  netting  (Monitoring 
Avian  Productivity  and  Survivorship  protocol; 
DeSante  et  al.  2000).  Sampling  occurred  at 
Flynn,  Ohm,  and  Sul  Norte  during  1993-1995 
and  at  Ohm  and  Phelan  Island  during  1995- 
2000.  Ten  12-m,  36-mm-mesh  mist  nets  were 
operated  at  each  study  site  for  5 (morning)  hr 
per  day  for  1 day  during  each  of  10  consec- 
utive 10-day  periods  (—500  net-hr/site).  Start- 
ing dates  were  approximately  1 May  and  op- 
eration continued  through  the  10— day  period 
ending  8 August.  Nets  were  opened  15  min 
after  sunrise  and  kept  open  for  5 hr  during 
each  day  of  net  operation.  Nets  were  checked 
every  20  to  45  min,  depending  on  weather 
conditions,  and  were  closed  when  water  ac- 
cumulated on  them,  or  when  wind  caused  net 
pockets  to  consistently  billow.  Because  of 
these  standardized  protocols,  effort  (net  hr) 
was  similar  among  sites  and  years.  Captured 
birds  were  banded  with  federal  bands,  mea- 
sured, and  released  immediately. 


180 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol  118,  No.  2,  June  2006 


Statistical  analyses. — We  used  capture/re- 
capture data  of  adult  birds  to  estimate  annual 
survival,  and  we  used  program  SURGE  4.3  to 
calculate  recapture  probabilities  (Lebreton  et 
al.  1992,  Cooch  et  al.  1996).  Recapture  prob- 
abilities of  transients  can  be  lower  than  recap- 
ture probabilities  of  site-faithful  individuals  in 
most  species  (Peach  et  al.  1991,  Chase  et  al. 
1997,  Pradel  et  al.  1997),  and  such  heteroge- 
neity in  recapture  probabilities  violates  an  as- 
sumption of  capture-recapture  methodology 
(see  Lebreton  et  al.  1992  for  discussion).  Var- 
ious methods  have  been  used  to  identify  site- 
faithful  individuals,  such  as  identifying  indi- 
viduals captured  at  least  twice  during  any 
breeding  season  and/or  captured  in  more  than 
one  year  (e.g.,  Chase  et  al.  1997,  Gardali  et 
al.  2000,  Nur  et  al.  2000).  In  our  study,  how- 
ever, we  did  not  recapture  enough  individuals 
meeting  these  criteria  to  allow  such  analyses 
except  for  (1)  Black-headed  Grosbeaks  at  Sul 
Norte  in  1993-1995  and  (2)  Black-headed 
Grosbeaks  at  Phelan  Island  in  1995-2000.  In 
these  two  cases,  individual  Black-headed 
Grosbeaks  captured  at  least  twice  during  any 
breeding  season  and  at  least  7 days  apart,  and/ 
or  those  captured  in  more  than  one  year  were 
considered  site-faithful  breeders.  We  com- 
pared survival  estimates  from  this  “high  site 
fidelity”  subset  with  those  from  the  full  da- 
taset, to  determine  whether  the  inclusion  of 
transient  Black-headed  Grosbeaks  biased  our 
survival  estimates.  All  other  analyses  were 
based  on  the  full  dataset  (site-faithful  breeders 
and  transients). 

For  both  species  and  each  dataset  (i.e., 
1993-1995  and  1995-2000),  we  evaluated 
four  models  with  time-constant  survival  (phi) 
and  recapture  (p)  probabilities  to  test  for  po- 
tential site-specific  variation:  ( 1 ) constant  sur- 
vival and  recapture  probability  across  sites, 
(2)  variable  survival  but  constant  recapture 
probability  across  sites,  (3)  variable  survival 
and  recapture  probability  across  sites,  and  (4) 
constant  survival  but  variable  recapture  prob- 
ability across  sites.  To  select  the  most  appro- 
priate model,  we  employed  Akaike’s  Infor- 
mation Criterion  (AIC)  and  chose  the  model 
with  the  lowest  AIC  value  (Lebreton  et  al. 
1992,  Burnham  and  Anderson  2002).  We  used 
differences  in  AIC  between  that  model  and 
other  models  to  evaluate  the  evidence  in  sup- 
port of  particular  models.  Models  with  AAIC 


TABLE  1.  Total  numbers  of  Black-headed  Gros- 
beaks and  Spotted  Towhees  captured,  by  site  and  time 
period,  Sacramento  Valley,  California,  1993-2000. 

Years 

(dataset) 

Site 

Black-headed 
Grosbeak  («) 

Spotted 
Towhee  (n) 

1993-1995 

Flynn 

29 

30 

Ohm 

39 

17 

Sul  Norte 

85 

37 

1995-2000 

Ohm 

56 

50 

Phelan  Island 

150 

33 

<2  can  be  said  to  exhibit  moderately  strong 
support  relative  to  the  preferred  model;  those 
with  2-4  have  less  support  and  those  with 
>10  have  none  (Burnham  and  Anderson 
2002). 

We  analyzed  the  data  as  a partial  time  series 
because  we  did  not  collect  data  at  all  netting 
sites  in  all  years.  Ohm  was  the  only  site  where 
mist  netting  was  conducted  over  the  course  of 
the  entire  study;  Llynn  and  Sul  Norte  were  run 
from  1993  to  1995  and  Phelan  Island  was  op- 
erated from  1995  to  2000. 

RESULTS 

Overall,  more  Black-headed  Grosbeaks 
were  captured  than  Spotted  Towhees  (Table 
1 ),  but  a slightly  higher  percentage  of  towhees 
was  recaptured  (17.1%  versus  12.3%);  most 
of  our  captures  were  presumed  to  be  tran- 
sients. Captures  were  greatest  for  both  species 
at  Sul  Norte  (1993  to  1995  data);  during  1995 
to  2000,  we  captured  more  grosbeaks  at  Phe- 
lan Island  but  more  towhees  at  Ohm  (Table 
1). 

Black-headed  Grosbeak. — The  model 
where  survival  differed  across  sites  while  re- 
capture probability  was  constant  performed 
best  (AAIC  = 0)  among  the  four  models 
(1993  to  1995  dataset;  Table  2).  Model  3 (both 
survival  and  recapture  probabilities  differed 
across  sites)  did  not  produce  maximum-like- 
lihood estimates;  boundary  estimates  were  1 .0 
for  either  survival  or  recapture  probability  due 
to  the  small  sample  size  (Cooch  et  al.  1996). 
There  was  also  support  for  model  4 (constant 
survival,  recapture  probability  differed  across 
sites;  AAIC  = 0.4;  Table  2).  The  best  model 
from  the  1995  to  2000  dataset  (model  2)  was 
also  the  one  that  supported  site-specific  vari- 
ation in  survival;  there  was  some  support  for 
model  4 as  well  (AAIC  = 1.14;  Table  2). 


Gardali  and  Nur  • SITE-SPECIFIC  SURVIVAL 


181 


TABLE  2.  Survival  and  recapture  probabilities  for  Black-headed  Grosbeaks,  1993-1995  (Flynn,  Ohm,  and 
Sul  Norte)  and  1995-2000  (Ohm  and  Phelan  Island),  Sacramento  Valley,  California.  Models  were  (1)  constant 
survival  and  recapture  probability  across  sites,  (2)  variable  survival  but  constant  recapture  probability  across 
sites,  (3)  variable  survival  and  recapture  probability  across  sites,  and  (4)  constant  survival  but  variable  recapture 
probability  across  sites.  All  = all  sites  combined.  AICM.  = AIC  weights. 


Model 

Survival  estimate 
(phi) 

SE 

Recapture  probability 
estimate  (p) 

SE 

AIC 

AAIC 

AIC„, 

1993- 

-1995 

1 

All:  0.813 

0.327 

All:  0.136 

0.040 

126.63 

1.77 

0.172 

2 

Flynn:  0.797 

0.496 

All:  0.181 

0.102 

124.86 

0 

0.417 

Ohm:  0.158 

0.191 

NAa 

NA 

NA 

NA 

NA 

Sul  Norte:  0.773 

0.131 

NA 

NA 

NA 

NA 

NA 

3 

Flynn:  1.000 

b 

Flynn:  0.1  18 

0.090 

128.42 

3.56 

0.070 

Ohm:  0.031 

0.053 

Ohm:  1.000 

— 

NA 

NA 

NA 

Sul  Norte:  0.692 

0.470 

Sul  Norte:  0.216 

0.134 

NA 

NA 

NA 

4 

All:  0.762 

0.624 

Flynn:  0.173 

0.280 

125.26 

0.40 

0.341 

NA 

NA 

Ohm:  0.032 

0.077 

NA 

NA 

NA 

NA 

NA 

Sul  Norte:  0.189 

0.172 

NA 

NA 

NA 

1995- 

2000 

1 

All:  0.642 

0.088 

All:  0.166 

0.046 

264.99 

13.80 

0.001 

2 

Ohm:  0.088 

0.090 

All:  0.205 

0.054 

251.19 

0 

0.510 

Phelan:  0.664 

0.111 

NA 

NA 

NA 

NA 

NA 

3 

Ohm:  0.020 

0.030 

Ohm:  1.000 

— 

253.06 

1.87 

0.200 

Phelan:  0.666 

0.139 

Phelan:  0.204 

0.054 

NA 

NA 

NA 

4 

All:  0.659 

0.349 

Ohm:  0.017 

0.020 

252.33 

1.14 

0.289 

NA 

NA 

Phelan:  0.208 

0.077 

NA 

NA 

NA 

a NA  = not  applicable. 

b Boundary  estimates  were  1 .0  for  either  survival  or  recapture  probability  due  to  small  sample  size. 


Black-headed  Grosbeak  l l Spotted  Towhee 

1.00 


0.75 

> 

| 0.50 
03 

0.25 


0.00 


FIG.  1.  Site-specific  survival  estimates  for  Black- 
headed Grosbeaks  and  Spotted  Towhees  at  four  sites 
in  the  Sacramento  Valley,  California,  over  two  time 
periods,  1993—1995  and  1995—2000.  For  both  species, 
survival  estimates  from  Model  2 (variable  survival,  but 
constant  recapture  probability  across  sites)  are  pre- 
sented (see  Tables  2 and  3).  Error  bars  are  ± SE. 


Flynn  Ohm  Sul  Norte  Ohm  Phelan  Island 

Site 


Overall,  we  found  site-specific  variation  in 
survival  for  Black-headed  Grosbeaks  within 
the  1993-1995  dataset  (Flynn  = 0.797  ± 
0.496,  Ohm  = 0.158  ± 0.191,  Sul  Norte  = 
0.773  ± 0.131)  and  the  1995-2000  dataset 
(Ohm  = 0.088  ± 0.090,  Phelan  Island  = 
0.664  ±0.111;  Fig.  1,  Table  2). 

To  investigate  whether  including  transients 
would  bias  the  full  datasets  (i.e.,  those  includ- 
ing both  transient  and  site-faithful  individu- 
als), we  analyzed  the  subset  of  data  that  met 
our  requirements  for  site-faithful  birds.  For 
this,  we  only  estimated  constant  survival  and 
constant  recapture  probabilities.  At  the  Sul 
Norte  site  (1993-1995  data),  the  estimate  for 
the  site-faithful  subset  (0.700  ± 0.271)  was 
similar  to  that  of  the  full  dataset  (0.773  ± 
0.131;  Table  2).  If  transients  were  biasing  re- 
sults, the  survival  estimate  for  the  site-faithful 
subset  would  have  been  greater  than  that  for 
the  full  dataset,  but  this  was  not  the  case.  For 
Phelan  Island  (1995-2000),  however,  the  dif- 
ferences in  survival  estimates  (site-faithful 


182 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol  118,  No.  2,  June  2006 


TABLE  3.  Survival  and  recapture  probabilities  for  Spotted  Towhees,  1993-1995  (Flynn,  Ohm,  and  Sul 
Norte)  and  1995-2000  (Ohm  and  Phelan  Island),  Sacramento  Valley,  California.  Models  are  (1)  constant  survival 
and  recapture  probability  across  sites,  (2)  variable  survival  but  constant  recapture  probability  across  sites,  (3) 
variable  survival  and  recapture  probability  across  sites,  and  (4)  constant  survival  but  variable  recapture  proba- 
bility across  sites.  All  = all  sites  combined.  AIC*,  = AIC  weights. 


Model 

Survival  estimate 
(phi) 

SE 

Recapture  probability 
estimate  (p) 

SE 

AIC 

AAIC 

AIC*, 

1993- 

1995 

1 

All:  0.602 

0.240 

All:  0.317 

0.173 

85.06 

0 

0.512 

2 

Flynn:  0.653 

0.365 

All:  0.340 

0.180 

86.56 

1.50 

0.242 

Ohm:  0.214 

0.253 

NAa 

NA 

NA 

NA 

NA 

Sul  Norte:  0.632 

0.258 

NA 

NA 

NA 

NA 

NA 

3 

Flynn:  0.557 

0.501 

Flynn:  0.431 

0.366 

90.18 

5.12 

0.040 

Ohm:  0.083 

0.140 

Ohm:  1.000 

b 

NA 

NA 

NA 

Sul  Norte:  0.752 

0.470 

Sul  Norte:  0.256 

0.212 

NA 

NA 

NA 

4 

All:  0.626 

0.532 

Flynn:  0.375 

0.482 

86.87 

1.81 

0.207 

NA 

NA 

Ohm:  0.104 

0.261 

NA 

NA 

NA 

NA 

NA 

Sul  Norte:  0.324 

0.377 

NA 

NA 

NA 

1995- 

2000 

1 

All:  0.245 

0.11 1 

All:  0.496 

0.271 

68.98 

0 

0.508 

2 

Ohm:  0.248 

0.165 

All:  0.496 

0.272 

70.97 

1.99 

0.188 

Phelan:  0.238 

0.241 

NA 

NA 

NA 

NA 

NA 

3 

Ohm:  0.296 

0.244 

Ohm:  0.378 

0.373 

72.13 

3.15 

0.105 

Phelan:  0.138 

0.170 

Phelan:  1.000 

— 

NA 

NA 

NA 

4 

All:  0.237 

0.241 

Ohm:  0.472 

0.453 

70.85 

1.87 

1.990 

NA 

NA 

Phelan:  0.598 

0.596 

NA 

NA 

NA 

a NA  = not  applicable. 

b Boundary  estimates  were  1.0  for  either  survival  or  recapture  probability  due  to  small  sample  size. 


subset  = 0.739  ± 0.276;  full  dataset  = 0.664 
± 0.1 1 1)  were  consistent  with  the  supposition 
that  transients  could  have  biased  our  survival 
estimates,  but  we  could  not  conclude  with 
confidence  that  this  was  the  case  (Table  2). 

Spotted  Towhee. — Model  1 (constant  sur- 
vival and  recapture  probabilities)  from  the 
1993  to  1995  dataset  received  the  most  sup- 
port (Table  3).  Models  assuming  site  differ- 
ences for  either  survival  or  recapture  proba- 
bilities (but  not  both)  also  received  some  sup- 
port (models  2 and  4).  The  magnitude  of  site 
variation  with  respect  to  survival  (model  2; 
Table  3)  was  large  (a  difference  of  0.41  to 
0.43,  when  comparing  Ohm  with  the  other 
two  sites),  but  the  standard  errors  were  large 
and  overlapping  (Fig.  1,  Table  3). 

Model  1 (constant  survival  and  constant  re- 
capture probability)  was  also  best  (AAIC  = 0) 
in  the  1995  to  2000  dataset  (Table  3).  Like 
those  of  the  earlier  dataset  (1993-1995),  mod- 
els with  site-specific  differences  in  either  sur- 
vival or  recapture  probabilities  (models  2 and 


4)  could  not  be  ruled  out  (Table  3).  The  sur- 
vival estimates  for  both  Ohm  and  Phelan  Is- 
land were  very  low  (model  2;  Fig.  1,  Table 
3).  Furthermore,  the  survival  estimates  for 
Ohm  were  similar  in  both  the  1993-1995  and 
1995-2000  time  periods,  indicating  within- 
site  consistency,  but  Phelan  Island  also  had  a 
low  survival  rate.  Overall,  the  pattern  of  site- 
specific  Spotted  Towhee  survival  observed  in 
the  1993  to  1995  data  was  not  manifest  in  the 
1995  to  2000  data. 

DISCUSSION 

In  addition  to  the  site-specific  variation  we 
found  in  survival  rates,  our  survival  estimates 
differed  from  those  published  elsewhere. 
Based  on  the  best-supported  models,  our  sur- 
vival estimates  for  Black-headed  Grosbeaks  at 
Flynn,  Sul  Norte,  and  Phelan  Island  (0.664  to 
0.797)  were  greater  than  those  calculated 
(with  a modified  Cormack-Jolly-Seber  meth- 
od) by  DeSante  and  O’ Grady  (2000)  from 
1992-1998  data  collected  at  51  mist-net  sta- 


Gardali  and  Nur  • SITE-SPECIFIC  SURVIVAL 


183 


tions  in  northwestern  North  America  (0.573  ± 
0.046  SE)  and  28  mist-net  stations  in  south- 
western North  America  (0.576  ± 0.05 1 SE). 
Survival  estimates  from  the  Ohm  site  (0.088 
and  0.158)  were  considerably  lower.  Of  mod- 
els assuming  constant  survival  across  sites, 
the  best  model  for  all  sites  combined  (model 
4;  both  time  periods)  also  estimated  notably 
higher  survival  rates  than  those  reported  by 
DeSante  and  O’ Grady  (2000).  For  Spotted  To- 
whees,  our  1993  to  1995  survival  estimate 
(0.602)  was  greater  than  those  found  by 
DeSante  and  O’ Grady  (2000)  for  the  north- 
western (34  sites)  and  southwestern  (17  sites) 
regions  (0.519  ± 0.047  SE  and  0.486  ± 0.043 
SE,  respectively),  whereas  our  1995  to  2000 
estimate  (0.245)  was  lower. 

Site  variation  in  survival  was  indicated  for 
Black-headed  Grosbeaks.  There  was  also 
some  evidence  that  Spotted  Towhee  survival 
varied  by  site,  although  the  variation  around 
several  of  these  estimates  was  large  and  over- 
lapping. For  both  species,  survival  estimates 
at  the  Ohm  site  were  low;  this  site  differs  from 
the  others  in  that  cattle  were  grazed  there  dur- 
ing the  entire  study  period.  Grazing  has  the 
potential  to  affect  habitat  quality  in  several 
ways,  which  may  influence  survival  and  em- 
igration probabilities  (Saab  et  al.  1995).  For 
example,  grazing  appeared  to  have  reduced 
the  amount  of  low  shrubby  vegetation  cover 
that  serves  as  protection  from  predators. 
Heightened  predation  pressure  could  negative- 
ly affect  survival  directly  via  adult  mortality, 
and/or  indirectly  via  nest  predation,  whereby 
there  is  a fitness  cost  for  individuals  that  re- 
nest relatively  more  than  other  individuals.  In- 
deed, reproductive  effort  can  affect  adult  sur- 
vival rates  in  some  landbird  species  (Nur 
1988a,  1988b;  McCleery  et  al.  1996;  Cichon 
et  al.  1998).  An  additional  cause  of  reduced 
survival  at  the  grazed  site  could  be  related  to 
food  resources:  fewer  insects  may  have  been 
available  because  of  diminished  or  modified 
foraging  substrates  (but  see  Haas  1998). 

In  general,  the  few  past  studies  conducted 
to  examine  over-summer  survival  (i.e.,  during 
the  breeding  period  itself)  have  revealed  rel- 
atively high  survival  rates  during  this  period 
(Smith  1995,  Lahti  et  al.  1998,  Powell  et  al. 
2000,  Sillett  and  Holmes  2002).  There  is, 
however,  some  evidence  that  subordinate  in- 
dividuals (e.g.,  young  birds)  experience  higher 


rates  of  mortality  during  this  period — primar- 
ily due  to  predation  (Geer  1982,  Smith  1995, 
Powell  et  al.  2000).  Additionally,  recent  evi- 
dence suggests  that  events  during  one  stage  in 
the  annual  cycle  may  influence  the  subsequent 
stage  (Marra  et  al.  1998,  Sillett  et  al.  2000), 
and  that  conditions  in  one  year  may  affect  re- 
productive performance  in  a subsequent  year 
(Nur  1988a).  Hence,  differences  in  survival 
caused  by  differences  in  conditions  on  the 
breeding  grounds  would  not  be  captured  in 
those  studies  limited  to  estimating  survival 
during  the  summer/breeding  season  months 
(Smith  1995,  Lahti  et  al.  1998,  Powell  et  al. 
2000,  Sillett  and  Holmes  2002). 

Spotted  Towhee  survival  was  low  at  both 
Ohm  and  Phelan  Island,  whereas  Black-head- 
ed Grosbeak  survival  was  low  at  Ohm  but  rel- 
atively high  at  Phelan  Island.  Although  we  do 
not  know  why  towhee  survival  was  low  at 
Phelan  Island,  this  is  a restoration  site  and 
some  vegetation  features,  or  other  component 
of  the  ecosystem,  may  have  reduced  survival. 

We  could  not  distinguish  true  mortality 
from  permanent  dispersal.  Therefore,  ob- 
served differences  in  apparent  survival  rates 
may  reflect  site-specific  differences  in  site  fi- 
delity instead  of  true  survival.  Such  differenc- 
es also  suggest  that  habitat  conditions  at  Ohm 
were  relatively  poor  for  grosbeaks,  and  per- 
haps towhees,  because  individuals  did  not  re- 
main faithful  to  that  site  for  multiple  years. 
Several  studies  have  shown  that  individuals 
are  less  likely  to  return  to  a territory  or  site  if 
reproductive  performance  at  that  location  was 
poor  (e.g.,  Harvey  et  al.  1979,  Haas  1998, 
Porneluzi  2003).  Interestingly,  Haas  (1997) 
found  that  Brown  Thrashers  ( Toxostoma  ruf- 
um)  return  significantly  more  to  grazed  sites 
than  to  ungrazed  sites  and  speculated  that 
thrashers  in  grazed  sites  were  able  to  forage 
more  effectively  (e.g.,  preferred  substrate)  and 
maintain  better  body  condition. 

It  is  possible  that  transients  may  have  dom- 
inated captures  at  Ohm,  but  not  at  other  sites. 
Perhaps  the  habitat  configuration  or  net  loca- 
tions at  Ohm  were  more  suitable  for  capturing 
migrating  grosbeaks.  However,  Spotted  Tow- 
hees are  year-round  residents,  yet  they  had 
low  survival  estimates  at  Ohm  as  well.  Also, 
the  Ohm  and  Flynn  sites  are  only  3.2  km 
apart,  making  it  difficult  to  imagine  that  more 
migrants  would  be  captured  at  Ohm.  Further- 


184 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


more,  the  differences  in  survival  rates  be- 
tween these  two  sites  were  so  large  that  it 
seems  unlikely  that  the  cause  would  have  been 
a preponderance  of  transients  at  Ohm  (Table 
2).  Although  transient  composition  may  have 
explained  part  of  the  site  differences  in  sur- 
vival, we  believe  that  the  differences  are  pri- 
marily due  to  true  differences  in  survival.  The 
fact  that  survival  estimates  of  both  species 
were  low  at  Ohm  makes  this  argument  more 
compelling,  and  it  is  likely  that  both  species 
are  being  affected  by  the  same  mortality  fac- 
tors at  that  site. 

The  large  differences  we  found  in  our  sur- 
vival estimates  have  strong  implications  for 
source-sink  dynamics  (Pulliam  1988).  Wheth- 
er a site  is  a source  or  a sink  depends  on  a 
combination  of  adult  survival  and  juvenile  re- 
cruitment. The  low  survival  rates  of  Spotted 
Towhees  at  Ohm  suggest  that  Ohm  is  a sink 
population.  Population  growth  rate  (lambda) 
is  equal  to  the  sum  of  adult  survival  and  net 
recruitment  rate  of  offspring,  which  itself  is  a 
product  of  the  number  of  female  offspring 
produced  per  adult  female  and  the  survival  of 
fledged  offspring  to  breeding  age  (Pulliam 
1988,  see  also  Nur  and  Sydeman  1999). 
Therefore,  a difference  in  adult  survival  of 
0.40  (such  as  that  found  for  Spotted  Towhees) 
will  lead  to  a difference  in  lambda  of  0.40  if 
the  other  parameter  values  are  the  same;  thus, 
if  a population  is  growing  at  10%  per  year  at 
a favorable  site  (lambda  = 1.1),  the  popula- 
tion would  be  declining  at  30%  per  year  at  an 
unfavorable  site,  such  as  Ohm  (lambda  = 
0.70). 

For  Black-headed  Grosbeaks,  the  differenc- 
es in  apparent  survival  between  Ohm  and  the 
other  sites  were  even  greater.  Adult  survival 
rates  of  0.77  to  0.80  at  Flynn  and  Sul  Norte 
(Model  2,  1993  to  1995)  may  be  consistent 
with  those  of  a source  population.  At  Ohm, 
with  an  adult  survival  rate  of  0.19,  it  would 
not  be  possible  for  a population  to  yield  a 
lambda  of  1.0  or  greater  (Model  2,  1993  to 
1995),  even  using  the  most  optimistic  param- 
eter values  for  reproductive  success  and  off- 
spring survival.  As  a result,  the  Ohm  popu- 
lation may  be  a sink  due  to  low  adult  survival, 
irrespective  of  reproductive  success.  Alterna- 
tively, emigration  rates  may  be  high  at  Ohm 
because  it  is  a reproductively  inferior  site  re- 
sulting in  a high  turnover  of  individuals.  In 


this  case,  survival  could  be  as  great  at  Ohm 
as  at  the  other  sites  and  Ohm  would  contribute 
to  the  overall  metapopulation  (Howe  et  al. 
1991). 

It  is  common  for  researchers  modeling  site-, 
treatment-,  or  habitat-specific  lambda  to  use  a 
single  survival  estimate  in  combination  with 
several  reproductive  estimates  (e.g.,  Donovan 
et  al.  1995,  Manolis  et  al.  2002).  The  practice 
has  been  to  use  a single  survival  estimate  from 
one  site  (not  necessarily  derived  from  the 
study  area),  regional  estimates  that  combine 
several  sites,  or  a mean  of  published  esti- 
mates. This  is  understandable  because  site- 
specific  survival  is  difficult  to  estimate, 
whereas  estimating  nest  survival  is  relatively 
easy.  Our  results,  however,  emphasize  the 
need  to  combine  site-,  treatment-,  or  habitat- 
specific  estimates  of  adult  survival  with  com- 
parable estimates  of  nest  survival  when  mod- 
eling population  viability. 

ACKNOWLEDGMENTS 

We  are  indebted  to  the  many  biologists  that  helped 
to  collect  these  data,  especially  A.  M.  King  and  S.  L. 
Small.  We  would  like  to  thank  PRBO  staff;  G.  R.  Geu- 
pel  and  G.  Ballard  both  contributed  to  this  study  in 
important  ways.  Funds  for  this  monitoring  project 
came  from  several  sources:  The  Nature  Conservancy, 
U.S.  Fish  and  Wildlife  Service,  the  David  and  Lucile 
Packard  Foundation,  the  William  and  Flora  Hewlett 
Foundation,  the  National  Fish  and  Wildlife  Founda- 
tion, CALFED  Bay/Delta  Program,  the  U.S.  Bureau  of 
Reclamation,  the  Natural  Resource  Conservation  Ser- 
vice, and  River  Partners.  We  are  deeply  grateful  to  all 
of  them.  We  especially  thank  our  many  partners  for 
long-term  support:  J.  Carlon,  D.  Efseaff,  G.  Golet,  S. 
Lawson,  G.  Mensik,  S.  Phelps,  J.  Silveira,  R.  Vega,  T. 
Zimmerman,  and  D.  Zeleke.  We  thank  G.  H.  Golet,  C. 
A.  Haas,  and  two  anonymous  reviewers  for  helpful 
comments  on  this  manuscript.  This  is  PRBO  contri- 
bution # 1297. 

LITERATURE  CITED 

Alpert,  R,  F.  T.  Griggs,  and  D.  R.  Peterson.  1999. 
Riparian  forest  restoration  along  large  rivers:  ini- 
tial results  from  the  Sacramento  River  Project. 
Restoration  Ecology  7:360-368. 

Askins,  R.  A.,  J.  F.  Lynch,  and  R.  Greenberg.  1990. 
Population  declines  in  migratory  birds  in  eastern 
North  America.  Current  Ornithology  7:1-57. 
Bollinger,  E.  K.  and  T.  A.  Gavin.  1989.  The  effects 
of  site  quality  on  breeding-site  fidelity  in  Bobo- 
links. Auk  106:584-594. 

Burnham,  K.  P.  and  D.  R.  Anderson.  2002.  Model 
selection  and  multimodel  inference:  a practical  in- 


Gardali  and  Nur  • SITE-SPECIFIC  SURVIVAL 


185 


formation-theoretic  approach,  2nd  ed.  Springer, 
New  York. 

Bryant,  D.  M.  1979.  Reproductive  costs  in  the  House 
Martin  ( Delichon  urbica).  Journal  of  Animal 
Ecology  48:655-675. 

Chase,  M.,  N.  Nur,  and  G.  R.  Geupel.  1997.  Survival, 
productivity,  and  abundance  of  a Wilson’s  Warbler 
population.  Auk  1 14:354-366. 

Cichon,  M.,  P.  Olejniczak,  and  L.  Gustafsson.  1998. 
The  effect  of  body  condition  on  the  cost  of  repro- 
duction in  female  Collared  Flycatchers  Ficedula 
albicollis.  Ibis  140:128-130. 

Cooch,  E.,  R.  Pradel,  and  N.  Nur.  1996.  A practical 
guide  to  capture/recapture  analysis  using  SURGE. 
Centre  d’Ecologie  Fonctionnelle  et  Evolutive- 
CNRS,  Montpellier,  France. 

DeSante,  D.  E,  K.  M.  Burton,  P.  Velez,  and  D. 
Froehlich.  2000.  MAPS  manual:  2000  protocol. 
The  Institute  for  Bird  Populations,  Point  Reyes 
Station,  California. 

DeSante,  D.  F.  and  D.  R.  O’Grady.  2000.  The  Mon- 
itoring Avian  Productivity  and  Survivorship 
(MAPS)  program  1997  and  1998  report.  Bird  Pop- 
ulations 5:49-102. 

Donovan,  T.  M.,  F.  R.  Thompson,  III,  J.  Faaborg,  and 
J.  R.  Probst.  1995.  Reproductive  success  of  mi- 
gratory birds  in  habitat  sources  and  sinks.  Con- 
servation Biology  9:1380-1395. 

Gardali,  T.,  G.  Ballard,  N.  Nur,  and  G.  R.  Geupel. 
2000.  Demography  of  a declining  population  of 
Warbling  Vireos  in  coastal  California.  Condor 
102:601-610. 

Gardali,  T.,  A.  L.  Holmes,  S.  L.  Small,  N.  Nur,  G. 
R.  Geupel,  and  G.  H.  Golet.  In  press.  Abundance 
patterns  of  landbirds  in  restored  and  remnant  ri- 
parian forests  on  the  Sacramento  River,  California, 
U.S.A.  Restoration  Ecology. 

Geer,  T.  A.  1982.  The  selection  of  tits  Parus  spp.  by 
Sparrowhawks  Accipiter  nisus.  Ibis  124:159-167. 

Haas,  C.  A.  1997.  What  characteristics  of  shelterbelts 
are  important  to  breeding  success  and  return  rates 
of  birds.  American  Midland  Naturalist  137:225- 
238. 

Haas,  C.  A.  1998.  Effects  of  prior  nesting  success  on 
site  fidelity  and  breeding  dispersal:  an  experimen- 
tal approach.  Auk  115:929-936. 

Harvey,  P.  H.,  P.  J.  Greenwood,  and  C.  M.  Perrins. 
1979.  Breeding  area  fidelity  of  Great  Tits  ( Parus 
major).  Journal  of  Animal  Ecology  48:305-313. 

Howe,  R.  W.,  G.  J.  Davis,  and  V.  Mosca.  1991.  The 
demographic  significance  of  ‘sink’  populations. 
Biological  Conservation  57:239-255. 

Lahti,  K.,  M.  Orell,  S.  Rytkonen,  and  K.  Koivula. 
1998.  Time  and  food  dependence  in  Willow  Tit 
winter  survival.  Ecology  79:2904-2916. 

Lebreton,  J.-D.,  K.  P.  Burnham,  J.  Clobert,  and  D. 
D.  Anderson.  1992.  Modeling  survival  and  test- 
ing biological  hypotheses  using  marked  animals: 
a unified  approach  with  case  studies.  Ecological 
Monographs  62:67-118. 

Manolis,  J.  C.,  D.  E.  Anderson,  and  F.  J.  Cuthbert. 


2002.  Edge  effect  on  nesting  success  of  ground 
nesting  birds  near  regenerating  clearcuts  in  a for- 
est-dominated landscape.  Auk  119:955-970. 

Marra,  P P,  K.  A.  Hobson,  and  R.  T.  Holmes.  1998. 
Linking  winter  and  summer  events  in  a migratory 
bird  by  using  stable-carbon  isotopes.  Science  282: 
1884-1886. 

McCleery,  R.  H.,  J.  Clobert,  R.  Julliard,  and  C.  M. 
Perrins.  1996.  Nest  predation  and  delayed  cost  of 
reproduction  in  the  Great  Tit.  Journal  of  Animal 
Ecology  65:96-104. 

Moore,  F.  R.,  S.  A.  Gauthreaux,  Jr.,  P.  Kerlinger, 
and  T.  R.  Simons.  1995.  Habitat  requirements 
during  migration:  important  link  in  conservation. 
Pages  121-144  in  Ecology  and  management  of 
Neotropical  migratory  birds:  a synthesis  and  re- 
view of  critical  issues  (T.  E.  Martin  and  D.  M. 
Finch,  Eds.).  Oxford  University  Press,  New 
York. 

Nur,  N.  1988a.  The  consequences  of  brood  size  for 
breeding  Blue  Tits.  III.  Measuring  the  cost  of  re- 
production: survival,  future  fecundity  and  differ- 
ential dispersal.  Evolution  42:351-362. 

Nur,  N.  1988b.  The  cost  of  reproduction  in  birds:  an 
examination  of  the  evidence.  Ardea  76:151-162. 

Nur,  N.  and  J.  Clobert.  1988.  Measuring  Darwinian 
fitness  in  birds:  a field  guide.  Pages  2121-2130  in 
Acta  XIX  Congressus  Internationalis  Ornitholo- 
gici  (H.  Ouellet,  Ed.).  University  of  Ottawa  Press, 
Ottawa,  Ontario,  Canada. 

Nur,  N.  and  G.  R.  Geupel.  1993.  Evaluation  of  mist- 
netting,  nest-searching  and  other  methods  for 
monitoring  demographic  processes  in  landbird 
populations.  Pages  237-244  in  Status  and  man- 
agement of  Neotropical  migrant  birds  (D.  Finch 
and  P.  Stangel,  Eds.).  General  Technical  Report 
RM-229,  USDA  Forest  Service,  Fort  Collins,  Col- 
orado. 

Nur,  N.,  G.  R.  Geupel,  and  G.  Ballard.  2000.  The 
use  of  constant-effort  mist-netting  to  monitor  de- 
mographic processes  in  passerine  birds:  annual 
variation  in  survival,  productivity  and  floaters. 
Pages  185—194  in  Strategies  for  bird  conserva- 
tion: the  partners  in  flight  planning  process  (R. 
Bonny,  D.  N.  Pashley,  R.  J.  Cooper,  and  L.  Niles, 
Eds.).  Proceedings  RMRS-P-16,  USDA  Forest 
Service,  Rocky  Mountain  Research  Station,  Og- 
den, Utah. 

Nur,  N.  and  W.  J.  Sydeman.  1999.  Demographic  pro- 
cesses and  population  dynamic  models  of  sea- 
birds: implications  for  conservation  and  restora- 
tion. Current  Ornithology  15:149-188. 

Peach,  W.  1993.  Combining  mark-recapture  data  sets 
for  small  passerines.  Pages  107-122  in  Marked 
individuals  in  the  study  of  bird  populations  (J.-D. 
Lebreton  and  P.  M.  North,  Eds.).  Birkhauser,  Ba- 
sel, Switzerland. 

Peach,  W.,  S.  Baillie,  and  L.  G.  Underhill.  1991. 
Survival  of  British  Sedge  Warblers  Acrocephalus 
schoenobaenus  in  relation  to  West  African  rain- 
fall. Ibis  133:300-305. 


186 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Porneluzi,  P.  A.  2003.  Prior  breeding  success  affects 
return  rates  of  territorial  male  Ovenbirds.  Condor 
105:73-79. 

Powell,  L.  A.,  J.  D.  Lang,  M.  J.  Conroy,  and  D.  G. 
Krementz.  2000.  Effects  of  forest  management  on 
density,  survival,  and  population  growth  of  Wood 
Thrushes.  Journal  of  Wildlife  Management  64: 1 1- 
23. 

Pradel,  R.,  J.  E.  Hines,  J.-D.  Lebreton,  and  J.  D. 
Nichols.  1997.  Capture-recapture  survival  models 
taking  account  of  transients.  Biometrics  53:60-72. 

Pulliam,  H.  R.  1988.  Sources,  sinks,  and  population 
regulation.  American  Naturalist  132:652-661. 

Rappole,  J.  H.  and  M.  V.  McDonald.  1994.  Cause 
and  effect  in  population  declines  of  migratory 
birds.  Auk  1 1 1 :652-660. 

Saab,  V.  A.,  C.  E.  Bock,  T.  D.  Rich,  and  D.  S.  Dob- 
kin.  1995.  Livestock  grazing  effects  in  western 
North  America.  Pages  311-353  in  Ecology  and 
management  of  Neotropical  migratory  birds:  a 


synthesis  and  review  of  critical  issues  (T.  E.  Mar- 
tin and  D.  M.  Finch,  Eds.).  Oxford  University 
Press,  New  York. 

Sedgwick,  J.  A.  2004.  Site  fidelity,  territory  fidelity, 
and  natal  philopatry  in  Willow  Flycatchers  ( Em - 
pidonax  traillii).  Auk  121:1103-1121. 

Sillett,  T.  S.  AND  R.  T.  Holmes.  2002.  Variation  in 
survivorship  of  a migratory  songbird  throughout 
its  annual  cycle.  Journal  of  Animal  Ecology  71: 
296-308. 

Sillett,  T.  S.,  R.  T.  Holmes,  and  T.  W.  Sherry.  2000. 
Impacts  of  a global  climate  cycle  on  population 
dynamics  of  a migratory  songbird.  Science  288: 
2040-2042. 

Smith,  S.  M.  1995.  Age-specific  survival  in  breeding 
Black-capped  Chickadees  ( Parus  atricapillus ). 
Auk  112:840-846. 

Yong,  W.,  D.  M.  Finch,  F.  R.  Moore,  and  J.  F.  Kelly. 
1998.  Stopover  ecology  and  habitat  use  of  migra- 
tory Wilson’s  Warblers.  Auk  115:829-842. 


The  Wilson  Journal  of  Ornithology  1 18(2):  187— 193,  2006 


PRE-MIGRATORY  FATTENING  AND  MASS  GAIN  IN 
FLAMMULATED  OWLS  IN  CENTRAL  NEW  MEXICO 

john  p.  Delong1  2 


ABSTRACT. — Hatching-year  (HY)  and  presumed  HY  Flammulated  Owls  (Otus  flammeolus)  were  captured 
during  a period  of  pre-migratory  activity  in  central  New  Mexico  from  2000  to  2003.  Mass  gains  were  evident 
through  the  pre-migratory  period.  Fat  deposition  was  an  important  component  of  these  mass  gains;  muscle 
growth  appeared  to  contribute  to  a lesser  degree.  Fat  scores  and  pectoral-muscle  scores  were  positively  related 
to  body  mass  and  to  each  other,  and,  from  first  to  last  capture,  most  recaptured  owls  showed  increases  in  body 
mass  that  were  accompanied  by  fat  deposition  and  growth  in  pectoral  muscles.  These  data  add  to  a growing 
body  of  research  indicating  that  pre-migration  increases  in  fat  and  muscle  mass  may  be  interdependent,  but  the 
magnitude  of  increased  muscle  mass  may  be  too  small  to  be  detected  at  certain  scales.  Received  4 February 
2005,  accepted  26  November  2005. 


Many  migratory  birds  show  substantial 
gains  in  body  mass  prior  to  migration  (King 
1972,  Bairlein  2002).  These  gains  typically 
represent  some  combination  of  growth  in  fat, 
muscle,  and  organ  tissues  (King  1972,  Lind- 
strom and  Piersma  1993,  Bairlein  2002).  Fat 
is  a major  component  of  internal  energy  re- 
serves and  it  can  be  catabolized  during  mi- 
gratory flights  (King  1972).  The  amount  of  fat 
stored  appears  to  vary  in  relation  to  the  ex- 
pected travel  distance,  opportunities  to  refuel, 
and  predation  risk  en  route  (King  1972,  Al- 
erstam  and  Lindstrom  1990,  Bairlein  2002). 
Increases  in  muscle  size  appear  to  have  a two- 
fold role:  to  increase  the  power  output  from 
the  wings  (specifically  for  pectoral  muscles) 
and  to  provide  a source  of  amino  acids  and 
water  as  they  are  catabolized  during  flight 
(Marsh  1984,  Pennycuick  1998,  Lindstrom  et 
al.  2000,  Bairlein  2002).  Increases  in  the  size 
of  digestive  organs  facilitate  more  rapid  up- 
take of  nutrients,  aiding  in  fat  storage  and  the 
growth  of  pectoral  and  other  muscles.  When 
not  in  use,  the  digestive  organs  themselves 
may  provide  additional  nutrient  sources  as 
they  are  catabolized  (Karasov  and  Pinshow 
1998,  Piersma  et  al.  1999). 

The  masses  of  fat  and  non-fat  tissues  often 
are  correlated  with  overall  body  mass,  but  it 
is  not  clear  that  changes  in  masses  of  fat  and 
lean  tissues  are  interdependent  (Gosler  1991; 


1 HawkWatch  International,  Inc.,  1800  S.  West  Tem- 
ple, Ste.  226,  Salt  Lake  City,  UT  84115,  USA. 

2 Current  address:  Eagle  Environmental,  Inc.,  2314 
Hollywood  Ave.  NW,  Albuquerque,  NM  87104,  USA; 
e-mail:  jpdelong@comcast.net 


Selman  and  Houston  1996;  Redfern  et  al. 
2000,  2004).  Because  changes  in  mass  are  re- 
lated to  foraging  and  behavioral  patterns  be- 
fore migration  and  during  migration  stop- 
overs, understanding  how  lean  and  fat  tissues 
contribute  to  changes  in  mass  in  migratory 
birds  may  help  to  elucidate  important  aspects 
of  migratory  bird  ecology  (Karasov  and  Pin- 
show  1998,  Bairlein  2002).  The  concurrent 
study  of  fat  deposition,  muscle  hypertrophy, 
and  mass  gain  prior  to  migration  has  received 
little  attention  in  field  studies,  probably  be- 
cause carcass  analysis  is  usually  required 
(e.g.,  Redfern  et  al.  2000).  Although  carcass 
analysis  can  provide  precise  measurements, 
samples  sizes  are  often  small  because  birds 
must  be  killed  for  analysis.  Scoring  body 
composition  does  not  require  killing  birds  and 
it  confers  the  possibility  of  adequate  sample 
sizes  (Redfern  et  al.  2004). 

Flammulated  Owls  ( Otus  flammeolus ) are 
small,  insectivorous  birds  that  breed  in  the 
montane  forests  of  western  North  America 
and  Mexico  (McCallum  1994).  The  species  is 
believed  to  winter  in  southern  Mexico  and 
Guatemala;  thus,  it  is  considered  by  most 
sources  to  be  a Neotropical  migrant,  undertak- 
ing potentially  long  flights  between  summer- 
ing and  wintering  areas  (McCallum  1994). 
During  the  falls  of  2000-2003,  I examined  the 
interrelationships  among  mass  gain,  fat  de- 
position, and  the  size  of  pectoral  muscles  in 
Flammulated  Owls  captured  in  central  New 
Mexico.  My  coworkers  and  I captured  Flam- 
mulated Owls  from  late  August,  when  hatch- 
ing-year (HY)  birds  become  independent  from 
their  parents  (Linkhart  and  Reynolds  1987), 


187 


188 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


through  October,  when  birds  begin  their 
southward  migration.  These  capture  efforts 
were  part  of  a larger  study  on  the  migration 
ecology  of  Flammulated  Owls  (DeLong  2004, 
DeLong  et  al.  2005).  Based  on  stable  hydro- 
gen isotope  analysis  of  feathers  and  the  stage 
of  the  preformative  molt,  most  of  the  owls  had 
not  traveled  far  from  their  natal  areas  (De- 
Long 2004,  DeLong  et  al.  2005).  During  latter 
stages  of  our  field  seasons,  we  captured  some 
migrants  that  had  come  from  latitudes  north 
of  central  New  Mexico,  but  they  were  few  in 
number.  Hence,  our  sampling  period  was  a 
post-independence/pre-migration  period  for 
owls  that  had  summered  in  central  New  Mex- 
ico. Using  this  sample,  I tested  the  hypothesis 
that  fat  and  muscle  tissue  growth  simulta- 
neously contribute  to  overall  mass  gain  in 
Flammulated  Owls  prior  to  their  southward 
migration. 

METHODS 

The  study  site  was  located  near  Capilla 
Peak  in  the  Manzano  Mountains  of  central 
New  Mexico  (34°42'N,  106°  24' W).  The 
Manzano  Mountains  are  part  of  an  important 
migratory  corridor  for  many  raptors  and  song- 
birds that  move  through  New  Mexico  during 
the  fall  (see  DeLong  and  Hoffman  [1999]  and 
DeLong  et  al.  [2005]  for  additional  details). 
My  coworkers  and  I set  up  two  mist-netting 
stations,  spaced  ~200  m apart,  one  on  each 
side  of  the  north-south  trending  Capilla  Peak 
ridge.  We  lured  owls  to  the  stations  by  broad- 
casting the  territorial  breeding-season  hoots  of 
the  male  Flammulated  Owl  from  within  arrays 
of  3-6  mist  nets  (60-mm  mesh).  From  18  Au- 
gust to  22  October,  we  opened  mist  nets  3—7 
nights/week,  depending  on  volunteer  support 
and  weather.  We  typically  began  netting  0—30 
min  after  sunset  and  continued  until  15-30 
min  before  sunrise.  We  closed  the  nets  when 
winds  exceeded  —24  km/hr  or  when  precipi- 
tation began  to  fall.  We  checked  nets  for  cap- 
tured owls  every  40-70  min. 

We  banded  owls  with  federal  aluminum  leg 
bands,  used  an  electronic  scale  to  determine 
their  mass  to  the  nearest  0.1  g,  and  used  a 
standard  wing  chord  ruler  to  measure  their  un- 
flattened wing  chords  to  the  nearest  1 mm.  To 
determine  whether  body  mass  and  other  pa- 
rameters of  males  and  females  differed,  we 
obtained  blood  samples  or  feather  shafts  from 


randomly  selected  HY  owls  and  sent  them  to 
Wildlife  Genetics,  Inc.  (Nelson.  British  Co- 
lumbia, Canada;  www.wildlifegenetics.com) 
for  DNA  analysis  (CHD  gene  method;  Grif- 
fiths et  al.  1998). 

Whenever  possible,  we  aged  owls  as  either 
HY  or  adult.  We  identified  HY  owls  by  the 
presence  of  retained  juvenal  plumage  (De- 
Long 2004)  or  by  uniform  fault-barring  (Pyle 
1997).  We  identified  adult  owls  by  the  pres- 
ence of  multiple  generations  of  flight  feathers. 
For  the  analyses  in  this  paper,  I excluded 
adults  because  their  body  mass  was  signifi- 
cantly greater  (HY  mean  mass  — 53.9  g,  n — 
124;  adult  mean  mass  — 59.9  g , n = 13;  t = 
4.7,  P < 0.001)  and  adults  were  not  captured 
frequently  enough  to  analyze  separately.  The 
analyses  included  both  confirmed  and  pre- 
sumed HY  owls.  I presumed  that  owls  of  un- 
known age  were  HY  birds  if  they  were  molt- 
ing their  contour  feathers,  had  only  a single 
generation  of  flight  feathers,  and  weighed  less 
than  the  mean  weight  for  adults.  Most  adult 
Flammulated  Owls  finish  molting  their  flight 
feathers  by  late  September  (Reynolds  and 
Linkhart  1987),  in  which  case  they  too  would 
have  had  a single  generation  of  flight  feathers 
during  our  study  period;  thus,  it  is  possible 
that  some  adult  birds  were  misidentified  as 
hatching-year  birds.  For  two  reasons,  howev- 
er, I believe  the  number  of  adults  included  in 
the  analyses  is  small.  First,  most  unknown-age 
owls  were  captured  before  October  (74%  of 
128  unknown-age  owls)  and  thus  would  likely 
show  multiple  generations  of  flight  feathers  if 
adult.  Second,  nearly  all  of  these  birds  were 
captured  before  we  were  able  to  use  the  re- 
tained-plumage  criterion  for  identifying  hatch- 
ing-year owls;  therefore,  these  owls  were  la- 
beled unknown-age  only  because  they  did  not 
show  multiple  generations  of  flight  feathers, 
not  because  they  lacked  retained  juvenal 
plumage.  We  did  not  know  to  look  for  these 
feathers  in  the  early  years,  but  learned  to  do 
so  as  the  study  progressed  (DeLong  2004).  As 
the  study  progressed,  it  became  clear  that 
adult  owls  were  rarely  captured  at  our  study 
site  (JPD  unpubl.  data). 

We  used  a 5-point  scoring  technique  to  vi- 
sually assess  the  size  of  pectoral  muscles.  The 
pectoral-muscle  score  was  based  on  thickness 
(roughly  a cross-section),  as  follows:  1 = 
muscle  very  concave  with  keel  of  sternum 


DeLong  • PRE-MIGRATORY  MASS  GAIN  IN  FLAMMULATED  OWLS 


189 


protruding  sharply,  2 = muscle  roughly  tri- 
angle-shaped  with  keel  protruding  sharply,  3 
= rounded  muscle  with  keel  still  protruding 
just  slightly  above  the  muscle  level,  4 = mus- 
cle rounded  and  flush  with  keel,  and  5 = mus- 
cle depth  exceeds  (bulges  beyond)  the  keel. 
The  cross-sectional  shape  of  pectoral  muscles 
is  positively  correlated  with  the  pectoral  mass 
in  small  birds  (Selman  and  Houston  1996); 
therefore,  visual  assessments  of  the  cross-sec- 
tional shape  of  pectoral  muscles  should  pro- 
vide a suitable  index  of  pectoral-muscle  size. 
A similar  approach  has  been  used  effectively 
in  studies  of  songbirds  (Gosler  1991). 

We  visually  assessed  furcular  fat  deposits 
(i.e.,  the  claviculo-coracoid  fat  body  described 
by  King  and  Famer  1965)  using  a 6-point 
scoring  technique  similar  to  that  of  Helms  and 
Drury  (1960).  The  furcular  fat  score  reflected 
the  depth  of  fat  in  the  furculum:  0 = no  fat, 
1 = furculum  1-5%  filled  with  fat,  2 = 5- 
33%  filled,  3 = 34-66%  filled,  4 = 67-100% 
filled,  and  5 = fat  bulging  above  furculum. 
Subcutaneous  fat  in  this  region  is  correlated 
with  overall  body  fat  in  small  birds — as  are 
fat-scoring  procedures,  which  are  based  at 
least  partly  upon  it  (Krementz  and  Pendleton 
1990;  Rogers  1991;  Redfern  et  al.  2000, 
2004).  We  assigned  pectoral-muscle  and  fur- 
cular fat  scores  to  recaptured  birds  without 
reference  to  original  capture  records. 

This  study  incorporated  data  from  350  cap- 
tures, including  9 birds  recaptured  in  the  same 
season;  however,  sample  sizes  for  some  anal- 
yses were  <350  because  we  did  not  record  all 
of  the  necessary  measurements  for  all  birds.  I 
used  r-tests  and  Kolmogorov-Smimov  tests  to 
evaluate  whether  males  and  females  differed 
in  body  composition  variables.  I used  linear 
regression  to  evaluate  the  effect  of  capture 
date  on  body  mass  and  fat  and  pectoral-mus- 
cle scores.  I used  analysis  of  covariance  (AN- 
COVA)  to  evaluate  the  relationship  of  fat 
score  and  body  mass,  with  wing  chord  length 
and  pectoral-muscle  score  as  covariates.  I also 
evaluated  the  relationship  of  pectoral-muscle 
score  and  body  mass,  with  wing-chord  length 
and  fat  score  as  covariates.  These  two  analy- 
ses allowed  me  to  produce  mass  estimates  for 
each  level  of  each  score,  having  controlled  for 
the  effects  of  the  other  tissue  type  and  size. 
Statistical  tests  were  conducted  with  NCSS 


FIG.  1 . Body  mass  (g)  of  hatching-year  Flammu- 
lated  Owls  increased  in  relation  to  capture  date,  show- 
ing the  gradual  gains  in  body  mass  through  the  pre- 
migration season.  The  dashed  lines  represent  the  95% 
confidence  interval  for  the  regression  line  (solid  line). 
Owls  were  captured  during  fall  at  Capilla  Peak,  New 
Mexico,  2000-2003. 

2004  (Hintze  2001)  and  considered  significant 
if  P < 0.05. 

RESULTS 

The  number  of  owls  captured  varied  annu- 
ally— 89  owls  were  captured  in  2000,  157  in 
2001,  85  in  2002,  and  19  in  2003.  Of  these 
350  owls,  our  first  capture  was  on  19  August 
and  our  last  capture  was  on  18  October,  with 
a median  capture  date  of  17  September. 

Of  the  88  owls  whose  sex  was  determined, 
37  were  female  and  51  were  male.  Females 
and  males  did  not  differ  in  body  mass  ( t = 
1.04,  P = 0.30,  n = 88),  fat  score  (Z  = -0.66, 
P = 0.51,  n = 85),  or  pectoral-muscle  score 
(Z  = 0.50,  P = 0.62,  n = 88).  Therefore,  I 
combined  data  for  males  and  females  in  all 
further  analyses. 

Body  mass  increased  through  the  season  in 
all  years,  but  capture  date  explained  only  a 
small  proportion  of  the  variation  in  body  mass 
(. R 2 = 0.06,  P < 0.001,  n = 350;  Fig.  1).  Body 
mass  was  significantly  lower  in  2000  than  in 
2001-2003  (F3<346  = 46.4,  P < 0.001,  n = 
350),  but  there  was  no  body  mass  X date  in- 
teraction and  no  effect  on  the  overall  pattern 
of  mass  change.  Fat  scores  also  increased 
through  the  season  ( R 2 = 0.19,  P < 0.001; 
Fig.  2).  There  was  a drop  in  fat  scores  in  mid- 


190 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


FIG.  2.  Mean  ± SE  fat  scores  (filled  circles)  and 
pectoral-muscle  scores  (unfilled  circles)  in  relation  to 
capture  date  for  Flammulated  Owls  captured  during 
fall  at  Capilla  Peak,  New  Mexico,  2000-2003.  Fat 
scores  increased  through  the  season,  but  pectoral-mus- 
cle scores  did  not.  Dates  were  grouped  into  5-day  pe- 
riods from  18  August  to  22  October. 

September  (Fig.  2),  but  fat  scores  continued 
to  increase  after  that  time.  Pectoral-muscle 
scores  did  not  change  through  the  season  ( R 2 
- 0.0,  P = 0.34;  Fig.  2). 

ANCOVA  revealed  that  fat  scores  and  pec- 
toral-muscle scores  were  both  related  posi- 
tively to  body  mass  (Table  1).  Wing  chord 
length  was  a strong  predictor  of  body  mass, 
and  fat  score  was  a stronger  predictor  of  body 
mass  than  pectoral-muscle  score  (Table  1). 
Based  on  least-square  means  of  fat  scores 
(ANCOVA),  increments  in  mass  from  one  fat 
score  to  the  next  ranged  from  1.0  to  1.8  g and 
spanned  7.0  g overall  (difference  in  least- 
square  mean  mass  of  fat  scores  0 and  5;  Table 
2).  Mass  increments  from  one  pectoral-muscle 
score  to  the  next  ranged  from  0.3  to  1.1  g but 


TABLE  2.  Least-square  mean  (as  derived  from 
ANCOVA,  see  Table  1)  body  mass  and  body  mass 
gain  from  one  score  to  the  next  for  furcular  fat  and 
pectoral-muscle  scores  of  Flammulated  Owls  at  Cap- 
illa Peak,  New  Mexico,  2000-2003. 


Scoring 

regime 

n 

Mass  (g) 

SE 

Gain  in 
mass  (g) 

Fat 

0 

3 

51.03 

1.97 

a 

1 

61 

52.45 

0.44 

1.4 

2 

50 

53.59 

0.48 

1.1 

3 

79 

54.63 

0.38 

1.0 

4 

61 

56.45 

0.44 

1.8 

5 

6 

58.07 

1.39 

1.6 

Muscleb 

2 

10 

53.29 

1.08 

— 

3 

83 

53.60 

0.37 

0.3 

4 

125 

54.73 

0.30 

1.1 

5 

42 

55.16 

0.53 

0.4 

a Gain  in  mass  not  calculated  for  lowest  fat  and  muscle  class. 
b No  birds  had  a pectoral-muscle  score  of  1 . 


spanned  only  1 .9  g overall  (difference  in  least- 
square  mean  mass  of  pectoral-muscle  scores  2 
and  5). 

Based  on  the  mean  body  mass  of  the  first 
10%  of  captured  owls  and  that  of  the  last  10% 
captured,  the  overall  mass  gain  from  the  be- 
ginning to  the  end  of  the  season  was  2.5  g,  or 
4.8%  of  initial  body  mass,  and  the  mean  fat 
score  increased  from  1.5  to  3.2.  Using  the  data 
in  Table  2,  I estimated  that  fat  mass  increased 
by  2 g over  the  sampling  period,  or  approxi- 
mately 80%  of  the  total  mass  increase  (i.e., 
body  mass  of  a bird  with  a fat  score  of  3.2 
[—55  g]  — body  mass  of  a bird  with  a fat 
score  of  1 .5  [—53  g]  = a 2-g  increase  in  fat). 
In  contrast,  pectoral-muscle  scores  averaged 


TABLE  1.  Results  of  analyses  of  covariance  evaluating  the  relationships  of  fat  and  pectoral-muscle  scores 
versus  body  mass  in  Flammulated  Owls  captured  during  fall  at  Capilla  Peak,  New  Mexico,  2000-2003. 


Analysis/Factor 

df 

F 

p 

Fat  score  as  main  factor 

Fat  score 

5 

9.45 

<0.001 

Pectoral-muscle  score  (covariate) 

1 

6.82 

0.009 

Wing  chord  length  (covariate) 

1 

33.29 

<0.001 

Pectoral-muscle  score  as  main  factor 

Pectoral-muscle  score 

3 

2.54 

0.057 

Fat  score  (covariate) 

1 

44.86 

<0.001 

Wing  chord  length  (covariate) 

1 

36.78 

<0.001 

DeLong  • PRE-MIGRATORY  MASS  GAIN  IN  FLAMMULATED  OWLS 


191 


TABLE  3.  Nine  within-season  recaptures  of  Flam- 
mulated  Owls  indicating  changes  in  mass  and  body 
condition  indices,  Capilla  Peak,  New  Mexico,  2000- 


2003. 

Year 

Initial 

capture 

date 

Days  to 
next 
capture 

Change  in 
mass  (g) 

Change  in 
fat  score 

Change 
in  muscle 
score 

2000 

9 Sep 

19 

+ 5.0 

+ 1 

— a 

2000 

10  Sep 

5 

+ 2.0 

+ 1 

— 

2000 

9 Sep 

18 

0.0 

— 

— 

2000 

30  Sep 

14 

+5.0 

— 

— 

2001 

2 Sep 

21 

-2.6 

0 

0 

2001 

25  Sep 

1 

+0.3 

0 

0 

2001 

30  Sep 

2 

+ 1.0 

+ 1 

+ 1 

2002 

19  Aug 

34 

+ 1.7 

+2 

+ 1 

2003 

5 Sep 

9 

+2.7 

+ 1 

+ 1 

a Data  not  available. 


3.5  among  both  the  first  10%  and  the  last  10% 
of  birds  captured. 

All  but  two  of  the  owls  recaptured  later  in 
the  same  season  ( n = 9)  increased  in  body 
mass  between  the  initial  and  second  capture, 
and  three  of  the  owls  exhibited  simultaneous 
increases  in  fat  and  pectoral-muscle  scores 
(Table  3).  In  addition,  scores  for  fat  and  pec- 
toral muscle  were  positively  correlated  (r  = 
0.37,  P < 0.001),  indicating  that  owls  with 
high  fat  scores  tended  to  have  high  pectoral- 
muscle  scores.  Owls  showed  nearly  every 
combination  of  fat  and  pectoral-muscle 
scores,  except  for  the  highest  pectoral-muscle 
score  being  paired  with  the  lowest  fat  score, 
or  vice  versa. 

DISCUSSION 

Body  mass  of  Flammulated  Owls  increased 
significantly  as  the  migration  season  ap- 
proached. This  result  is  consistent  with  data 
showing  that  migratory  birds  often  increase 
their  total  body  mass  prior  to  migration  (Bair- 
lein  2002).  Such  patterns  have  been  shown  for 
songbirds,  shorebirds,  and  even  some  diurnal 
raptors,  but  little  information  is  available  on 
pre-migration  gain  in  mass  among  owls  (Ges- 
saman  1979,  Bairlein  2002).  In  Colorado, 
Linkhart  and  Reynolds  (1987)  found  mass 
gain  in  one  radio-tracked  adult  Flammulated 
Owl  during  the  month  of  September.  In  the 
present  study,  I confirmed  this  pattern  for  a 
large  number  of  owls,  but  I also  found  that 
capture  date  explained  only  a small  amount  of 
variation  in  the  mass  of  captured  owls.  This 


latter  pattern  is  not  surprising  given  the  ex- 
pected variation  in  hatching  dates  and  that 
owls  of  different  ages  likely  gain  mass  at  dif- 
ferent rates. 

I evaluated  the  relationship  of  pectoral- 
muscle  size  and  fat  stores  to  the  seasonal  in- 
crease in  body  mass  in  three  ways.  First,  using 
recapture  data,  I found  that  there  were  con- 
current increases  in  fat  scores,  muscle  scores, 
and  body  mass  for  most  individuals.  Second, 
scores  of  furcular  fat  and  pectoral  muscles 
were  closely  tied  to  body  mass,  but  fat  scores 
were  better  predictors  of  body  mass  than  pec- 
toral-muscle scores.  Third,  fat  scores  in- 
creased through  the  season  along  with  total 
body  mass,  but  pectoral-muscle  scores  did 
not.  Taken  together,  these  three  results  indi- 
cate that  fat  stores  are  an  important  compo- 
nent of  the  overall  mass  gain  in  Flammulated 
Owls  prior  to  migration,  but  pectoral-muscle 
size  is  not  as  important. 

Recently,  the  question  of  whether  fat  stores 
and  muscle  tissues  develop  independently  has 
been  raised.  For  example,  Redfern  et  al. 
(2000,  2004)  found  a general  interdependence 
in  fat  stores  and  muscle  mass  for  Sedge  War- 
blers (. Acrocephalus  schoenobaenus ) and  Red- 
wings ( Turdus  iliacus ).  My  data  also  support 
the  hypothesis  that  fat  and  pectoral-muscle 
scores  are  interdependent  because  (1)  there 
were  concurrent  increases  in  fat  scores,  mus- 
cle scores,  and  body  mass  for  most  recaptured 
birds;  (2)  there  was  a positive  correlation  be- 
tween the  variables;  and  (3)  there  were  no 
owls  having  high  scores  for  one  parameter 
without  also  having  high  scores  for  the  other. 

There  appeared  to  be  a non-fat  component 
to  the  season-long  mass  gain  that  was  unre- 
lated to  pectoral-muscle  size.  About  20%  of 
the  season-long  mass  gain  was  not  explained 
by  increases  in  fat  mass  or  pectoral  muscle. 
These  increases  in  mass  may  have  been  relat- 
ed to  increased  sizes  of  internal  organs,  which 
may  have  been  necessary  to  facilitate  the  ob- 
served accumulation  of  muscle  and  fat  re- 
serves. Such  changes  have  been  observed  in 
other  migratory  birds  as  fat  reserves  were  re- 
plenished. For  example,  Karasov  and  Pinshow 
(1998)  found  that  internal  organ  size  increased 
and  contributed  to  gains  in  body  mass  among 
foraging  Blackcaps  ( Sylvia  atricapilla)  cap- 
tured at  a stopover  site  in  Israel  during  north- 
bound-migration. 


192 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


These  data  add  to  the  growing  body  of 
work  showing  that  both  fat  deposition  and 
muscle  growth  are  associated  with  migration- 
related  mass  gains  and  that  the  two  processes 
are  somewhat  interdependent.  The  implication 
of  these  studies  is  that  birds  getting  ready  to 
migrate  or  already  migrating  may  have  spe- 
cific nutrient  needs  when  foraging.  This  work 
may  help  to  improve  our  understanding  of  for- 
aging ecology  and  site  selection  before  and 
during  migration — two  concerns  becoming  in- 
creasingly important  for  the  conservation  of 
migratory  birds. 

ACKNOWLEDGMENTS 

Primary  financial  support  was  provided  by  the 
USDA  Forest  Service  (Cibola  National  Forest  and  Re- 
gion 3),  the  New  Mexico  Game  and  Fish  Department 
(Share  with  Wildlife  Program),  Public  Service  Com- 
pany of  New  Mexico,  and  the  New  Mexico  Ornitho- 
logical Society.  Additional  funds  for  the  project  came 
from  J.  and  J.  Humber.  I gratefully  thank  all  of  our 
supporters  for  their  generosity.  B.  G.  deGruyter,  H.  P. 
Gross,  and  J.  P.  Smith  were  instrumental  in  arranging 
funding  and  logistical  support  for  this  project.  I ben- 
efited greatly  from  working  with  S.  L.  Stock,  who  in- 
troduced me  to,  and  trained  me  in,  fat-  and  muscle- 
scoring techniques  for  owls.  The  company  and  assis- 
tance of  the  Manzano  Mountains’  field  crew  and  vol- 
unteers from  both  HawkWatch  International  and  Rio 
Grande  Bird  Research  were  greatly  appreciated.  In  par- 
ticular, W.  L.  Beard,  Z.  M.  Hurst,  W.  M.  King,  and  E. 
A.  Snyder  made  significant  contributions  to  the  field 
effort.  I appreciate  discussions  with  J.  F.  Kelly  about 
migration  ecology  and  analytical  techniques.  Thanks 
to  J.  V.  Jewell  for  support  in  so  many  ways.  This  man- 
uscript benefited  from  thoughtful  reviews  by  J.  F.  Kel- 
ly, J.  P.  Smith,  and  three  anonymous  reviewers. 

LITERATURE  CITED 

Alerstam,  T.  and  A.  Lindstrom.  1990.  Optimal  bird 
migration:  the  relative  importance  of  time,  energy, 
and  safety.  Pages  331-351  in  Bird  migration:  the 
physiology  and  ecophysiology  (E.  Gwinner,  Ed.). 
Springer- Verlag,  Berlin,  Germany. 

Bairlein,  F.  2002.  How  to  get  fat:  nutritional  mecha- 
nisms of  seasonal  fat  accumulation  in  migratory 
songbirds.  Naturwissenshaften  89:1-10. 

DeLong,  J.  P.  2004.  Age  determination  and  prefor- 
mative  molt  in  hatch-year  Flammulated  Owls  dur- 
ing the  fall.  North  American  Bird  Bander  29: 1 1 1- 
115. 

DeLong,  J.  P.,  S.  W.  Cox,  and  N.  S.  Cox.  2005.  A 
comparison  of  avian  use  of  high  and  low  elevation 
sites  during  autumn  migration  in  central  New 
Mexico.  Journal  of  Field  Ornithology  76:326— 
333. 

DeLong,  J.  P.  and  S.  W.  Hoffman.  1999.  Differential 


autumn  migration  of  Sharp-shinned  and  Cooper’s 
hawks  in  western  North  America.  Condor  101: 
674-678. 

DeLong,  J.  P,  T.  D.  Meehan,  and  R.  B.  Smith.  2005. 
Investigating  fall  movements  of  hatch-year  Flam- 
mulated Owls  ( Otus  flammeolus)  in  central  New 
Mexico  using  stable  hydrogen  isotopes.  Journal  of 
Raptor  Research  39:19-25. 

Gessaman,  J.  A.  1979.  Premigratory  fat  in  the  Amer- 
ican Kestrel.  Wilson  Bulletin  91:625-626. 

Gosler,  A.  G.  1991.  On  the  use  of  greater  covert 
moult  and  pectoral  muscle  as  measures  of  condi- 
tion in  passerines  with  data  for  the  Great  Tit  Pams 
major.  Bird  Study  38:1-9. 

Griffiths,  R.,  M.  C.  Double,  K.  Orr,  and  R.  J.  G. 
Dawson.  1998.  A DNA  test  to  sex  most  birds. 
Molecular  Ecology  7:1071-1075. 

Helms,  C.  W.  and  W.  H.  Drury,  Jr.  1960.  Winter  and 
migratory  weight  and  fat  field  studies  on  some 
North  American  buntings.  Bird-banding  31:1-40. 

Hintze,  J.  2001.  NCSS  and  PASS:  Number  Cruncher 
Statistical  Systems.  Kaysville,  Utah. 

Karasov,  W.  H.  and  B.  Pinshow.  1998.  Changes  in 
lean  mass  and  in  organs  of  nutrient  assimilation 
in  a long-distance  passerine  migrant  at  a spring- 
time stopover  site.  Physiological  Zoology  71:435- 
448. 

King,  J.  R.  1972.  Adaptive  periodic  fat  storage  by 
birds.  Proceedings  of  the  International  Ornitholog- 
ical Congress  15:200-217. 

King,  J.  R.  and  D.  S.  Farner.  1965.  Studies  of  fat 
deposition  in  migratory  birds.  Annals  of  the  New 
York  Academy  of  Science  131:422-440. 

Krementz,  D.  G.  and  G.  W.  Pendleton.  1990.  Fat 
scoring:  sources  of  variability.  Condor  92:500- 
507. 

Lindstrom,  A.,  A.  Kvist,  T.  Piersma,  A.  Dekinga,  and 
M.  W.  Dietz.  2000.  Avian  pectoral  muscle  size 
rapidly  tracks  body  mass  changes  during  flight, 
fasting,  and  fuelling.  Journal  of  Experimental  Bi- 
ology 203:913-919. 

Lindstrom,  A.  and  T.  Piersma.  1993.  Mass  changes 
in  migrating  birds:  the  evidence  for  fat  and  protein 
storage  re-examined.  Ibis  135:70-78. 

Linkhart,  B.  D.  and  R.  T.  Reynolds.  1987.  Brood 
division  and  postnesting  behavior  of  Flammulated 
Owls.  Wilson  Bulletin  99:240-243. 

Marsh,  R.  L.  1984.  Adaptations  of  the  Gray  Catbird 
Dumetella  carolinensis  to  long-distance  migra- 
tion: flight  muscle  hypertrophy  associated  with  el- 
evated body  mass.  Physiological  Zoology  57:105- 
117. 

McCallum,  D.  A.  1994.  Flammulated  Owl  ( Otus  flam- 
meolus). The  Birds  of  North  America,  no.  93. 

Pennycuick,  C.  J.  1998.  Computer  simulation  of  fat 
and  muscle  bum  in  long-distance  bird  migration. 
Journal  of  Theoretical  Biology  191:47-61. 

Piersma,  T.,  G.  A.  Gudmundsson,  and  K.  Lellien- 
dahl.  1999.  Rapid  changes  in  the  size  of  different 
functional  organ  and  muscle  groups  during  refu- 


DeLong  • PRE-MIGRATOR  Y MASS  GAIN  IN  FLAMMULATED  OWLS 


193 


eling  in  a long-distance  migrating  shorebird.  Phys- 
iological and  Biochemical  Zoology  72:405-415. 

Pyle,  P.  1997.  Identification  guide  to  North  American 
birds,  part  I.  Slate  Creek  Press,  Bolinas,  Califor- 
nia. 

Redfern,  C.  P.  F.,  A.  E.  J.  Slough,  B.  Dean,  J.  L. 
Brice,  and  P.  H.  Jones.  2000.  Fat  and  body  con- 
dition in  migrating  Redwings  Turdus  iliacus.  Jour- 
nal of  Avian  Biology  31:197-205. 

Redfern,  C.  P.  E,  V.  J.  Topp,  and  P.  Jones.  2004.  Fat 
and  pectoral  muscle  in  migrating  Sedge  Warblers 
Acrocephalus  schoenobaenus.  Ringing  and  Mi- 
gration 22:24-34. 

Reynolds,  R.  T.  and  B.  D.  Linkhart.  1987.  The  nest- 


ing biology  of  Flammulated  Owls  in  Colorado. 
Pages  239-248  in  Biology  and  conservation  of 
northern  forest  owls  (R.  W.  Nero,  R.  J.  Clark,  R. 
J.  Knapton,  and  R.  H.  Hamre,  Eds.).  General 
Technical  Report  RM-142,  USDA  Forest  Service, 
Rocky  Mountain  Forest  and  Range  Experiment 
Station,  Fort  Collins,  Colorado. 

Rogers,  C.  M.  1991.  An  evaluation  of  the  method  of 
estimating  body  fat  in  birds  by  quantifying  visible 
subcutaneous  fat.  Journal  of  Field  Ornithology  62: 
349-356. 

Selman,  R.  G.  and  D.  C.  Houston.  1996.  A technique 
for  measuring  lean  pectoral  muscle  mass  in  live 
small  birds.  Ibis  138:348-350. 


The  Wilson  Journal  of  Ornithology  1 18(2):  194— 207,  2006 


MORPHOLOGICAL  VARIATION  AND  GENETIC  STRUCTURE  OF 
GALAPAGOS  DOVE  {ZEN AID  A GALAPAGOENSIS)  POPULATIONS: 
ISSUES  IN  CONSERVATION  FOR  THE  GALAPAGOS  BIRD  FAUNA 

DIEGO  SANTIAGO- ALARCON,1 3 SUSAN  M.  TANKSLEY,2  AND 
PATRICIA  G.  PARKER1 2 3 


ABSTRACT. — Island  species,  particularly  endemics,  tend  to  have  lower  genetic  diversity  than  their  continental 
counterparts.  The  low  genetic  variability  of  endemic  species  and  small  populations  has  a direct  impact  on  the 
evolutionary  potential  of  those  organisms  to  cope  with  changing  environments.  We  studied  the  genetic  population 
structure  and  morphological  differentiation  among  island  populations  of  the  Galapagos  Dove  ( Zenaida  galapa- 
goensis).  Doves  were  sampled  from  five  islands:  Santa  Fe,  Santiago,  Genovesa,  Espanola,  and  Santa  Cruz.  Five 
microsatellite  markers  were  used  to  determine  genetic  diversity,  population  structure,  gene  flow,  and  effective 
population  sizes.  jFsx  and  Rsr  values  did  not  differ  among  populations;  in  general,  populations  with  greater 
geographical  separation  were  not  more  genetically  distinct  than  those  closer  to  one  another,  and  estimated  gene 
flow  was  high.  There  were  no  significant  differences  in  allelic  richness  and  gene  diversity  among  populations. 
Although  there  was  extensive  morphological  overlap  among  individuals  from  different  island  populations  for 
both  males  and  females,  we  found  significant  differences  in  overall  body  size  only  between  populations  on  Santa 
Fe  and  Santa  Cruz  (males  and  females)  and  between  Espanola  and  Santa  Fe  (males  only).  Significant  differences 
in  body  size  between  populations  undergoing  high  rates  of  gene  flow  indicate  that  differentiation  may  be  due 
to  either  phenotypic  plasticity  or  ecotypic  differentiation.  Based  on  the  results  of  previously  conducted  disease 
surveys,  we  discuss  the  conservation  implications  for  the  Galapagos  Dove  and  other  endemics  of  the  archipelago; 
we  also  discuss  the  possible  effects  of  wind  currents  on  gene  flow.  Received  24  January  2005,  accepted  28 
November  2005. 


Historically,  islands  are  places  where  the 
most  dramatic  morphological  and  genetic  dif- 
ferentiations have  occurred  (Grant  1998, 
2001).  Geographic  isolation  between  popula- 
tions is  expected  to  promote  differentiation  of 
both  morphological  and  genetic  characters, 
due  to  either  drift  or  different  selective  re- 
gimes (Slatkin  1985,  Bohonak  1999).  This 
may  reflect  population  divergence  due  to  in- 
sufficient gene  flow  that  would  counteract  the 
effects  of  drift  and  selection  (Slatkin  1985, 
Hutchison  and  Templeton  1999,  Coleman  and 
Abbott  2003).  Isolation  leads  to  the  formation 
of  geographical  races,  which  is  considered  one 
of  the  initial  stages  of  speciation  (Grant  2001). 
However,  factors  independent  of  geographical 
isolation  (e.g.,  microclimate,  resources,  habi- 
tat structure)  may  be  acting  to  create  differ- 
ences between  sympatric  populations  or  pop- 
ulations undergoing  high  gene  flow  (e.g., 
Schluter  2001,  Ogden  and  Thorpe  2002). 
There  is  also  the  possibility  that  morphologi- 


1  Dept,  of  Biology,  Univ.  of  Missouri-St.  Louis, 
8001  Natural  Bridge  Rd„  St.  Louis,  MO  63121,  USA. 

2 Dept,  of  Animal  Science,  Kleberg  Center,  Texas 
A&M  Univ.,  College  Station,  TX  77843-2471,  USA. 

3 Corresponding  author;  e-mail:  onca77@yahoo.com 


cal  differences  may  be  observed — either  im- 
mediately or  within  a few  generations — at  dif- 
ferent geographic  locations  (different  popula- 
tions) without  corresponding  genetic  differ- 
entiation (phenotypic  plasticity;  e.g.,  James 
1983,  Losos  et  al.  1997,  Trussell  and  Etter 
2001). 

Island  species  have  served  as  models  for 
studies  of  evolution  due  to  the  discrete  nature 
of  island  archipelagos  and  the  isolation  be- 
tween different  island  populations  of  the  same 
species.  Several  Galapagos  archipelago  en- 
demics have  very  limited  inter-island  move- 
ment, resulting  in  morphological  differences 
(e.g.,  Bollmer  2000,  Grant  2001).  Columbi- 
formes  on  the  other  hand  are  strong  fliers  able 
to  move  long  distances  (Goodwin  1977,  Bap- 
tista  et  al.  1997).  Because  of  the  proximity  of 
several  islands  in  the  archipelago,  we  expect- 
ed high  gene  flow  among  populations  of  the 
Galapagos  Dove  {Zenaida  galapagoensis ) and 
no  morphological  differentiation. 

The  Galapagos  Dove  is  an  endemic  species 
whose  biology  and  ecology  are  poorly  under- 
stood. Our  knowledge  of  this  species  is  re- 
stricted to  taxonomic  relationships  (Goodwin 
1977,  Johnson  and  Clayton  2000),  morpho- 
logical descriptions  (Ridgway  1897,  Gifford 


194 


Santiago-Alarcon  et  al.  • MORPHOLOGY  AND  GENETICS  OF  THE  GALAPAGOS  DOVE  195 


1913,  Prestwich  1959),  and  more  recently,  to 
some  aspects  of  its  breeding  and  feeding  ecol- 
ogy on  Genovesa  Island  (Grant  and  Grant 
1979).  Morphological  and  ecological  studies 
of  bird  species  in  the  Galapagos  archipelago 
have  been  mostly  restricted  to  Darwin’s  finch- 
es (Bowman  1961;  Boag  1981,  1983;  Grant  et 
al.  1985;  Grant  2001),  Galapagos  mocking- 
birds (Nesomimus  spp.;  Curry  1988,  1989; 
Curry  and  Grant  1989),  and  the  Galapagos 
Hawk  ( Buteo  galapagoensis’,  de  Vries  1973, 
1975;  Bollmer  et  al.  2003).  Measurements  and 
a general  description  of  Galapagos  Doves  are 
provided  by  Ridgway  (1897),  Gifford  (1913), 
and  Swarth  (1931).  Gifford  (1913)  suggested 
that  doves  inhabiting  the  northern-most  is- 
lands— Wolf  (formerly  Wenman)  and  Darwin 
(formerly  Culpepper) — are  larger  than  those 
located  within  the  main  cluster  of  islands;  for 
this  reason,  dove  populations  were  classified 
as  two  subspecies:  Z.  g.  exsul  (on  Wolf  and 
Darwin)  and  Z.  g.  galapagoensis  (Swarth 
1931,  Baptista  et  al.  1997).  To  assess  levels 
of  population  structure  and  morphological 
variation,  our  study  focused  on  populations  of 
the  southern  subspecies  (Z.  g.  galapagoensis). 

Island  species,  particularly  endemics,  tend 
to  have  lower  genetic  diversity  than  their  con- 
tinental counterparts,  especially  when  such 
species  inhabit  small  islands  (Frankham  1996, 

1997) .  Maintaining  genetic  diversity  and  un- 
derstanding patterns  of  genetic  diversity  in 
natural  populations  is  a central  issue  in  con- 
servation genetics  (Frankham  1996,  1997, 

1998) .  Populations  are  not  equivalent  in  their 
capacity  to  adapt  to  changing  environmental 
conditions,  and  genetic  diversity  maximizes 
the  potential  evolutionary  responses  of  con- 
served populations  (Petit  et  al.  1998,  Hedrick 
2001).  Species  inhabiting  islands  are  consid- 
ered behaviorally  and  physiologically  naive; 
thus,  they  might  be  affected  more  severely 
than  mainland  species  by  the  introduction  of 
predators  and  diseases  (Mack  et  al.  2000).  De- 
mographic and  environmental  stochasticity 
can  be  accentuated  in  small  island  populations 
with  little  genetic  variability,  increasing  their 
risk  of  extinction  (Frankham  1996,  1997, 
1998). 

The  introduction  of  exotic  organisms  to  is- 
lands is  one  of  the  most  important  factors  in 
the  extinction  of  endemic  species  (Wikelski  et 
al.  2004).  Because  of  the  negative  impact  of 


pathogens  on  the  avian  endemics  in  several 
other  archipelagos,  preventing  the  introduc- 
tion of  avian  diseases  is  a conservation  pri- 
ority in  the  Galapagos  archipelago  (Padilla  et 
al.  2004,  Wikelski  et  al.  2004).  Some  diseases 
common  to  Columbiformes,  such  as  Tricho- 
monas gallinae,  might  be  transmitted  to  Ga- 
lapagos Doves  by  other  Columbiformes,  such 
as  the  exotic  Rock  Pigeon  ( Columba  livia ) and 
the  transient  (from  South  America)  Eared 
Dove  (Z.  auriculata ; Harmon  et  al.  1987,  Cur- 
ry and  Stoleson  1988,  McQuistion  1991,  Mete 
et  al.  2001,  Padilla  et  al.  2004).  Padilla  et  al. 
(2004)  have  reported  a >85%  prevalence  of 
Haemoproteus  malaria  in  Galapagos  Doves 
and  infections  of  Chlamydophila  psittaci  in 
doves  inhabiting  the  island  of  Espanola.  Buck- 
ee  et  al.  (2004)  have  shown  theoretically  that 
host  spatial  structure  directly  affects  pathogen 
diversity  and  strain  structure.  Thus,  it  is  a con- 
servation priority  to  understand  the  movement 
patterns  of  those  species  that  could  serve  as 
vectors  or  reservoirs  of  diseases  with  inter- 
specific infection  potential.  We  have  shown 
how  lice  from  Galapagos  Doves  can  be  trans- 
mitted to  Galapagos  Hawks  when  they  prey 
on  doves;  predation  may  represent  a route  of 
transmission  for  several  infectious  agents 
transmitted  by  lice  (Whiteman  et  al.  2004). 

Among  the  islands  sampled  in  this  study, 
only  Santa  Cruz  was  inhabited  by  humans, 
and  it  holds  the  largest  human  population  of 
the  inhabited  islands  in  the  archipelago.  Es- 
panola was  the  most  isolated  island,  lying  at 
the  southeastern  extreme  of  the  archipelago. 
Santa  Fe  and  Genovesa  were  the  smallest  is- 
lands, and  Genovesa  was  the  northern-most 
island  (Fig.  1).  The  Galapagos  islands  selected 
for  this  study — Santiago,  Santa  Cruz,  Santa 
Fe,  Genovesa,  and  Espanola — were  chosen  to 
represent  the  maximum  geographic  isolation 
between  populations  (e.g.,  Espanola  versus 
Genovesa)  and  widest  (east-west  and  north- 
south)  coverage  of  the  archipelago  that  our 
budget  and  logistical  restrictions  could  accom- 
modate. In  this  study,  we  (1)  used  principal 
components  analysis  (PCA)  to  examine  mor- 
phological variation,  (2)  used  five  microsat- 
ellite loci  to  describe  the  population  structure 
and  genetic  diversity,  and  (3)  estimated  effec- 
tive population  sizes  and  gene  flow  of  Z.  ga- 
lapagoensis on  five  islands  of  the  Galapagos 
archipelago:  Santiago,  Santa  Cruz,  Santa  Fe, 


196 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


91°  W 90“' W 

t 0 20  40  60  60  km 

Pin, a I i i i I i i ' I 


FIG.  1 . Map  of  the  Galapagos  archipelago,  Ecua- 
dor, showing  the  five  islands  (in  dark  gray)  where  Ga- 
lapagos Doves  were  sampled  in  2002  and  2004.  The 
Galapagos  Dove  occurs  on  all  the  major  islands  of  the 
archipelago. 

Genovesa,  and  Espanola.  Specifically,  we 
asked  (1)  are  there  significant  morphological 
differences  among  island  populations  of  the 
Galapagos  Dove,  (2)  are  these  populations 
isolated,  and  (3)  is  there  evidence  of  low  ge- 
netic variability  in  the  Galapagos  Dove? 

METHODS 

Field  methods. — We  conducted  our  study  in 
the  Galapagos  archipelago  from  May  through 
July  2002  and  from  June  through  July  2004. 
Following  the  guidelines  described  in  Ralph 
et  al.  (1996),  we  captured  Galapagos  Doves 
by  using  hand  nets  and  mist  nets.  We  took 
blood  samples  (50  pd  each)  by  venipuncture 
of  the  brachial  vein  from  25  birds  each  on 
Santa  Cruz,  Santa  Fe,  and  Espanola,  and  30 
birds  each  on  Santiago  and  Genovesa  islands 
(Fig.  1).  Samples  were  mixed  with  500-700 
pj  of  lysis  buffer  (100  mM  Tris  pH  8.0,  100 
mM  EDTA,  10  mM  NaCl,  0.5%  SDS;  Long- 
mire  et  al.  1988).  We  also  measured  25  birds 
each  from  Santa  Cruz,  Santa  Fe,  and  Espanola 
islands  and  30  each  from  Santiago  and  Gen- 
ovesa islands  (Fig.  1).  During  the  2002  study 
season,  we  sampled  doves  on  San  Cristobal 
Island,  but  due  to  the  small  sample  size  ( n = 
2)  they  were  not  included  in  our  analysis.  En- 
demics on  San  Cristobal  are  rare,  and  the  Ga- 
lapagos Dove  seems  to  be  among  the  rarest. 

In  order  to  quantify  inter-population  differ- 
ences in  morphology,  we  took  the  following 
measurements  to  the  nearest  0. 1 mm  from  the 


right  side  of  each  individual:  (1)  tarsus  length, 
(2)  tail  length,  (3)  length  of  exposed  culmen 
(from  terminus  of  the  feathering  to  the  bill’s 
tip),  (4)  bill  width  (calipers  were  oriented  at  a 
90°  angle  to  the  axis  of  the  bill  and  measure- 
ment was  taken  at  the  terminus  of  the  feath- 
ering), and  (5)  bill  depth  (at  the  terminus  of 
the  feathering  and  again  at  a 90°  angle  to  the 
axis  of  the  bill).  Using  a ruler  with  a brass 
perpendicular  stop,  we  also  measured  wing 
chord  length  (unflattened,  from  carpal  joint  to 
the  tip  of  the  longest  primary)  to  the  nearest 
0.5  mm.  We  used  Pesola  scales  (100  and  300 
g)  to  measure  mass  to  the  nearest  0.1  g.  Bird 
measurements  were  taken  by  DSA  on  all  the 
islands  but  Santa  Fe,  where  J.  L.  Bollmer  con- 
ducted the  sampling. 

Using  plumage  patterns,  we  identified  birds 
as  adults  or  juveniles:  adults  have  brighter  col- 
oration, and  juveniles  are  much  duller  in  color 
(Ridgway  1 897).  Because  individual  adults  of 
some  dove  species  do  not  have  completely  os- 
sified skulls  (Pyle  1997),  and  because  the  use 
of  cranium  calcification  (pneumatization)  for 
aging  doves  is  not  well  developed  (Pyle 
1997),  any  captured  individual  with  incom- 
plete calcification  and  adult  coloration  was 
considered  an  adult.  Although  it  is  possible  to 
identify  males  and  females  in  the  field  by  their 
plumage  coloration  and  body  size  (males  and 
females  have  similar  coloration  patterns,  but 
males  tend  to  be  brighter  than  females  and  are 
larger;  Ridgway  1897,  Gifford  1913;  DSA  and 
PGP  unpubl.  data),  this  technique  is  not  al- 
ways reliable  due  to  individual  variation. 
Therefore,  we  used  a polymerase  chain  reac- 
tion- (PCR)  based  technique  for  sexing  every 
individual  (Fridolfsson  and  Ellegren  1999). 
Birds  were  released  within  40  m of  capture 
location. 

Morphology 

Statistical  analyses. — We  used  Principal 
Component  Analysis  (PCA)  to  describe  mor- 
phological variation  among  islands  (SPSS, 
Inc.  2001).  Prior  to  PCA,  variables  were 
checked  for  outliers  (standardizing  to  zero 
mean  and  unit  variance);  four  values  with 
standard  deviations  >2.5  were  eliminated.  Al- 
though all  variables  (raw  data)  were  normally 
distributed  (Kolmogorov-Smirnov  test,  P ^ 
0.06)  and  have  the  same  scale  and  dimension 
(except  mass),  they  were  log-transformed  in 


Santiago-Alarcon  et  al  • MORPHOLOGY  AND  GENETICS  OF  THE  GALAPAGOS  DOVE  1 97 


TABLE  1.  Microsatellite  primers  and  number  of  alleles  scored  for  Galapagos  Doves  from  five  islands 
sampled  in  2002  and  2004,  Galapagos  Islands,  Ecuador  (n  = 134). 


Locus 

Primer 

sequence  5 '-3' 

TAa 

No.  alleles 

WU7al  17F 

CTC 

AGT 

GTA 

AAT 

ATG 

GCA 

GGG 

AAT  C 

54 

7 

WU7al  17R 

CAG 

GTC 

TTT 

TTG 

GTG 

GAT 

GTC 

AC 

WUa38F 

GGA 

GGG 

CAC 

CAG 

AGT 

TG 

55 

7 

WUa38R 

GAT 

AAG 

ACC 

CGA 

CTT 

TCA 

GC 

WUelF 

CAG 

TGT 

GGC 

AGG 

TAC 

TTC 

A 

54 

3 

WUelR 

CTC 

ATT 

AGT 

GGA 

CCT 

TGG 

AC 

WUj22F 

CAG 

GAG 

CCA 

TCG 

TAC 

ACA 

T 

56 

5 

WUj22R 

TGA 

ATT 

ACC 

CCA 

TCA 

ACA 

AG 

ClipT17 

See  Traxler  et  al.  2000 

55 

11 

a Annealing  temperature  (°C). 


order  to  examine  proportional  contributions  of 
large  and  small  measurements  equally.  We 
used  PCA  on  the  correlation  matrix  because 
one  of  the  variables  (mass)  did  not  have  the 
same  dimension,  and  because  a PCA  on  a cor- 
relation matrix  applied  to  transformed  data  is 
equivalent  to  a variance-covariance  matrix 
analysis  (McGarigal  et  al.  2000).  Furthermore, 
a PCA  from  a variance-covariance  matrix  ap- 
plied to  untransformed  (raw)  data  will  give 
more  weight  to  variables  with  large  variance, 
which  will  have  a larger  influence  on  the  PCA 
(McGarigal  et  al.  2000).  Because  males  are 
larger  than  females,  analyses  describing  the 
morphological  variation  among  islands  were 
conducted  separately  for  each  sex  to  prevent 
the  variance  due  to  sexual  dimorphism  from 
masking  variation  among  populations.  For 
each  PCA,  principal  component  scores  were 
normally  distributed  (Kolmogorov-Smirnov 
test,  P ^ 0.74).  Communalities  (total  variation 
extracted  from  each  variable)  are  reported  for 
each  PCA.  All  components  with  eigenvalues 
>1  were  retained  for  subsequent  analyses.  Ei- 
genvectors were  rotated  using  varimax  rota- 
tion and  retained  when  the  explained  variance 
was  higher  than  that  of  unrotated  components 
or  when  the  interpretation  of  PCs  was  easier. 
After  conducting  a PCA  for  females,  we  did 
not  find  significant  differences  between  adult 
and  juvenile  females  (/46  = -0.69,  P = 0.48); 
thus,  we  retained  both  groups  in  the  PCA. 
However,  we  did  find  significant  differences 
between  adult  and  juvenile  males  (/67  = 4.23, 
P < 0.001)  and  removed  juveniles  (15)  from 
the  male  pool.  We  excluded  female  bill  depth 
from  the  analyses  for  inter-island  comparisons 
because  only  one  such  record  was  available 


for  Santiago  Island.  We  used  /-tests  and  AN- 
OVAs  on  PC  scores  for  group  comparisons 
and  Tukey  post-hoc  tests  any  time  an  ANOVA 
was  significant.  In  every  case,  variances  of  PC 
scores  were  homogeneous  between  and 
among  groups  (Levene’s  test,  P > 0.25).  All 
/-tests  were  independent  and  two-tailed. 

Genetics 

DNA  isolation  and  amplification. — DNA 
extractions  by  phenol-chloroform  were  fol- 
lowed by  dialysis  in  IX  TNE2  (10  mM  Tris- 
HC1,  10  mM  NaCl,  2 mM  EDTA)  and  diluted 
to  a working  concentration  of  20  ng/pl.  Integ- 
rity and  concentration  of  each  DNA  sample 
was  determined  by  spectrophotometry  and 
electrophoresis  in  0.8%  agarose  gels  run  in 
1 X TBE.  Individuals  were  scored  at  four  poly- 
morphic microsatellite  loci  (Table  1)  original- 
ly developed  for  White-winged  Doves  (Z. 
asiatica ; accession  numbers  for  WU7all7, 
WUel,  WUa38,  and  WUj22  are  AF260574, 
AF260573,  AY428751,  and  AY428752,  re- 
spectively) and  one  locus  developed  for  Rock 
Pigeon  (Traxler  et  al.  2000).  We  prepared  PCR 
reactions  of  10  pi  that  included  50  ng  of 
whole  genomic  DNA,  1 mM  dNTP’s,  10X  re- 
action buffer,  25  mM  MgCl2,  0.5  pg  of  each 
primer,  0.1  pi  of  DMSO,  and  0.5  units  of  Taq 
DNA  polymerase  (SIGMA).  PCR  conditions 
were  as  follows:  initial  denaturation  at  94°  C 
for  3 min  followed  by  35  cycles  of  denatur- 
ation at  94°  C for  30  sec;  annealing  from  54 
to  56°  C (see  Table  1)  for  1 min  and  extension 
at  72  C for  1 min;  and  a final  extension  at  72° 
C for  10  min.  PCR  products  were  separated 
in  non-denaturing  7.5%  polyacrylamide  gels 
run  on  BioRad  sequencing  rigs.  Gels  were 


198 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  2.  Principal  component  (PC)  scores  and  communalities  for  seven  morphological  variables  of  male 
(n  = 50)  and  female  ( n = 52)  Galapagos  Doves  sampled  from  five  islands  in  2002  and  2004,  Galapagos  Islands, 
Ecuador.  PC  scores  represent  the  correlations  of  each  variable  with  the  principal  components;  communalities 
represent  the  sums  of  squares  of  correlation  coefficients  on  the  first  two  PCs  or  the  proportion  of  variance 
extracted  from  each  variable. 


Males 


Females 


Variable 

PCI 

PC2 

Communalities 

PCI 

PC2 

Communalities 

Culmen 

0.626 

-0.212 

0.508 

0.614 

0.515 

0.678 

Bill  width 

0.331 

0.734 

0.762 

0.172 

0.639 

0.918 

Bill  depth 

0.492 

0.272 

0.550 

a 

— 

— 

Tarsus 

0.367 

0.644 

0.888 

0.720 

0.331 

0.629 

Tail 

0.786 

-0.101 

0.692 

0.421 

-0.642 

0.820 

Wing 

0.674 

-0.256 

0.604 

0.790 

-0.006 

0.739 

Weight 

0.779 

-0.294 

0.696 

0.606 

-0.644 

0.787 

a Not  included. 


stained  with  0.05%  ethidium  bromide  (EtBr) 
and  visualized  using  a Kodak  UV  digital  im- 
ager (KODAK  image  station  440CF). 

Statistical  analyses. — We  calculated  genetic 
diversity  using  Nei’s  unbiased  estimator  (Nei 
1973),  which  is  the  probability  that  two  alleles 
randomly  sampled  from  a population  are  dif- 
ferent. We  analyzed  allelic  richness  through 
rarefaction  analysis  as  implemented  by  El 
Mousadik  and  Petit  (1996)  and  Petit  et  al. 
(1998). 

^ST  estimates  outperform  RSJ  counterparts 
under  some  circumstances  (e.g.,  when  there 
are  allele  size  constraints  in  a microsatellite 
marker,  size  differences  cannot  be  used  to  re- 
flect distances  among  alleles),  even  under  the 
stepwise  mutation  model  (SMM).  Further- 
more, Rst  can  be  less  accurate  at  reflecting 
population  differentiation  due  to  its  greater  as- 
sociated variance.  Even  a small  number  of 
random  mutation  events  tends  to  erase  part  of 
the  memory  of  the  mutation  process  that  is  the 
base  of  the  SMM,  which  makes  RSJ  estimates 
superior  to  ^ST  only  when  the  mutation  pro- 
cess follows  the  SMM  exactly  (Gaggiotti  et 
al.  1999,  Balloux  et  al.  2000,  Balloux  and  Lu- 
gon-Moulin  2002).  Due  to  the  uncertainty  of 
the  mutation  process  of  microsatellites  (Prim- 
mer and  Ellegren  1998,  Goldstein  and  Schlot- 
terer  1999),  we  decided  to  use  F-statistics 
(Weir  and  Cockerham  1984)  for  our  analysis. 
For  the  sake  of  comparison,  we  also  calculat- 
ed RSJ  across  samples,  and  the  significance  of 
population  differentiation  based  on  ^ST  was 
evaluated  using  a G-test  and  1,000  randomi- 
zations (Goudet  et  al.  1996).  We  used  pairwise 


^ST  values  and  geographic  distance  matrices 
to  test  for  isolation  by  distance  (Slatkin  1993, 
Hutchison  and  Templeton  1999);  significance 
was  evaluated  with  a Mantel  test  (Mantel 
1967)  and  distance  was  log-transformed  be- 
fore analysis.  Geographical  distance  was  mea- 
sured as  the  closest  distance  between  islands. 

Data  were  analyzed  for  linkage  disequilib- 
rium and  Hardy-Weinberg  equilibrium  using 
FIS,  and  testing  was  conducted  via  G-test  and 
randomization  procedures  (Goudet  et  al.  1996, 
Goudet  1999).  Bonferroni  corrections  were 
applied  when  appropriate  (Rice  1989).  Loci 
proved  to  be  in  linkage  equilibrium  after  200 
permutations  ( P > 0.08,  Bonferroni  corrected 
P-value  at  a = 0.05  was  0.005).  Samples  were 
under  Hardy-Weinberg  equilibrium  after  500 
randomizations,  except  for  one  locus/popula- 
tion (WU7all7,  P = 0.002  for  Santiago  Is- 
land, Bonferroni  corrected  P-value  at  a = 
0.05  was  0.002).  Therefore,  we  tested  for  pop- 
ulation differentiation  without  assuming  H-W 
equilibrium.  Analyses  were  conducted  using 
FSTAT  (Goudet  2002). 

Because  gene  flow  and  effective  population 
size  estimates  based  on  ^ST  depend  on  many 
unrealistic  assumptions  (Waples  1998,  Whit- 
lock and  McCauley  1999),  we  used  a coales- 
cent-based  approach  to  calculate  migration 
rates  (Nm)  and  theta  (0  = 4Nep,,  which  is  a 
genetic  diversity  parameter  related  to  the  ef- 
fective population  size  [Ne]  from  which  Ne 
can  be  estimated)  using  the  program  MI- 
GRATE (Beerli  and  Felsenstein  1999,  2001). 
Unlike  Fsx,  this  program  accounts  for  direc- 
tional gene  flow  and  for  differences  in  popu- 


Santiago-Alarcon  et  al.  • MORPHOLOGY  AND  GENETICS  OF  THE  GALAPAGOS  DOVE  199 


PCI 


PCI 

FIG.  2.  (A)  Morphological  ordination  space  be- 

tween islands  for  adult  male  Galapagos  Doves.  PCI  is 
an  axis  of  overall  body  size  and  PC2  is  a vector  re- 
flecting bill  size  and  tarsus  length  ( n = 50).  Sample 
sizes  per  island  were  as  follows:  Santiago  (18),  Santa 
Cruz  (15),  Espanola  (11),  Santa  Fe  (15),  and  Genovesa 
(20).  (B)  Morphological  ordination  space  between  is- 
lands for  female  Galapagos  Doves.  PCI  is  an  axis  of 
overall  body  size  and  PC2  is  a vector  reflecting  bill 
size  and  tarsus  length  ( n = 52).  Sample  sizes  per  island 
were  as  follows:  Santiago  (12),  Santa  Cruz  (10),  Es- 
panola (14),  Santa  Fe  (10),  and  Genovesa  (10).  Ellip- 
ses represent  the  95%  confidence  interval  for  the  dif- 
ferent islands. 


lation  size.  We  ran  the  program  five  times  us- 
ing the  estimates  of  each  run  as  starting  pa- 
rameters for  the  next  one.  We  assumed  equal 
mutation  rates  among  loci,  which  is  an  unre- 
alistic assumption  (Goldstein  and  Schlotterer 
1999);  however,  it  provides  better  estimates  of 
parameters  than  when  using  variable  mutation 
rates  among  loci,  which  increase  the  variance 
(Beerli  and  Felsenstein  1999).  We  estimated 
parameters  for  the  first  run,  since  using  an  ^ST 
initial  estimate  produced  an  attraction  to  the 
area  of  the  likelihood  surface  of  the  generated 
Fst  values,  thus  preventing  the  program  from 


searching  efficiently  throughout  the  likelihood 
surface  (P.  Beerli  pers.  comm.).  Ten  short 
chains  and  two  long  chains  were  used  to  cal- 
culate parameters.  We  sampled  500  genealo- 
gies for  each  short  chain  and  5,000  for  each 
long  chain;  increments  were  set  to  20  for  the 
short  chains  and  to  100  for  the  long  chains; 
an  initial  stabilizing  period  (burn-in)  was  set 
to  10,000  genealogies.  We  computed  multiple 
estimation  of  parameters  using  the  two  long 
chains  of  each  run.  Because  MIGRATE  cal- 
culates historical  migration  rates,  we  used  the 
assignment/exclusion  method  of  Comuet  et  al. 
(1999),  implemented  in  the  program  GENE- 
CLASS  (Piry  et  al.  2004),  to  estimate  current 
levels  of  gene  flow.  This  method  is  appropri- 
ate to  use  when  all  possible  sources  of  mi- 
grants (populations)  have  not  been  sampled 
(Comuet  et  al.  1999,  Berry  et  al.  2004).  We 
used  the  “leave  one  out”  criterion,  which  re- 
moves the  individual  for  which  probabilities 
of  assignment/exclusion  to  a specific  popula- 
tion are  calculated  (Berry  et  al.  2004).  We 
used  the  simulation  algorithm  of  Paetkau  et 
al.  (2004)  to  estimate  assignment/exclusion 
probabilities  (a  = 0.05,  10,000  simulated  in- 
dividuals). 

RESULTS 

Morphological  variation  of  males  among 
islands. — We  retained  the  first  two  principal 
components.  PCI,  representing  an  overall  size 
dimension,  explained  36%  of  the  variance. 
PC2,  a bill-  (width  and  depth)  and  tarsus- 
length  component,  explained  17%  of  the  var- 
iance. The  variance  extracted  from  each  var- 
iable was  >50%  (Table  2).  There  were  sig- 
nificant differences  among  islands  in  the 
doves’  overall  body  size  (PCI,  F4A5  = 4.99, 
P = 0.002;  Fig.  2a),  but  not  bill  size  (PC2, 
F445  = 1.53,  P = 0.21).  Based  on  PCI,  Santa 
Cruz  and  Espanola  doves  were  significantly 
larger  than  Santa  Fe  doves  (Tukey-test,  HSD 
= 1.16,  P = 0.033  and  HSD  = 1.23,  P = 
0.019,  respectively).  There  is  overlap,  how- 
ever, among  individuals  of  these  three  islands, 
as  well  as  those  from  the  other  islands  (Fig. 
2a). 

Morphological  variation  of  females  among 
islands. — We  retained  the  first  two  principal 
components.  PCI,  which  represents  an  overall 
size  dimension,  explained  37%  of  the  variance 
(Table  2).  PC2,  a bill-  (culmen  length  and 


200 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  3.  Genetic  diversity  (Nei  1973)  and  allelic  richness  for  Galapagos  Dove,  as  estimated  by  rarefaction 
analysis  (Petit  et  al.  1998)  per  locus  and  population.  Samples  were  collected  from  five  islands  in  2002  and  2004, 
Galapagos  Islands,  Ecuador. 


Locus 

Genetic  diversity 

SFb 

E 

SC 

s 

G 

Wu7al  17 

0.75 

0.73 

0.66 

0.69 

0.72 

Wua38 

0.56 

0.71 

0.52 

0.55 

0.67 

Wuel 

0.35 

0.42 

0.24 

0.24 

0.31 

Wuj22 

0.49 

0.55 

0.62 

0.56 

0.61 

Cli|xT17 

0.79 

0.84 

0.78 

0.84 

0.79 

Mean  ± SD 

0.59  ±0.18 

0.65  ± 0.16 

0.56  ± 0.20 

0.58  ± 0.22 

0.62  ±0.18 

a /?t  = estimated  allelic  richness  for  all  islands. 

b SF  = Santa  Fe,  E = Espahola,  SC  = Santa  Cruz,  S = Santiago,  G = Genovesa. 


width)  and  tarsus-length  component,  ex- 
plained 23%  of  the  variance.  The  variance  ex- 
tracted from  each  variable  was  >62%  (Table 
2).  There  were  significant  differences  among 
islands  in  overall  body  size  (PCI,  F441  = 3.14, 
P = 0.023;  Fig.  2b),  but  not  in  the  second 
component  (PC2,  F441  = 0.84,  P = 0.51).  Dif- 
ferences in  overall  body  size  were  found  only 
among  doves  from  Santa  Cruz  and  Santa  Fe, 
where  Santa  Cruz  females  were  larger  than 
those  from  Santa  Fe  (Tukey-test,  HSD  = 1.53, 
P = 0.005);  otherwise  there  was  extensive 
overlap  among  individuals  from  the  different 
islands  (Fig.  2b). 

Population  structure  and  genetic  diversi- 
ty.— We  scored  33  alleles  for  five  polymorphic 
microsatellite  loci  from  25  doves  on  Santa 
Cruz,  Santa  Fe,  and  Espanola,  30  on  Santiago, 
and  29  on  Genovesa.  Santa  Fe  doves  had  the 
fewest  alleles  (23);  Espanola  and  Santiago  had 
29  each,  Genovesa  had  25,  and  Santa  Cruz 
had  26.  The  populations  with  the  richest  al- 
lelic composition  (Santiago  and  Espanola)  had 
86%  ([29  - 5]/[33  - 5])  of  the  allelic  diver- 
sity (excluding  the  five  alleles  that  were  au- 
tomatically present  because  there  are  five 
loci).  Rarefaction  analysis  showed  the  same 
tendency  in  allelic  richness  among  popula- 
tions; allelic  richness  across  loci  and  samples 
was  27  (Table  3).  Genetic  diversity  was  great- 
est among  doves  from  Espanola  and  lowest 
among  those  from  Santa  Cruz;  however,  there 
were  no  significant  differences  among  islands 
for  either  allelic  richness  or  genetic  diversity 
(both  P > 0.19). 

Estimates  of  Fsx  (0.01,  P > 0.43)  and  RSJ 
(0.0057,  P > 0.43)  across  samples  showed  no 
genetic  structure.  The  95%  bootstrap  confi- 


dence intervals  of  the  overall  F ST  estimate 
were  —0.001  and  0.02.  No  pairwise  F ST  values 
were  significantly  different  (all  P > 0.025, 
Bonferroni  corrected  P-value  at  a = 0.05  was 
0.005;  Table  4),  and  we  failed  to  detect  iso- 
lation by  distance  in  our  data  set  (Mantel  test 
after  2,000  randomizations,  P > 0.25). 

We  estimated  high  levels  of  historical  gene 
flow  between  populations  of  the  Galapagos 
Dove  (Table  5).  The  highest  estimated  number 
of  migrants  per  generation  was  71  (Espanola 
to  Genovesa),  which  was  surprising  consid- 
ering that  they  are  separated  by  the  largest 
geographic  distance  (—200  km)  compared 
with  distances  between  the  other  islands  sam- 
pled. Genovesa  Island  had  the  highest  theta 
value  (1.91)  and  Santa  Fe  had  the  lowest 
(0.18).  The  high  theta  for  Genovesa  is  sur- 
prising because  it  is  the  smallest  island  of 
those  included  in  the  study;  however,  Santa 
Cruz,  the  largest  island,  had  the  second  lowest 
theta  value  (0.4).  If  we  assume  that  microsat- 
ellite markers  have  a mutation  rate  of  10  4 
events  per  locus  per  generation  (Goldstein  and 
Schlotterer  1999),  and  that  this  mutation  rate 
is  the  same  for  each  locus,  the  effective  pop- 
ulation sizes  are  as  follows:  Santa  Fe  463;  Es- 
panola 3,600;  Santa  Cruz  1,000;  Santiago 
4,600;  and  Genovesa  4,775.  The  current  high 
rate  of  gene  flow,  as  estimated  with  GENE- 
CLASS,  suggests  that  doves  are  moving 
among  islands.  The  assignment  analysis  cor- 
rectly allocated  27.6%  (37)  of  the  individuals 
(P  < 0.009),  but  most  (34  of  37)  had  likeli- 
hoods lower  than  the  threshold  value  of  being 
assigned  to  another  population.  The  difficul- 
ties of  assigning  individuals  suggest  high  cur- 
rent gene  flow  among  populations.  Analyses 


Santiago-Alarcon  et  al.  • MORPHOLOGY  AND  GENETICS  OF  THE  GALAPAGOS  DOVE  20 1 


TABLE  3. 

Extended. 

Allelic  richness 

SF 

E 

SC 

s 

G 

/?Ta 

5 

7 

6 

6.75 

4.98 

6.1 1 

6 

6 

5 

3.97 

4.96 

5.46 

2 

3 

2 

2.99 

2.00 

2.56 

4 

4 

4 

4.97 

4.98 

4.61 

6 

9 

9 

9.63 

7.70 

8.25 

4.6  ± 1.67 

5.8  ± 2.38 

5.2  ± 2.58 

5.8  ± 2.77 

5.0  ± 2.12 

5.4  ± 2.08 

to  detect  first  generation  (F0)  migrants  detect- 
ed 15  migrants  (P  < 0.05;  Table  6). 

DISCUSSION 

In  this  study,  we  present  evidence  that  pop- 
ulations of  Galapagos  Doves  are  morpholog- 
ically and  genetically  similar,  which  must  be, 
in  part,  the  result  of  high  rates  of  gene  flow 
among  islands.  However,  our  results  also  in- 
dicate that  there  are  morphological  differences 
between  doves  from  some  island  pairs.  This 
might  be  due  to  different  abiotic  and  biotic 
pressures  operating  on  different  islands  (see 
below)  and  to  the  degree  of  connectedness 
(gene  flow)  between  some  island  pairs  (Table 
5).  For  example,  Santa  Cruz  and  Santa  Fe 
doves  differ  in  body  size  (both  males  and  fe- 
males) and  gene  flow  estimates  for  these  is- 
lands are  low  (see  Table  5)  even  though  they 
are  the  closest  among  all  the  island  pairs  (17.5 
km).  Genovesa,  the  island  with  the  largest  ef- 
fective population  size,  is  the  smallest  island 
of  those  sampled  and  is  also  the  one  receiving 
the  largest  number  of  migrants  from  the  other 
islands.  In  addition,  it  is  remarkable  that  the 
lowest  F ST  value  and  highest  numbers  of  mi- 
grants coming  to  Genovesa  are  from  Espan- 


ola,  which  is  the  island  most  distant  from 
Genovesa  (Fig.  1,  Tables  4 and  5).  Dove  pop- 
ulations on  both  Genovesa  and  Espanola, 
which  are  small  and  relatively  isolated  com- 
pared with  the  central  islands  (Fig.  1),  are  the 
two  populations  with  the  greatest  genetic  di- 
versities, largest  estimated  population  sizes, 
and  highest  rates  of  gene  flow  (Tables  3 and 
5). 

Environmental  factors  such  as  wind  cur- 
rents might  be  influencing  the  travel  routes  se- 
lected by  doves  from  different  islands,  thus 
affecting  the  degree  of  connectivity  among  is- 
land populations.  Several  phylogeographic  re- 
constructions of  other  vertebrate  endemics  of 
the  archipelago  have  shown  that  present  and 
historical  wind  and  ocean  currents  have  had  a 
south-southeast  to  north-northeast  effect  on 
the  evolutionary  history  of  organisms  (e.g., 
Caccone  et  al.  1999,  2002;  B.  S.  Arbogast  un- 
publ.  data).  However,  it  is  difficult  to  believe 
that  wind  currents  are  the  main  reason  for 
movements  of  Galapagos  Doves  among  is- 
lands. Even  though  there  is  a high  rate  of  gene 
flow  in  a south-to-north  direction  (e.g.,  Espan- 
ola to  Genovesa  [71.4],  Espanola  to  Santa 
Cruz  [17.85]),  gene  flow  is  also  high  in  the 


TABLE  4.  Estimates  of  genetic  differentiation  for  Galapagos  Doves  sampled  from  five  islands  in  2002  and 
2004,  Galapagos  Islands,  Ecuador.  Pairwise  Fst  values  are  above,  and  P- values  are  below,  the  dashes  (geographic 
distances  in  km  are  given  in  parentheses).  No  values  were  significant  (Bonferroni  corrected  P-value  at  a = 0.05 
was  0.002). 


Island 

Santa  Fe 

Espanola 

Santa  Cruz 

Santiago 

Genovesa 

Santa  Fe 

— 

0.0028 

0.0033 

-0.0036 

0.0090 

Espanola 

0.22  (74) 

— 

0.0264 

0.0159 

-0.0003 

Santa  Cruz 

0.42  (18) 

0.16  (99) 

— 

-0.0096 

0.0372 

Santiago 

0.10  (76) 

0.34  (161) 

0.66  (24) 

— 

0.0160 

Genovesa 

0.035  (135) 

0.20  (204) 

0.025  (103) 

0.095  (100) 

— 

TABLE  5.  Bi-directional  gene  flow  estimates  and  theta  values  (95%  Cl),  estimated  with  MIGRATE  (Beerli  and  Felsenstein  1999,  2001),  for  Galapagos  Doves 
from  five  islands,  2002  and  2004,  Galapagos  Islands,  Ecuador. 


202 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 . No.  2,  June  2006 


P 

in 

vO 

>n 

*0 

in 

(N 

rn 

(N 

CO 

P 

7 

P 

1 

X 

<N 

1 

U 

> 

v© 

r- 

vO  | 

o 

fi 

2 

O 

P 

vO 

o^  1 

in 

VO 

in 

1-^ 

o 

>n 

<N 

Ov 

CN 

t"- 

— 

m 

in 

vd 

vd 

— 

<N 

P 

/^s 

(N 

r- 

00 

CM 

oc 

00 

; 

p 

— 

1 

o 

X 

1 

I 

Ov 

5-0 

CO 

CN 

1 ^ 

2 

rn 

Ov 

r- 

1 Tf 

§ 

rn 

© 

Ov 

C/3 

' — - 

N — ' 

' — ^ 



(N 

00 

ON 

rn 

vq 

in 

rn 

- 

vd 

P 

m 

m 

ov 

(N 

ov 

m 

£ £ 

N 

in 

1 

vd  i 

9 

X 

u 

_ 

c 

vO 

1 

y—t  O 

z 

co 

2 

00 

1 

^ <N 

B 

on 

m 

— 

C/2 

s— * 

v — ' 

' — ” ^ ' 

<N 

r- 

vO  00 

Ov 

vq  oc 

-'t 

vd 

in  rn 

m 

00 

/—S  O 

. — 1 

o 

vO 

<N 

OV 

2 vd 

o r- 

in 

1 

— I 

22 

X 

oc 

1 — 

JC 

3 

c. 

2 

r- 

1 

m 

vd 

Tt  — i 

^ P 

<N 

— 

00  vO 

uJ 

'' — ' 

s ' 

^ 

m 

ov  r- 

vq 

00 

m 

P 

P 

Ov  — 

r- 

00 

m 

rn 

— v 

m 

m 

vC 

m 

- 2° 

© 

P 7 

rr 

X 

o 

i - 

Q 

1 

(N 

Ov 

c 

1 

P 

~ 

7 p 

C0 

, 

(N 

o 

in  — 

C/3 

' — ' 

rf 

in 

vc  m 

(N 

rn 

o — 

O^ 

© 

vd  vd 

(N 

o 

P 

-'t 

rn  — 

d. 

(N 

«n 

O <N 

2 

Q 

P 

o 

P P 

vO 

r- 

1 1 
r-  r- 

** 

m 

m 

vO  vC 

05 

© 

p 

d 

P P 

<D 

c* 

00 

Tt  — 

H 

^t 

00  OV 

d 

d 

N 

3 

0> 

O 1/5 

T3 

C 

P 

ca 

tC 

u 

00  « 
■2  o 

"7 

c 

o. 

C 

C C 

y. 

UQ 

S3 

C/5 

cn 

on  O 

P 

(N 

m 

tj-  in 

E I 

3 c£ 


■s  8. 


CN  00 

o ^ 


e 2 

x m 
u 

c 


<L>  3 

E - 
R a. 


ji E 

$ a; 


2 oo 
= E 


2 2 
oo  u 
E a 
00  ^ 
.E  £ 


opposite  direction  (e.g.,  Santa  Cruz  to  Espan- 
ola  [36.9],  Santa  Fe  to  Espanola  [29.2],  Gen- 
ovesa  to  Santiago  [21.5],  Genovesa  to  Espan- 
ola [14.2];  Table  5).  Hence,  wind  currents 
might  not  completely  account  for  movements 
among  islands.  Perhaps  the  lack  of  any  clear 
pattern  in  dove  movement  among  islands  is 
due  to  the  strong  flight  capabilities  of  Co- 
lumbiformes  and  the  short  distances  between 
some  islands  (<20  km).  Doves  may  simply 
move  between  islands  to  track  food  resources 
and  suitable  environmental  conditions.  The 
lack  of  any  pattern  in  isolation  by  distance 
among  populations  supports  the  idea  that 
doves  can  move  in  any  direction. 

Low  genetic  differentiation  among  dove 
populations  might  also  be  accounted  for  either 
by  a recent  population  expansion  or  by  the 
presence  of  alleles  shared  due  to  common  an- 
cestry (e.g..  Grant  et  al.  2005),  rather  than  by 
frequent  dispersal  between  populations.  Rapid 
population  expansion  could  explain  reduced 
within-population  diversity  (versus  global  di- 
versity linked  to  founder  events;  Hedrick 
2000,  McCoy  et  al.  2003).  In  our  study,  esti- 
mates of  genetic  diversity  were  similar  among 
populations,  which  would  support  a gene  flow 
explanation  instead  of  a recent  expansion.  The 
possible  effect  of  shared  alleles  due  to  com- 
mon ancestry  might  be  ruled  out  by  the  results 
obtained  with  GENECLASS,  which  estimated 
that  current  rates  of  gene  flow  are  high.  More- 
over, if  the  Galapagos  Dove  colonized  the  ar- 
chipelago between  2.5  and  3 mya,  as  proposed 
by  Johnson  and  Clayton  (2000),  we  should 
have  detected  a genetic  signature  of  diver- 
gence, given  isolation  (by  distance)  between 
populations. 

Morphological  variation  among  islands. — 
Altitudinal  and  latitudinal  patterns  of  morpho- 
logical variation  within  islands  have  been  con- 
firmed for  Darwin's  finches,  but  some  patterns 
are  not  consistent  among  islands  (Grant  et  al. 
1985).  For  a given  finch  species,  individuals 
are  larger  at  higher  elevations  within  any  one 
island,  but  size  variation  among  island  popu- 
lations is  not  systematically  related  to  either 
latitude  or  longitude.  However,  this  is  not  the 
case  for  other  endemic  species  of  the  archi- 
pelago, such  as  Galapagos  Hawks,  where 
there  is  a clear  north-  (smaller  size)  to-south 
(larger  size)  trend  in  morphological  variation 
(Bollmer  et  al.  2003).  Body  size  variation  in 


Santiago-Alarcon  et  al.  • MORPHOLOGY  AND  GENETICS  OF  THE  GALAPAGOS  DOVE  203 


TABLE  6.  Gene  flow  estimates  of  first  generation  migrants  (F0),  calculated  with  GENECLASS  (Piry  et  al. 
2004),  for  Galapagos  Doves  on  five  islands,  2002  and  2004,  Galapagos  Islands,  Ecuador.  P-values  are  given  in 
parentheses. 

Nm 

Island 

Santa  Fe 

Espanola 

Santa  Cruz 

Santiago 

Geneovesa 

1 to  Xa 

2 to  X 

3 to  X 

4 to  X 

5 to  X 

1:  Santa  Fe 

— 

1 (0.039) 

1 (0.028) 

0 

0 

2:  Espanola 

1 (0.025) 

— 

1 (0.016) 

0 

0 

3:  Santa  Cruz 

2 (0.027) 

1 (0.003) 

— 

0 

1 (0.002) 

4:  Santiago 

1 (0.026) 

0 

2 (0.026) 

— 

1 (0.004) 

5:  Genovesa 

0 

1 (0.006) 

1 (0.035)  1 

(0.012) 

a The  population  receiving  migrants  = x,  and  the  number  preceding  x is  the  population  from  where  migrants  come.  For  example,  in  row  1 : Population 
2 (Espanola)  provides  1 migrant  per  generation  to  Population  1 (Santa  Fe);  Population  3 (Santa  Cruz)  provides  1 migrant;  Population  4 (Santiago)  provides 
0 migrants;  and  Population  5 (Genovesa)  provides  0 migrants  per  first  generation  to  Population  1 . 


the  Galapagos  Dove,  however,  did  not  show 
geographical  patterns  among  the  group  of  is- 
lands studied  here,  most  likely  because  (1)  en- 
vironmental characteristics  on  the  different  is- 
lands do  not  vary  geographically  in  a simple 
manner  (Grant  et  al.  1985),  and  (2)  gene  flow 
for  doves  among  islands  is  greater  than  it  is 
for  finches  or  hawks  (see  below).  Moreover, 
the  dove’s  omnivorous  diet  (see  Grant  and 
Grant  1979)  could  further  impede  extensive 
morphological  differentiation  between  island 
populations — a situation  similar  to  that  of  Ga- 
lapagos mockingbirds  (B.  S.  Arbogast  unpubl. 
data)  and  Hawaiian  thrushes  ( Myadestes  spp.; 
Lovette  et  al.  2001). 

Population  structure  and  conservation. — 
The  lack  of  population  structure  and  the  high 
levels  of  gene  flow  and  genetic  variation  are 
in  stark  contrast  with  results  reported  for  other 
species  in  the  archipelago,  which  are  charac- 
terized by  divergence  among  different  island 
populations  and  low  genetic  diversity  (e.g.. 
Grant  2001,  Bollmer  2000,  Bollmer  et  al. 
2003).  Allelic  richness  of  the  Galapagos  Dove 
for  the  five  microsatellite  loci  genotyped  in 
this  study  was  similar  to  the  values  reported 
for  its  continental  relatives.  White-winged 
Dove  (Tanksley  2000)  and  Mourning  Dove  (Z. 
macroura ; L.  M.  Reichart  unpubl.  data),  and 
in  some  cases  it  was  greater. 

Tanksley  (2000)  used  microsatellite  mark- 
ers and  reported  no  genetic  structure  in  White- 
winged Doves  sampled  at  a broader  geograph- 
ic scale  in  North  America;  mtDNA  revealed 
slight  differentiation  between  populations  ac- 
cording to  a historical  east-west  division  of  its 
distribution  (Pecos  River  in  Texas)  that  is  cur- 
rently disappearing  due  to  the  species’  range 


expansion  (Pruett  et  al.  2000).  Pruett  et  al. 
(2000)  suggested  that  the  White-winged 
Dove’s  range  expansion  is  due  to  urban  de- 
velopment, which  provides  water,  food,  and 
nesting  sites.  Urban  development  also  might 
be  affecting  Galapagos  Dove  populations,  at 
least  on  the  two  inhabited  islands  visited  in 
this  study  (Santa  Cruz  and  San  Cristobal). 
Santa  Cruz  doves  had  the  third  lowest  number 
of  alleles,  second  lowest  effective  population 
size,  and  the  lowest  genetic  diversity.  On  San 
Cristobal,  extensively  surveyed  for  3 days,  we 
saw  and  captured  only  two  doves.  Population 
declines  of  other  endemic  bird  species  on  San 
Cristobal  have  been  reported  (Vargas  1996). 
The  rarity  of  doves  and  population  declines  of 
other  endemic  bird  species  on  San  Cristobal 
seem  to  be  due  to  the  large  number  of  intro- 
duced species  and  to  the  longer  history  of  hu- 
man settlement  (Vargas  1996).  These  results 
provide  some  support  for  a negative  impact  of 
urban  development  on  Galapagos  Doves. 

Harmon  et  al.  (1987)  reported  Galapagos 
Doves  infected  with  Trichomonas  gallinae 
(believed  to  have  been  transmitted  by  Rock 
Pigeons)  on  Santa  Cruz  Island,  and  Padilla  et 
al.  (2004)  reported  infected  Rock  Pigeons,  but 
no  infected  Galapagos  Doves.  Galapagos 
Doves  on  Espanola  were  infected  with  Chla- 
mydophila  psittaci.  The  prevalence  of  Hae- 
moproteus  spp.  in  Galapagos  Doves  was 
found  to  be  >85%  on  five  islands  (Padilla  et 
al.  2004).  The  presence  of  infectious  diseases 
and  mosquitoes  of  the  genus  Culex  (Wikelski 
et  al.  2004,  Whiteman  et  al.  2005) — the  vector 
of  some  malaria  species — poses  serious 
threats  to  endemic  species.  The  fact  that  in- 
fectious diseases  have  resulted  in  epidemics  or 


204 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


epizootics  (e.g.,  C.  psittaci  and  T.  gallinae ) in 
Columbidae  and  other  bird  taxa  suggests  that 
regular  population  and  disease  surveys  are 
needed  for  Galapagos  Doves.  High  rates  of 
gene  flow  in  Galapagos  Doves  could  contrib- 
ute to  the  endangerment  of  native  and  endem- 
ic species  prone  to  the  effects  of  introduced 
pathogens  that  can  be  transmitted  across  spe- 
cies (e.g.,  Galapagos  Dove  lice  being  trans- 
mitted to  Galapagos  Hawks  during  predation; 
Whiteman  et  al.  2004).  We  recommend  that 
the  Galapagos  Dove  be  considered  a focal 
species  for  disease  research  in  the  archipelago 
because  it  could  serve  as  a reservoir/vector  for 
some  infectious  diseases  (Padilla  et  al.  2004). 

Morphology  and  dispersal. — That  we  found 
morphological  differences  between  some  is- 
land pairs  is  not  congruent  with  low  genetic 
differentiation  and  high  rates  of  gene  flow 
among  islands.  Lack  of  concordance  between 
morphology  and  genetics,  however,  is  not  un- 
common; through  the  use  of  mtDNA  markers, 
it  has  been  reported  for  other  groups,  such  as 
reptiles  (Schmitt  et  al.  2000,  Brehm  et  al. 
2001),  mollusks  (Mukaratirwa  et  al.  1998),  in- 
sects (Baranyi  et  al.  1997),  and  birds  (Seutin 
et  al.  1993,  1994;  Zink  and  Dittmann  1993; 
Freeman-Gallant  1996). 

One  might  expect  that  morphological  dif- 
ferences would  have  been  erased  by  the  con- 
nectedness between  populations.  However,  be- 
cause genes  under  selective  pressure  likely 
control  morphological  traits,  and  because  ^ST 
assumes  neutral  markers,  selectively  neutral 
markers  might  not  track  morphological  differ- 
ences among  populations.  We  do  not  believe 
that  processes  such  as  genetic  drift  are  impor- 
tant in  determining  the  morphological  differ- 
ences in  Galapagos  Doves,  since  they  require 
that  gene  flow  be  restricted  among  popula- 
tions. Alternatively,  morphological  characters 
can  be  very  plastic  and  might  vary  within  spe- 
cies, depending  on  the  environmental  charac- 
teristics of  an  area.  Many  studies  have  shown 
that  environmental  factors  are  sufficient  to 
produce  morphological  changes,  either  im- 
mediately or  within  a few  generations  (James 
1983,  Losos  et  al.  1997,  Trussed  and  Etter 
2001).  In  other  words,  environmentally  in- 
duced differences  among  populations  are  in- 
dependent of  genetic  differences.  Another 
possibility  is  that  even  where  dove  popula- 
tions are  sympatric  and/or  affected  by  high 


rates  of  gene  flow,  there  may  be  an  ecotypic- 
differentiation  process  driven  by  divergent  se- 
lection (Schluter  2001).  This  has  been  report- 
ed in  several  studies  and  for  different  taxa 
(Schluter  2001,  Ogden  and  Thorpe  2002). 
Based  on  the  estimated  effective  population 
sizes  for  the  different  islands  (from  —400  on 
Santa  Fe  to  —4,800  on  Genovesa),  the  migra- 
tion rates  (0  to  —70  individuals  per  genera- 
tion) represent  —2%  of  the  effective  size  of 
the  population  on  the  different  islands.  At  this 
level  of  migration,  the  genetic  influx  might 
not  completely  counteract  the  effects  of  selec- 
tion (Conner  and  Haiti  2004),  which  could  ac- 
count for  the  morphological  differences  ob- 
served in  our  study. 

ACKNOWLEDGMENTS 

We  thank  all  who  provided  help  during  the  different 
stages  of  the  field  season,  particularly  N.  K.  Whiteman, 
J.  L.  Bollmer,  G.  Jimenez,  J.  Merkel,  J.  Rabenold,  and 
N.  Gottdenker.  We  thank  the  staff  of  the  Charles  Dar- 
win Research  Station  for  their  invaluable  help  and  lo- 
gistical support  during  the  course  of  this  study,  espe- 
cially P.  Robayo.  We  also  thank  N.  Freire  and  J.  Mi- 
randa who  helped  with  dove  sampling  at  Tortuga  Bay, 
Santa  Cruz.  Permits  for  sample  collection  were  pro- 
vided by  Galapagos  National  Park.  We  thank  B.  A. 
Loiselle,  R.  E.  Ricklefs,  B.  T.  Ryder,  A.  Cohen,  and  J. 
L.  Bollmer  for  helpful  comments  and  suggestions  on 
earlier  versions  of  this  manuscript.  We  thank  H.  Vargas 
for  sharing  his  knowledge  of  Galapagos  Doves  from 
islands  not  visited  in  this  study.  We  thank  P.  Beerli 
who  provided  guidance  on  using  program  MIGRATE. 
We  thank  R.  L.  Curry  and  two  anonymous  reviewers 
for  comments  that  greatly  improved  the  manuscript. 
Financial  support  was  provided  by  The  International 
Center  for  Tropical  Ecology,  The  Saint  Louis  Zoo,  and 
E.  Des  Lee  Collaborative  Vision  in  Zoological  Re- 
search. 

LITERATURE  CITED 

Balloux,  E,  H.  Brunner,  N.  Lugon-Moulin,  J.  Haus- 
ser,  and  J.  Goudet.  2000.  Microsatellites  can  be 
misleading:  an  empirical  and  simulation  study. 
Evolution  54:1414-1422. 

Balloux,  F.  and  N.  Lugon-Moulin.  2002.  The  esti- 
mation of  population  differentiation  with  micro- 
satellite markers.  Molecular  Ecology  11:155-165. 
Baptista,  L.  E,  P.  W.  Trail,  and  H.  M.  Horblit.  1997. 
Family  Columbidae  (pigeons  and  doves).  Pages 
60-243  in  Handbook  of  the  birds  of  the  world, 
vol.  4:  sandgrouse  to  cuckoos  (J.  del  Hoyo,  A. 
Elliot,  and  J.  Sargatal,  Eds.).  Lynx  Edicions,  Bar- 
celona, Spain. 

Baranyi,  C.,  G.  Gollmann,  and  M.  Bobin.  1997.  Ge- 
netic and  morphological  variability  in  roach  Ru- 


Santiago-Alarcon  et  al.  • MORPHOLOGY  AND  GENETICS  OF  THE  GALAPAGOS  DOVE  205 


tilus  rutilus,  from  Austria.  Hydrobiologia  350:13- 
23. 

Beerli,  P.  and  J.  Felsenstein.  1999.  Maximum-like- 
lihood estimation  of  migration  rates  and  effective 
population  numbers  in  two  populations  using  a 
coalescent  approach.  Genetics  152:763-773. 

Beerli,  P.  and  J.  Felsenstein.  2001.  Maximum  like- 
lihood estimation  of  a migration  matrix  and  effec- 
tive population  sizes  in  n subpopulations  by  using 
a coalescent  approach.  Proceedings  of  the  Nation- 
al Academy  of  Sciences,  USA  98:4563-4568. 

Berry,  O.,  M.  D.  Tocher,  and  S.  D.  Sarre.  2004. 
Can  assignment  tests  measure  dispersal?  Molec- 
ular Ecology  13:551-561. 

Boag,  P.  T.  1981.  Morphological  variation  in  the  Dar- 
win’s finches  ( Geospizinae ) of  Daphne  Major  Is- 
land, Galapagos.  Ph.D.  dissertation,  McGill  Uni- 
versity, Montreal,  Canada. 

Boag,  P.  T.  1983.  The  heritability  of  external  mor- 
phology in  Darwin’s  ground  finches  ( Geospiza ) on 
Isla  Daphne  Major,  Galapagos.  Evolution  37:877- 
894. 

Bohonak,  A.  J.  1999.  Dispersal,  gene  flow,  and  pop- 
ulation structure.  Quarterly  Review  of  Biology  74: 
21-45. 

Bollmer,  J.  L.  2000.  Genetic  and  morphologic  differ- 
entiation among  island  populations  of  Galapagos 
Hawks  ( Buteo  galapagoensis ).  M.Sc.  thesis,  Ohio 
State  University,  Columbus. 

Bollmer,  J.  L.,  T.  Sanchez,  M.  D.  Cannon,  D.  San- 
chez, B.  Cannon,  J.  C.  Bednarz,  T.  de  Vries,  M. 
S.  Struve,  and  P.  G.  Parker.  2003.  Variation  in 
morphology  and  mating  system  among  island 
populations  of  Galapagos  Hawks.  Condor  105: 
428-438. 

Bowman,  R.  L.  1961.  Morphological  differentiation 
and  adaptation  in  the  Galapagos  finches.  Univer- 
sity of  California  Publications  in  Zoology,  no.  58. 

Brehm,  A.,  M.  Khadem,  J.  Jesus,  P.  Andrade,  and  L. 
Vicente.  2001.  Lack  of  congruence  between  mor- 
phometric evolution  and  genetic  differentiation 
suggests  a recent  dispersal  and  local  habitat  ad- 
aptation of  the  Madeiran  lizard  Lacerta  dugesii. 
Genetics  Selection  and  Evolution  33:671-685. 

Buckee,  C.  O’.  F.,  K.  Koelle,  M.  J.  Mustard,  and  S. 
Gupta.  2004.  The  effects  of  host  contact  network 
structure  on  pathogen  diversity  and  strain  struc- 
ture. Proceedings  of  the  National  Academy  of  Sci- 
ences, USA  101:10839-10844. 

Caccone,  A.,  G.  Gentile,  J.  P.  Gibbs,  T.  H.  Fritts,  H. 
L.  Snell,  J.  Betts,  and  J.  R.  Powell.  2002.  Phy- 
logeography  and  history  of  giant  Galapagos  tor- 
toises. Evolution  56:2052-2066. 

Caccone,  A.,  J.  P.  Gibbs,  V.  Ketmaier,  El  Suatoni, 
and  J.  R.  Powell.  1999.  Origin  and  evolutionary 
relationships  of  giant  Galapagos  tortoises.  Pro- 
ceedings of  the  National  Academy  of  Sciences, 
USA  96:13223-13228. 

Coleman,  M.  and  R.  J.  Abbott.  2003.  Possible  causes 
of  morphological  variation  in  an  endemic  Mor- 
occan groundsel  ( Senecio  leucanthemifolius  var. 


casablancae ):  evidence  from  chloroplast  DNA 
and  random  amplified  polymorphic  DNA  markers. 
Molecular  Ecology  12:423-434. 

Conner,  J.  K.  and  D.  L.  Hartl.  2004.  A primer  of 
ecological  genetics.  Sinauer,  Sunderland,  Massa- 
chusetts. 

CORNUET,  J.-M.,  S.  PlRY,  G.  LUIKART,  A.  ESTOUP,  AND 
M.  Solignac.  1999.  New  methods  employing 
multilocus  genotypes  to  select  or  exclude  popu- 
lations as  origins  of  individuals.  Genetics  153: 
1989-2000. 

Curry,  R.  L.  1988.  Group  structure,  within  group  con- 
flict and  reproductive  tactics  in  cooperatively 
breeding  Galapagos  Mockingbirds,  Nesomimus 
parvulus.  Animal  Behaviour  36:1708-1728. 

Curry,  R.  L.  1989.  Geographic  variation  in  social  or- 
ganization of  Galapagos  Mockingbirds:  ecological 
correlates  of  group  territoriality  and  cooperative 
breeding.  Behavioral  Ecology  and  Sociobiology 
25:147-160. 

Curry,  R.  L.  and  P.  R.  Grant.  1989.  Demography  of 
the  cooperatively  breeding  Galapagos  Mocking- 
bird, Nesomimus  parvulus , in  a climatically  vari- 
able environment.  Journal  of  Animal  Ecology  58: 
441-463. 

Curry,  R.  L.  and  S.  H.  Stoleson.  1988.  New  bird 
records  from  the  Galapagos  associated  with  the  El 
Nino-Southern  Oscillation.  Condor  90:505-507. 

de  Vries,  T.  J.  1973.  The  Galapagos  Hawk,  an  eco- 
geographical  study  with  special  reference  to  its 
systematic  position.  Ph.D.  thesis,  Vrije  University, 
Amsterdam,  Holland. 

de  Vries,  T.  J.  1975.  The  breeding  biology  of  the  Ga- 
lapagos Hawk,  Buteo  galapagoensis.  Le  Gerfaut 
65:29-57. 

El  Mousadik,  A.  and  R.  J.  Petit.  1996.  High  level  of 
genetic  differentiation  for  allelic  richness  among 
populations  of  the  argan  tree  [Argania  spinosa 
(L.)  Skeels]  endemic  of  Morocco.  Theoretical  and 
Applied  Genetics  92:832-839. 

Frankham,  R.  1996.  Relationship  of  genetic  variation 
to  population  size  in  wildlife.  Conservation  Biol- 
ogy 10:1500-1508. 

Frankham,  R.  1997.  Do  island  populations  have  less 
genetic  variation  than  mainland  populations?  He- 
redity 78:311-327. 

Frankham,  R.  1998.  Inbreeding  and  extinction:  island 
populations.  Conservation  Biology  12:665-675. 

Freeman-Gallant,  C.  R.  1996.  Microgeographic  pat- 
terns of  genetic  and  morphological  variation  in 
Savannah  Sparrows  ( Passerculus  sandxvichensis). 
Evolution  50:1631-1637. 

Fridolfsson,  A.  K.  and  H.  Ellegren.  1999.  A simple 
and  universal  method  for  molecular  sexing  of  non- 
ratite  birds.  Journal  of  Avian  Biology  30:1 16-121. 

Gaggiotti,  O.  E.,  O.  Lange,  K.  Rassmann,  and  C. 
Gliddon.  1999.  A comparison  of  two  indirect 
methods  for  estimating  average  levels  of  gene 
flow  using  microsatellite  data.  Molecular  Ecology 
8:1513-1520. 

Gifford,  E.  W.  1913.  Expedition  of  the  California 


206 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Academy  of  Sciences  to  the  Galapagos  Islands, 
1905-1906.  Proceedings  of  the  California  Acad- 
emy of  Sciences  2:1-132. 

Goldstein,  D.  B.  and  C.  Schlotterer.  1999.  Micro- 
satelites:  evolution  and  applications.  Oxford  Uni- 
versity Press,  New  York. 

Goodwin,  D.  1977.  Pigeons  and  doves  of  the  world. 
Cornell  University  Press,  Ithaca,  New  York. 

Goudet,  J.  1999.  An  improved  procedure  for  testing 
key  innovations.  American  Naturalist  53:549- 
555. 

Goudet,  J.  2002.  FSTAT,  ver.  2. 9. 3. 2.  University  of 
Lausanne,  Lausanne,  Switzerland,  www2.unil.ch/ 
popgen/softwares/fstat.htm  (accessed  September 
2004). 

Goudet,  J.,  M.  Raymond,  T.  Demeeus,  and  F.  Rous- 
set.  1996.  Testing  genetic  differentiation  in  dip- 
loid populations.  Genetics  144:1933-1940. 

Grant,  P.  R.  1998.  Evolution  on  islands.  Oxford  Uni- 
versity Press,  Oxford,  United  Kingdom. 

Grant,  P.  R.  2001.  Reconstructing  the  evolution  of 
birds  on  islands:  100  years  of  research.  Oikos  92: 
385-403. 

Grant,  P.  R.,  I.  Abbott,  D.  Schluter,  R.  L.  Curry, 
and  L.  K.  Abbott.  1985.  Variation  in  the  size  and 
shape  of  Darwin’s  finches.  Biological  Journal  of 
the  Linnean  Society  25:1-39. 

Grant,  P.  R.,  B.  R.  Grant,  and  K.  Petren.  2005.  Hy- 
bridization in  the  recent  past.  American  Naturalist 
166:56-67. 

Grant,  P.  R.  and  K.  T.  Grant.  1979.  Breeding  and 
feeding  ecology  of  the  Galapagos  Dove.  Condor 
81:397-403. 

Harmon,  W.  M.,  W.  A.  Clark,  A.  C.  Hawbecker,  and 
M.  Stafford.  1987.  Trichomonas  gallinae  in  col- 
umbiform  birds  from  the  Galapagos  Islands.  Jour- 
nal of  Wildlife  Diseases  23:492-494. 

Hedrick,  P.  W.  2000.  Genetics  of  populations.  Jones 
and  Bartlett,  Sudbury,  Massachusetts. 

Hedrick,  P.  W.  2001.  Conservation  genetics:  where  are 
we  now?  Trends  in  Ecology  and  Evolution  16: 
629-636. 

Hutchison,  D.  W.  and  A.  R.  Templeton.  1999.  Cor- 
relation of  pairwise  genetic  and  geographic  dis- 
tance measures:  inferring  the  relative  influences  of 
gene  flow  and  drift  on  the  distribution  of  genetic 
variability.  Evolution  53:1898-1914. 

James,  F.  C.  1983.  Environmental  component  of  mor- 
phological differentiation  in  birds.  Science  221: 
184-186. 

Johnson,  K.  P.  and  D.  H.  Clayton.  2000.  A molecular 
phylogeny  of  the  dove  genus  Zenaida : mitochon- 
drial and  nuclear  DNA  sequences.  Condor  102: 
864-870. 

Longmire,  J.  L.,  A.  K.  Lewis,  N.  C.  Brown,  J.  M. 
Buckingham,  L.  M.  Clark,  M.  D.  Jones,  L.  J. 
Meincke,  et  al.  1988.  Isolation  and  characteriza- 
tion of  a highly  polymorphic  centromeric  tandem 
repeat  in  the  family  Falconidae.  Genomics  2:14- 
24. 

Losos,  J.  B.,  K.  I.  Warheit,  and  T.  W.  Schoener. 


1997.  Adaptive  differentiation  following  experi- 
mental island  colonization  in  Anolis  lizards.  Na- 
ture 387:70-73. 

Lovette,  I.  J.,  E.  Bermingham,  and  R.  E.  Ricklefs. 
2001.  Clade-specific  morphological  diversification 
and  adaptive  radiation  in  Hawaiian  songbirds. 
Proceedings  of  the  Royal  Society  of  London,  Se- 
ries B 269:37-42. 

Mack,  R.  N.,  D.  Simberloff,  W.  M.  Lonsdale,  H. 
Evans,  M.  Clout,  and  F.  A.  Bazzaz.  2000.  Biotic 
invasions:  causes,  epidemiology,  global  conse- 
quences, and  control.  Ecological  Applications  10: 
689-710. 

Mantel,  N.  1967.  The  detection  of  disease  clustering 
and  a generalized  regression  approach.  Cancer  Re- 
search 27:209-220. 

McCoy,  K.  D.,  T.  Boulinier,  C.  Tirard,  and  Y.  Mich- 
alakis.  2003.  Host-dependent  genetic  structure  of 
parasite  populations:  differential  dispersal  of  sea- 
bird tick  host  races.  Evolution  57:288-296. 

McGarigal,  K.,  S.  Cushman,  and  S.  Stafford.  2000. 
Multivariate  statistics  for  wildlife  and  ecology  re- 
search. Springer- Verlag,  New  York. 

McQuistion,  T.  E.  1991.  Eimeria  palumbi,  a new  coc- 
cidian  parasite  (Apicomplexa:  Eimeriidae)  from 
the  Galapagos  Dove  {Zenaida  galapagoensis). 
Transactions  of  the  American  Microscopical  So- 
ciety 110:178-181. 

Mete,  A.,  G.  H.  A.  Borst,  and  G.  M.  Dorrestein. 
2001.  Atypical  poxvirus  lesions  in  two  Galapagos 
Doves  {Nesopelia  g.  galapagoensis).  Avian  Pa- 
thology 30:159-162. 

Mukaratirwa,  S.,  T.  K.  Kristensen,  H.  R.  Siegis- 
mund,  and  S.  K.  Chandiwana.  1998.  Genetic  and 
morphological  variation  of  populations  belonging 
to  the  Bulinus  truncatusl tropicus  complex  (Gas- 
tropoda: Planorbidae)  in  south  western  Zimbabwe. 
Journal  of  Molluscan  Studies  64:435-446. 

Nei,  M.  1973.  Analysis  of  gene  diversity  in  subdivided 
populations.  Proceedings  of  the  National  Acade- 
my of  Sciences,  USA  70:3321-3323. 

Ogden,  R.  and  R.  S.  Thorpe.  2002.  Molecular  evi- 
dence for  ecological  speciation  in  tropical  habi- 
tats. Proceedings  of  the  National  Academy  of  Sci- 
ences, USA  99:13612-13615. 

Padilla,  L.  R.,  D.  Santiago-Alarcon,  J.  Merkel,  E. 
Miller,  and  P.  G.  Parker.  2004.  Survey  for  Hae- 
moproteus  spp..  Trichomonas  gallinae,  Chlamy- 
dophila  psittaci  and  Salmonella  spp.  in  Galapagos 
Islands  Columbiformes.  Journal  of  Zoo  and  Wild- 
life Medicine  35:60-64. 

Paetkau,  D.,  R.  Slade,  M.  Burdens,  and  A.  Estoup. 
2004.  Genetic  assignment  methods  for  the  direct, 
real-time  estimation  of  migration  rate:  a simula- 
tion-based exploration  of  accuracy  and  power. 
Molecular  Ecology  13:55-65. 

Petit,  R.  J.,  A.  El  Mousadik,  and  O.  Pons.  1998. 
Identifying  populations  for  conservation  on  the 
basis  of  genetic  markers.  Conservation  Biology 
12:844-855. 

Piry,  S.,  A.  Alapetite,  J.-M.  Cornuet,  D.  Paetkau, 


Santiago-Alarcon  et  al.  • MORPHOLOGY  AND  GENETICS  OF  THE  GALAPAGOS  DOVE  207 


L.  Baudouin,  and  A.  Estoup.  2004.  GENE- 
CLASS2:  a software  for  genetic  assignment  and 
first-generation  migrant  detection.  Journal  of  He- 
redity 95:536-539. 

Prestwich,  A.  A.  1959.  The  Galapagos  Dove  in  free- 
dom and  captivity.  Avicultural  Magazine  65:66- 
76. 

Primmer,  C.  R.  and  H.  Ellegren.  1998.  Patterns  of 
molecular  evolution  in  avian  microsatellites.  Mo- 
lecular Biology  and  Evolution  15:997-1008. 

Pruett,  C.  L.,  S.  E.  Henke,  S.  M.  Tanksley,  M.  E. 
Small,  K.  M.  Hogan,  and  J.  Roberson.  2000. 
Mitochondrial  DNA  and  morphological  variation 
of  White- winged  Doves  in  Texas.  Condor  102: 
871-880. 

Pyle,  P.  1997.  Identification  guide  to  North  American 
birds,  part  I.  Slate  Creek  Press,  Bolinas,  Califor- 
nia. 

Ralph,  C.  J.,  G.  R.  Geupel,  P.  Pyle,  T.  E.  Martin,  D. 
F.  DeSante,  and  B.  Mila.  1996.  Manual  de  me- 
todos  de  campo  para  el  monitoreo  de  aves  terres- 
tres.  General  Technical  Report  159,  USDA  Forest 
Service,  Pacific  Southwest  Research  Station,  Al- 
bany, California. 

Rice,  W.  R.  1989.  Analyzing  tables  of  statistical  tests. 
Evolution  43:223-225. 

Ridgway,  R.  1897.  Birds  of  the  Galapagos  archipela- 
go. Proceedings  of  the  U.S.  National  Museum  19: 
459-670. 

Schluter,  D.  2001.  Ecology  and  the  origin  of  species. 
Trends  in  Ecology  and  Evolution  16:372-380. 

Schmitt,  L.  H.,  R.  A.  How,  S.  Hisheh,  J.  Goldberg, 
and  I.  Maryanto.  2000.  Geographic  patterns  in 
genetic  and  morphological  variation  in  two  skink 
species  along  the  banda  arcs,  southeastern  Indo- 
nesia. Journal  of  Herpetology  34:240-258. 

Seutin,  G.  J.,  J.  Brawn,  R.  E.  Ricklefs,  and  E.  Ber- 
mingham.  1993.  Genetic  divergence  among  pop- 
ulations of  a tropical  passerine,  the  Streaked  Sal- 
tator  ( Saltator  albicollis).  Auk  110:1 17-126. 

Seutin,  G.  J.,  N.  K.  Klein,  R.  E.  Ricklefs,  and  E. 
Bermingham.  1994.  Historical  biogeography  of 
the  Bananaquit  ( Coereva  flaveola ) in  the  Carib- 
bean region:  a mitochondrial  DNA  assessment. 
Evolution  48:1041-1061. 

Slatkin,  M.  1985.  Gene  flow  in  natural  populations. 
Annual  Review  of  Ecology  and  Systematics  16: 
393-30. 

Slatkin,  M.  1993.  Isolation  by  distance  in  equilibrium 


and  non-equilibrium  populations.  Evolution  47: 
264-279. 

SPSS,  Inc.  2001.  SPSS,  release  11.0.1.  SPSS,  Inc., 
Chicago,  Illinois. 

Swarth,  H.  S.  1931.  The  avifauna  of  the  Galapagos 
Islands.  Occasional  Papers  of  the  California  Acad- 
emy of  Sciences,  no.  18. 

Tanksley,  S.  M.  2000.  Analysis  of  genetic  differen- 
tiation in  White-winged  Doves.  Ph.D.  dissertation, 
Texas  A&M  University,  College  Station. 

Traxler,  B.,  G.  Brem,  M.  Muller,  and  R.  Achmann. 
2000.  Polymorphic  DNA  microsatellites  in  the  do- 
mestic pigeon,  Columba  livia  var.  domestica.  Mo- 
lecular Ecology  9:365-378. 

Trussell,  G.  C.  and  R.  J.  Etter.  2001.  Integrating 
genetic  and  environmental  forces  that  shape  the 
evolution  of  geographic  variation  in  a marine 
snail.  Genetica  112:321-337. 

Vargas,  H.  1996.  What  is  happening  with  the  avifauna 
of  San  Cristobal?  Noticias  de  Galapagos  57:23- 
24. 

Waples,  R.  S.  1998.  Separating  the  wheat  from  the 
chaff:  patterns  of  genetic  differentiation  in  high 
gene  flow  species.  Journal  of  Heredity  89:438- 
450. 

Weir,  B.  S.  and  C.  C.  Cockerham.  1984.  Estimating 
F-statistics  for  the  analysis  of  population  struc- 
ture. Evolution  38:1358-1370. 

Whiteman,  N.  K.,  S.  J.  Goodman,  B.  J.  Sinclair,  T. 
Walsh,  A.  A.  Cunningham,  L.  D.  Kramer,  and 
P.  G.  Parker.  2005.  Detection  of  the  avian  disease 
vector  Culex  quinquefasciatus  Say  1823  (Diptera: 
Culicidae)  on  the  Galapagos  Islands,  Ecuador,  af- 
ter a 14-year  interval.  Ibis  147:844-847. 

Whiteman,  N.  K.,  D.  Santiago-Alarcon,  K.  P.  John- 
son, and  P.  G.  Parker.  2004.  Differences  in  strag- 
gling rates  between  two  genera  of  dove  lice  (In- 
secta:  Phthiraptera)  reinforce  population  genetic 
and  cophylogenetic  patterns.  International  Journal 
for  Parasitology  34: 113-119. 

Whitlock,  M.  C.  and  D.  E.  McCauley.  1999.  Indirect 
measures  of  gene  flow  and  migration:  Fsx  1/ 
(4 Am  + 1).  Heredity  82:117-125. 

Wikelski,  M.,  J.  Foufopoulos,  H.  Vargas,  and  H. 
Snell.  2004.  Galapagos  birds  and  diseases:  inva- 
sive pathogens  as  threats  for  island  species.  Ecol- 
ogy and  Society  9(1  ):5.  www.ecologyandsociety. 
org/vol9/issl/art5  (accessed  5 January  2005). 

Zink,  R.  M.  and  D.  L.  Dittmann.  1993.  Gene  flow, 
refugia,  and  evolution  of  geographic  variation  in 
the  Song  Sparrow  ( Melospiza  melodia).  Evolution 
47:717-729. 


The  Wilson  Journal  of  Ornithology  1 18(2):208— 217,  2006 


BREEDING  ECOLOGY  OF  AMERICAN  AND  CARIBBEAN 
COOTS  AT  SOUTHGATE  POND,  ST.  CROIX: 

USE  OF  WOODY  VEGETATION 

DOUGLAS  B.  McNAIR1’34  AND  CAROL  CRAMER-BURKE2 3 4 


ABSTRACT. — American  ( Fulica  americana ) and  Caribbean  ( F . caribaea)  coots  nested  colonially  at  brackish 
Southgate  Pond,  St.  Croix,  United  States  Virgin  Islands  (USVI),  following  a 50-year  rainfall  event  in  mid- 
November  2003.  Breeding  occurred  during  three  time  periods:  seven  pairs  bred  from  6 December  to  2 January 
(early),  seven  from  17  January  to  15  February  (middle),  and  eight  from  26  April  to  19  May  (late)  (range  of 
clutch  initiation  dates  = 165  days).  Hatching  success  was  high  (65.3%),  but  overall  reproductive  success  was 
low  (27%)  owing  to  poor  brood  survival.  Coots  built  all  but  2 of  22  nests  at  the  water  line  in  sturdy  crotches 
of  small,  live  white  mangroves  ( Laguncularia  racemosaf  two  late  nests  were  built  on  remnant  stubs  of  dead 
white  mangroves  after  water  levels  had  sharply  declined.  Early  pairs  nested  in  manglars  (islets  of  one  or  more 
mangroves  without  solid  land)  farther  away  from  shore  and  in  deeper  water  than  middle  or  late  pairs  (65.6 
versus  42.1  and  29.0  cm,  respectively).  Southgate  Pond  remains  the  preferred  breeding  site  for  coots  on  St. 
Croix  and  the  USVI.  Coots  have  also  recently  nested  on  St.  Croix  at  seven  semi -permanent  or  permanent,  man- 
made, freshwater  ponds  where  they  have  probably  been  overlooked,  as  coots  respond  rapidly  to  changes  in 
water  levels  at  semi-permanent  or  permanent  wetlands.  Predominance  of  non-assortative  pairing  at  Southgate 
Pond  suggests  that  American  and  Caribbean  coots  are  morphs  of  one  species.  Received  7 February  2005, 
accepted  7 November  2005. 


The  Caribbean  Coot  ( Fulica  caribaea)  is 
not  globally  threatened  (Taylor  1996),  but  the 
species  is  listed  as  locally  endangered  in  the 
United  States  Virgin  Islands  (USVI;  Indige- 
nous and  Endangered  Species  Act  of  1990) 
and  is  considered  threatened  throughout  the 
West  Indies,  especially  breeding  populations 
(Raffaele  et  al.  1973,  1998).  Caribbean  and 
American  (F.  americana)  coots  are  two  of  the 
rarest  bird  species  that  nest  in  wetlands  of  the 
USVI,  including  St.  Croix  (Beatty  1930,  Raf- 
faele 1989),  and  their  breeding  ecology  in  the 
Caribbean  is  poorly  known  (Taylor  1996, 
Brisbin  et  al.  2002).  In  North  America,  Amer- 
ican Coots  are  associated  with  freshwater 
marshes  and  low-salinity  brackish  wetlands 
(Kantrud  1985). 

Following  the  largest  rainfall  event  in  over 
50  years,  we  studied  the  breeding  ecology  of 
Caribbean  and  American  coots  at  Southgate 
Pond,  the  largest  seasonal  brackish  pond  on 
St.  Croix.  Although  degraded  by  previous 


1 Div.  of  Fish  and  Wildlife,  Dept,  of  Planning  and 
Natural  Resources,  45  Mars  Hill,  Frederiksted,  St. 
Croix,  USVI  00840,  USA. 

2 St.  Croix  Environmental  Assoc.,  Arawak  Bldg., 
Ste.  #3,  Christiansted,  USVI  00820,  USA. 

3 Current  address:  Sapphos  Environmental,  Inc.,  133 
Martin  Alley,  Pasadena,  CA  91 105,  USA. 

4 Corresponding  author;  e-mail: 
dmcnair@sapphosenvironmental.com 


coastal  development,  Southgate  Pond  is  still 
one  of  the  most  productive  mangrove  wet- 
lands for  birds  on  St.  Croix  (Scott  and  Car- 
bonell  1986,  Sladen  1992;  DBM  and  CCB  un- 
publ.  data).  We  describe  coot  breeding  adap- 
tations in  use  of  woody  vegetation  as  nest 
sites  (Sugden  1979),  and  provide  information 
on  phenology,  clutch  size,  and  breeding  suc- 
cess. We  also  present  recent  breeding  infor- 
mation (since  2002)  on  coots  for  seven  other 
sites  on  St.  Croix,  formulate  management 
strategies  (especially  for  Southgate  Pond),  and 
assess  the  taxonomic  significance  of  pairing 
between  the  two  species. 

METHODS 

During  2003-2004,  we  studied  American 
and  Caribbean  coots  at  Southgate  Pond,  a 
17.9-ha  wetland  (17°  45'  29.6"  N,  64°  39' 
45.9"  W)  on  St.  Croix,  USVI.  We  used  the  cri- 
teria of  Roberson  and  Baptista  (1988)  to  dis- 
tinguish American  (types  A and  B)  from  Ca- 
ribbean coots  (types  C,  D,  and  E)  in  the  field. 
A small  percentage  (<1.4%)  of  the  males  with 
broad,  high,  and  bulbous  shields  may  be 
white-shielded  morphs  of  American  Coots 
(Roberson  and  Baptista  1988).  Types  A and  B 
have  a dark  chestnut  or  red-brown  corneous 
callus,  whereas  types  C,  D,  and  E lack  a cal- 
lus. After  becoming  familiar  with  vocal  dif- 


208 


McNair  and  Cramer-Burke  • COOTS  AT  SOUTHGATE  POND,  ST.  CROIX 


209 


ferences  between  the  sexes  (Gullion  1950),  we 
also  identified  the  genders  of  some  coots  at 
their  nests.  Males  were  larger  than  females 
and  had  larger  shields  and  bills,  regardless  of 
species,  which  agrees  with  expectations  based 
on  size  and  hormonal  differences  between  the 
sexes  (Gullion  1951;  Fredrickson  1968,  1970). 

We  visited  Southgate  Pond  twice  a week 
after  the  first  nest  was  discovered  in  early  Jan- 
uary 2004.  Nests  were  marked  with  numbered 
flagging  and  the  location  of  each  nest  was  re- 
corded with  a Global  Positioning  System 
(GPS)  unit  and  plotted  on  a map  using 
Arc  View  3.2.  We  recorded  the  coot  species 
associated  with  each  nest  and  coot  behavior 
during  each  nest  visit.  Some  individuals  were 
not  identified  to  species  because  of  their  elu- 
sive behavior.  Dates  of  clutch  initiation  for 
nests  found  during  laying  were  calculated  by 
backdating  and  assuming  that  one  egg  was 
laid  per  day  (Gorenzel  et  al.  1982,  Brisbin  et 
al.  2002).  Assuming  a 23-day  incubation  pe- 
riod (Brisbin  et  al.  2002),  initiation  dates  for 
nests  found  after  laying  were  estimated  based 
on  hatch  dates  minus  1 day  (the  day  on  which 
the  first  egg  hatched).  For  failed  nests,  we  ad- 
justed hatch  date  for  incomplete  or  under-re- 
corded clutch  sizes  based  on  the  mean  clutch 
size  and  backdating  from  the  midpoint  be- 
tween the  first  and  last  egg  dates.  Because  our 
potential  renest  intervals  were  long,  renests 
were  not  assigned  to  any  one  pair  of  coots 
(based  on  criteria  in  Arnold  1993). 

We  used  the  method  of  Mayfield  (1975),  as 
modified  by  Johnson  (1979),  to  calculate 
hatching  success  (based  on  a 23-day  incuba- 
tion period).  To  determine  reproductive  suc- 
cess, we  followed  the  fate  of  individuals  and 
broods  until  they  were  fully  grown  and  inde- 
pendent (60-70  days;  Taylor  1996).  Young 
coots  leave  the  nest  on  the  day  of  hatching 
and  broods  are  difficult  to  count  accurately 
when  young  birds  hide  in  emergent  vegetation 
(Gullion  1956);  however,  emergent  vegetation 
was  scarce  at  Southgate  Pond.  As  young  ac- 
quired juvenal  plumage  (~3  weeks  old)  they 
left  the  breeding  area  for  deeper  water  along 
the  northwestern  shore  of  Southgate  Pond, 
where  different  broods  coalesced  into  larger 
flocks  and  were  easier  to  see  and  count.  All 
nesting  attempts  had  known  outcomes  and  we 
calculated  reproductive  success  (number  of 
young  fledged/number  of  eggs  laid)  by  (1) 


multiplying  the  number  of  active  nests  by 
mean  clutch  size  to  derive  an  estimate  of  the 
total  number  of  eggs  laid,  and  (2)  dividing  the 
number  of  fully  grown  and  independent  young 
(not  broods  per  se)  by  eggs  laid.  Fledging  suc- 
cess (number  of  young  fledged/number  of 
eggs  hatched)  was  determined  by  dividing  re- 
productive success  by  hatching  success.  The 
number  of  breeding  pairs  was  based  on  the 
number  of  active  nests.  Coot  nest  density  at 
Southgate  Pond  and  the  seven  man-made 
freshwater  ponds  was  calculated  based  on 
pond  area  and  the  number  of  nests  or  pairs 
simultaneously  active  at  each  pond.  Assess- 
ment of  intraspecific  brood  parasitism  (“nest- 
dumping”) followed  the  criteria  of  Post  and 
Seals  (2000). 

We  recorded  the  following  parameters  at 
each  active  nest  and  nest  site:  nest  height  from 
the  water  line  to  the  top  of  the  nest  rim  (cm), 
length  and  width  of  outer  nest  cup  (cm), 
length  and  width  of  inner  lining  (cm),  water 
depth  below  the  nest  (cm),  above-water  height 
(cm)  and  greatest  breadth  (m)  of  the  white 
mangrove,  distance  to  nearest  white  mangrove 
(m),  distance  to  nearest  shoreline  (m),  distance 
to  nearest  active  nest  (m),  and  distance  to 
nearest  active  or  inactive  nest  (m).  For  each 
pair  of  coots,  four  variables  (water  depth  be- 
low the  nest,  distance  from  the  water  line  to 
the  top  of  the  nest  rim,  height  of  white  man- 
grove above  water,  distance  to  nearest  shore- 
line) were  adjusted  to  the  date  of  clutch  ini- 
tiation. We  also  noted  whether  white  man- 
groves that  contained  nests  were  isolated 
manglars  (islets  of  one  or  more  mangroves 
without  solid  land)  or  formed  a line  of  con- 
nected manglars  away  from  the  shoreline.  We 
used  a bathymetric  map  of  Southgate  Pond  to 
adjust  distances  between  nests  and  the  shore- 
line by  taking  the  mean  value  of  four  distance 
measurements  from  the  —15.25  to  30.5  cm 
contour  (—0.5  to  1 foot)  centered  on  the  main 
breeding  area.  We  then  used  sine/cosine  func- 
tions to  calculate  an  angle  of  0.026  degrees, 
which  translated  to  a 1.9-m  change  in  shore- 
line distance  per  cm  drop  (or  rise)  in  water 
levels.  Baseline  water  level  data  (in  cm)  were 
recorded  in  situ  from  several  2-m  sticks 
placed  in  the  lowest  bed  of  the  flat-bottomed 
pond.  The  water  level  decline  was  nearly  con- 
stant throughout  the  study  period  (mean  of 
0.58  cm/day),  except  for  one  heavy  rainfall 


210 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


event  when  water  levels  rose  17  cm  from  16 
to  18  April.  We  obtained  monthly  measure- 
ments of  salinity  at  Southgate  Pond  using  a 
temperature-compensated  refractometer  (ac- 
curate to  within  ± 1 ppt).  From  December 
2003  through  July  2004,  salinity  increased 
from  4 to  32  ppt  as  water  levels  dropped. 

To  systematically  sample  coot  breeding 
habitat,  we  established  a grid  of  56  line  tran- 
sects, spaced  8 m apart  along  north-south  car- 
dinal directions  from  15  m west  of  the  south- 
western shoreline  and  extending  to  the  eastern 
point  of  Southgate  Pond  just  beyond  the  main 
coot  colony.  We  randomly  selected  sample 
points  ( n = 436)  every  8 m along  each  tran- 
sect. The  last  point  along  each  transect  was  a 
point  in  open  water  beyond  the  vegetation  far- 
thest from  shore.  Water  depth  (cm),  vegetation 
present  or  absent  (open  water),  and  species 
composition  (if  vegetation  present)  were  sam- 
pled at  each  point.  We  used  a random  number 
generator  to  assign  numbers  1 to  22  (i.e.,  cor- 
responding to  the  number  of  coot  nests  we 
found)  to  sample  points  of  water  depth.  Water 
depth  at  each  sampled  point  that  contained 
vegetation  was  then  adjusted  to  reflect  water 
depth  at  the  observed  or  estimated  date  of 
clutch  initiation  for  each  coot  nest  represented 
by  each  random  number  (e.g.,  random  number 
one  represents  coot  nest  one,  which  initiated 
incubation  on  6 December).  This  procedure 
removed  the  effects  of  declining  water  levels 
so  that  vegetation  data  would  be  comparable 
to  nest  data. 

To  assess  differences  in  water  depths  be- 
tween vegetation  and  open  water  and  among 
species  of  plants,  we  used  two-tailed  f-tests 
and  one-way  ANOVA  (Zar  1999,  StatSoft 
2002).  Because  the  sample  sizes  for  five  of 
the  eight  vegetative  species/types  recorded 
were  small  (total  n = 15),  we  did  not  include 
them  in  the  ANOVA.  We  used  simple  linear 
regression  to  assess  the  relationship  between 
water  depth  below  nests  and  the  date  of  clutch 
initiation.  We  used  a Mann-Whitney  U- test  to 
assess  whether  water  depths  at  coot  nests  dif- 
fered from  random  and  to  examine  whether 
phenological  or  habitat  variables  were  related 
to  hatching  outcome  (success/fail).  We  used 
nonparametric  tests  (Mann-Whitney  U , Krus- 
kal-Wallis  //,  and  Spearman’s  rank  correlation 
rs ) when  sample  sizes  were  small  and  data  did 
not  otherwise  meet  assumptions  of  the  normal 


distribution,  including  homogeneity  of  vari- 
ances and  distribution  of  residuals.  For  all 
tests,  we  used  an  a value  of  0.05.  Means  are 
reported  ± SD. 

RESULTS 

We  located  22  active  coot  nests  at  South- 
gate  Pond  during  winter  and  spring  of  2003- 
2004.  Dates  of  clutch  initiation  ranged  from  6 
December  to  19  May  (165  days),  with  breed- 
ing occurring  during  three  periods:  early  (6 
December  to  2 January;  27  days),  middle  (17 
January  to  15  February;  29  days),  and  late  (26 
April  to  19  May;  23  days).  One  nesting  at- 
tempt during  the  late  period  was  overlooked 
(see  below).  We  identified  one  pair  of  Carib- 
bean Coots  and  five  Caribbean  X American 
coot  pairs  (hereafter  mixed  pairs)  during  the 
early  period,  two  pairs  of  Caribbean  Coots 
and  four  mixed  pairs  during  the  middle  period 
(two  male  American  and  two  female  Carib- 
bean coots  were  sexed  in  two  of  these  four 
mixed  pairs),  and  three  pairs  of  Caribbean 
Coots  and  two  mixed  pairs  during  the  late  pe- 
riod (both  males  were  American  and  both  fe- 
males were  Caribbean  coots).  One  of  the 
American  Coots  of  one  mixed  pair  during 
each  of  the  first  two  periods  was  type  B (in- 
termediate, sensu  Roberson  and  Baptista 
1988).  The  other  American  Coots  appeared  to 
be  type  A birds.  Of  the  three  coots  whose 
mates  were  not  identified,  two  were  Caribbean 
and  one  was  American. 

Clutch  size  decreased  as  the  nesting  season 
progressed  ( rs  = —0.56,  P = 0.025)  and  av- 
eraged 6.88  ± 1.41  eggs  (range  = 5-9,  n — 
16).  Seventeen  of  22  nests  (77%)  hatched  at 
least  one  chick,  and  only  5 of  the  130  eggs 
(3.8%)  that  remained  unbroken  in  the  nest 
bowl  throughout  the  normal  incubation  period 
failed  to  hatch.  Daily  nest  survival  (5)  was 
0.982  ± 0.008  SE  and  hatching  success  was 
65.3%  (Mayfield  method).  Hatching  success 
was  not  related  to  clutch  initiation  dates  (U  = 
26,  Z = 1 .29,  P = 0.20)  or  any  other  pheno- 
logical or  habitat  variable,  although  successful 
nests  generally  began  earlier  and  had  larger 
clutches,  greater  water  depths,  and  were  far- 
ther away  from  shore  than  failed  nests.  Forty- 
one  young  became  fully  grown  and  indepen- 
dent 60-70  days  after  hatching.  This  excludes 
three  young — attended  by  a pair  of  Caribbean 
Coots — that  fledged  from  a ninth  nest  over- 


McNair  and  Cramer-Burke  • COOTS  AT  SOUTHGATE  POND,  ST.  CROIX 


21  1 


TABLE  1.  Measurements  of  14  nest  and  nest-site  parameters  for  22  coot  nests  built  in  white  mangroves  at 
Southgate  Pond,  St.  Croix,  U.S.  Virgin  Islands,  during  winter  and  spring  of  2003-2004. 


Parameter 

Mean  ± 

SD 

Range 

Nest  height  from  water  line  to  top  of  nest  rim  (cm) 

13.5 

6.9 

4.8-35.5 

Length  of  outer  nest  cup  (cm) 

35.9 

± 

8.8 

25.4-61.0 

Width  of  outer  nest  cup  (cm) 

28.6 

± 

4.2 

20.3-36.2 

Length  of  inner  lining  (cm) 

19.8 

+ 

2.2 

15.5-25.4 

Width  of  inner  lining  (cm) 

18.0 

± 

2.1 

14.0-22.9 

Water  depth  below  nest  (cm) 

44.8 

± 

17.6 

15.0-78.5 

Above-water  height  of  white  mangrove  (m)a 

1.9 

± 

0.6 

1. 0-3.7 

Greatest  breadth  of  white  mangrove  (m)b 

3.9 

± 

1.1 

1.5-5. 8 

Distance  to  nearest  white  mangrove  (m)c 

3.2 

-t- 

2.4 

0.0-8. 5 

Distance  to  nearest  shoreline  (m) 

48.4 

-t- 

26.0 

10.4-98.1 

Distance  to  nearest  active  nest  (m) 

60.4 

+ 

59.6 

18.7-308.2 

Distance  to  nearest  active  nest  (m)d  (excluding  three  isolated  nests) 

42.7 

-4- 

13.5 

18.7-60.2 

Distance  to  nearest  active  or  inactive  nest  (m) 

42.3 

± 

59.9 

10.2-283.9 

Distance  to  nearest  active  or  inactive  nest  (m)d  (excluding  three  isolated  nests) 

23.1 

-I- 

11.3 

10.2-50.5 

a One  outlier  excluded  (dead  white  mangrove:  nest  17;  height  <20  cm). 

bTwo  outliers  excluded  (one  dead  white  mangrove  and  one  live  white  mangrove:  nests  17,  21;  breadth  not  measured  and  = 55.7  m,  respectively). 
c One  outlier  excluded  (live  white  mangrove:  nest  22;  distance  = 127.9  m). 

d One  isolated  nest  excluded  from  each  of  early,  middle,  and  late  nesting  periods  (nests  7,  13,  and  17). 


looked  during  the  late  period  (date  of  clutch 
initiation  was  later  than  19  May).  The  largest 
single  brood  observed  comprised  five  young 
(from  a mixed  pair),  and  there  were  six  broods 
(from  four  mixed  pairs  and  two  Caribbean 
Coot  pairs)  with  four  young.  Reproductive 
success  was  27%,  and  fledging  success  was 
41.3%. 

Nests  were  built  along  the  water  line  in  par- 
tially submerged,  small,  live  white  mangroves 
( Laguncularia  racemosa\  Table  1).  Most  nests 
were  placed  either  in  the  central  crotch  (early 
and  middle  periods)  or  in  smaller  crotches  of 
outside  branches  (late  period);  two  nests  dur- 
ing the  late  period  were  also  placed  either  on 
remnants  of  dead  white  mangroves  under  live 
vegetation  or  on  unconcealed  dead  white  man- 
groves. All  nests  during  the  early  and  middle 
periods  had  short  or  long  ramps,  while  only 
two  nests  during  the  late  period  had  ramps. 
Nests  were  in  isolated  manglars  (n  = 18)  or 
in  rows  of  manglars  {n  = 4),  but  away  from 
mangroves  that  formed  the  outer  fringes  of 
Southgate  Pond’s  vegetated  shoreline.  Nests 
were  located  close  to  nest  materials,  the  bulk 
of  which  (excluding  sticks  and  twigs  of  man- 
groves) consisted  of  shoreline  sea  purslane 
( Sesuvium  portulacastrum ),  a perennial  suc- 
culent forb  also  used  to  construct  most  of  the 
ramps.  Seed  pods  of  Sesbania  sericea,  a short- 
lived shrub,  composed  the  inner  nest  lining  of 
several  nests.  The  dominant  submerged  plant 


(forb)  of  Southgate  Pond  was  widgeon  grass 
( Ruppia  maritima),  but  this  species  was  not 
used  as  nest  material.  Most  manglars,  both 
white  and  black  ( Avicennia  germinans ) man- 
groves, were  located  at  the  east  end  of  the 
pond,  where  most  nests  were  concentrated 
(Fig.  1).  Two  rather  isolated  nests  (7,  13)  were 
near  the  southwestern  shoreline,  and  the  most 
isolated  nest  (17)  was  near  the  northwestern 
shoreline.  The  density  of  coot  nests  during  the 
three  periods  was  0.39-0.45  nests/ha. 

Mean  water  depth  at  nests  was  44.8  cm  (Ta- 
ble 1)  and  declined  throughout  the  breeding 
season  (early  period:  65.6  cm  ± 11.0;  middle 
period:  42.1  cm  ± 3.8;  late  period:  29.0  cm 
± 9.5;  Kruskal-Wallis  H = 15.14,  P < 0.001; 
Fig.  2).  Early  nests  were  also  farther  away 
from  the  shoreline  than  middle  or  late  nests 
(early  period:  70.0  m ± 29.7;  middle  period: 
49.6  m ± 11.2;  late  period:  28.5  m ± 15.8; 
Kruskal-Wallis  H = 8.81,  P = 0.010).  Other 
comparisons  of  nest  or  nest-site  variables  be- 
tween early,  middle,  and  late  periods  were  not 
significantly  different. 

Vegetation  sampled  at  random  points  along 
line  transects  composed  34.6%  (n  = 151)  of 
breeding  habitat;  the  remainder  was  open  wa- 
ter ( n = 285),  where  mean  water  depth  was 
significantly  greater  than  in  vegetated  areas 
(open  water:  45.8  cm  ± 27.3;  vegetation:  37.4 
cm  ± 26.2;  t = 3.11,  df  434,  P = 0.002). 
Live  white  and  black  mangroves  and  dead 


212 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


A 


60  120  180  240  300  360  420  Meters 


FIG.  1 . The  location  of  22  coot  nests  at  Southgate  Pond,  St.  Croix,  U.S.  Virgin  Islands,  during  winter  and 
spring  of  2003-2004. 


white  mangroves  dominated  vegetation  types 
within  breeding  habitat,  and  mean  water 
depths  at  dead  and  live  white  mangroves  were 
significantly  deeper  than  at  black  mangroves 
(F2134  = 8.28,  P < 0.001;  Table  2).  Mean  wa- 
ter depths  at  live  white  mangroves  with  and 


90: 


80:  . 
70:* 


y = 61 .84  - 0.23x,  r2  = 0.68,  P < 0.001 


E 

60: 


20: 


10 1 — - 

0 20  40  60  80  100  120  140  160  180 

Date  of  clutch  initiation  (1=6  December) 


FIG.  2.  Relationship  between  water  depth  below 
22  coot  nests  and  the  date  of  clutch  initiation  at  South- 
gate  Pond,  St.  Croix,  U.S.  Virgin  Islands,  during  winter 
and  spring  of  2003-2004. 


without  coot  nests  were  similar  (without  nests: 
39.3  cm  ± 25.7;  with  nests:  44.8  cm  ± 17.6; 
Mann-Whitney  U = 434.5,  P = 0.24). 

Freshwater  ponds. — Since  2002,  1—3  pairs 
of  Caribbean  Coots  and  mixed  pairs  have  bred 
intermittently  year-round  at  seven  man-made, 
freshwater  ponds  on  St.  Croix,  which  range  in 
size  from  0.1  to  2.9  ha.  The  mean  coot  density 
at  all  sites  combined  for  all  breeding  sequenc- 
es over  the  4 years  was  4.2  pairs/ha  (range  = 
0.3-10.0,  n = 17)  and  apparent  hatching  suc- 
cess was  high  (16  of  24  nests  based  on  hatch 
rates  of  the  proportion  of  nests  found).  Most 
breeding  records  occurred  after  the  50-year 
rainfall  event  of  mid-November  2003  filled 
the  ponds.  This  event  followed  a dry  period, 
when  a variety  of  plant  species  had  colonized 
the  bottom  of  many  dry,  or  nearly  dry,  ponds. 
In  addition,  the  Virgin  Islands  Agricultural 
Station  Middle  Pond  (see  McNair  2006  for  list 
of  pond  names  and  their  locations  on  St. 
Croix)  was  deliberately  drained  during  winter 
2002-2003.  Water  levels  varied  between  years 
at  several  sites  when  nesting  occurred,  espe- 


McNair  and  Cramer-Burke  • COOTS  AT  SOUTHGATE  POND,  ST.  CROIX 


213 


TABLE  2.  Mean  water  depth  (cm)  for  eight  vegetation 

types  at 

Southgate  Pond,  St.  Croix,  U.S.  Virgin 

Islands,  during  winter  and  spring  of  2003—2004. 

Vegetation  type 

n 

Mean  ± SDh 

Dead  Laguncularia  racemosa 

54 

45.4  ± 26.9  A 

Live  Laguncularia  racemosa 

48 

39.3  ± 25.7  A 

Live  Avicennia  germinans 

34 

23.2  ± 21.2  B 

Dead  Avicennia  germinans 

3 

38.6  ± 20.6C 

Sesbania  sericea 

4 

43.9  ± 17.0C 

Sesuvium  portulacastrum 

4 

27.8  ± 28. 8C 

Sesuvium  portulacastrum  on  dead  L.  racemosa 

2 

52.7  ± 20.2C 

Sporobolus  virginicus a 

2 

6.8  ± 29.  lc 

All  vegetation 

151 

37.4  ± 26.2 

Open  water 

285 

45.8  ± 27.3 

a Seashore  rush  grass. 

b Overall  ^2,134  = 8.28,  P < 0.001;  rows  with  different  letters  (A,  B)  are  significantly  different  (Tukey’s  unequal  n HSD  post-hoc  tests:  P = 0.026  for 
live  Avicennia  germinans  versus  live  Laguncularia  racemosa',  P < 0.001  for  live  A.  germinans  versus  dead  L.  racemosa). 
c Sample  size  too  small  to  test. 


cially  at  the  Virgin  Islands  Agricultural  Sta- 
tion Middle  Pond.  Live  creeping  burrhead 
(Echinodorus  berteroi ) was  almost  absent 
there  in  2004,  when  the  pond  was  not  used  by 
coots  and  emergent  vegetation  was  restricted 
to  the  shoreline  when  the  water  level  was 
higher.  Although  coots  nested  in  a variety  of 
live  (five  species)  and  dead  (two  species)  veg- 
etation, woody  (especially  remnant  S.  sericea, 
at  four  ponds)  vegetation  rather  than  perennial 
herbaceous  vegetation  was  the  predominant 
nest  substrate  (18  of  27,  67%).  Nests  ranged 
from  4 to  33.5  m away  from  the  shoreline, 
and  water  depths  below  nests  were  generally 
greater  for  nests  built  in  woody  vegetation, 
especially  S.  sericea  (usually  1.25-2.25  m). 
The  bulky,  conspicuous  nests  composed  of 
sticks  of  S.  sericea  (—90  X 65  cm)  were  su- 
perficially shaped  like  the  above-water  portion 
of  a beaver  lodge.  Anthropogenic  disturbance 
at  these  seven  ponds  was  negligible  except 
around  Carlton  North  Pond,  where  all  vege- 
tation except  that  fringing  the  shoreline  was 
cleared  for  a housing  development  in  early 
October  2004;  however,  coots  continue  to 
breed  at  Carlton  North  Pond. 

DISCUSSION 

Because  of  a drought  on  St.  Croix  that  be- 
gan in  2002,  the  bottom  of  Southgate  Pond 
was  dry  in  2003  until  water  from  heavy  rains 
began  to  fill  the  pond  in  late  August.  None- 
theless, the  basin  was  only  about  one-quarter 
full  until  a 50-year  rainfall  event  during  10- 
14  November  2003  caused  Southgate  Pond  to 


overflow.  Coots  colonized  the  pond  and  began 
laying  eggs  within  2-3  weeks  after  this  sea- 
sonal wetland  filled  with  water,  typical  of 
coots  after  arrival  on  their  breeding  grounds 
(Alisauskas  and  Arnold  1994). 

When  conditions  are  suitable,  Southgate 
Pond  is  probably  the  preferred  breeding  site 
for  coots  on  St.  Croix  (and  in  the  USVI; 
McNair  2006),  even  though  freshwater  ponds, 
each  with  a small  number  of  birds,  support 
higher  breeding  densities  (this  study).  Al- 
though Southgate  Pond  is  brackish,  the  num- 
ber of  breeding  pairs  during  three  consecutive 
periods  from  December  to  May  did  not  de- 
cline as  salinity  increased  from  low  to  mod- 
erately brackish;  elsewhere,  breeding  densities 
typically  decline  as  salinity  increases  (Kantrud 
1985,  Arnold  1993).  Regardless,  semi-per- 
manent or  seasonal  wetlands  are  generally 
preferred  habitat  for  American  Coots  in  North 
America  (Kantrud  1985,  Arnold  1993,  Ali- 
sauskas and  Arnold  1994).  Nests  at  Southgate 
Pond,  which  generally  lacked  emergent  her- 
baceous vegetation,  were  built  in  woody  veg- 
etation. In  Saskatchewan,  small,  isolated,  par- 
tially submerged  willow  ( Salix  spp.)  clumps 
were  used  as  nest  sites  for  a substantial  per- 
centage (22%)  of  American  Coot  nests  during 
a wet  year  (Sugden  1979),  although  willows 
were  not  used  as  nest  materials.  This  is  dif- 
ferent from  what  we  observed  at  Southgate 
Pond,  where  white  mangroves  served  as  nest 
sites  and  as  nest-building  material;  remnant  or 
live  woody  plants  (especially  remnant  S.  ser- 
icea) at  freshwater  ponds  on  St.  Croix  were 


214 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


used  similarly.  In  another  Saskatchewan 
study,  coot  nests  were  composed  of  the  same 
plant  species  that  provided  support  for  the  nest 
(cf.  Sutherland  and  Maher  1987). 

Water  depths  below  many  nests  in  fresh- 
water ponds  on  St.  Croix  were  much  deeper 
than  water  depths  below  nests  in  white  man- 
groves at  Southgate  Pond.  Apart  from  South- 
gate  Pond,  the  most  suitable  freshwater  breed- 
ing site  for  coots  on  St.  Croix  has  been  Gran- 
ard  South  Pond,  where  three  pairs  nested  in 
remnant  Sesbania  and  other  nest  sites.  Unlike 
mangroves  at  Southgate  Pond,  suitable  rem- 
nant woody  vegetation  at  freshwater  ponds 
usually  becomes  available  only  when  these 
ponds  dry  up  and  then  refill  with  water,  which 
kills  the  colonizing  shrubs.  Emergent  vegeta- 
tion suitable  for  nests  at  some  of  these  ponds 
can  be  scarce,  even  when  water  levels  are  low. 
Nests  in  perennial  emergent  forbs  were  float- 
ing platforms  built  amongst  this  vegetation, 
which  is  typical  of  coot  nests  in  marshes 
(Fredrickson  1970,  Sugden  1979,  Gorenzel  et 
al.  1982,  Kantrud  1985,  Post  1990,  Alisauskas 
and  Arnold  1994,  Frost  and  Massiah  2001). 

At  Southgate  Pond,  water  depths  at  coot 
nests  during  each  period  were  typical  of  those 
observed  at  American  Coot  nests  on  the  North 
American  mainland  (Sugden  1979,  Gorenzel 
et  al.  1982,  Sutherland  and  Maher  1987,  Post 
1990,  Arnold  1993),  although  depths  during 
the  third  period  were  rather  shallow.  Even 
though  coots  on  the  North  American  mainland 
frequently  nest  in  residual  emergent  vegeta- 
tion (Gorenzel  et  al.  1982,  Alisauskas  and  Ar- 
nold 1993),  in  our  study  they  probably  avoid- 
ed using  dead  white  mangroves  as  nest  sites 
in  deeper  water  at  Southgate  Pond  because 
nests  in  these  sites  would  have  been  exposed. 
Were  it  not  for  the  effects  of  hurricanes  Hugo 
and  Marilyn  in  1989  and  1996,  which  killed 
many  white  mangroves  farther  from  shore, 
several  more  pairs  of  coots  may  have  used 
these  mangroves  as  nest  sites.  Coots  also 
avoided  nesting  in  black  mangroves,  which 
are  generally  located  closer  to  shore  than  the 
live  white  mangroves  they  used.  Water  depths 
at  nests  in  white  mangroves  during  the  late 
period  were  similar  to  mean  depths  at  black 
mangroves,  suggesting  that  water  depth  at 
black  mangroves  was  otherwise  acceptable  to 
coots.  However,  coots  generally  prefer  deeper 
water  farther  from  shore  (Sutherland  and 


Maher  1987,  Post  1990,  Arnold  1993).  Fur- 
thermore, white  mangroves  offer  superior 
structural  support  for  nests  (black  mangroves 
lack  the  sturdy  bowl-shaped  central  crotch  and 
low  lateral  branches)  and  greater  concealment. 
For  similar  reasons,  American  Coots  in  Sas- 
katchewan nested  in  live  willows  but  not 
quaking  aspens  ( Populus  tremuloides ) (Sug- 
den 1979). 

As  water  levels  declined,  nest-site  selection 
changed;  by  the  late  period,  the  central  crotch- 
es of  white  mangroves  were  no  longer  suitable 
(too  far  above  water).  Nonetheless,  inter-nest 
distances  remained  similar  during  all  three  pe- 
riods, suggesting  that  territory  sizes  (which 
were  not  measured)  also  remained  similar.  In- 
ter-nest distances  between  simultaneously  ac- 
tive (or  inactive)  nests  during  all  three  periods 
were  typical  of  those  observed  for  coots  else- 
where, although  published  data  are  unavail- 
able for  nests  limited  to  woody  vegetation. 
Unlike  what  has  been  observed  at  many  North 
American  sites  characterized  by  emergent 
vegetation,  coots  at  our  study  site  built  few 
non-nesting  platforms  (six  in  white  man- 
groves), and  the  distribution  and  structure  of 
nesting  cover  at  Southgate  Pond  did  not 
change  over  the  breeding  season.  Given  the 
fixed  number  of  live  white  mangroves  as  po- 
tential nest  sites  for  coots  at  Southgate  Pond, 
territorial  behavior  probably  prevented  any 
additional  coot  pairs  from  breeding  at  the  site. 
The  location  of  coot  nests  is  mainly  controlled 
by  territorial  spacing,  distance  from  shore,  and 
the  distribution  and  structure  of  nesting  cover 
(Gullion  1953,  Sugden  1979,  Sutherland  and 
Maher  1987).  Water  depth,  although  correlat- 
ed with  distance  from  shore  in  this  study,  was 
probably  a less  important  factor  in  nest-site 
selection. 

Nest  concealment  in  woody  vegetation 
must  have  been  effective  because  hatching 
success  at  Southgate  Pond  was  high.  Apparent 
hatching  success  was  also  high  at  freshwater 
sites,  which  is  typical  of  American  Coots 
(Gorenzel  et  al.  1982,  Alisauskas  and  Arnold 
1994,  Brisbin  et  al.  2002).  Intraspecific  nest 
parasitism  was  not  observed  at  Southgate 
Pond  or  at  the  freshwater  ponds.  Fledging  suc- 
cess at  Southgate  Pond,  although  not  consis- 
tently associated  with  differences  in  water 
depth,  was  low  (<41%).  This  contrasts  with 
apparent  fledging  success  at  freshwater  sites 


McNair  and  Cramer-Burke  • COOTS  AT  SOUTHGATE  POND,  ST.  CROIX 


215 


(this  study),  and  that  in  North  America,  which 
is  generally  high  (>50%;  Alisauskas  and  Ar- 
nold 1994).  Most  broods  observed  at  South- 
gate  Pond  consisted  of  2-3  birds,  lower  than 
the  number  typically  observed  at  freshwater 
ponds  (7  of  13  broods  had  >4  fledged  young). 
Thus,  we  speculate  that  brood  losses  within  5 
days  after  hatching  exceeded  50%  at  South- 
gate  Pond.  Low  survivorship  of  young  also 
occurred  during  the  early  brood  period  for 
White-cheeked  Pintails  {Anas  bahamensis ) at 
Humacao,  Puerto  Rico  (F.  J.  Vilella  pers. 
comm.),  where  most  losses  were  attributed  to 
rats  ( Rattus  spp.),  Great  Egrets  (Ardea  alba), 
and  Black-crowned  {Nycticorax  nycticorax ) 
and  Yellow-crowned  {Nyctanassa  violacea ) 
night-herons.  All  of  these  potential  predators 
were  present  at  Southgate  Pond. 

Despite  low  reproductive  success  at  South- 
gate  Pond,  the  long  intervals  between  breed- 
ing periods  and  the  similar  number  of  pairs 
breeding  during  each  period  suggest  that  some 
middle  and  late  period  nests  were  probably 
second  or  third  broods  rather  than  renests. 
Presumed  success  of  second  nesting  attempts 
also  occurred  at  three  of  the  seven  freshwater 
ponds.  Nesting  during  the  late  period  at 
Southgate  Pond  appeared  to  be  possible  be- 
cause of  heavy  rainfall  that  occurred  from  16 
to  17  April,  when  water  levels  rose  17  cm, 
allowing  coots  to  reset  their  breeding  clock 
despite  an  overall  drop  in  water  level  (14  cm) 
since  the  middle  breeding  period.  Before  the 
50-year  rainfall  event  of  mid-November  2003, 
coots  probably  last  nested  at  Southgate  Pond 
in  2001,  following  the  previous  torrential  rain- 
fall event  of  8 May  when  water  filled  the  pond 
(CCB  unpubl.  data).  This  opportunistic,  multi- 
brooded  breeding  response  to  aquatic  periods 
resulting  from  torrential  vernal  and  autumnal 
rainfalls  in  an  otherwise  semi-arid  environ- 
ment may  allow  coots  to  overcome  generally 
low  reproductive  success  on  St.  Croix.  Nev- 
ertheless, three  breeding  periods  during  one 
aquatic  phase  is  probably  exceptional  (DBM 
and  CCB  unpubl.  data).  How  frequently  and 
successfully  coots  breed  at  Southgate  Pond 
and  freshwater  sites  on  St.  Croix  in  the  future 
is  currently  being  determined  through  an  on- 
going wetlands  bird-monitoring  scheme. 

Management  recommendations. — South- 
gate  Pond  (now  part  of  the  Southgate  Coastal 
Reserve  owned  by  the  St.  Croix  Environmen- 


tal Association)  remains  favorable  habitat  for 
nesting  coots,  even  though  environmental 
degradation  has  diminished  this  brackish  pond 
to  <50%  of  its  original  size  (Gaines  2004, 
Gaines  and  Gladfelter  2004).  The  most  diffi- 
cult task  at  Southgate  Pond  is  to  maintain  ap- 
propriate water  levels  for  coot  nest  initiation 
during  seasons  and  years  when  rainfall  is  in- 
sufficient. We  endorse  Gaines  and  Gladfelter’s 
(2004:54-56)  two  major  recommendations  for 
water  management  to  prolong  the  aquatic 
phase  of  Southgate  Pond:  (1)  divert  water  into 
the  pond,  and  (2)  raise  the  maximum  water 
depth  from  —103  to  —138  cm.  Manipulation 
of  water  levels  should  favor  nesting  coots  and 
other  wetland  birds,  although  it  may  eliminate 
species  that  nest  in  terrestrial  sites.  During  its 
dry  phase,  two  species  of  conservation  con- 
cern on  St.  Croix  may  nest  at  Southgate  Pond: 
Wilson’s  Plover  {Charadrius  wilsonia)  and 
Least  Tern  {Sterna  antillarum).  However,  both 
species  breed  at  more  than  10  sites  and  are 
not  as  rare  as  coots.  Furthermore,  Southgate 
Pond  is  the  best-documented  site  for  coots  in 
the  eastern  Caribbean  (McNair  2006).  When 
water  levels  are  sufficient,  the  brackish  habitat 
at  Southgate  Pond  may  be  similar  to  that  of 
brackish  impoundments  along  the  northern 
Gulf  coast  of  the  United  States  (e.g.,  an  abun- 
dance of  sea  purslane  and  widgeon  grass), 
where  coots  are  abundant  (Swiderek  et  al. 
1988). 

At  the  seven  man-made,  freshwater  ponds, 
piped  water  is  generally  the  best  management 
option  to  maintain  stable,  generally  high  water 
levels.  The  most  suitable  freshwater  site  in  the 
eastern  Caribbean  (Barbados)  is  man-made 
Marshall’s  Pond,  which  is  dominated  by  Echi- 
nodorus  berteroi  (Frost  and  Massiah  2001;  M. 
D.  Frost  pers.  comm.),  the  herbaceous  species 
used  most  frequently  for  nest  sites  on  St. 
Croix.  Maintaining  stable  water  levels  at  the 
best  site  on  St.  Croix  (Granard  South  Pond), 
as  well  as  at  the  other  ponds,  should  generally 
favor  E.  berteroi  and  other  species  with  sim- 
ilar vegetative  characteristics.  Woody  vegeta- 
tion would  no  longer  compose  the  dominant 
nest  sites  because  stable  water  levels  would 
generally  prevent  woody  plants  such  as  S.  ser- 
icea  from  becoming  established  except  along 
the  immediate  shorelines  of  these  ponds. 

Caribbean  Coot  systematic s. — The  taxo- 
nomic status  of  the  Caribbean  Coot  requires 


216 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


further  investigation  (Roberson  and  Baptista 
1988;  also  Gullion  1951,  Phillips  1967,  Payne 
and  Master  1983,  Clark  1985,  Taylor  1996). 
Apparent  non-assortative  pairing  of  coots  pre- 
vailed at  Southgate  Pond,  where  both  types  of 
coots  occurred.  One-half  of  the  pairs  at  fresh- 
water sites  on  St.  Croix  were  paired  non-as- 
sortatively.  Furthermore,  at  least  some  mixed 
pairs  successfully  raised  young,  especially  at 
Southgate  Pond,  indicating  that  the  two  types 
of  coots  can  produce  viable  offspring  (Gill 
1964,  Payne  and  Master  1983,  Bond  1984). 
Thus,  American  and  Caribbean  coots  may 
compose  one  species  with  variant,  intergraded 
phenotypes  of  which  A and  E birds  represent 
the  extreme  types.  Although  some  birds  can 
be  individually  recognized  in  the  field,  an  ac- 
curate assessment  of  phylogenetic  relation- 
ships and  the  taxonomic  status  of  American 
and  Caribbean  coots  will  require  studies  based 
on  morphological  and  genetic  analyses  along 
with  observations  of  mating  behavior  and  pair 
bonds  of  marked  birds.  This  will  also  require 
confirming  identification  of  shield  character- 
istics and  correlating  them  with  other  mor- 
phological measurements. 

ACKNOWLEDGMENTS 

The  U.S.  Fish  and  Wildlife  Service  provided  partial 
financial  support  to  the  Division  of  Fish  and  Wildlife, 
U.S.  Virgin  Islands  (Federal  Aid  Program,  Pittman- 
Robertson  Wetlands  Project,  W15).  S.  L.  Fromer  and 
L.  D.  Yntema  shared  their  bird  observations;  A.  G. 
Gaines  contributed  data  on  salinities  at  Southgate  Pond 
and  provided  a digitized  bathymetric  map  of  this  site; 
D.  B.  Nelthrop  and  D.  M.  Schuster,  among  other  land- 
owners,  provided  access  to  their  freshwater  ponds;  J. 
A.  Collazo  and  F.  J.  Vilella  reviewed  a penultimate 
draft  of  the  manuscript,  and  T.  W.  Arnold,  F.  E.  Hayes, 
and  an  anonymous  individual  reviewed  earlier  versions 
of  the  paper. 

LITERATURE  CITED 

Alisauskas,  R.  T.  and  T.  W.  Arnold.  1994.  American 
Coot.  Pages  126-143  in  Migratory  shore  and  up- 
land game  bird  management  in  North  America  (T. 
C.  Tacha  and  C.  E.  Braun,  Eds.).  International  As- 
sociation of  Fish  and  Wildlife  Agencies,  Allen 
Press,  Lawrence,  Kansas. 

Arnold,  T.  W.  1993.  Factors  affecting  renesting  in 
American  Coots.  Condor  95:273-281. 

Beatty,  H.  A.  1930.  Birds  of  St.  Croix.  Journal  of  the 
Department  of  Agriculture  of  Puerto  Rico  14:1 35 — 
150. 

Bond,  J.  1984.  Twenty-fifth  supplement  to  the  check- 
list of  birds  of  the  West  Indies  (1956).  Academy 


of  Natural  Sciences  of  Philadelphia,  Philadelphia, 
Pennsylvania. 

Brisbin,  I.  L.,  Jr.,  H.  D.  Pratt,  and  T.  B.  Mowbray. 
2002.  American  Coot  ( Fulica  americana ) and  Ha- 
waiian Coot  ( F . alai).  The  Birds  of  North  Amer- 
ica, no.  697. 

Clark,  C.  T.  1985.  Caribbean  Coot?  Birding  17:84- 
88. 

Fredrickson,  L.  H.  1968.  Measurements  of  coots  re- 
lated to  sex  and  age.  Journal  of  Wildlife  Manage- 
ment 32:409-41 1. 

Fredrickson,  L.  H.  1970.  Breeding  biology  of  Amer- 
ican Coots  in  Iowa.  Wilson  Bulletin  82:445-458. 

Frost,  M.  D.  and  E.  B.  Massiah.  2001.  Caribbean 
Coot  ( Fulica  caribaea ):  the  return  of  a former 
breeding  resident  bird.  Journal  of  the  Barbados 
Museum  and  Historical  Society  47:85-90. 

Gaines,  A.  G.,  Jr.  2004.  The  Southgate  watershed:  ge- 
ology and  hydrology  of  an  arid  landscape.  SCR 
Technical  Report,  no.  2.  The  Coast  & Harbor  In- 
stitute, Woods  Hole,  Massachusetts. 

Gaines,  A.  G.,  Jr.,  and  E.  H.  Gladfelter.  2004.  The 
Southgate  Coastal  Reserve:  a strategy  for  man- 
agement and  implementation.  SCR  Technical  Re- 
port, no.  1.  The  Coast  & Harbor  Institute,  Woods 
Hole,  Massachusetts. 

Gill,  F.  B.  1964.  The  shield  color  and  relationships  of 
certain  Andean  coots.  Condor  66:209-211. 

Gorenzel,  W.  R,  R.  A.  Ryder,  and  C.  E.  Braun. 
1982.  Reproduction  and  nest  site  characteristics  of 
American  Coots  at  different  altitudes  in  Colorado. 
Condor  84:59-65. 

Gullion,  G.  W.  1950.  Voice  differences  between  sexes 
in  the  American  Coot.  Condor  52:272-273. 

Gullion,  G.  W.  1951.  The  frontal  shield  of  the  Amer- 
ican Coot.  Wilson  Bulletin  63:157-166. 

Gullion,  G.  W.  1953.  Territorial  behavior  of  the 
American  Coot.  Condor  55:169-186. 

Gullion,  G.  W.  1956.  An  observation  concerning  the 
validity  of  coot  brood  counts.  Journal  of  Wildlife 
Management  20:465-466. 

Johnson,  D.  H.  1979.  Estimating  nest  success:  the 
Mayfield  method  and  an  alternative.  Auk  96:651- 
661. 

Kantrud,  H.  A.  1985.  American  Coot  habitat  in  North 
Dakota.  Prairie  Naturalist  17:23-32. 

Mayfield,  H.  F.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

McNair,  D.  B.  2006.  Reviews  of  the  status  of  Amer- 
ican Coot  ( Fulica  americana)  and  Caribbean  Coot 
( Fulica  caribaea ) in  the  United  States  Virgin  Is- 
lands. North  American  Birds  59:678-684. 

Payne,  R.  T.  and  L.  L.  Master.  1983.  Breeding  of  a 
mixed  pair  of  white-shielded  and  red-shielded 
American  Coots  in  Michigan.  Wilson  Bulletin  95: 
467-469. 

Phillips,  A.  R.  1967.  Some  Antillean  coots  ( Fulica ) in 
the  Cambridge  University  and  British  Museums. 
Bulletin  of  the  British  Ornithological  Club  87:35- 
36. 

Post,  W.  1990.  American  Coots  nest  in  South  Carolina 


McNair  and  Cramer-Burke  • COOTS  AT  SOUTHGATE  POND,  ST.  CROIX 


217 


after  a 35-year  interlude,  and  a summary  of  South 
Carolina  coot  nidiology.  Chat  54:9-11. 

Post,  W.  and  C.  A.  Seals.  2000.  Breeding  biology  of 
the  Common  Moorhen  in  an  impounded  cattail 
marsh.  Journal  of  Field  Ornithology  71:437-442. 

Raffaele,  H.  1989.  A guide  to  the  birds  of  Puerto  Rico 
and  the  Virgin  Islands,  2nd  ed.  Princeton  Univer- 
sity Press,  Princeton,  New  Jersey. 

Raffaele,  H.  A.,  M.  J.  Velez,  R.  Cotte,  J.  J.  Wehlan, 
E.  R.  Keil,  and  W.  Cumpiano.  1973.  Rare  and 
endangered  animals  of  Puerto  Rico — a committee 
report.  U.S.  Department  of  Agriculture-Soil  Con- 
servation Service  and  Department  of  Natural  Re- 
sources, Commonwealth  of  Puerto  Rico. 

Raffaele,  H.,  J.  Wiley,  O.  Garrido,  A.  Keith,  and  J. 
Raffaele.  1998.  A guide  to  the  birds  of  the  West 
Indies.  Princeton  University  Press,  Princeton,  New 
Jersey. 

Roberson,  D.  and  L.  F.  Baptista.  1988.  White-shield- 
ed coots  in  North  America:  a critical  evaluation. 
American  Birds  42:1241-1246. 

Scott,  D.  A.  and  M.  Carbonell.  1986.  A directory 
of  Neotropical  wetlands.  International  Union  for 
Conservation  of  Nature  and  Natural  Resources, 
Cambridge,  United  Kingdom,  and  International 
Waterfowl  and  Wetlands  Research  Bureau,  Slim- 
bridge,  United  Kingdom. 


Sladen,  F.  W.  1992.  Abundance  and  distribution  of 
waterbirds  in  two  types  of  wetlands  on  St.  Croix, 
U.S.  Virgin  Islands.  Ornitologia  Caribena  3:35- 
42. 

StatSoft,  Inc.  2002.  Statistica  base,  ver.  6. 1 for  Win- 
dows. StatSoft,  Inc.,  Tulsa,  Oklahoma. 

Sugden,  L.  G.  1979.  Habitat  use  by  nesting  American 
Coots  in  Saskatchewan  parklands.  Wilson  Bulletin 
91:599-607. 

Sutherland,  J.  M.  and  W.  J.  Maher.  1987.  Nest-site 
selection  of  the  American  Coot  in  the  aspen  park- 
lands  of  Saskatchewan.  Condor  89:804-810. 

Swiderek,  P.  K.,  A.  S.  Johnson,  P.  E.  Hale,  and  R.  L. 
Joyner.  1988.  Production,  management,  and  wa- 
terfowl use  of  sea  purslane.  Gulf  Coast  muskgrass, 
and  widgeongrass  in  brackish  impoundments. 
Pages  441-457  in  Waterfowl  in  winter  (M.  W. 
Weller,  Ed.).  University  of  Minnesota  Press,  Min- 
neapolis. 

Taylor,  P.  B.  1996.  Family  Rallidae  (rails,  gallinules, 
and  coots).  Pages  108-209  in  Handbook  of  the 
birds  of  the  world,  vol.  3:  Hoatzin  to  auks  (J.  del 
Hoyo,  A.  Elliott,  and  J.  Sargatal,  Eds.).  Lynx  Ed- 
icions,  Barcelona,  Spain. 

Zar,  J.  H.  1999.  Biostatistical  analysis,  4th  ed.  Pren- 
tice-Hall, Upper  Saddle  River,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  1 1 8(2):2 1 8 — 224,  2006 


INSULAR  AND  MIGRANT  SPECIES,  LONGEVITY  RECORDS, 
AND  NEW  SPECIES  RECORDS  ON  GUANA  ISLAND, 
BRITISH  VIRGIN  ISLANDS 

CLINT  W.  BOAL,1 4 FRED  C.  SIBLEY,1 2  TRACY  S.  ESTABROOK,3 4  AND 

JAMES  LAZELL2 


ABSTRACT. — We  conducted  mist  netting  each  October  from  1994  to  2004  on  Guana  Island,  British  Virgin 
Islands,  and  recorded  bird  sightings  to  develop  a more  complete  inventory  of  the  island’s  resident  and  migrant 
species.  During  our  study,  we  recorded  four  new  species  for  the  British  Virgin  Islands:  Magnolia  Warbler 
{Dendroica  magnolia ; 1996),  Golden-winged  Warbler  ( Vermivora  chrysoptera\  1997),  Swainson’s  Thrush  ( Ca - 
tharus  ustulatus ; 2000),  and  Red-necked  Phalarope  ( Phalaropus  lobatus ; 2004).  Blackpoll  Warbler  ( Dendroica 
striata)  was  the  most  frequently  captured  Neotropical  migrant  landbird,  despite  only  being  first  detected  in  the 
region  in  1989.  Captures  and  detections  of  other  Neotropical  migrant  landbirds  suggest  that  many  species  may 
be  more  common  in  the  region  than  previously  believed,  or,  as  speculated  by  other  researchers,  that  migrant 
routes  may  be  shifting  eastward  due  to  habitat  degradation  on  western  Caribbean  islands.  We  also  used  recapture 
data  to  establish  longevity  records  of  resident  species,  including  Caribbean  Elaenia  ( Elaenia  martinica\  >7 
years),  Bananaquit  ( Coereba  flaveola-,  7 years),  Black-faced  Grassquit  ( Tiaris  bicolor,  >9  years),  and  Zenaida 
Dove  {Zenaida  aurita\  5 years).  Longevities  of  other  resident  species  were  similar  to,  or  slightly  less  than,  those 
reported  elsewhere.  Received  22  February  2005,  accepted  30  November  2005. 


Ornithological  research  conducted  in  the 
West  Indies  has  covered  an  array  of  topics, 
including  avian  species  occurrence  and  distri- 
bution, ecology  of  individual  species,  effects 
of  hurricanes  on  island  bird  populations,  mi- 
gration patterns,  and  community  dynamics 
(Wiley  2000).  In  the  Virgin  Islands  region,  re- 
searchers have  addressed  avifaunal  occurrence 
and  distribution  (LaBastille  and  Richmond 
1973,  Mirecki  et  al.  1977,  Norton  et  al.  1989), 
and  species  ecologies  (Askins  and  Ewert 
1991,  Chipley  1991,  Mayer  and  Chipley  1992, 
McNair  et  al.  2002);  however,  considerably 
less  ornithological  study  has  been  conducted 
in  the  Virgin  Islands — especially  the  British 
Virgin  Islands  (BVI) — than  in  other  areas  of 
the  West  Indies.  In  a bibliography  consisting 
of  1 1 ,648  entries  for  ornithological  work  con- 
ducted in  the  West  Indies  from  1750  to  1994, 
only  7.5%  of  the  entries  included  information 
for  the  Virgin  Islands;  only  the  extralimital  is- 
lands of  San  Andres,  Providencia,  and  the 


1 U.S.  Geological  Survey,  Texas  Coop.  Fish  and 
Wildlife  Research  Unit,  Dept,  of  Range,  Wildlife  and 
Fisheries  Management,  Texas  Tech  Univ.,  Lubbock, 
TX  79409-2120,  USA. 

2 The  Conservation  Agency,  6 Swinburne  St., 
Jamestown,  RI  02835,  USA. 

3 5529  90th  St.,  Lubbock,  TX  79424,  USA. 

4 Corresponding  author;  e-mail:  clint.boal@ttu.edu 


Swans  have  received  less  attention  (Wiley 

2000). 

We  conducted  mist  netting  on  Guana  Is- 
land, BVI,  each  October  from  1994  to  2004. 
To  our  knowledge,  the  Guana  Island  station  is 
the  only  current  and  consistently  operated 
banding  station  in  the  British  Virgin  Islands 
and  one  of  only  three  in  the  eastern  Caribbean 
(St.  Martin  and  Barbados  being  the  others). 
Previously,  information  from  the  island  has 
proven  important  in  developing  a better  un- 
derstanding of  Neotropical  migrant  bird  use  of 
the  region  during  the  autumn  migration 
(McNair  et  al.  2002).  However,  our  data  on 
species  frequency  of  occurrence,  which  could 
be  helpful  in  this  effort,  have  not  been  made 
available  until  now.  For  example,  Faaborg  and 
Terborgh  (1980)  considered  the  Red-eyed  Vir- 
eo  ( Vireo  olivaceus ) as  a rare  transient  mi- 
grant encountered  only  in  the  Greater  Antilles. 
In  a status  review  of  migrant  landbirds  in  the 
Caribbean,  Arendt  (1992)  did  not  list  Red- 
eyed Vireos  as  even  occurring  in  the  British 
Virgin  Islands.  Indeed,  Norton  (1996)  noted 
an  account  of  a Red-eyed  Vireo  in  Puerto  Rico 
as  one  of  only  a few  confirmed  records  for  the 
species  on  the  Puerto  Rico  Bank.  The  regular 
occurrences  of  Red-eyed  Vireos  at  Guana  Is- 
land (CWB  and  FCS  unpubl.  data),  however, 
suggest  that  the  species  uses  the  Virgin  Is- 
lands as  a migration  stopover  more  than  pre- 
viously believed. 


218 


Boal  et  al.  • BIRD  SPECIES  AND  LONGEVITY  ON  GUANA  ISLAND,  BVI 


219 


Here,  we  present  an  account  of  resident  and 
migrant  species  banded  during  October  each 
year  for  1 1 years  on  Guana  Island.  For  some 
species,  we  report  longevity  records  based  on 
recaptures  of  banded  individuals.  Additional- 
ly, we  provide  accounts  of  new  or  rarely  re- 
ported species  based  on  both  banding  and  site 
records. 

METHODS 

The  Virgin  Islands,  including  both  the  U.S. 
Virgin  Islands  and  the  BVI,  are  a chain  of 
approximately  76  islands  and  cays  located 
100-150  km  east  of  Puerto  Rico.  Guana  Is- 
land (18°  30'  N,  64°  30'  W)  lies  immediately 
north  of  Tortola,  the  largest  of  the  BVI  is- 
lands. Within  the  BVI,  Guana  Island  is  rela- 
tively small  (3  km2)  compared  with  other  in- 
habited islands,  such  as  Tortola  (54  km2).  Vir- 
gin Gorda  (21  km2),  and  Jost  Van  Dyke  (10 
km2).  The  BVI  has  a subtropical  climate  tem- 
pered by  northeasterly  trade  winds,  with  tem- 
peratures normally  ranging  from  28  to  33°  C, 
and  fairly  constant  humidity  levels  (—78%) 
throughout  the  year  (Lazell  2005).  Annual 
mean  rainfall  for  Guana  Island  is  estimated  at 
92  cm  (Lazell  2005),  but  data  are  limited  and 
the  long-term  average  may  be  lower. 

Guana  Island  is  topographically  rugged, 
with  elevations  ranging  from  sea  level  to  246 
m.  Approximately  90%  of  the  island  is  cov- 
ered by  subtropical  dry  forest,  with  ghut  for- 
ests (mesic  forest;  5%)  present  in  some  drain- 
ages; miscellaneous  covers  include  human-al- 
tered areas  (3%),  mangroves  (1%),  and  beach 
(1%)  (Lazell  1996;  CWB  unpubl.  data).  Lazell 
(1996)  lists  the  primary  native  vegetation  on 
Guana  Island  as  tabebuia  ( Tabebuia  hetero- 
phylla ),  gumbo-limbo  ( Bursera  simaruba), 
loblolly  ( Pisonia  subcordata),  buttonwood 
{Conocarpus  erectus),  frangipani  ( Plumeria 
alba),  acacia  {Acacia  muricata ),  and  sea  grape 
{Coccoloba  uvifera).  Tam-tam  {Leucaena  leu- 
cocephela)  is  common  in  disturbed  areas.  In- 
troduced species  include  coconut  {Cocos  nu- 
cifera),  tamarind  {Tamarindus  indica),  and 
royal  poinciana  {Delonix  regia). 

We  operated  a mist-netting  station  each  Oc- 
tober from  1994  to  2004.  Nets  were  located 
primarily  along  a northeast-southwest  ridge 
and  southeast-facing  slope  of  a mountain  on 
the  island’s  west  side.  The  majority  of  nets 
were  in  subtropical  dry  forest  areas,  but  each 


year  we  placed  2—3  nets  in  human-altered  ar- 
eas along  the  ridge,  all  at  approximately  100- 
m elevation.  For  one  afternoon  each  year,  we 
also  netted  along  the  shore  of  a salt  pond  to 
sample  the  shorebirds  present.  We  attempted 
to  use  the  same  net  locations  each  year,  but 
during  the  earlier  years  of  the  project  we  con- 
ducted some  “exploratory  netting”  in  other 
areas.  Duration  of  mist-netting  operations  and 
number  of  nets  operated  were  subject  to  local 
weather  conditions,  the  number  of  assistants 
available,  and  the  amount  of  time  we  were 
allowed  access  to  the  island  by  its  owners; 
thus,  the  number  of  nets  used  (mean  = 8.1  ± 
0.9  SE)  and  mist-netting  days  (mean  = 8.8  ± 

1 .3  SE)  varied  annually.  Weather  permitting, 
nets  were  opened  at  06:30  AST  and  closed 
between  10:00  and  11:00;  occasionally,  mist- 
netting  was  also  conducted  in  the  afternoon. 

We  identified  all  birds  captured  to  the  spe- 
cies level,  and,  when  possible,  determined 
their  sex  and  age  (Raffaele  1989,  Pyle  1997, 
Raffaele  et  al.  2003).  We  recorded  weight  (g), 
length  of  wing  chord  (mm),  and  banded  each 
bird  with  a federal  aluminum  leg  band.  We  did 
not  conduct  systematic  avian  surveys  (e.g., 
point  counts),  but  we  did  record  species  en- 
countered while  engaged  in  other  studies  and 
activities  on  the  island.  Combined,  our  obser- 
vation records  and  mist-netting  efforts  al- 
lowed us  to  compile  an  annual  species  list  for 
the  island  and  document  occurrences  of  spe- 
cies previously  unrecorded  on  the  island  and / 
or  the  BVI.  We  compiled  recapture  records  to 
determine  longevity  for  both  resident  and  mi- 
grant species.  We  considered  all  after-hatch- 
ing-year  birds  (AHY)  to  be  1 year  old  at  time 
of  initial  capture. 

RESULTS 

Banding. — We  conducted  mist  netting  for  a 
mean  of  252  ± 53  SE  net-hr  each  October 
from  1994  through  2004.  During  the  study  pe- 
riod, we  captured  1,410  birds,  188  (13%)  of 
which  were  recaptures  of  birds  banded  in  pre- 
vious years  (Table  1).  These  numbers  do  not 
include  captures  of  birds  that  we  did  not  band, 
such  as  the  Green-throated  Carib  {Eulampis 
holosericeus)  and  the  Antillean  Crested  Hum- 
mingbird {Orthorhyncus  cristatus).  We  cap- 
tured 44  species,  the  most  common  of  which 
was  the  resident  Bananaquit  {Coereba  flav- 
eola;  676  captures).  Other  frequently  captured 


220 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


c 

‘5b 

> 

SZ 

C/3 

CB 

T3 

C 

3 


90 

Cl 

00 


OONOooOOr-OOOOO  — 3-00  — OOOO-'tOOOOOOOOOOOOOOOOOc'lO 

— (N  _____  <N  — 


ooococNi/n  — io3-<ncn  — — — onm  — — cncnco-  3-  — c<-3  — (Nunr-cs  — NOonc^- NOcN»noo  — o 

— <N  — _____  (NO-)  3-  — — On  00  r-  3-  — 

_ NO  _ 3- 


OCN  — O — Oc*nO  — 0000m0(Nin003-0>/'30<N0  — OOI^OOO  — O — OcoOcnOOt^ 

— 3-  co  oi 


Nto  - OOOOcnOCNOO  — — 0 0X0  — (NOOO  — OOOCMO  — (NO  — — (N  — — — Ocn 

— — C"  — 3- 


— — oo  — oocn  — ooo  — 3-00  — — ooooooooooooo  — ooooomooooox 

3-  r-' 


0(N0(N0000  — OOOmNO  — — — O — OOOOOONOOOOOOOO  — OONOOt^ 

(N  (N  — co  — m 


^tmOOOO(Nir)(NOO  — ONOOO  — — cnOCNOc^OOOOOOfT',  — — Or^(NOO»nOO00  — 

— — 00  3"  — — 

— CO 


0000000000000  — 00 


OOOOOOOOCOOO  — 00000000<300000000000000CM00  — or" 


r*  to  >;  K 

£ 5 -2  „ a. 


•2  -c  £ 

w Cl 
93  -P  O 


(3 

“93  -g 

•2  £. 


§>.2 

■ c 


^ -s; 


< 5 


> _ 

0 X 

UT) 

1 ’S 

3 

c o 

_ 90 

73  c 


c 3 

90  — 

m oa 


•p  ‘r*  „C 


C/3 


ec-5  £ 
th  ’vJ  ^ 

r-<  ^ ^ 
w ^ 


g>.8-  8.  g 

> a a o 

f.  ^ o C/3 

in  Js  <u 

8.3 1? 

ac* 


90  u u 

J c/d  go 


S..O. 

a S- 

T3  ^ 

§ § 
co  CO 
•a  nO 

§ g. 


2 £ 
90  !> 
in  > 


- ^ s > 

5 cQ 

§ S -a 

w 00  c 


90  CL.  q 

l-s  6 


5.1 


* g b | | 

1 8 | s | 

§ io  o C3  r^N>  ^ 
N 3 2 e w 

S-eI  I 

g 1 U UJ  2 o > 

II |l SI i 

90  90  J3  3 i-  90  90 

N > 2 U O > & 


N ; C 
2?  | § 


~ 'd  ^ 

5 « Q 


2 J2 

PS  £> 


.93  a 

a '2 

■gs 

V.  w 


V-  _ C3 

SZ  T3  — 
f—  90  3 

r bo  h 


s 2 5;  ^ ? £ 

JO 

'— 

$ 


&0  ~ 
p .2 


=2  2 
P--2 

2 ^ 


90  — rn  « 


p 

> .2 
£ 1 

£o!2S 


s ~ ft* 

>N  ^ C 

V ' £ 

' C a) 
>*«  ^ 
3 ^ ~ 


90 

90  J2 

-e  m 

J3  T3 


11 

U CB 


!l 

O CO 

OQ  DQ 


to  in  s on 

w U Q OC 

t;  -P  ^ 2 

cd  c3  jS  £ 

5^.1  S 

§ 2 *2  9 


'a  * 

| § 


90  _ 

e-8^ 

90 


j=  3 


_90  u 
3 _90 

>> 

*r-n 

O 


« rt  £ .O  > o 


c ^ > 
< ^ O 


90 

TO 

c c 

90  O 

^ X 


^ c 

2 x> 


ca  o 
OQ  OC 


C3  P 

<43  (J  C30 

§ -3  g 
e-£d 

93  .2  SZ 

’I  w C 
*>  __  cc 
5 a = 

£f« 

<9  C 
00  rb  «J 
cO  U 

'c  ^ o 

3 « C 

CQ  .in  < 


33  - 
.3  U m cU 

2d3  Jh 


Boat  et  al.  • BIRD  SPECIES  AND  LONGEVITY  ON  GUANA  ISLAND,  BVI 


221 


TABLE  2.  Longevity  records  for  species 

>4 

years  old  on  Guana  Island,  British  Virgin  Islands, 

1994-2004. 

Species 

Agea 

Sex 

Year 

captured 

Last 

recapture 

No.  of 
recaptures 

Minimum 
age  (years) 

Wilson’s  Plover 

AHY 

F 

1996 

1999 

2 

4 

AHY 

M 

1996 

1999 

1 

4 

AHY 

U 

1996 

1999 

1 

4 

AHY 

M 

1996 

1999 

2 

4 

Black-necked  Stilt 

AHY 

F 

1997 

2001 

1 

5 

Spotted  Sandpiper 

HY 

U 

1998 

2004 

2 

6 

Common  Ground-Dove 

AHY 

F 

1998 

2001 

1 

4 

Zenaida  Dove 

AHY 

M 

1997 

2001 

2 

5 

AHY 

M 

1998 

2001 

1 

4 

AHY 

M 

2001 

2004 

1 

4 

Caribbean  Elaenia 

Unk 

U 

1996 

2003 

1 

7 

Unk 

U 

1996 

2001 

1 

5 

Pearly-eyed  Thrasher 

AHY 

U 

1998 

2001 

1 

4 

Black-faced  Grassquit 

AHY 

F 

1996 

2004 

2 

9 

AHY 

F 

1998 

2004 

1 

7 

AHY 

M 

1998 

2003 

1 

6 

HY 

U 

1998 

2003 

1 

5 

AHY 

F 

2000 

2004 

2 

5 

AHY 

M 

1996 

2000 

1 

5 

Bananaquit 

AHY 

M 

1995 

2001 

3 

7 

AHY 

M 

1997 

2003 

2 

7 

HY 

F 

1998 

2004 

2 

6 

AHY 

M 

1998 

2002 

2 

5 

AHY 

M 

1997 

2001 

2 

5 

AHY 

F 

1997 

2001 

1 

5 

HY 

M 

1998 

2003 

2 

5 

AHY 

M 

2001 

2004 

3 

5 

AHY 

M 

1997 

2000 

1 

4 

AHY 

M 

1995 

1998 

2 

4 

AHY 

M 

1995 

1998 

3 

4 

HY 

F 

1997 

2001 

2 

4 

AHY 

M 

2000 

2004 

2 

4 

AHY 

M 

2001 

2004 

2 

4 

AHY 

M 

1994 

1997 

2 

4 

HY 

F 

1998 

2002 

2 

4 

HY 

M 

1998 

2002 

2 

4 

AHY 

M 

1998 

2001 

1 

4 

AHY 

M 

1998 

2001 

1 

4 

a AHY  = after-hatching-year,  HY  = hatching-year,  Unk  = unknown  age. 


resident  species  were  Black-faced  Grassquit 
(Tiaris  bicolor,  148  captures)  and  Pearly-eyed 
Thrasher  ( Margarops  fuscatus\  93  captures). 
These  three  species  are  among  the  most  abun- 
dant residents  on  Guana  Island.  We  also  cap- 
tured 20  species  of  Neotropical  migrant  land- 
birds,  the  majority  of  which  were  warblers 
(Table  1).  The  Neotropical  migrant  captured 
most  frequently  was  the  Blackpoll  Warbler 
C Dendroica  striata ; 185  captures),  followed 
by  the  Red-eyed  Vireo  (12  captures,  multiple 
additional  sightings).  Other  Neotropical  mi- 
grants encountered  included  many  species 


(e.g.,  Yellow-throated  Vireo,  Vireo  flavifrons; 
Table  1)  previously  reported  only  from  the 
western  Greater  Antilles  or  for  which  there 
were  no  records  from  the  BVI  or  the  Lesser 
Antilles  (Faaborg  and  Terborgh  1980,  Arendt 
1992). 

Longevity. — We  determined  longevity  for 
all  species  recaptured  on  the  island,  and  pro- 
vide data  for  those  older  than  3 years  (Table 
2).  Among  shorebirds,  the  longevity  records 
were  5 years  for  Black-necked  Stilt  ( Himan - 
topus  mexicanus ),  6 years  for  Spotted  Sand- 
piper ( Actitis  macularius ),  and  4 years  for 


222 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Wilson’s  Plover  ( Charadrius  wilsonia );  how- 
ever, our  recapture  rate  for  these  species  was 
low  and  we  suspect  that  our  longevity  esti- 
mates, especially  for  the  resident  Wilson’s 
Plover,  may  be  substantially  lower  than  actual 
longevity.  Among  Columbiformes,  our  lon- 
gevity records  were  4 years  for  Common 
Ground-Dove  ( Columbina  passerina)  and  5 
years  for  Zenaida  Dove  ( Zenaida  aurita). 
Among  resident  passerines,  we  recaptured  Ca- 
ribbean Elaenias  ( Elaenia  martinica ) that 
were  >7  and  >5  years  old,  and  we  recaptured 
a 4-year-old  Pearly-eyed  Thrasher.  Among  the 
19  recaptured  Bananaquits,  two  were  7 years 
old,  one  was  6 years  old,  and  the  others  were 
5 and  4 years  old.  The  oldest  bird  recaptured 
was  a >9-year-old  female  Black-faced  Grass- 
quit;  we  also  recaptured  one  6-year-old  and 
three  5-year-old  grassquits. 

New  species  records. — During  the  course  of 
our  netting  operations  and  surveys,  we  ob- 
tained species  records  for  Guana  Island  and, 
in  some  cases,  the  British  Virgin  Islands.  Our 
captures  of  a Magnolia  Warbler  ( Dendroica 
magnolia)  in  1996  and  a Golden-winged  War- 
bler ( Vermivora  chrysoptera ) in  1997  were 
first  records  for  the  BVI.  More  significant, 
however,  was  our  capture  of  a Swainson’s 
Thrush  ( Catharus  ustulatus)  in  2000,  the  first 
record  for  the  Virgin  Islands  and  only  the  sec- 
ond from  east  of  Cuba  (McNair  et  al.  2002). 
In  2003,  we  captured  another  Swainson’s 
Thrush  and  obtained  a visual  sighting  of  a sec- 
ond, unbanded  individual.  Finally,  our  obser- 
vation of  a hatching-year  Red-necked  Phala- 
rope  ( Phalaropus  lobatus ) on  the  salt  pond  of 
Guana  Island  in  October  2004  represented  a 
first  record  for  that  species  in  the  Virgin  Is- 
lands. 

DISCUSSION 

Deriving  longevity  estimates  from  survi- 
vorship models  is  preferable  to  using  simple 
longevity  records  (Krementz  et  al.  1989).  The 
reliability  of  survival  estimates,  however,  de- 
pends upon  robust  recapture  data  (e.g.,  Burn- 
ham et  al.  1987),  which  often  are  not  available 
for  many  species.  Longevity  records,  there- 
fore, are  still  valuable  for  providing  some  ba- 
sic life-history  information  on  little-studied 
species.  This  may  be  especially  true  for  island 
settings,  where  longer-lived  species  are  at 
lower  risk  of  localized  extinction  (Newton 


1998).  Although  longevity  records  have  been 
reported  for  many  North  American  bird  spe- 
cies (e.g.,  Kennard  1975,  Klimkiewicz  et  al. 
1983),  little  information  is  available  on  the 
life  spans  of  tropical  birds  (Snow  and  Lill 
1974,  Faaborg  and  Winters  1979,  Johnston  et 
al.  1997).  The  few  Caribbean  bird  species  for 
which  there  are  longevity  records  are  primar- 
ily Puerto  Rican  (Faaborg  and  Winters  1979, 
Woodworth  et  al.  1999),  and  there  is  virtually 
no  published  information  on  the  longevity  of 
birds  in  the  eastern  Caribbean.  Thus,  our  data 
provide  new  age  records  for  several  Caribbean 
species.  In  Puerto  Rico,  Faaborg  and  Winters 
(1979)  recaptured  36  of  219  Bananaquits,  the 
oldest  of  which  was  4 years  and  7 months. 
Outside  of  the  Caribbean,  de  Souza  Lopes  et 
al.  (1980)  reported  a 4-year,  8-month-old  Ba- 
nanaquit  from  their  study  in  Brazil.  Our  lon- 
gevity record  of  7 years  for  Bananaquits  ex- 
ceeds previous  reports  by  a minimum  of  2 
years.  Furthermore,  our  Bananaquit  data  sug- 
gest that  ages  of  4 and  5 years  are  not  uncom- 
mon. Perhaps  most  unusual  is  our  9-year-old 
age  record  for  a Black-faced  Grassquit,  with 
additional  individuals  aged  6 and  5 years. 
These  far  exceed  the  previous  report  of  2 
years  and  1 1 months  (Faaborg  and  Winters 
1979).  The  4-year-old  Common  Ground-Dove 
in  our  study  is  similar  to  the  longevity  records 
of  4 years  and  4 years  and  1 month  from 
Puerto  Rico  (Faaborg  and  Winters  1979). 
However,  the  5-year,  5-month-old  Pearly-eyed 
Thrasher  reported  by  Faaborg  and  Winters 
(1979)  exceeds  our  oldest  known  thrasher  by 
1 to  2 years.  We  found  no  reports  of  longevity 
for  Caribbean  Elaenia  with  which  to  compare 
our  records;  however,  our  records  of  7-  and  5- 
year-old  Caribbean  Elaenia  are  similar  to 
those  reported  for  unspecified  Elaenia  spp.  in 
Brazil  (6  years  and  3 months,  and  5 years;  de 
Souza  Lopes  et  al.  1980)  and  substantially  ex- 
ceed ages  recorded  for  Yellow-bellied  Elaenia 
(E.  flavogaster ; 2 years  and  1 1 months)  and 
Mountain  Elaenia  ( E . frantzii ; 3 years  and  9 
months)  in  Panama  (Loftin  1975).  We  believe 
that  the  5-year-old  Zenaida  Dove  from  our 
study  also  represents  a longevity  record  for 
that  species,  as  we  could  find  no  reports  with 
which  to  compare  our  data. 

Many  of  the  Neotropical  migrants  captured 
or  sighted  during  our  study  are  known  to  oc- 
casionally occur  in  the  BVI.  Some  of  our 


Boat  et  al.  • BIRD  SPECIES  AND  LONGEVITY  ON  GUANA  ISLAND,  BVI 


223 


sightings  and  captures,  such  as  Hooded  War- 
blers ( Wilsonia  citrina ) and  Worm-eating 
Warblers  ( Helmitheros  vermivorum),  are  un- 
usual for  the  BVI.  Still  others,  including  Mag- 
nolia Warbler,  Golden-winged  Warbler, 
Swainson’s  Thrush,  and  Red-necked  Phala- 
rope,  provide  new  records  for  the  BVI.  De- 
tections of  Swainson’s  Thrush  and  Red- 
necked Phalarope  were  particularly  interest- 
ing. Within  the  Caribbean  region,  Raffaele  et 
al.  (2003)  indicated  that  Swainson’s  Thrush 
was  found  only  rarely  in  the  western  Greater 
Antilles  and  only  during  migration;  thus,  de- 
tections of  Swainson’s  Thrush  in  2 different 
years  on  Guana  Island  was  notable.  Raffaele 
et  al.  (2003)  also  indicated  that  Red-necked 
Phalarope  is  a very  rare  migrant  in  the  Ba- 
hamas, Cuba,  and  Hispaniola  (e.g..  Greater 
Antilles);  in  Puerto  Rico,  the  species  has  been 
recorded  only  twice  (Raffaele  1989).  In  Sep- 
tember 2003,  however,  a Red-necked  Phala- 
rope was  reported  on  Guadaloupe  Island  (Nor- 
ton et  al.  2003),  which  lies  400  km  southeast 
of  Guana  Island. 

Our  detections  of  Blackpoll  Warbler  and 
Red-eyed  Vireo,  and  our  consistent  detections 
of  other,  less  common  species — such  as  Yel- 
low-throated Vireo,  Swainson’s  Thrush,  Indi- 
go Bunting  ( Passerina  cyanea ),  and  numerous 
warbler  species — indicate  that  they  may  be 
more  common  in  the  eastern  Caribbean  during 
migration  than  previously  believed  due  to  a 
lack  of  searching  or  banding  efforts  in  that 
region.  For  example,  Blackpoll  Warbler,  the 
most  common  warbler  encountered  on  Guana 
Island  and  the  second-most  frequently  cap- 
tured species  overall,  was  not  reported  in  the 
BVI  until  1989  (Norton  1990);  it  had  been 
considered  a common  Neotropical  migrant 
through  the  Greater  Antilles  but  uncommon  to 
rare  on  other  islands  (Arendt  1992,  Raffaele 
et  al.  2003).  Similarly,  Red-eyed  Vireo  was 
thought  to  be  very  uncommon  or  vagrant  in 
the  Lesser  Antilles  (Faaborg  and  Terborgh 
1980,  Arendt  1992,  Norton  1996,  Raffaele  et 
al.  2003);  however,  our  regular  sightings  and 
captures  of  Red-eyed  Vireos  suggest  that  the 
species  may  be  a more  common  migrant  in 
the  BVI  than  previously  believed. 

Overall,  our  detections  of  species  previous- 
ly believed  to  be  uncommon  or  not  present 
within  the  BVI  may  have  been  due  to  a lack 
of  field  surveys  and  banding  efforts  through- 


out most  of  the  Virgin  Islands  and  Lesser  An- 
tilles. Alternatively,  our  detections  may  be  re- 
lated to  changes  in  habitat  conditions  in  the 
western  Caribbean  islands.  As  habitat  avail- 
ability decreases  in  the  western  islands,  some 
migrant  species  might  be  shifting  their  migra- 
tion routes  eastward  (Arendt  1992).  Regard- 
less of  possible  shifts  in  migration  routes,  it 
appears  that  Guana  Island — a functional  eco- 
system protected  as  a nature  preserve  (Lazell 
1996) — provides  important  habitat  for  both 
resident  and  transient  migrant  species.  A low- 
occupancy,  private  resort  occupies  less  than 
2%  of  the  surface  area  of  Guana  Island;  the 
remainder  of  the  island  is  almost  completely 
free  of  direct  human  impacts  and  exists  in  a 
near-natural  state  (Lazell  1996).  Furthermore, 
exotic  herbivores  and  carnivores,  which  are  a 
severe  problem  throughout  much  of  the  Ca- 
ribbean, occur  at  very  low  densities  and  are 
heavily  controlled  on  the  island. 

As  larger  islands  in  the  Virgin  Islands  (e.g., 
Tortola,  St.  John,  Virgin  Gorda)  continue  to 
undergo  deforestation  and  development  (e.g., 
Arendt  1992),  smaller  islands  maintained  in 
primarily  natural  states  are  likely  to  become 
increasingly  important  for  conservation  of 
both  resident  and  migrant  birds.  However, 
small  islands,  such  as  Guana  Island,  may  not 
provide  a full  range  of  landscape  characteris- 
tics required  for  some  migrant  or  wintering 
Neotropical  songbirds.  For  example.  Northern 
Parula  ( Parula  americana ) and  American 
Redstart  ( Setophaga  ruticilla),  both  common 
nonbreeding  residents  in  the  Virgin  Islands 
(Raffaele  et  al.  2003),  are  seldom  detected  on 
Guana.  Further  examination  of  resource  use 
and  spatial  needs  of  Neotropical  songbirds  mi- 
grating through  or  wintering  in  the  BVI  is 
needed  to  facilitate  conservation  efforts. 

ACKNOWLEDGMENTS 

We  give  our  sincere  thanks  to  G.  Jarecki,  H.  Jarecki, 
and  the  staff  of  Guana  Island  for  their  support  and 
facilitation  of  this  research.  We  thank  W.  J.  Arendt,  A. 
Olivieri,  G.  Perry,  O.  Perry,  J.  Richardson,  P.  Sibley, 
A.  Sutton,  S.  Valentine,  and  T.  Willard  for  assisting 
with  banding  operations  and  other  logistics.  Support 
for  the  study  was  provided  by  The  Conservation  Agen- 
cy through  a grant  from  the  Falconwood  Foundation, 
and  by  the  U.S.  Geological  Survey  Texas  Cooperative 
Fish  and  Wildlife  Research  Unit.  This  manuscript  ben- 
efited from  the  reviews  and  constructive  comments  of 
G.  Perry,  R.  L.  Norton,  and  three  anonymous  reviewers. 


224 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


LITERATURE  CITED 

Arendt,  W.  J.  1992.  Status  of  North  American  migrant 
landbirds  in  the  Caribbean  region:  a summary. 
Pages  143-171  in  Ecology  and  conservation  of 
Neotropical  migrant  landbirds  (J.  M.  Hagan,  III, 
and  D.  W.  Johnson,  Eds.).  Smithsonian  Institution 
Press,  Washington,  D.C. 

Askins,  R.  A.  and  D.  N.  Ewert.  1991.  Impact  of  Hur- 
ricane Hugo  on  bird  populations  on  St.  John,  U.S. 
Virgin  Islands.  Biotropica  23:481-487. 

Burnham,  K.  R,  D.  R.  Anderson,  G.  C.  White,  C. 
Brownie,  and  K.  H.  Pollock.  1987.  Design  and 
analysis  methods  for  fish  survival  experiments 
based  on  release-recapture.  American  Fisheries 
Society  Monograph,  no.  5. 

Chipley,  R.  M.  1991.  Notes  on  the  biology  of  the  Bri- 
dled Quail-Dove  ( Geotrygon  mystacea).  Caribbe- 
an Journal  of  Science  27:180-184. 

de  Souza  Lopes,  O.,  L.  de  Abreu  Sacchetta,  and  E. 
Dente.  1980.  Longevity  of  wild  birds  obtained 
during  a banding  program  in  Sao  Paulo,  Brasil. 
Journal  of  Field  Ornithology  51:144-148. 

Faaborg,  J.  and  J.  W.  Terborgh.  1980.  Patterns  of 
migration  in  the  West  Indies.  Pages  157-163  in 
Neotropics:  ecology,  behavior,  distribution,  and 
conservation  (A.  Keast  and  E.  S.  Morton,  Eds.). 
Smithsonian  Institution  Press,  Washington,  D.C. 

Faaborg,  J.  and  J.  E.  Winters.  1979.  Winter  resident 
returns  and  longevity  and  weights  of  Puerto  Rican 
birds.  Bird-Banding  50:216-223. 

Johnston,  J.  P,  W.  J.  Peach,  R.  D.  Gregory,  and  S. 
A.  White.  1997.  Survival  rates  of  tropical  and 
temperate  passerines:  a Trinidadian  perspective. 
American  Naturalist  150:771-789. 

Kennard,  J.  H.  1975.  Longevity  records  of  North 
American  birds.  Bird-Banding  46:55-73. 

Klimkiewicz,  M.  K.,  R.  B.  Clapp,  and  A.  G.  Futcher. 
1983.  Longevity  records  of  North  American  birds: 
Remizidae  through  Parulinae.  Journal  of  Field  Or- 
nithology 54:287-294. 

Krementz,  D.  G..  J.  R.  Sauer,  and  J.  D.  Nichols. 
1989.  Model-based  estimates  of  annual  survival 
are  preferable  to  observed  maximum  lifespan  sta- 
tistics for  use  in  comparative  life-history  studies. 
Oikos  56:203-208. 

LaBastille,  A.  and  M.  Richmond.  1973.  Birds  and 
mammals  of  Anegada  Island.  Caribbean  Journal 
of  Science  13:91-109. 

Lazell,  J.  1996.  Guana  Island:  a natural  history  guide. 
Conservation  Agency  Occasional  Paper,  no.  1 . 
Jamestown.  Rhode  Island. 


Lazell,  J.  2005.  Island:  fact  and  theory  in  nature.  Uni- 
versity of  California  Press,  Berkeley. 

Loftin,  H.  1975.  Recaptures  and  recoveries  of  banded 
native  Panamanian  birds.  Bird-Banding  46:19-27. 

Mayer,  G.  C.  and  R.  M.  Chipley.  1992.  Turnover  in 
the  avifauna  of  Guana  Island,  British  Virgin  Is- 
lands. Journal  of  Animal  Ecology  61:561-566. 

McNair,  D.  B.,  F.  Sibley,  E.  B.  Massiah,  and  M.  D. 
Frost.  2002.  Ground-based  Nearctic-Neotropic 
landbird  migration  during  autumn  in  the  eastern 
Caribbean.  Pages  86-103  in  Studies  in  Trinidad 
and  Tobago  ornithology  honouring  Richard 
ffrench  (F.  E.  Hays  and  S.  A.  Temple,  Eds.).  Oc- 
casional Paper,  no.  1 1 . University  of  West  Indies, 
St.  Augustine,  Trinidad  and  Tobago. 

Mirecki,  D.  N.,  J.  M.  Hutton,  C.  M.  Pannell,  T.  J. 
Stowe,  and  R.  W.  Unite.  1977.  Report  of  the 
Cambridge  ornithological  expedition  to  the  British 
Virgin  Islands  1976.  Churchill  College,  Cam- 
bridge, United  Kingdom. 

Newton,  I.  1998.  Population  limitation  in  birds.  Aca- 
demic Press,  San  Diego,  California. 

Norton,  R.  L.  1990.  West  Indies  region.  American 
Birds  44:168-169. 

Norton,  R.  L.  1996.  West  Indies  region.  Audubon 
Field  Notes  50:120-122. 

Norton,  R.  L.,  R.  M.  Chipley,  and  J.  D.  Lazell,  Jr. 
1989.  A contribution  to  the  ornithology  of  the 
British  Virgin  Islands.  Caribbean  Journal  of  Sci- 
ence 25:115-118. 

Norton,  R.  L.,  A.  White,  and  A.  Dobson.  2003.  West 
Indies  and  Bermuda.  North  American  Birds  57: 
417-419. 

Pyle,  P 1997.  Identification  guide  to  North  American 
birds,  part  1.  Slate  Creek  Press,  Bolinas,  Califor- 
nia. 

Raffaele,  H.  A.  1989.  Guide  to  the  birds  of  Puerto 
Rico  and  the  Virgin  Islands.  Princeton  University 
Press,  Princeton,  New  Jersey. 

Raffaele,  H.,  J.  Wiley,  O.  Garrido,  A.  Keith,  and  J. 
Raffaele.  2003.  Birds  of  the  West  Indies.  Prince- 
ton University  Press,  Princeton,  New  Jersey. 

Snow,  D.  W.  and  A.  Lill.  1974.  Longevity  records  for 
some  Neotropical  land  birds.  Condor  76:262-267. 

Wiley,  J.  W.  2000.  A bibliography  of  ornithology  in 
the  West  Indies.  Proceedings  of  the  Western  Foun- 
dation of  Vertebrate  Zoology,  no.  7.  Camarillo, 
California. 

Woodworth,  B.  L.,  J.  Faaborg,  and  W.  J.  Arendt. 
1999.  Survival  and  longevity  of  the  Puerto  Rican 
Vireo.  Wilson  Bulletin  1 1 1 :376-380. 


The  Wilson  Journal  of  Ornithology  1 1 8(2):225-236,  2006 


REPRODUCTIVE  BEHAVIOR  OF  THE  YELLOW-CROWNED 
PARROT  C AMAZONA  OCHROCEPHALA)  IN  WESTERN  PANAMA 

ANGELICA  M.  RODRIGUEZ  CASTILLO1 3 AND  JESSICA  R.  EBERHARD245 


ABSTRACT. — We  studied  the  breeding  biology  of  the  Panamanian  subspecies  of  the  Yellow-crowned  Parrot, 
Amazona  ochrocephala  panamensis,  during  1997—1999  in  the  province  of  Chiriqm,  Panama,  to  provide  basic 
information  regarding  the  breeding  behavior  and  reproductive  success  of  these  parrots  in  their  natural  habitat. 
We  recorded  parrot  behaviors  throughout  the  reproductive  period,  monitored  nest  success,  and  characterized 
occupied  and  non-occupied  tree  cavities.  All  breeding  attempts  involved  a male-female  pair.  Clutch  size  ranged 
from  2 to  4 eggs,  which  were  incubated  only  by  the  female,  beginning  when  the  first  egg  was  laid.  Incubation 
averaged  25  days  and  the  eggs  hatched  asynchronously.  During  the  incubation  period,  females  remained  inside 
the  nest  for  long  periods  of  time,  though  they  often  departed  from  the  nest  area  during  early  mornings  and  late 
afternoons,  presumably  to  forage;  during  this  period,  males  were  not  observed  entering  the  nest,  though  they 
often  remained  nearby.  During  the  nestling  period,  males  contributed  significantly  to  feeding  the  offspring.  Pairs 
nested  in  trees  that  were  in  good  or  fair  condition,  and  did  not  favor  cavities  in  any  one  tree  species.  As  found 
in  many  other  field  studies  of  parrots,  breeding  success  was  low.  Only  10%  (1997-1998)  and  14%  (1998-1999) 
of  the  nests  survived  poaching  and  natural  predation.  Because  nest  poaching  was  the  primary  cause  of  breeding 
failure  and  poses  a serious  threat  to  population  viability,  we  also  present  data  on  poaching  techniques  and  the 
local  trade  of  nestling  parrots.  Overall,  the  pool  of  breeding  adults  is  likely  made  up  of  aging  individuals  that 
are  not  being  replaced,  setting  the  stage  for  a rapid  population  decline.  Received  13  January  2005,  accepted  23 
November  2005. 


The  genus  Amazona  consists  of  3 1 species 
distributed  throughout  the  Neotropics  (Juniper 
and  Parr  1998);  however,  the  breeding  biology 
of  only  a few  species  has  been  studied  (see 
below).  Nest  poaching  and  the  capture  of 
adult  birds  for  the  pet  trade,  together  with  hab- 
itat loss  due  to  deforestation,  have  contributed 
to  the  precipitous  decline  of  Amazona  popu- 
lations in  Central  America,  South  America, 
and  the  Caribbean  region  (Forshaw  1989,  Ju- 
niper and  Parr  1998,  Wright  et  al.  2001).  Like 
many  of  the  eight  other  subspecies  that  form 
the  Yellow-crowned  Parrot  complex  (Juniper 
and  Parr  1998,  Eberhard  and  Bermingham 
2004),  Amazona  ochrocephala  panamensis 
has  not  escaped  these  pressures  (Asociacion 
Nacional  para  la  Conservacion  de  la  Natural- 
eza  1995,  Autoridad  Nacional  del  Ambiente 
1995a).  In  Panama,  the  population  of  this  sub- 
species has  declined  considerably  due  to  nest 


1 Escuela  de  Biologia,  Univ.  Autonoma  de  Chiriqm, 
David,  Panama. 

2 Smithsonian  Tropical  Research  Inst.,  Apdo.  2072, 
Balboa,  Panama. 

3 Current  address:  Estafeta  Universitaria,  UNACHI, 
David,  Panama. 

4 Current  address:  Biological  Sciences  Dept,  and 
Museum  of  Natural  Science,  202  Life  Sciences,  Lou- 
isiana State  Univ.,  Baton  Rouge,  LA  70803,  USA. 

5 Corresponding  author;  e-mail:  eberhard@lsu.edu 


poaching  (Ridgely  1981)  and  the  loss  of  nest- 
ing habitat  to  agricultural  and  cattle-grazing 
activities  (Autoridad  Nacional  del  Ambiente 
1995a,  1995b). 

The  breeding  biology  of  a few  Amazona 
species  has  been  studied  in  the  wild;  many  of 
these  studies  occurred  on  Caribbean  islands 
(Snyder  et  al.  1987,  Gnam  1991,  Rojas-Suarez 
1994,  Wilson  et  al.  1995)  while  others  provide 
information  on  mainland  species  (Enkerlin- 
Hoeflich  1995,  Enkerlin-Hoeflich  and  Hogan 
1997,  Renton  and  Salinas-Melgoza  1999,  and 
Seixas  and  Mourao  2002).  Additional  data  on 
breeding  behavior  come  from  studies  of  cap- 
tive A.  albifrons  (Skeate  1984)  and  A.  viridi- 
genalis  (Wozniak  and  Lanterman  1984).  Over- 
all, the  studies  have  revealed  that  females  typ- 
ically spend  long  periods  inside  the  nest  dur- 
ing the  incubation  and  early  nestling  periods, 
and  depend,  at  least  to  some  degree,  on  being 
fed  by  their  mates.  Four  Amazona  species  in 
Mexico  apparently  select  nest  sites  based  on 
tree  species,  size,  cavity  height,  and  entrance 
size  (Enkerlin-Hoeflich  1995,  Renton  and  Sa- 
linas-Melgoza 1999). 

Wright  et  al.  (2001)  summarized  data  from 
many  field  studies  and  showed  that  nest 
poaching  is  a principal  cause  of  reproductive 
failure  in  Neotropical  parrots,  with  poaching 
rates  being  higher  at  mainland  sites  than  on 


225 


226 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118 , No.  2,  June  2006 


islands,  and  lower  in  protected  areas  (e.g.,  na- 
ture reserves).  While  the  impact  of  nest 
poaching  on  parrot  reproductive  success  is 
clear,  there  are  few  studies  that  provide  infor- 
mation on  specific  techniques  used  by  poach- 
ers. 

To  date,  there  have  been  no  published  stud- 
ies of  the  reproductive  behavior  of  A.  ochro- 
cephala  in  Panama  or  other  parts  of  its  range. 
Here,  we  report  our  observations  of  the  spe- 
cies’ breeding  behavior,  describe  the  charac- 
teristics of  nest  sites  and  nest  trees,  and  quan- 
tify reproductive  success  during  two  breeding 
seasons.  We  also  present  data  regarding  the 
poaching  techniques  used  in  the  study  area. 

METHODS 

Study  area. — Fieldwork  was  conducted  dur- 
ing the  dry  season  (December-April)  of 
1997-1998  and  1998-1999  in  the  lowlands  of 
Corregimiento  de  San  Juan  (San  Lorenzo  dis- 
trict) of  the  province  of  Chiriquf  in  western 
Panama.  The  natural  vegetation  in  the  area  is 
tropical  dry  forest  (following  Holdridge’s 
[1967]  life  zone  classification)  and  mangrove, 
but  in  many  places  it  has  been  cleared  for  ag- 
riculture and  cattle  grazing.  Annual  rainfall  is 
~ 1,000  mm;  mean  annual  temperature  is 
—30°  C,  with  mean  temperatures  of  35°  and 
28°  C during  the  dry  and  rainy  seasons,  re- 
spectively (Instituto  de  Recursos  Hidraulicos 
y Electrificacion  1998,  1999).  The  study  area 
was  located  at  ~8°  17'  15"  N,  82°  3'  10"  W 
and  encompassed  an  area  of  ~8,800  ha; 
~3,875  ha  had  been  partially  cleared  for  ag- 
riculture and  cattle  grazing  (on  haciendas  Mir- 
aflores,  El  Tekal,  and  Los  Asentamientos  de 
San  Juan),  and  the  remaining  4,925  ha  were 
mangrove.  The  partially  cleared  areas  still 
contained  remnant  patches  of  tropical  dry  for- 
est dominated  by  Gliricidia  sepium  and  Ery- 
thrina  fusca  trees,  the  lower-statured  Curatel- 
la  americana , and  palms  belonging  to  the  gen- 
era Roystonea  and  Acrocomia  (Acosta  1996). 

Characterization  of  nest  sites. — During  the 
first  breeding  season  (1997-1998),  we  only 
studied  nests  found  in  the  mangrove  habitat; 
in  the  following  season  (1998-1999),  we  ex- 
tended our  nest  monitoring  to  include  those 
found  in  the  partially  cleared  dry  forest  hab- 
itat. We  found  21  active  nests  during  the  1st 
year  and  42  during  the  2nd  year.  Of  the  nests 
found  in  the  2nd  year,  14  had  been  used  by 


parrots  during  the  previous  breeding  season; 
therefore,  to  avoid  pseudoreplication,  our  data 
on  cavity  and  nest-tree  characteristics  repre- 
sent 49  (and  not  63)  active  nests.  In  the  sec- 
ond breeding  season,  20  of  the  nests  were 
found  in  mangrove  habitat,  and  the  remaining 
22  in  the  partially  cleared  dry  forest. 

To  find  nest  cavities,  we  searched  for  trees 
with  cavities,  observed  parrots  flying  and  vo- 
calizing in  the  area,  and  interviewed  local  res- 
idents and  field  laborers  for  information  about 
nesting  parrots.  Nests  were  considered  active 
if  they  contained  A.  o.  panamensis  eggs  or 
nestlings. 

To  determine  the  availability  of  cavities,  we 
searched  for  additional  tree  cavities  near  nest 
trees.  By  searching  the  area  surrounding  an 
occupied  nest  tree,  we  attempted  to  control  for 
larger-scale  habitat  variation  (e.g.,  vegetation 
density,  canopy  height,  distance  to  feeding  ar- 
eas) that  might  have  influenced  cavity  choice. 
All  trees  within  100  m of  each  nest  tree  were 
examined  for  the  presence  of  large  cavities 
(i.e.,  cavities  similar  in  size  to  those  occupied 
by  parrots).  For  a given  nest  tree,  two  of  the 
surrounding  trees  found  to  contain  cavities 
were  selected  at  random  for  inclusion  in  the 
sample  of  unoccupied  cavities.  If  a selected 
tree  contained  more  than  one  cavity,  we  se- 
lected one  of  them  at  random  to  provide  data 
on  cavity  location  and  orientation.  In  the  par- 
tially cleared  dry  forest  habitat,  we  extended 
two  of  these  searches  beyond  100  m (108  and 
1 16  m)  in  order  to  find  trees  with  large  cavi- 
ties. Determining  that  a cavity  was  similar  in 
size  to  occupied  cavities  was  admittedly  sub- 
jective; therefore,  we  do  not  present  any  anal- 
yses comparing  the  dimensions  of  occupied 
versus  unoccupied  cavities. 

We  used  leaf,  flower,  and/or  fruit  samples 
to  identify  the  genus  and  species  (where  pos- 
sible) of  trees  containing  cavities.  For  each 
cavity  we  measured  horizontal  and  vertical 
width  of  the  cavity  opening,  inside  vertical 
depth  and  cavity  diameter  (measured  at  the 
cavity  floor),  and  distance  from  the  ground  to 
the  lower  edge  of  the  cavity  opening  (see 
Saunders  et  al.  1982).  Measurements  were 
made  using  a 30-m  tape  to  a precision  of  0.5 
cm,  and  were  used  to  calculate  the  areas  of 
the  cavity  entrance  and  cavity  floor.  For  each 
cavity,  we  noted  its  location  relative  to  the 
tree’s  structure — branch  (cavity  completely 


Rodriguez  and  Eberhard  • REPRODUCTIVE  BEHAVIOR  OF  AMAZONA  OCHROCEPHALA  227 


contained  within  a branch),  trunk  (cavity  com- 
pletely contained  within  the  main  trunk),  and 
branch/trunk  (cavity  at  the  intersection  of  a 
branch  and  the  trunk).  We  determined  the  ori- 
entation of  the  cavity  opening  using  a com- 
pass, and  measured  each  tree’s  height  using  a 
clinometer.  We  classified  the  physical  condi- 
tion of  each  tree — good,  fair,  poor,  or  dead — 
using  the  scheme  outlined  by  Sauad  et  al. 
(1991;  see  also  Saunders  et  al.  1982). 

Behavioral  observations. — We  monitored 
63  nests  during  the  two  breeding  seasons:  21 
during  1997-1998  and  42  during  1998-1999. 
Of  the  63  nests,  5 were  selected  each  year  for 
detailed  behavioral  observations  of  parrots 
(hereafter  referred  to  as  focal  nests).  In  the 
first  field  season,  focal  nests  were  chosen  at 
random;  during  the  second  field  season,  nests 
were  selected  on  the  basis  of  their  accessibil- 
ity. 

We  made  preliminary  observations  early  in 
the  breeding  season  (prior  to  egg-laying)  at 
each  of  the  focal  nests.  An  observation  period 
lasted  13  hr  (06:00  to  19:00  UTC-5).  Each 
year,  we  watched  three  of  the  five  focal  nests 
for  two  preliminary  observation  periods,  and 
the  other  two  were  watched  for  a single  ob- 
servation period.  In  most  cases  (9  of  16  ob- 
servation periods),  we  conducted  preliminary 
observations  prior  to  capture  of  the  focal  in- 
dividuals. 

To  identify  the  sex  of  focal  individuals,  we 
used  nets  (set  up  at  dawn)  to  capture  one  or 
both  members  of  each  focal  pair  early  in  the 
field  season  (prior  to  the  onset  of  breeding). 
We  used  nylon  (4.5  X 15  m)  and  cotton  (6  X 
8 m)  fishing  nets  (mist  nets  were  not  avail- 
able) and  suspended  them  using  ropes  and/or 
poles  over  the  nest  opening  or  across  a flyway 
used  by  the  birds.  In  both  years,  the  sex  of 
each  captured  individual  was  identified  in  the 
field  by  a veterinarian  (R.  De  Obaldia)  using 
a laparoscope.  We  then  marked  the  female  on 
the  upper  chest  with  Rhodamine  B,  so  that  she 
could  be  distinguished  from  the  male  in  sub- 
sequent observations.  Because  the  Rhodamine 
B marks  faded  after  several  weeks,  the  birds 
were  subsequently  marked  passively  by  ap- 
plying dye  to  the  nest  opening  (see  Eberhard 
1998).  This  passive  marking  was  done  before 
the  prior  markings  had  faded  completely,  so 
that  the  identity  of  the  newly  marked  birds 
was  known.  With  this  technique,  the  birds  in- 


variably marked  themselves  on  different  parts 
of  the  body  with  unique  patterns,  so  the  male 
and  female  could  be  distinguished  from  one 
another. 

For  the  remainder  of  each  focal  pair’s 
breeding  attempt,  we  made  behavioral  obser- 
vations at  ~3-day  intervals.  We  observed  dur- 
ing 3-hr  periods  when  the  parrots  were  most 
active  (either  06:30-09:30  or  15:45-18:45), 
following  the  methodology  used  in  other  par- 
rot studies  (e.g.,  Eberhard  1998,  Renton  and 
Salinas-Melgoza  1999).  The  results  reported 
here  are  based  on  859  hr  of  nest  observation 
(208  hr  were  preliminary  observations).  We 
observed  nests  with  the  aid  of  binoculars  from 
a distance  of  —15  m (the  parrots  quickly  ha- 
bituated to  the  observer’s  presence).  During 
each  observation  period,  we  noted  the  follow- 
ing: time  spent  by  the  adults  inside  the  nest; 
time  spent  in  the  nest  area  (defined  as  being 
in  visual  range  of  the  observer,  which  was  ap- 
proximately 50-75  m in  the  mangrove  habitat 
and  approximately  100  m in  the  partially 
cleared  dry  forest  areas);  number  of  other  par- 
rots traveling  with  the  focal  individual  when 
approaching  or  departing;  and  presence  of 
other  humans  in  the  nesting  area.  Other  gen- 
eral observations  (allofeeding,  allogrooming, 
vocal  and  plumage  displays,  appearances  of 
nestlings  at  the  cavity  opening,  age  at  which 
young  left  the  nest)  were  noted  ad  lib.  When 
adults  made  short  visits  to  the  nest,  presum- 
ably to  feed  young,  we  recorded  total  time  in 
the  nest  cavity.  Focal  nest  observations  were 
made  until  6 days  after  the  last  chick  fledged, 
or  6 days  after  a nest  was  depredated  or 
poached. 

Three  of  the  focal  nests  observed  during  the 
first  breeding  season  (1997-1998)  were  in 
cavities  that  were  re-occupied  in  the  following 
breeding  season,  and  were  considered  focal 
nests  during  the  2nd  year  of  the  study.  Be- 
cause it  is  possible  that  pairs  used  the  same 
cavity  in  consecutive  years,  our  data  might  in- 
clude some  year-to-year  pseudoreplication  in 
the  focal-nest  behavioral  observations.  The 
adults  were  not  permanently  marked,  so  it  was 
impossible  to  determine  whether  this  oc- 
curred. 

For  the  analysis  of  behavioral  data,  we  di- 
vided the  breeding  season  into  four  stages: 
pre-laying,  laying,  incubation,  and  nestling 
periods.  The  laying  period  began  with  the  lay- 


228 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


ing  of  the  first  egg  and  extended  until  the  last 
egg  was  laid;  the  incubation  period  began  with 
the  laying  of  the  last  egg  and  extended  until 
the  last  egg  hatched  (in  fact,  incubation  began 
when  the  first  egg  was  laid,  but  for  our  data 
presentation  and  analyses,  we  defined  the  in- 
cubation period  as  described  here  to  avoid 
overlap  of  data  from  the  laying  and  incubation 
periods);  the  nestling  period  began  with  the 
hatching  of  the  last  egg  and  extended  until  the 
last  nestling  had  fledged,  or  the  nest  was 
poached  or  depredated. 

Nest  checks. — During  the  laying  and  incu- 
bation periods,  each  focal  and  nonfocal  nest 
was  checked  daily  and  its  contents  inspected; 
during  the  nestling  period,  we  reduced  the  fre- 
quency of  checks  to  once  per  week.  On  days 
when  a focal  nest  was  the  object  of  behavioral 
observations,  the  nest  was  checked  at  the  con- 
clusion of  the  observation  period,  or  at  least 
2 hr  before  the  start  of  an  observation  period. 
This  was  done  to  minimize  disruption  of  the 
adults’  behavior.  At  each  nest  check,  we  noted 
the  presence  of  any  new  eggs  (eggs  were 
numbered  with  a pencil),  used  calipers  to 
measure  the  dimensions  (length  and  width)  of 
new  eggs,  and  noted  laying  and  hatching 
dates.  During  the  nestling  period,  we  noted 
morphological  characteristics  of  the  hatch- 
lings and  the  emergence  and  locations  of  new 
feathers,  and  recorded  fledging  dates.  We  also 
noted  evidence  of  cavity  enlargement  by  the 
parrots  and  presence  of  a nest  lining.  Al- 
though the  frequency  of  nest  checks  was  re- 
duced during  the  nestling  period,  we  visited 
nest  trees  2 to  3 times  per  day  in  order  to 
maintain  a presence  that,  we  hoped,  would  re- 
duce the  likelihood  that  our  study  nests  would 
be  poached. 

Poaching  interviews. — We  obtained  infor- 
mation on  the  techniques  used  by  parrot 
poachers  in  the  San  Juan  area  through  anon- 
ymous interviews  of  individuals  actively  en- 
gaged in  the  capture  and  sale  of  A.  o.  pana- 
mensis.  Poachers  were  contacted  with  the  help 
of  an  area  resident  who  is  familiar  with  the 
parrot  trade  around  San  Juan.  A consistent  set 
of  questions  or  talking  points  was  included  in 
each  interview,  but  the  respondents  were  en- 
couraged to  offer  any  information  that  they 
might  have  regarding  the  parrots.  The  inter- 
view questions  focused  on  the  poaching  of  A. 
o.  panamensis ; however,  additional  informa- 


tion on  other  species  was  noted  whenever 
mentioned  by  the  respondents.  All  interviews 
were  conducted  by  AMRC. 

Statistical  analyses. — Descriptive  statistics 
(mean  ± SD,  range,  percentage)  are  presented 
for  nest  site  and  behavioral  data.  Data  from 
the  2 years  are  presented  separately  in  tables, 
since  the  2nd  year  included  data  from  nests  in 
both  partially  cleared  dry  forest  and  mangrove 
habitats;  however,  the  descriptive  statistics 
presented  in  text  summarize  both  years’  data. 
We  used  the  Lilliefors  (Kolmogorov-Smirnov) 
test  to  check  for  normality  prior  to  performing 
parametric  tests.  We  performed  chi-square 
tests  of  independence  to  test  the  hypothesis 
that  parrots  prefer  cavities  in  certain  tree  spe- 
cies. For  each  habitat,  we  compared  the  num- 
ber of  nests  (occupied  cavities)  in  different 
tree  species  with  the  number  of  unoccupied 
cavities  in  those  species.  Chi-square  tests  of 
independence  were  also  used  to  determine 
whether  parrots  showed  a preference  for  trees 
in  relatively  good  condition.  We  used  circular 
statistics  (Batschelet  1981)  to  analyze  the  ori- 
entation of  nest-cavity  openings,  and  per- 
formed Rayleigh  tests  to  determine  whether 
the  orientations  of  occupied  and  unoccupied 
cavities  were  random.  These  tests  were  per- 
formed using  R (R  Development  Core  Team 
2005);  cavity  openings  facing  upward  were 
excluded  from  the  orientation  analyses.  We 
performed  a discriminant  function  analysis  to 
determine  whether  there  were  significant  dif- 
ferences between  the  dimensions  of  trees  and 
cavities  containing  successful  nests  versus  the 
dimensions  of  those  with  nests  that  were 
poached  or  depredated.  Discriminant  function 
analysis  determines  which  variables  (in  our 
case,  nest  dimensions)  discriminate  between 
two  or  more  groups  (successful  versus  unsuc- 
cessful nests),  and  identifies  those  variables 
that  contribute  most  to  the  differences  be- 
tween groups  (Huberty  1994,  Silva  and  Stam 
1995).  We  employed  a forward  stepwise  pro- 
cedure to  select  among  nine  nest  dimensions 
(see  Table  1),  with  entry  and  removal  P-v al- 
ues  of  0.05.  We  used  linear  regression  to  as- 
sess the  degree  to  which  time  spent  by  the 
females  in  the  nest  changed  through  the  nest- 
ling period.  For  analyses  of  data  that  were  not 
normally  distributed,  we  used  Mann- Whitney 
U- tests  and  Wilcoxon  tests.  Statistical  analy- 
ses (with  the  exception  of  circular  statistics) 


Rodriguez  and  Eberhard  • REPRODUCTIVE  BEHAVIOR  OF  AMAZONA  OCHROCEPHALA  229 


TABLE  1.  Dimensions  of  occupied  cavities  ( n — 
of  San  Juan,  Chiriqui,  western  Panama,  1997-1999. 

49)  of  Amazona  ochrocephala  panamensis  in  the  lowlands 

Measurement 

Mean  ± SD 

Range 

Vertical  depth  (cm) 

99.2  ±71.2 

34.8-445.0 

Internal  width  (cm) 

26.8  ± 4.3 

18.1-34.0 

Internal  length  (cm) 

26.8  ± 4.5 

16.5-36.0 

Area  of  cavity  floor  (cm2) 

575.9  ± 175.1 

257.3-907.9 

Area  of  cavity  entrance  (cm2) 

229.7  ± 63.0 

149.8-380.1 

Horizontal  diameter  of  cavity  entrance  (cm) 

15.6  ± 2.7 

10.9-19.8 

Vertical  diameter  of  cavity  entrance  (cm) 

17.2  ± 2.8 

12.0-22.5 

Height  of  cavity  entrance  (m) 

12.4  ± 2.7 

9.2-16.5 

Height  of  nest  tree  (m) 

19.2  ± 3.1 

10.7-26.1 

were  performed  using  Statistica  6.0  (StatSoft, 
Inc.  1998).  For  all  tests,  statistical  significance 
was  set  at  a = 0.05  and  means  are  presented 
± SD. 

RESULTS 

Characterization  of  nest  sites. — In  our 
study  area,  A.  o.  panamensis  used  a diversity 
of  tree  species  for  nesting.  In  mangrove  hab- 
itat, active  nest  cavities  were  found  in  five  tree 
species:  Rhizophora  mangle,  R.  brevistyla, 
Avicennia  bicolor , Pelliciera  rhizophorae, 
Mora  oleifera.  In  partially  cleared  dry  forest 
habitat,  parrots  were  found  nesting  in  two  spe- 
cies of  palms,  Roystonea  regia.  Cocos  nuci- 
fera,  and  in  Ficus  insipida  trees.  The  most  fre- 
quently used  tree  species  were  R.  regia  (18  of 
49  nests)  and  R.  mangle  (13  of  49  nests).  The 
tree  species  used  least  frequently  were  A.  bi- 
color and  F.  insipida,  each  of  which  was  used 
only  once.  Overall,  parrots  showed  no  pref- 
erence for  nesting  cavities  in  any  one  tree  spe- 
cies in  either  mangrove  or  dry  forest  habitat 
(mangrove:  x2  = 0.813,  df  = 4,  P = 0.94,  n 
= 27  nests;  dry  forest:  x2  = 0.039,  df  = 2,  P 
= 0.98,  n = 22  nests).  Rather,  the  use  of  tree 
species  for  nesting  was  proportional  to  cavity 
availability  in  those  species.  We  found  no  ev- 
idence in  either  habitat  type  that  any  one  tree 
species  is  more  likely  to  develop  cavities  than 
the  others  (mangrove:  x2  = 0.257,  df  = 4,  P 
= 0.99,  n = 41  cavities;  dry  forest:  x2  = 
0.666,  df  = 2 ,P  = 0.72,  n = 25  cavities). 

Characteristics  of  occupied  and  unoccupied 
cavities. — Breeding  pairs  preferred  cavities 
that  were  relatively  high  above  the  ground  and 
with  dimensions  similar  to  those  reported  for 
other  Amazona  species  (see  Table  1).  The  ori- 
entation of  occupied  cavity  openings  was  non- 


random (Rayleigh  test:  r = 0.4408,  P < 0.001, 
n = 39),  with  a bias  toward  the  northeast 
quadrant  (25  of  39  occupied  nests  had  orien- 
tations between  250°  and  360°).  In  contrast, 
the  orientations  of  unoccupied  cavities  were 
randomly  distributed  (Rayleigh  test;  r — 
0.1495,  P = 0.26,  n = 61). 

In  both  habitat  types,  we  found  that  A.  o. 
panamensis  preferred  trees  with  single  cavi- 
ties (x2  = 41.49,  df  = 2,  P < 0.001),  possibly 
because  trees  with  more  than  one  cavity  were 
in  poorer  condition  than  those  with  single  cav- 
ities. Indeed,  parrots  preferred  trees  in  rela- 
tively good  condition.  Forty-two  of  49  (86%) 
occupied  trees  were  in  good  or  fair  condition, 
while  56  of  98  (57%)  unoccupied  trees  were 
in  poor  condition  or  they  were  dead  (x2  — 
24.5,  df  = 1,  P < 0.001).  Comparing  the  lo- 
cation of  cavities  (branch,  branch/trunk,  or 
trunk)  in  occupied  versus  unoccupied  trees  in- 
dicated that  the  parrots  had  no  preference  for 
any  particular  cavity  location  (x2  — 0.807,  df 
= 2 , P = 0.67). 

Pre-laying  period. — We  observed  pairs  of 
A.  o.  panamensis  prospecting  for  nest  sites 
early  in  each  field  season  (13-30  December 
1997,  21  December  1998-5  January  1999). 
Both  members  of  the  breeding  pair  participat- 
ed in  nest  prospecting.  On  four  occasions,  we 
observed  one  of  the  two  birds  apparently  take 
the  lead  in  cautiously  approaching  and  inves- 
tigating the  cavity  while  its  partner  remained 
perched  in  a nearby  tree.  Once  a nest  tree  was 
selected,  but  before  egg-laying  began,  the  fe- 
male (sex  was  known  for  focal  pairs  once  they 
had  been  captured  and  marked)  spent  long  pe- 
riods of  time  within  the  nest  cavity,  while  the 
male  remained  perched  at  the  entrance  or 
nearby.  On  two  occasions  in  the  mangrove 


230 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


habitat,  a focal  female  was  seen  taking  a twig 
into  her  nest  cavity.  In  a third  instance,  an 
individual  (sex  unknown)  took  a leafy  twig 
into  its  nest  cavity  in  a M.  oleifera  tree.  Inside 
6 of  the  49  monitored,  occupied  cavities,  we 
found  wood  chips  and  leaves — materials  that 
were  a result  of  the  parrots’  chewing  activities 
and/or  brought  in  from  outside  the  cavity. 

Throughout  the  breeding  season,  pairs  often 
perched  together,  grooming  each  other’s  neck, 
head,  and  wings.  We  observed  no  copulations 
or  copulation  attempts  during  our  study.  Prior 
to  the  onset  of  egg-laying,  the  male  occasion- 
ally entered  the  nest  cavity  with  the  female 
and  remained  inside  for  several  minutes 
(mean  time  inside  = 3.40  ± 0.44  min,  n = 
17).  During  these  visits,  it  is  likely  that  he  was 
feeding  the  female,  but  it  is  also  possible  that 
copulations  occurred.  As  the  egg-laying  peri- 
od approached,  the  female  increased  the 
amount  of  time  that  she  spent  inside  the  nest 
cavity,  emerging  for  a few  minutes  at  intervals 
of  1. 5-2.5  hr  to  stretch  her  wings  and  legs 
before  returning  to  the  cavity.  During  this  pe- 
riod, we  observed  eight  instances  in  which  the 
male  presented  his  mate  with  flowers  of  Ery- 
thrina  fusca  or  Gliricidia  sepium,  which  the 
female  subsequently  consumed. 

Egg-laying  and  incubation  periods. — Egg- 
laying  in  the  monitored  nests  (focal  and  non- 
focal)  occurred  from  15  December  to  3 Jan- 
uary (1997-1998)  and  24  December  to  13 
January  (1998-1999).  Clutch  size  averaged 
3.08  ± 0.77  eggs  over  both  years  of  the  study 
(Table  2),  with  no  significant  difference  be- 
tween years  (Mann-Whitney  (7-test:  Z = 
-0.584,  P = 0.56,  n = 63  clutches).  The 
mean  laying  interval  was  2.16  ± 0.92  days 
(Table  2).  Incubation  began  when  the  first  egg 
was  laid  and  was  conducted  exclusively  by  the 
female.  The  incubation  period  lasted  25.14  ± 
1.77  days. 

During  egg-laying  and  incubation,  females 
spent  most  of  their  time  inside  the  nest  or 
perched  nearby,  and  males  were  never  seen 
entering  the  nest  cavity,  although  they  often 
remained  perched  nearby  (Fig.  1).  The  amount 
of  time  the  male  spent  with  the  female  during 
the  egg-laying  period  (the  female’s  fertile  pe- 
riod) versus  during  the  incubation  period  did 
not  differ  (Wilcoxon  test:  Z = 1.48,  P = 0.14, 
n = 10  breeding  attempts).  During  incubation, 
the  female  occasionally  emerged  from  the  nest 


5b  On 

5 On 

a ON 


a Q- 

“I  < 
lot 
§■3 


S <u 
K a 
o <u 

y a 

^ s 


cm  m 
o U 
m m 
I I 
o 

r-’  <n 

CM  cn 


+ 1 +1 


n cm 
o oo 


t"-  VO 
CM  CO 


Tt  00 

r-  r- 
o o 
+1  +1 
m in 


jz  *-> 


T3 

C/5 

OJO  ^ 
OX) 

PQ 


Ov  VO 

in  cm  ^ oo 

I I I I I I I 
CM  CM  — O O 

CM  t-~ 


r--  o vo  vo  o-  ov  oo 
r-;  in  — ov  — oc 
d d ^ -h  o — < ro 

+1  +1  +1  +1  +1  +1  +1 

cm  Nt  - n in  oo  on 

— cn  <n  — - ov  q <n 

cn  cm  in  cm  cm  d oo 

cm 


00  — 

(\|  in  t h 
I I I I I I I 
CM  — CM  — o O ov 
cm  in 


r-  sc  — -t 
h-  O ON  CM  h 

o o 

+i  +i  +i  +i  +i 

O cn  ov  — vo 

o o o cm  r- 


+1  +1 
oo  o 

CO  nO 

o oo 
vo 


>N 

3 

-o 


<D  >>  <D 


> 3 

•a  2 

C/5 

_ w e 

CO  8 O 
OX)  g 


OX)  32 


OX)  ^ rt 
0)  £> 
. C 3 
o (U  u 
C IX  c 

- £ r 

<0  O o 

N X) 


o £ 


a ° = 

c <5 
- 2.  ox) 

3—30 

.2  s:  <u 

3 l « *= 
>2  <U  0X)  • 

Q £ UJ  Z 


Rodriguez  and  Eberhard  • REPRODUCTIVE  BEHAVIOR  OF  AMAZON  A OCHROCEPHALA  23 1 


■ Inside  nest  0 Inside  nest  area  □ Outside  nest  area 


Laying  Incubation  Nestling 

FIG.  1.  Mean  time  (out  of  180  min)  spent  by  adult 
Amazona  ochrocephala  panamensis  parrots  inside  or 
near  the  nest,  by  nesting  stage.  Data  are  presented  sep- 
arately for  males  (M)  and  females  (F);  error  bars  cor- 
respond to  standard  deviations.  Sample  sizes  refer  to 
the  number  of  3-hr  observation  periods.  The  number 
of  pairs  observed  was  as  follows:  1997-1998,  n — 5 
during  laying,  incubation,  and  nestling  stages;  1998- 
1999,  n — 5 during  laying  and  incubation  stages,  and 
n = 4 during  the  nestling  stage.  Total  observation 
times  were  27,  219,  and  405  hr  during  the  laying,  in- 
cubation, and  nestling  stages,  respectively. 


for  a short  time  (8—17  min)  to  perch  at  the 
cavity  entrance  or  on  a nearby  branch,  some- 
times engaging  in  allogrooming  with  the  male. 
On  10  occasions,  the  male  was  observed  feed- 
ing the  female  near  the  nest.  Males  spent 
much  of  their  time  in  the  nest  area,  and  typi- 
cally departed  on  two  foraging  trips  per  day — 
one  in  the  morning  and  the  other  in  the  late 
afternoon.  Early  in  the  morning  and  late  in  the 
afternoon,  the  female  often  left  the  nest  area — 
possibly  to  forage  with  the  male  (Fig.  1) — and 
remained  out  of  the  nest  area  for  85.6  ± 11.5 
min  (range  = 61-110  min).  For  27  of  83  de- 
partures, the  pair  departed  with  small  groups 
of  two  to  four  other  parrots — conspecifics 
and/or  Amazona  autumnalis.  Upon  their  re- 
turn, the  pairs  often  flew  in  the  company  of 
other  parrots  (30  of  83  arrivals).  At  the  end 
of  the  day,  the  male  usually  departed  from  the 
nest  area  (65%  of  late  afternoon  observation 
periods  in  1997-1998,  and  76%  in  1998- 
1999),  either  alone  or  with  other  parrots  as 
they  passed  by.  On  1 1 occasions  (involving 
10  different  nests),  we  made  nocturnal  nest 
checks  during  the  incubation  period  and  ex- 
amined the  nest  area  with  a flashlight;  on  only 
three  (27%)  of  the  checks  did  we  see  the  male 
perched  in  the  nest  tree. 

Nestling  period. — Chicks  hatched  with  their 


eyes  closed,  and  their  bodies  were  covered 
with  a sparse  white  down  that  was  later  re- 
placed by  a gray  down,  as  described  by  For- 
shaw  (1989).  Nestlings  spent  just  over  2 
months  in  their  nests  before  fledging,  varying 
somewhat  between  the  2 years  of  the  study 
(mean  age  at  fledging  in  1997-1998  = 68.6 
± 5.36  days,  n = 5 fledglings;  mean  age  at 
fledging  in  1998-1999  = 78.3  ± 3.88  days,  n 
= 7 fledglings).  Young  fledged  between  22 
March  and  5 April  in  1998,  and  between  6 
and  24  April  in  1999;  those  hatched  during 
the  1 st  year  fledged  in  less  time  than  did  those 
hatched  during  the  2nd  year  (Mann-Whitney 
U-test:  Z = —2.94,  n = 12  fledglings,  P = 
0.003). 

We  made  six  nocturnal  visits  to  nests  and 
on  five  of  the  visits  (83%)  the  male  was  found 
in  the  nest  area,  but  never  inside  the  nest  cav- 
ity; in  three  cases,  he  was  perched  near  the 
nest  entrance,  and  twice  he  was  perched  a few 
meters  away  in  a nearby  tree.  On  all  six  visits, 
the  female  was  inside  the  nest  (on  four  of 
these  occasions,  the  female  briefly  came  to  the 
nest  entrance  to  look  out  and  then  quietly 
went  back  inside;  on  the  other  two  visits,  she 
exited  the  nest  and  returned  —30  min  later). 

During  the  nestling  period,  we  typically 
found  the  female  inside  the  nest  as  we  began 
each  observation  period  (114  of  135  obser- 
vation periods).  She  would  spend  much  of  her 
time  brooding  recently  hatched  young,  but  as 
the  nestling  period  progressed,  she  decreased 
the  amount  of  time  spent  in  the  nest  (linear 
regression:  Fl  l32  = 419.08,  P < 0.001,  R2  = 
0.76,  b = -2.44).  When  outside  of  the  nest, 
she  perched  nearby  and  engaged  in  allo- 
grooming with  her  mate  and/or  she  left  the 
nest  area,  presumably  to  forage  (Fig.  1).  Each 
day  during  this  period,  both  the  male  and  fe- 
male would  follow  their  foraging  trips  with 
two  to  four  short  visits  to  the  nest  cavity,  pre- 
sumably to  feed  the  nestlings  (males:  mean 
duration  - 5.5  ± 1.3  min,  range  = 2. 2-8. 3, 
n = 135  observation  periods;  females:  mean 
duration  = 5.1  ± 1.2  min,  range  - 2.4-7. 5, 
n — 135  observation  periods). 

Nestlings  acquired  their  plumage  rather 
slowly.  The  flight  feathers  were  the  first  to 
appear,  with  pin  feathers  for  the  remiges  be- 
ginning to  emerge  when  nestlings  were  16  to 
28  days  old.  Green  contour  pin  feathers  on  the 
wings  and  yellow  contour  feathers  on  the  head 


232 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


began  to  unsheath  at  26-30  days,  and  contour 
feathers  on  the  legs  and  back  began  to  un- 
sheath at  35-38  days.  At  —40-42  days,  the 
green  feathers  on  the  head  and  red  feathers  at 
the  bend  of  the  wing  began  to  unsheath.  Fi- 
nally, at  —49-52  days,  the  tail  feathers  were 
completely  unsheathed.  About  2 weeks  before 
leaving  the  nest,  the  nestlings  began  to  perch 
at  the  cavity  opening.  Fledging  was  asynchro- 
nous, and  the  age  at  which  young  left  the  nest 
ranged  from  59  to  86  days  (see  Table  2).  We 
observed  the  nestlings’  first  flight  from  the 
nest  on  seven  occasions;  five  flights  occurred 
in  the  morning  and  two  in  the  late  afternoon. 
The  first  flights  were  relatively  short  (mean 
distance  = 34.6  ± 8.0  m,  range  = 25.0-48.5, 
n = 1 fledglings),  low,  and  quiet,  and  the 
young  were  accompanied  by  one  or  both 
adults.  After  the  last  chick  in  a clutch  had 
fledged,  neither  the  young  nor  the  adults  en- 
tered the  nest  cavity  again;  for  at  least  6 more 
days,  however,  the  adults  continued  to  visit 
the  nest  area.  Breeding  pairs  whose  nests  were 
poached  by  humans  or  failed  due  to  natural 
predation  did  not  make  a second  breeding  at- 
tempt in  the  same  cavity  that  year;  however, 
they  continued  to  visit  the  nest  area  for  at  least 
6 days  following  nest  failure. 

Breeding  success. — We  obtained  productiv- 
ity data  for  63  breeding  attempts  (Table  2). 
Overall,  breeding  success  of  A.  o.  panamensis 
was  very  low.  Over  both  breeding  seasons, 
only  12.7%  (8  of  63)  of  nests  fledged  young. 
Of  the  remaining  nests,  9.5%  (6  of  63)  failed 
due  to  natural  predation  at  the  nestling  stage, 
all  of  which  we  visually  confirmed  as  preda- 
tion by  boas  ( Boa  constrictor).  The  principal 
cause  of  breeding  failure  was  nest  poaching 
by  humans.  A total  of  77.8%  of  nests  (49  of 
63)  were  poached  or  presumed  to  have  been 
poached.  Poachers  accessed  nest  contents  by 
chopping  holes  in  trunks  at  the  level  of  the 
nest  cavity  (17  of  49  poached  nests),  climbing 
trees  to  reach  nests  (27  of  49),  and  less  fre- 
quently by  felling  trees  (5  of  49).  The  disap- 
pearance of  nestlings  often  coincided  with  ev- 
idence of  machete  cutting  of  understory  veg- 
etation near  the  nest  tree  (17  of  49  poached 
nests). 

Fourteen  of  the  19  cavities  (74%)  contain- 
ing nests  that  failed  due  to  predation  or  poach- 
ing during  the  1st  year  were  reused  during  the 
following  breeding  season.  Only  8 of  49  cav- 


ities monitored  during  one  or  both  breeding 
seasons  housed  nests  that  successfully  fledged 
young,  but  we  found  no  evidence  of  a rela- 
tionship between  breeding  success  and  the  di- 
mensions of  nest  trees  or  nest  cavities.  Dis- 
criminant function  analysis  indicated  that  the 
dimensions  of  trees  and  cavities  containing 
successful  versus  failed  (poached  or  depredat- 
ed) nests  did  not  differ  (Wilks’  Lambda  = 
0.7468,  F9>39  = 1.47,  P = 0.19,  n = 49  nests); 
only  cavity  depth  contributed  significantly  to 
the  discriminant  function  (Wilks’  Lambda  = 
0.8953,  P = 0.008). 

Poaching  techniques  and  illegal  trade. — 
Eighteen  parrot  poachers  were  interviewed, 
and  they  described  a range  of  poaching  strat- 
egies that  included  the  removal  of  unhatched 
eggs,  newly  hatched  nestlings,  fully  feathered 
nestlings,  and  the  capture  of  recently  fledged 
juveniles.  The  majority  of  poachers  (13  of  18) 
preferred  fully  feathered  nestlings  —40  days 
old  and  only  one  of  the  poachers  took  newly 
hatched  young  (3  to  8 days  old).  Relatively 
few  poachers  (2  of  18)  took  eggs  from  the 
nest,  and  the  remainder  (2  of  1 8)  preferred  to 
capture  juveniles  that  had  already  fledged. 
None  of  the  poachers  targeted  adult  parrots. 

More  than  three  quarters  of  the  poachers 
(14  of  18)  considered  the  demand  for  A.  o. 
panamensis  nestlings  to  be  very  high,  and  said 
that  they  always  had  customers  lined  up  to 
purchase  birds  even  before  they  had  been  tak- 
en from  their  nests.  Many  of  the  poached  birds 
are  sold  locally  to  customers  in  Chiriqui,  but 
poachers  indicated  that  vacationers  from  Pan- 
ama City  and  truck  drivers  involved  in  the 
transport  of  merchandise  between  Panama  and 
Costa  Rica  pay  the  highest  prices  (as  much  as 
US$100  for  a fully  feathered  and  healthy  par- 
rot chick).  Half  of  the  poachers  said  that  they 
typically  sold  parrot  nestlings  for  $40  or  more; 
most  of  the  others  (8  of  18,  44.4%)  sold  nest- 
lings for  $30-39,  and  only  one  of  the  poachers 
sold  nestlings  for  $20-29.  Poachers  were  not 
asked  to  reveal  total  annual  earnings  from 
poaching,  but  five  volunteered  this  informa- 
tion: four  indicated  that  they  typically  made 
$200-350  per  year  and  one  said  that  he  never 
earned  less  than  $200  annually  and  sometimes 
made  as  much  as  $750  per  year.  For  compar- 
ison, the  typical  monthly  salary  for  a farm  la- 
borer in  the  area  is  $130.  According  to  a 1990 
census  (Direccion  General  de  Estadistica  y 


Rodriguez  and  Eberhard  • REPRODUCTIVE  BEHAVIOR  OF  AMAZONA  OCHROCEPHALA  233 


Censo  de  Panama  1991),  the  human  popula- 
tion in  the  study  area  was  approximately 
2,358,  but  the  number  of  people  involved  in 
poaching  activities  is  difficult  to  estimate. 
Poaching  of  parrot  nestlings  is  punishable  by 
fines  of  up  to  $1,000,  but  poachers  indicated 
that  if  they  were  caught,  the  authorities  typi- 
cally seized  the  nestlings  and  did  not  impose 
any  further  punishment. 

Fifteen  of  the  poachers  interviewed  (83%) 
said  that  they  usually  collected  6-9  nestlings 
per  breeding  season,  and  the  remaining  indi- 
viduals typically  collected  2-5  nestlings  per 
season.  Most  of  the  poachers  (13  of  18)  said 
that  they  have  been  collecting  and  selling  par- 
rot nestlings  for  7-13  years,  and  the  others 
have  done  so  for  1-6  years.  Eleven  poachers 
(61%)  noted  that,  in  the  past,  they  had  also 
taken  A.  autumnalis  nestlings,  but  no  longer 
did  so  because  this  species  is  not  a good  im- 
itator of  human  speech  and  therefore  is  much 
less  marketable  than  A.  o.  panamensis.  Eight 
poachers  said  that  both  of  these  species  were 
hunted  for  food  in  eastern  Chiriqui,  but  five 
of  the  men  indicated  that  this  practice  is  no 
longer  common,  especially  in  the  case  of  A. 
o.  panamensis , which  could  be  sold  for  a rel- 
atively high  price.  Recently,  some  poachers 
have  begun  to  use  yellow  dye  on  the  forehead 
feathers  of  A.  autumnalis  and  even  Aratinga 
pertinax  (both  of  which  are  less  desirable  than 
A.  ochrocephala  in  the  pet  trade),  in  order  to 
sell  them  to  unsuspecting  buyers  as  A.  och- 
rocephala (AMRC  pers.  obs.).  These  data  in- 
dicate that  poaching  of  A.  o.  panamensis  is  not 
a new  phenomenon  and  has  likely  impacted 
resident  populations  of  the  species  by  reduc- 
ing recruitment  of  juveniles. 

DISCUSSION 

Characterization  of  nesting  habitats  and 
cavities  used  for  breeding. — Breeding  pairs  of 
A.  o.  panamensis  preferred  relatively  large 
cavities  high  up  in  trees  and  palms.  The  di- 
mensions of  the  cavities  used  by  these  parrots 
was  within  the  range  of  those  reported  for  oth- 
er Amazona  species,  such  as  A.  vittata  (Snyder 
et  al.  1987),  A.  leucocephala  bahamensis 
(Gnam  1991)  and  A.  barbadensis  (Rojas-Sua- 
rez  1994). 

We  found  no  evidence  that  A.  o.  panamen- 
sis prefers  to  nest  in  any  one  species  of  tree; 
the  frequency  of  nests  in  different  tree  species 


reflected  the  frequency  of  cavity  occurrence  in 
those  species.  Saunders  (1979)  found  a similar 
lack  of  preference  in  a study  of  Calyptorhyn- 
chus  baudinii  latirostris\  in  three  of  four  nest- 
ing areas  studied,  the  dominant  tree  species 
housed  the  majority  of  nests.  Snyder  et  al. 
(1987)  found  that  most  A.  vittata  nests  are  in 
palo  Colorado  ( Cyrilla  racemiflora ),  but  this 
was  due  to  the  scarcity  of  cavities  in  other  tree 
species  found  in  the  parrots’  habitat.  In  our 
study,  breeding  pairs  preferred  trees  in  good 
or  fair  condition.  This  contrasts  with  the  find- 
ing of  Sauad  et  al.  (1991),  who  found  that 
72%  of  A.  aestiva  nests  were  in  trees  that  were 
in  poor  condition  or  dead.  Similarly,  Calypto- 
rhynchus  magnificus  tended  to  nest  in  dead 
trees  more  often  than  expected  by  chance 
(Saunders  et  al.  1982). 

In  our  study  area  in  western  Panama,  we 
found  that  openings  of  cavities  occupied  by 
breeding  A.  o.  panamensis  tended  to  be  ori- 
ented toward  the  northeast.  A similar  prefer- 
ence for  certain  orientations  has  been  docu- 
mented for  several  other  parrots  (Rodrfguez- 
Vidal  1959,  Saunders  1979;  but  see  Saunders 
et  al.  1982,  Sauad  et  al.  1991). 

Breeding  behavior. — The  breeding  behavior 
of  A.  o.  panamensis  is  similar  to  that  reported 
for  other  psittacids.  Pairs  are  socially  monog- 
amous and  both  members  of  the  pair  contrib- 
ute significantly  to  nest  defense  and  caring  of 
young.  Allofeeding  of  the  female  by  her  mate, 
which  we  observed  on  several  occasions,  is 
typical  of  breeding  parrots  (Skeate  1984,  Sny- 
der et  al.  1987,  Gnam  1991,  Eberhard  1998), 
especially  early  in  the  breeding  cycle  (Snyder 
et  al.  1987,  Eberhard  1998).  Nevertheless,  fe- 
male A.  o.  panamensis  did  not  appear  to  de- 
pend on  their  mates  for  food;  they  regularly 
left  the  nest  area  with  their  mates,  presumably 
to  forage. 

Females  typically  laid  eggs  at  2-day  inter- 
vals, as  reported  for  other  congeners  (A.  vit- 
tata, Snyder  et  al.  1987;  A.  leucocephala  ba- 
hamensis, Gnam  1991;  and  A.  barbadensis, 
Rojas-Suarez  1994),  though  they  occasionally 
laid  eggs  on  successive  days  or  at  intervals  of 
up  to  5 days.  Clutch  size  varied  from  two  to 
four  eggs,  as  reported  for  A.  vittata  (Snyder 
et  al.  1987),  and  the  duration  of  incubation 
was  similar  to  that  reported  for  other  Amazona 
parrots  (Low  1972,  Skeate  1984,  Snyder  et  al. 
1987,  and  Rojas-Suarez  1994).  As  in  many 


234 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


other  parrots  (Forshaw  1989),  incubation  be- 
gan when  the  first  egg  was  laid,  resulting  in 
asynchronous  hatching,  and  the  female  was 
responsible  for  incubation.  During  incubation, 
the  female  occasionally  emerged  from  the  nest 
for  a few  minutes  at  a time  to  stretch,  groom, 
and  participate  in  nest  defense;  in  the  early 
morning  and  late  afternoon  she  often  departed 
for  longer  times,  possibly  to  forage.  The  sub- 
stantial proportion  of  time  spent  outside  of  the 
nest  during  this  period  was  greater  than  that 
reported  for  other  Amazona  parrots  (e.g.,  Sny- 
der et  al.  1987,  Wilson  et  al.  1995,  Renton  and 
Salinas-Melgoza  1999).  In  A.  vittata,  low  nest 
attendance  and  long  recesses  by  female  par- 
rots were  associated  with  failed  nesting  at- 
tempts (Wilson  et  al.  1997);  we  observed  sim- 
ilar behaviors  in  A.  o.  panamensis,  but  they 
did  not  appear  to  negatively  impact  breeding 
success,  and  the  duration  of  incubation  and 
the  number  of  eggs  hatched  per  clutch  in  our 
study  were  similar  to  those  reported  for  other 
Amazona  parrots  (Low  1972,  Snyder  et  al. 
1987,  Gnam  1991,  Rojas-Suarez  1994).  The 
long  departures  by  A.  o.  panamensis  females 
might  be  due  to  habitat  fragmentation  in  our 
study  area,  which  in  turn  has  disrupted  the 
parrots’  foraging  patterns,  as  observed  by 
Saunders  (1990)  in  a study  of  Carnaby’s 
Cockatoo  ( Calyptorhynchus  funereus  latiros- 
tris ) in  agricultural  areas. 

During  incubation,  we  never  saw  the  male 
enter  the  nest;  this  contrasts  with  observations 
of  A.  albifrons  (Skeate  1984)  and  A.  vittata 
(Snyder  et  al.  1987),  in  which  males  occa- 
sionally enter  the  nest  during  this  period.  In 
A.  o.  panamensis , the  male  spent  much  of  his 
time  near  the  nest  while  his  mate  was  incu- 
bating, possibly  to  alert  her  to  approaching 
predators  or  prevent  extra-pair  copulations  by 
his  mate  with  other  males;  however,  the  time 
the  male  spent  in  the  nest  area  did  not  differ 
between  the  egg-laying  (when  the  female  is 
fertile)  and  incubation  periods,  suggesting  that 
he  was  not  mate-guarding. 

The  female  was  apparently  responsible  for 
feeding  the  newly  hatched  chicks,  but  a few 
days  after  the  eggs  had  hatched,  the  male  be- 
gan to  enter  the  nest  regularly,  presumably  to 
feed  the  young.  This  also  has  been  reported 
for  other  Amazona  parrots,  including  A.  albi- 
frons (Skeate  1984),  A.  1.  bahamensis  (Gnam 
1991),  and  A.  vittata  (Snyder  et  al.  1987,  Wil- 


son et  al.  1995).  As  the  nestlings  grew,  the 
female  gradually  reduced  the  amount  of  time 
she  spent  in  the  nest  with  them.  She  ceased 
brooding  the  young  during  the  day  when  the 
oldest  nestling  was  18  to  25  days  old,  similar 
to  that  observed  in  other  Amazona  species 
(e.g.,  Snyder  et  al.  1987,  Enkerlin-Hoeflich 
and  Hogan  1997,  Renton  and  Salinas-Melgoza 
1999). 

Chicks  of  a single  clutch  usually  fledged  on 
different  days,  as  reported  for  A.  vittata  (Sny- 
der et  al.  1987)  and  A.  1.  bahamensis  (Gnam 
1991).  Mean  age  at  fledging  was  greater  than 
that  reported  by  Snyder  et  al.  (1987)  for  A. 
vittata , by  Rojas-Suarez  (1994)  for  A.  barba- 
densis , and  by  Renton  and  Salinas-Melgoza 
(1999)  for  A.  finschi.  As  described  for  A.  1. 
bahamensis  (Gnam  1991),  fledglings  were  ac- 
companied by  one  or  both  parents  on  their 
first  flight,  but  the  flights  of  A.  o.  panamensis 
fledglings  were  shorter.  After  leaving  the  nest, 
A.  o.  panamensis  fledglings  were  very  quiet, 
probably  to  avoid  attracting  the  attention  of 
predators;  similar  cryptic  behavior  has  been 
observed  in  A.  vittata  (Snyder  et  al.  1987). 

Breeding  success. — In  our  study  area,  the 
breeding  success  of  A.  o.  panamensis  was  low, 
principally  due  to  poaching,  and  to  a lesser 
extent  to  natural  predation  by  boas.  Habitat 
loss  due  to  deforestation,  which  often  involves 
felling  of  the  largest  trees,  has  been  cited  as 
an  important  cause  of  population  declines 
among  parrots  (Juniper  and  Parr  1998).  How- 
ever, in  the  case  of  A.  o.  panamensis  in  west- 
ern Panama,  our  results  indicate  that  breeding 
is  not  limited  by  the  availability  of  nesting 
sites,  even  though  much  of  the  area  has  been 
partially  cleared.  The  very  low  rate  of  breed- 
ing success  is  instead  due  to  extremely  high 
poaching  rates  fueled  by  demands  of  the  local 
pet  trade.  Low  salaries  and  the  lack  of  em- 
ployment opportunities  in  the  San  Juan  area 
undoubtedly  drive  individuals  to  poach  parrot 
nestlings.  Although  the  activity  is  illegal  and 
punishable  by  fines  of  up  to  $1,000,  anti- 
poaching laws  are  only  weakly  enforced.  Be- 
cause favored  poaching  techniques  are  fo- 
cused on  collecting  nestlings,  recruitment  into 
the  A.  o.  panamensis  population  is  severely 
reduced,  and  the  population  is  in  danger  of  a 
rapid  and  precipitous  decline  as  the  adults  age 
and  are  not  replaced  by  individuals  from 
younger  age  classes. 


Rodriguez  and  Eberhard  • REPRODUCTIVE  BEHAVIOR  OF  AMAZONA  OCHROCEPHALA  235 


ACKNOWLEDGMENTS 

We  thank  the  owners  and  laborers  of  Hacienda  El 
Tekal,  Hacienda  Miraflores,  and  Hacienda  Los  Asen- 
tamientos  de  San  Juan  for  their  collaboration  and  for 
allowing  access  to  their  properties.  Partial  financial 
support  was  provided  by  the  German  Technical  Co- 
operation Agency  through  its  Proyecto  Agroforestal 
Ngobe-ANAM-GTZ.  We  are  grateful  that  some  of  the 
parrot  poachers  were  willing  to  reveal  information 
about  their  activities,  which  may  help  in  developing 
future  conservation  measures.  We  also  thank  the  resi- 
dents of  the  Corregimiento  de  San  Juan,  especially 
those  engaged  in  the  collection  and  sale  of  parrots  in 
the  local  pet  trade,  for  sharing  information  on  their 
poaching  activities.  Two  poachers  were  helpful  in  lo- 
cating nests  as  well  as  establishing  contacts  with 
poachers  and  others  involved  in  the  local  parrot  trade. 
We  thank  K.  Harms  for  assistance  with  performing  sta- 
tistical tests  in  R,  and  F.  Gomez  and  E.  de  Morris  for 
their  comments  on  an  early  version  of  the  manuscript; 
the  manuscript  also  benefited  from  extensive  com- 
ments provided  by  three  anonymous  reviewers. 

LITERATURE  CITED 

Acosta,  J.  L.  1996.  Inventario  floristico  en  dos  co- 
munidades  del  Distrito  de  San  Lorenzo,  Chiriqui. 
Tesis  de  Licenciatura,  Escuela  de  Biologfa,  Univ- 
ersidad  de  Panama,  Ciudad  de  Panama,  Panama. 
Asociacion  Nacional  Para  la  Conservacion  de  la 
Naturaleza.  1995.  Fauna  silvestre  en  peligro  de 
extincion.  Grupo  Editorial  del  Istmo,  Ciudad  de 
Panama,  Panama. 

Autoridad  Nacional  del  Ambiente.  1995a.  Especies 
animates  en  peligro  de  extincion  en  Panama.  Gru- 
po Editorial  del  Istmo,  Ciudad  de  Panama,  Pana- 
ma. 

Autoridad  Nacional  del  Ambiente.  1995b.  Informes 
de  los  operativos  de  decomiso  de  aves  en  peligro 
de  extincion  en  Panama.  Direccion  Regional  de  la 
Provincia  de  Chiriqui,  David,  Chiriqui,  Panama. 
Batschelet,  E.  1981.  Circular  statistics  in  biology. 

Academic  Press,  New  York. 

Direccion  General  de  Estadistica  y Censo.  1991. 
Censos  nacionales  de  poblacion  y vivienda,  13  de 
mayo  de  1990.  Resultados  finales  basicos:  total 
del  pais.  Contralorfa  General  de  la  Republica, 
Ciudad  de  Panama,  Panama. 

Eberhard,  J.  R.  1998.  Breeding  biology  of  the  Monk 
Parakeet.  Wilson  Bulletin  110:463-473. 
Eberhard,  J.  R.  and  E.  Bermingham.  2004.  Phytog- 
eny and  biogeography  of  the  Amazona  ochroce- 
phala  complex.  Auk  121:318-332. 
Enkerlin-Hoeflich,  E.  C.  1995.  Comparative  ecology 
and  reproductive  biology  of  three  species  of  Ama- 
zona parrots  in  northeastern  Mexico.  Ph.D.  dis- 
sertation, Texas  A&M  University,  College  Station. 
Enkerlin-Hoeflich,  E.  C.  and  K.  M.  Hogan.  1997. 
Red-crowned  Parrot  ( Amazona  viridigenalis).  The 
Birds  of  North  America,  no.  292. 

Forshaw,  J.  M.  1989.  Parrots  of  the  world,  3rd  (re- 


vised) ed.  Lansdowne  Editions,  Melbourne,  Aus- 
tralia. 

Gnam,  R.  S.  1991.  Nesting  behaviour  of  the  Bahama 
Parrot  (. Amazona  leucocephala  bahamensis ) on 
Abaco  Island,  Bahamas.  Acta  Congressus  Inter- 
nationalis  Ornithologici  20:673-680. 

Holdridge,  L.  R.  1967.  Life  zone  ecology.  Tropical 
Science  Center,  San  Jose,  Costa  Rica. 

Huberty,  C.  J.  1994.  Applied  discriminant  analysis. 
John  Wiley  and  Sons,  Athens,  Georgia. 

Instituto  de  Recursos  Hidraulicos  y Electrifica- 
cion.  1998.  Informe  anual  de  temperatura  y pre- 
cipitacion  pluvial  en  la  provincia  de  Chiriqui.  De- 
partamento  de  Hidrometeorologfa-Seccion  Chiri- 
qui, David,  Panama. 

Instituto  de  Recursos  Hidraulicos  y Electrifica- 
cion.  1999.  Informe  tecnico:  evaluacion  anual  de 
precipitacion  pluvial,  temperatura  y humedad  re- 
lativa  de  las  zonas  bajas  de  la  provincia  de  Chi- 
riqui. Departamento  de  Hidrometeorologfa-Sec- 
cion Chiriqui,  David,  Panama. 

Juniper,  T.  and  M.  Parr.  1998.  Parrots:  a guide  to  the 
parrots  of  the  world.  Yale  University  Press,  New 
Haven,  Connecticut. 

Low,  R.  1972.  The  parrots  of  South  America.  John 
Gifford,  London,  United  Kingdom. 

R Development  Core  Team.  2005.  R:  a language  and 
environment  for  statistical  computing.  R Founda- 
tion for  Statistical  Computing,  Vienna,  Austria. 

Renton,  K.  and  A.  Salinas-Melgoza.  1999.  Nesting 
behavior  of  the  Lilac-crowned  Parrot.  Wilson  Bul- 
letin 111:488-493. 

Ridgely,  R.  S.  1981.  The  current  distribution  and  sta- 
tus of  mainland  Neotropical  parrots.  Pages  233- 
384  in  Conservation  of  New  World  parrots  (R.  F. 
Pasquier,  Ed.).  International  Council  for  Bird  Pres- 
ervation Technical  Bulletin,  no.  1.  Smithsonian 
Press,  Washington,  D.C. 

Rodriguez- Vidal,  J.  A.  1959.  Puerto  Rican  parrot 
study.  Monographs  of  the  Department  of  Agricul- 
ture and  Commerce,  Puerto  Rico,  no.  1. 

Rojas-Suarez,  F.  1994.  Biologfa  reproductiva  de  la 
Cotorra  Amazona  barbadensis  (Aves:  Psittacifor- 
mes)  en  la  Peninsula  de  Macanao,  Estado  de  Nue- 
va  Esparta.  Pages  73-87  in  Biologfa  y conserva- 
cion de  los  psitacidos  de  Venezuela  (G.  Morales, 
I.  Novo,  D.  Bigio,  A.  Luy,  and  F.  Rojas-Suarez, 
Eds.).  Editorial  Giavimar,  Caracas,  Venezuela. 

Sauad,  J.  J.,  V.  Nunez,  J.  L.  Garrido,  S.  Mosa,  M. 
E.  Calzon,  and  Z.  M.  Chorolque.  1991.  Am- 
bientes  de  nidificacion  del  Loro  Hablador  Ama- 
zona aestiva.  Salta,  Argentina.  III.  Caracterfsticas 
de  los  arboles  nido.  Publicacion  Tecnica,  no.  5, 
Manejo  de  Fauna.  Universidad  Nacional  de  Salta, 
Salta,  Argentina. 

Saunders,  D.  A.  1979.  The  availability  of  tree  hollows 
for  use  as  nests  sites  by  White-tailed  Black  Cock- 
atoos. Australian  Wildlife  Research  6:205-216. 

Saunders,  D.  A.  1990.  Problems  of  survival  in  an  ex- 
tensively cultivated  landscape:  the  case  of  Car- 


236 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


naby’s  Cockatoo  (Calypto rhyn chus  funereus  lati- 
rostris ).  Biological  Conservation  54:277-290. 

Saunders,  D.  A.,  G.  T.  Smith,  and  I.  Rowley.  1982. 
The  availability  and  dimensions  of  tree  hollows 
that  provide  nest  sites  for  cockatoos  (Psittacifor- 
mes)  in  Western  Australia.  Australian  Wildlife  Re- 
search 9:541-556. 

Seixas,  G.  H.  F.  and  G.  M.  Mourao.  2002.  Nesting 
success  and  hatching  survival  of  the  Blue-fronted 
Amazon  ( Amazona  aestiva ) in  the  Pantanal  of 
Mato  Grosso  do  Sul,  Brazil.  Journal  of  Field  Or- 
nithology 73:399-409. 

Silva,  A.  P.  D.  and  A.  Stam.  1995.  Discriminant  anal- 
ysis. Pages  277—318  in  Reading  and  understand- 
ing multivariate  statistics  (L.  G.  Grimm  and  P.  R. 
Yamold,  Eds.).  American  Psychological  Associa- 
tion, Washington,  D.C. 

Skeate,  S.  T.  1984.  Courtship  and  reproductive  behav- 
iour of  captive  White-fronted  Amazon  Parrots 
(. Amazona  albifrons ).  Bird  Behaviour  5:103—109. 

Snyder,  N.  F.  R.,  J.  W.  Wiley,  and  C.  B.  Kepler. 


1987.  The  parrots  of  Luquillo:  natural  history  and 
conservation  of  the  Puerto  Rican  Parrot.  Western 
Foundation  of  Vertebrate  Zoology,  Los  Angeles, 
California. 

StatSoft,  Inc.  1998.  Statistica  6.0.  StatSoft,  Inc.,  Tul- 
sa, Oklahoma. 

Wilson,  K.  A.,  R.  Field,  and  M.  H.  Wilson.  1995. 
Successful  nesting  behavior  of  Puerto  Rican  Par- 
rots. Wilson  Bulletin  107:518-529. 

Wilson,  K.  A.,  M.  H.  Wilson,  and  R.  Field.  1997. 
Behavior  of  Puerto  Rican  Parrots  during  failed 
nesting  attempts.  Wilson  Bulletin  109:490-503. 

Wozniak,  S.  and  W.  Lanterman.  1984.  Breeding  the 
Green-cheeked  Amazon  Parrot  Amazona  viridi- 
genalis  at  the  Ornithological  Institute,  Oberhau- 
sen,  Germany.  Avicultural  Magazine  90:195-197. 

Wright,  T.  F,  C.  A.  Toft,  E.  Enkerlin-Hoeflich,  J. 
Gonzalez-Elizondo,  M.  Albornoz,  A.  Ro- 
driguez-Ferraro,  F.  Rojas-Suarez,  et  al.  2001. 
Nest  poaching  in  Neotropical  parrots.  Conserva- 
tion Biology  15:710-720. 


The  Wilson  Journal  of  Ornithology  1 1 8(2):237-243,  2006 


GREGARIOUS  NESTING  BEHAVIOR  OF  THICK-BILLED  PARROTS 
(. RHYNCHOPSITTA  PACHYRHYNCHA)  IN  ASPEN  STANDS 

TIBERIO  C.  MONTERRUBIO-RICO,1’3  JAVIER  CRUZ-NIETO,2 
ERNESTO  ENKERLIN-HOEFLICH,2  DIANA  VENEGAS-HOLGUIN,2 
LORENA  TELLEZ-GARCIA,1  AND  CONSUELO  MARIN-TOGO1 


ABSTRACT. — We  studied  Thick-billed  Parrot  ( Rhynchopsitta  pachyrhyncha)  nest-site  density  and  social  nest- 
ing behavior  from  1998  to  2001  in  Madera,  Chihuahua,  Mexico.  The  species  formed  high-density  nesting 
clusters;  45  nesting  attempts  (30%)  involved  nesting  pairs  sharing  nest  trees,  with  a maximum  of  three  nesting 
pairs  per  tree.  The  majority  of  nest  trees  were  live  or  dead  quaking  aspens  ( Populus  tremuloides ).  Clusters 
contained  a mean  of  11.5  breeding  pairs  (5  nests/ha).  The  highly  social  nesting  behavior  of  Thick-billed  Parrots 
may  have  important  implications  for  management  and  conservation  of  their  breeding  habitat.  Received  31  March 
2005,  accepted  8 January  2006. 


Approximately  13%  of  all  bird  species  nest 
in  colonies  (Gill  1990).  Colonial  or  gregarious 
nesting  behavior  provides  important  advantag- 
es for  birds,  including  mate  access,  reduced 
probability  of  nest  predation,  improved  detec- 
tion and  defense  against  aerial  predators  while 
feeding,  and  enhanced  foraging  efficiency 
(Siegel-Causey  and  Kharitonov  1990,  Dan- 
chin  and  Wagner  1997,  Eberhard  2002).  De- 
spite the  advantages  of  colonial  nesting,  nest- 
site  availability  may  be  a limiting  factor  for 
social  species,  especially  those  that  nest  in 
tree  cavities  (Eberhard  2002). 

Colonial  nesting  is  uncommon  in  tree-cav- 
ity nesting  species  and  it  is  particularly  rare 
among  Neotropical  parrots  for  two  reasons: 
(1)  closely  spaced  tree  cavities  with  suitable 
characteristics  for  nesting  are  rare,  and  (2) 
most  parrot  species  are  territorial  around  nest 
sites  (Forshaw  1989,  Munn  1992,  Inigo-Elias 
1996).  Of  the  231  parrot  species,  Eberhard 
(2002)  reported  that  only  3 breed  colonially. 

In  Mexico,  20  parrot  species  nest  in  tree 
cavities,  4 of  which  (genus  Aratinga ) also  nest 
in  termitaries  (Hardy  1963,  Forshaw  1989, 
Howell  and  Webb  1995,  Rodrfguez-Estrella  et 
al.  1995).  Both  Maroon-fronted  Parrots  ( Rhyn- 
chopsitta terrisi)  and  Military  Macaws  ( Ara 


1 Lab.  de  Manejo  y Conservation  de  Fauna  Silves- 
tre,  Facultad  de  Biologfa,  Univ.  Michoacana  de  San 
Nicolas  de  Hidalgo,  Morelia,  Michoacan,  Mexico. 

2 Centro  de  Calidad  Ambiental,  Inst.  Tecnologico  y 
de  Estudios  Superiores  de  Monterrey,  Monterrey,  Nue- 
vo Leon,  Mexico. 

3 Corresponding  author;  e-mail: 
tiberio@zeus.umich.mx 


militaris ) sometimes  nest  at  high  densities  in 
cliff  crevices,  thus  forming  nesting  colonies 
(Forshaw  1989,  Macfas-Caballero  1998).  Pri- 
or to  our  study,  cavity-nesting  parrot  species 
in  Mexico  were  not  thought  to  nest  colonially 
in  tree  cavities,  nor  had  there  been  reports  of 
multiple  pairs  nesting  within  the  same  tree 
(Enkerlin-Hoeflich  1995,  Renton  and  Salinas- 
Melgoza  1999).  Although  most  parrots  are  so- 
cial and  have  been  considered  “suppressed 
colonial  breeders”  (Ward  and  Zahavi  1973), 
the  relative  density  of  suitable  tree  cavities  is 
low  and  competition  for  cavities  is  high 
(Munn  1992,  Gibbs  et  al.  1993). 

The  Thick-billed  Parrot  ( Rhynchopsitta  pa- 
chyrhyncha) is  a highly  social  species  that 
breeds  at  elevations  from  2,000  to  2,700  m in 
mature  and  old-growth  coniferous  forests  in 
the  northern  portions  of  the  Sierra  Madre  Oc- 
cidental, northwestern  Mexico.  Social  behav- 
iors include  the  formation  of  foraging  flocks, 
sentinel  posting  during  foraging,  simultaneous 
courtship  and  copulations  of  several  pairs  in 
neighboring  trees,  loud  vocalizations  of  neigh- 
boring nesting  pairs,  synchronized  defense 
against  raptors,  and  the  formation  of  large,  no- 
madic flocks  in  winter  (Lanning  and  Shiflett 
1983;  Snyder  et  al.  1994,  1999).  Even  when 
distances  among  nests  are  substantial,  males 
of  neighboring  nesting  pairs  communicate  and 
wait  for  each  other  when  forming  foraging 
flocks  (Snyder  et  al.  1999,  Monterrubio-Rico 
and  Enkerlin-Hoeflich  2004b). 

After  decades  of  intensive  logging,  few 
large  fragments  of  old-growth  forest  remain  in 
the  Sierra  Madre  Occidental;  thus,  the  number 


237 


238 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol  118,  No.  2,  June  2006 


110°  100°  90° 


FIG.  1.  Thick-billed  Parrot  study  area  near  Madera,  Chihuahua,  Mexico,  1998-2001. 


and  quality  of  Thick-billed  Parrot  nesting  ar- 
eas has  been  reduced  and  the  availability  of 
food  resources  has  probably  been  altered 
(Lanning  and  Shiflett  1983,  Benkman  1993, 
Lammertink  et  al.  1996).  Only  five  nesting  ar- 
eas are  known  to  remain  in  the  species’  breed- 
ing range,  and  two  of  them  (Cebadillas  de  Ya- 
huirachi  and  Madera)  encompass  >70%  of  the 
known  nesting  trees  (Monterrubio-Rico  and 
Enkerlin-Hoeflich  2004a).  Our  objectives 
were  to  evaluate  nest-site  use,  nest-tree  distri- 
bution, density  of  nesting  pairs,  and  tree  shar- 
ing by  nesting  pairs. 

METHODS 

The  study  area  was  near  Madera,  Chihua- 
hua, at  the  eastern  edge  of  the  Sierra  Madre 
Occidental,  (29°  19'  N,  108°  1 1'  W;  Fig.  1). 
Common  tree  species  included  Douglas-fir 
( Pseudotsuga  menziesii ),  white  fir  {Abies  con- 
color ).  Mexican  white  pine  ( Pinus  ayacahui- 
te),  and  quaking  aspen  {Populus  tremuloides). 
We  monitored  breeding  activity  from  July  to 
late  October  in  1998-2001.  The  total  area  sur- 
veyed for  nests  increased  each  year  from  5 ha 


in  1998  to  75  ha  in  2001.  Because  time,  per- 
sonnel, and  access  to  nesting  areas  were  lim- 
ited, however,  we  were  unable  to  completely 
sample  and  map  the  distributions  of  aspen 
stands  and  nest  trees. 

We  found  nests  by  conducting  intensive 
searches  during  the  prospecting  and  courtship 
phases  of  the  nesting  cycle.  A tree  cavity  was 
considered  a potential  nest  site  if  a nesting 
pair  was  observed  entering  the  cavity  during 
the  egg-laying  period  (late  July).  When  pos- 
sible, we  used  climbing  equipment  to  confirm 
presence  of  eggs  or  nestlings;  inaccessible  tree 
cavities  were  confirmed  as  nesting  cavities 
when  nestlings  could  be  heard  or  adult  parrots 
were  observed  feeding  nestlings.  A tree  cavity 
was  considered  a roost  site  if  it  was  used  by 
the  parrots  but  never  contained  eggs.  Because 
nesting  parrots  were  not  individually  marked, 
it  is  likely  that  some  birds  were  sampled  in 
multiple  years;  thus,  we  report  our  results  in 
terms  of  nesting  attempts  rather  than  number 
of  pairs. 

For  each  nest  tree,  we  documented  species, 
condition  (live  or  dead),  diameter  at  breast 


Monte rrubio-Rico  et  al.  • THICK-BILLED  PARROT  GREGARIOUS  NESTING 


239 


TABLE  1.  Use  of  nest  and  roost  trees  by  Thick-billed  Parrots,  Madera,  Chihuahua,  Mexico,  1998-2001. 

Parameter 

Number  per  year 

1998 

1999 

2000 

2001 

1998-2001 

Nesting  pairs  (attempts) 

20 

24 

30 

73 

147 

Nest  trees 

17 

23 

28 

55 

123a 

Nest  trees  used  by  one  pair 

14 

22 

26 

40 

102 

Nest  trees  with  >1  pair 

3 

1 

2 

15 

21 

Nesting  pairs  sharing  a tree 

6 

2 

4 

33 

45 

Trees  used  as  roost  sites 

3 

3 

9 

15 

30 

a72  different  nest  trees:  40  used  once,  18  used  twice,  9 used  three  times,  5 used  four  times  = 123. 


height  (dbh),  cavity  height,  and  tree  height. 
The  coordinates  of  each  nest  tree  were  ob- 
tained with  a Geographic  Positioning  System 
(GPS),  and  locations  were  plotted  on  topo- 
graphic maps  (scale  1:50,000).  Distances  be- 
tween neighboring  nesting  trees  were  mea- 
sured with  a 50-m  tape  or  determined  from 
GPS  coordinates  (for  trees  >100  m apart). 
Nest-tree  distribution  was  analyzed  with  Geo- 
graphic Information  System  software  (GIS; 
Arc  View  3.3)  using  geographic  coordinates 
with  six  decimals.  GIS  was  also  used  to  gen- 
erate a map  and  analyze  nest  distribution. 

We  defined  a “colony”  as  an  aggregation 
of  interacting  neighboring  groups  of  nesting 
pairs.  We  used  the  minimum  convex  polygon 
criterion  to  define  nesting  clusters,  where  a 
cluster  consisted  of  >3  nests,  each  <150  m 
from  any  other  nest;  the  significance  level  was 
set  at  a = 0.05  and  means  are  presented  ± 
SD.  Statistical  analyses  were  performed  with 
SAS  (SAS  Institute,  Inc.  1985). 

RESULTS 

During  4 years  of  study,  we  documented 
147  nesting  attempts  in  72  different  trees;  we 
also  documented  10  different  trees  used  as 
roost  sites.  We  monitored  48  of  the  nest  trees 
for  at  least  2 nesting  seasons  and  found  that 
33  (68%)  were  reused  in  subsequent  years; 
mean  annual  reuse  was  62  ± 0.08%  (range  = 
56-71%).  Eighty  of  the  82  trees  used  for  nest- 
ing or  roosting  were  aspen,  and  2 were  Mex- 
ican white  pine.  Aspen  snags  (n  = 39)  and 
live  aspen  ( n = 41)  were  used  with  similar 
frequency,  and  25%  ( n = 18)  of  the  snags 
were  severely  deteriorated  (total  absence  of 
bark).  The  majority  of  all  147  nesting  attempts 
(86%)  occurred  in  tree  cavities  that  appeared 
to  be  old  woodpecker  holes,  but  20  nesting 
attempts  (14%)  occurred  in  natural  cavities 


formed  by  tree  decay  and  detachment  of  large 
branches.  We  also  recorded  30  cavities  used 
as  roost  sites  (Table  1). 

Sixty-nine  percent  (102)  of  the  nesting  at- 
tempts involved  only  one  pair  of  parrots  per 
nest  tree.  The  45  remaining  attempts  (30%), 
however,  involved  more  than  one  pair  per  tree: 
18  attempts  involved  two  nesting  pairs  using 
different  cavities  in  the  same  nest  tree,  and 
three  times  we  observed  three  nesting  pairs 
sharing  different  cavities  in  the  same  tree  (Ta- 
ble 1).  We  found  more  nesting  pairs  in  2001 
C n = 73)  than  in  other  years  (Table  1),  but  that 
was  also  the  year  in  which  the  greatest  area 
(75  ha)  was  searched  for  nests. 

Overall,  the  parameters  of  trees  containing 
multiple  cavities  did  not  differ  significantly 
from  those  containing  only  one  cavity.  Nest 
trees  containing  >1  active  nest  did  not  differ 
in  dbh  (Wilcoxon  Z = 0.38,  P = 0.70;  mul- 
tiple-nest  trees:  57.0  ± 12.2  cm;  single-nest 
trees:  57.8  ±11.9  cm)  or  tree  height  (Wilcox- 
on Z = 1.82;  P = 0.068;  multiple  nest  trees: 
28.0  ± 5.4  m;  single  nest  trees:  24.7  ± 6.0 
m).  Vertical  distance  between  nest  cavities  in 
multiple-nest  trees  ranged  from  1 to  1 1 m 
(mean  = 4.3  ± 2.9  m).  Nest-cavity  height 
ranged  from  6.5  to  31  m above  ground  in  sin- 
gle-nest  trees  and  from  9 to  21  m (lowest  cav- 
ity) in  multiple-nest  trees. 

Most  nest  trees  used  by  Thick-billed  Parrots 
showed  a clumped  distribution  pattern,  form- 
ing aggregations  (nest  clusters)  in  aspen 
stands  (Fig.  2).  Mean  nest  cluster  area  was  2.3 
± 1.7  ha,  (range  = 0.04-4.4  ha),  and  mean 
within-cluster  nest  density  was  20.9  ± 32.6 
per  ha  (range  = 2.4-100  nests/ha;  Table  2). 
Mean  within-cluster  distance  between  active 
nests  was  31.9  ± 26.4  m (range  = 1.8-146 
m,  n = 147),  and  mean  distance  between  clus- 


240 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


1998  1999 


• Single  nests  * 

\ 500  0 500  1,000 

▲ Multiple  nests  [\ 


LIG.  2.  Diagram  showing  the  distribution  of  nesting  trees,  by  year,  in  1 1 Thick-billed  Parrot  nesting  clusters 
in  Madera,  Chihuahua,  Mexico.  Nesting  clusters  were  defined  using  the  minimum  convex  polygon  criterion. 
Capital  letters  indicate  the  different  clusters,  filled  circles  represent  nest  trees  with  one  nest,  and  triangles 
represent  nest  trees  with  two  or  three  nests. 


ters  was  325  ± 125  m (range  = 185-458  m, 
n — 11).  The  mean  number  of  nesting  at- 
tempts per  cluster  was  11.5  ± 8.1  (range  = 
3-31). 

No  agonistic  behavior  was  observed  among 
nesting  pairs.  Neighboring  nesting  pairs  were 
in  permanent  contact:  synchronized  foraging 
flocks  formed  every  morning  and  communi- 
cation among  pairs  occurred  with  loud  vocal- 
izations and  visual  contact.  We  also  observed 
five  events  of  collective  responses  to  raptors, 
in  which  parrots  rapidly  formed  a flock  after 
sharp  alarm  calls  had  been  emitted  by  the  par- 
rots that  first  detected  the  raptors. 

DISCUSSION 

Nest  site  density. — The  Thick-billed  Parrot 
is  a social  species  that  tolerates  other  nesting 
pairs,  often  in  the  same  nest  tree.  Previously, 


Lanning  and  Shiflett  (1983)  had  observed  two 
active  nests  (only  2 m apart)  in  a pine  snag. 
They  also  observed  two  pairs  nesting  in  a 
large  aspen  within  215  m of  three  other  nests 
and  within  1 km  of  six  additional  nests.  How- 
ever, we  observed  considerably  greater  nest 
density  and  number  of  nesting  pairs  sharing 
nest  trees  than  those  reported  by  Lanning  and 
Shiflett  (1983)  and  Snyder  et  al.  (1999).  The 
mean  distance  between  active  nests  (31.9  m) 
and  the  shortest  distance  (1.8  m)  between 
nesting  pairs  of  Thick-billed  Parrots  were  the 
smallest  values  reported  for  any  cavity-nesting 
parrot  species  in  Mexico.  For  example,  the 
same  values  for  Lilac-crowned  Parrot  ( Ama - 
zona  finschi ) in  the  tropical  subdeciduous  for- 
ests of  the  Chamela-Cuixmala  Biosphere  Re- 
serve were  948  m and  25  m,  respectively 
(Renton  and  Salinas-Melgoza  1999).  The 


TABLE  2.  Characteristics  of  Thick-billed  Parrot  nest  clusters  in  aspen  stands,  Madera,  Chihuahua,  Mexico,  1998-2001. 


Monterrubio-Rico  et  al.  • THICK-BILLED  PARROT  GREGARIOUS  NESTING 


241 


7 3 4 

nod 

m tn  a\ 
— ^ ni  o 


<n  rn 


vq  — < 00 
m os  cn  in 


^ o 
T o 
^ o 

2 7 

^ o o 
non 

cn  in  q 
n d d n 
— cn 


m in 

1 > 
I 

o it 
o (N 


n ^ °o 
m «n  id 

CO 


(N  in 
^ ni  in 
LJ  i i 
i T T 
oo  — < in 

o O)  Tt 
o — : in 


2 g? 


o c 
o « 

2 S 


O J3 
GO 

.5  c 
S 

c jC 

l+H  >> 

O .3 


<D  <D 

s s 


clumped  nest  distribution  and  multiple  nests 
per  tree  of  Thick-billed  Parrots  observed  in 
Madera  may  be  explained  by  (1)  the  existence 
of  adequate  tree  cavities  at  high  densities  and 
(2)  the  species’  high  level  of  sociality  and  tol- 
erance of  neighboring  nesting  pairs.  It  also 
may  be  that  nesting  pairs  experience  lower 
rates  of  predation  by  selecting  tree  cavities 
near  other  pairs. 

Conservation  and  management  recommen- 
dations.— In  addition  to  high  nest  density,  we 
also  documented  >50%  reuse  of  cavities  by 
Thick-billed  Parrots.  Lanning  and  Shiflett 
(1983)  recorded  lower  nest  densities  and  a 
lower  level  of  cavity  reuse  (1  of  12  nesting 
cavities  in  good  condition  were  reused).  High 
nest  density  and  reuse  of  cavities  may  indicate 
a scarcity  of  adequate  nesting  cavities  in  the 
surrounding  conifer  forests.  Several  authors 
have  addressed  the  alarming  reduction  in  the 
extent  of  old-growth  conifer  forests  in  the  Si- 
erra Madre  Occidental  and  its  negative  impact 
on  the  Thick-billed  Parrot  (Lanning  and  Shi- 
flett 1983,  Lammertink  et  al.  1996,  Snyder  et 
al.  1999).  As  a result  of  habitat  loss,  most  of 
the  nesting  activity  is  now  concentrated  in  two 
areas  (Cebadillas  de  Yahuirachi  and  Madera), 
making  the  species  vulnerable  to  the  effects 
of  illegal  logging,  forest  fires,  and  conifer  crop 
failures  (Benkman  1993,  Snyder  et  al.  1994, 
Monterrubio-Rico  and  Enkerlin-Hoeflich 
2004a). 

A fundamental  conservation  goal  for  Thick- 
billed Parrots  should  be  to  increase  the  num- 
ber of  nesting  areas.  This  can  be  achieved  by 
protecting  stands  of  old-growth  in  all  current 
and  historical  nesting  areas,  especially  those 
in  the  high-elevation  (2,000-3,000  m)  forests 
of  Durango  and  Chihuahua.  Because  trees 
large  enough  to  support  suitable  cavities  may 
take  4 decades  or  more  to  form,  nest  boxes 
should  be  erected  to  augment  nest-site  avail- 
ability. Although  it  remains  unknown  whether 
Thick-billed  Parrots  will  use  nest  boxes  in  the 
wild,  they  are  known  to  use  them  in  captivity 
(Snyder  et  al.  1994).  Retaining  old-growth  co- 
niferous forest  will  also  ensure  seed  avail- 
ability (Benkman  1993)  and  nesting  opportu- 
nities for  other  obligate  cavity  nesters,  such  as 
Eared  Quetzal  ( Euptilotis  neoxenus),  and 
Mexican  Spotted  Owl  (Strix  occidentalis  lu- 
cida ) — species  that  nest  in  the  same  habitats 
as  Thick-billed  Parrot. 


242 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


Areas  managed  for  Thick-billed  Parrots 
should  include  tree  species  commonly  used  as 
nest  sites,  such  as  Mexican  white  pine,  Doug- 
las-fir,  and  aspen.  In  addition,  pine  species 
such  as  Durango  pine  ( Pinus  durangensis ), 
teocote  pine  ( Pinus  teocote ),  Chihuahuan  pine 
( Pinus  leiophylla ),  and  Apache  pine  ( Pinus 
engelmannii),  should  be  included  to  provide  a 
constant  cone  crop  (Snyder  et  al.  1999).  Al- 
though stands  of  large  aspen  are  uncommon 
in  conifer  forests  of  the  Sierra  Madre  Occi- 
dental, aspens  can  be  planted  in  more  humid 
areas  selected  for  restoration. 

Populations  of  Thick-billed  Parrots  are  rel- 
atively small,  even  in  the  most  important  nest- 
ing areas.  Thus,  the  species’  recovery  will  re- 
quire sustained  periods  of  high  nesting  suc- 
cess and  productivity.  This  can  be  achieved 
only  by  providing  parrot  populations  with  ad- 
equate nesting  opportunities  across  the  land- 
scape. 

ACKNOWLEDGMENTS 

We  appreciate  the  help  and  advice  of  N.  F.  R.  Sny- 
der, R.  B.  Hamilton,  K.  Renton,  M.  S.  Hafner,  C.  Ma- 
cias, M.  A.  Cruz-Nieto,  C.  Valle,  and  J.  Shiflett.  Fi- 
nancial support  was  provided  by  CONABIO  (Mexico 
Biodiversity  Commission),  Sacramento  Zoo,  National 
Fish  and  Wildlife  Foundation,  Arizona  Game  and  Fish 
Department,  USFWS/SEMARNAT  program  in  Biodi- 
versity Conservation  for  NAFTA  initiative.  Wildlife 
Preservation  Trust,  Fondo  Sectorial  Ambiental  SE- 
MARNAT-CONACYT,  Centro  de  Calidad  Ambiental 
at  Instituto  Tecnologico  y Estudios  Superiores  de  Mon- 
terrey, and  the  School  of  Renewable  Natural  Resources 
at  Louisiana  State  University.  We  thank  the  Coordi- 
nacion  de  Investigation  Cientifica  and  Facultad  de 
Biologfa  at  Universidad  Michoacana  de  San  Nicolas 
de  Hidalgo  for  their  continued  logistical  and  economic 
support,  and  the  Direction  General  de  Vida  Silvestre 
at  SEMARNAT,  which  provided  research  authoriza- 
tions. We  also  thank  J.  W.  Wiley  and  two  anonymous 
reviewers  who  made  important  suggestions  that  im- 
proved this  manuscript. 

LITERATURE  CITED 

Benkman,  C.  W.  1993.  Logging,  conifers,  and  the  con- 
servation of  crossbills.  Conservation  Biology  7: 
473-479. 

Danchin,  E.  and  R.  H.  Wagner.  1997.  The  evolution 
of  coloniality:  the  emergence  of  new  perspectives. 
Trends  in  Ecology  and  Evolution  12:342-347. 
Eberhard,  J.  R.  2002.  Cavity  adoption  and  the  evo- 
lution of  coloniality  in  cavity  nesting  birds.  Con- 
dor 104:240-247. 

Enkerlin-Hoeflich,  E.  C.  1995.  Comparative  ecology 
and  reproductive  biology  of  three  species  of  Ama- 


zona  parrots  in  northeastern  Mexico.  Ph.D.  dis- 
sertation, Texas  A&M  University,  Kingsville. 

Forshaw,  J.  M.  1989.  Parrots  of  the  world,  3rd  (revised) 
ed.  Lansdowne  Editions,  Melbourne,  Australia. 

Gibbs,  J.  R,  M.  L.  Hunter,  Jr.,  and  S.  M.  Melvin. 
1993.  Snag  availability  and  communities  of  cavity 
nesting  birds  in  tropical  versus  temperate  forests. 
Biotropica  25:236-240. 

Gill,  F.  B.  1990.  Ornithology,  2nd  ed.  W.  H.  Freeman, 
New  York. 

Hardy,  J.  W.  1963.  Epigamic  and  reproductive  behav- 
ior of  the  Orange-fronted  Parakeet.  Condor  65: 
169-199. 

Howell,  S.  N.  G.  and  S.  Webb.  1995.  A guide  to  the 
birds  of  Mexico  and  northern  Central  America. 
Oxford  University  Press,  New  York. 

Inigo-Elias,  E.  1996.  Ecology  and  breeding  biology 
of  the  Scarlet  Macaw  ( Ara  macao)  in  the  Usu- 
macinta  drainage.  Ph.D.  dissertation.  University  of 
Florida,  Gainsville. 

Lammertink,  J.  M.,  J.  A.  Rojas-Tome,  F.  M.  Casillas- 
Orona,  and  R.  L.  Otto.  1996.  Status  and  con- 
servation of  old-growth  forests  and  endemic  birds 
in  the  pine-oak  zone  of  the  Sierra  Madre  Occi- 
dental, Mexico.  Institute  for  Systematics  and  Pop- 
ulation Biology  (Zoological  Museum),  University 
of  Amsterdam,  Amsterdam,  Netherlands. 

Lanning,  D.  V.  and  J.  T.  Shiflett.  1983.  Nesting  ecol- 
ogy of  Thick-billed  Parrots.  Condor  85:66-73. 

Macias-Caballero,  C.  1998.  Comportamiento  de  an- 
idacion  y monitoreo  de  la  productividad  de  la  Co- 
torra  Serrana  Oriental  ( Rhynchopsitta  terrisi)  en  el 
norte  de  Mexico  e implicaciones  para  su  conser- 
vation. M.Sc.  thesis,  Centro  de  Calidad  Ambien- 
tal, Instituto  Tecnologico  y de  Estudios  Superiores 
de  Monterrey,  Monterrey,  Nuevo  Leon,  Mexico. 

Monterrubio-Rico,  T.  and  E.  C.  Enkerlin-Hoeflich. 
2004a.  Present  use  and  characteristics  of  Thick- 
billed Parrot  nest  sites  in  northwestern  Mexico. 
Journal  of  Field  Ornithology  75:96-103. 

Monterrubio-Rico,  T.  C.  and  E.  C.  Enkerlin-Hoe- 
flich. 2004b.  Variation  anual  en  la  actividad  de 
anidacion  y productividad  de  la  Cotorra  Serrana 
Occidental  ( Rhynchopsitta  pachyrhyncha).  Anales 
del  Instituto  de  Biologfa,  Universidad  Nacional 
Autonoma  de  Mexico,  serie  Zoologfa  75:341-354. 

Munn,  C.  A.  1992.  Macaw  biology  and  ecotourism,  or 
“When  a bird  in  the  bush  is  worth  two  in  the 
hand.”  Pages  47-72  in  New  World  parrots  in  cri- 
sis (S.  R.  Beissinger  and  N.  F.  R.  Snyder,  Eds.). 
Smithsonian  Institution  Press,  Washington,  D.C. 

Renton,  K.  and  A.  Salinas-Melgoza.  1999.  Nesting 
behavior  of  the  Lilac-crowned  Parrot.  Wilson  Bul- 
letin 111:488-493. 

Rodriguez-Estrella,  R.,  L.  Rivera-Rodriguez,  and 
F.  Anguiano.  1995.  Nest-site  characteristics  of  the 
Socorro  Green  Parakeet.  Condor  97:575-577. 

Sas  Institute,  Inc.  1985.  SAS/STAT:  guide  for  per- 
sonal computers,  ver.  6.  SAS  Institute,  Inc.,  Cary, 
North  Carolina. 

Siegel-Causey,  D.  and  S.  P.  Kharitonov.  1990.  The 


Monterrubio-Rico  et  al.  • THICK-BILLED  PARROT  GREGARIOUS  NESTING 


243 


evolution  of  coloniality.  Current  Ornithology  7: 
285-330. 

Snyder,  N.  F.  R.,  E.  C.  Enkerlin-Hoeflich,  and  M. 
A.  Cruz-Nieto.  1999.  Thick-billed  Parrot  ( Rhyn - 
chopsitta  pachyrhyncha ).  The  Birds  of  North 
America,  no.  411. 


Snyder,  N.  F.  R.,  S.  E.  Koenig,  J.  Koschman,  H.  A. 
Snyder,  and  T.  B.  Johnson.  1994.  Thick-billed 
Parrot  releases  in  Arizona.  Condor  96:845-862. 
Ward,  P.  and  A.  Zahavi.  1973.  The  importance  of 
certain  assemblages  of  birds  as  information  cen- 
ters for  food  finding.  Ibis  115:517-534. 


Short  Communications 


The  Wilson  Journal  of  Ornithology  1 18(2):244- 247,  2006 

No  Extra-pair  Fertilization  Observed  in  Nazca  Booby 
{Sula  granti ) Broods 

David  J.  Anderson134  and  Peter  T.  Boag1 2 3 4 


ABSTRACT. — Nazca  Booby  ( Sula  granti)  broods  in 
the  Galapagos  Islands  showed  0%  extra-pair  fertiliza- 
tion, based  on  multilocus  band-sharing  values.  The 
95%  Cl  of  this  estimate  for  all  chicks  was  0-0.098, 
and  for  all  broods  it  was  0-0.139.  These  are  the  first 
data  on  extra-pair  paternity  to  be  reported  for  a mem- 
ber of  the  family  Sulidae.  Received  6 September  2005, 
accepted  22  February  2006. 


The  frequency  of  extra-pair  paternity  (EPP) 
among  bird  species  varies  widely,  from  0%  in 
some  seabirds  (e.g.,  Chinstrap  Penguins,  Py- 
goscelis  antarctica ; Moreno  et  al.  2000),  the 
Acorn  Woodpecker  (Melanerpes  formicivo- 
rus\  Dickinson  et  al.  1995,  Haydock  et  al. 

2001) ,  and  other  taxa  (Griffith  et  al.  2002)  to 
72%  in  Superb  Fairy-wrens  ( Malurus  cy- 
aneus\  Mulder  et  al.  1994,  Double  and  Cock- 
burn  2000).  Application  of  new  molecular  ge- 
netic techniques  has  enabled  the  recent  explo- 
sion in  availability  of  parentage  data  from 
birds,  and  estimates  of  EPP  exist  for  at  least 
186  species  in  at  least  39  families  (Griffith  et 
al.  2002,  Spottiswoode  and  Mpller  2004).  Of 
particular  interest  is  the  minority  (25%)  of  so- 
cially monogamous  taxa  in  which  EPP  is  ab- 
sent, or  nearly  so  (Griffith  et  al.  2002).  In  de- 
parting from  the  general  trend  in  birds,  these 
taxa  may  experience  selection  pressures,  or 
phylogenetic  constraints,  that  differ  from 
those  of  most  species,  and  they  can  provide 
insight  into  the  evolution  of  the  vast  diversity 
of  mating  systems  in  birds.  The  majority  of 
the  diversity  in  EPP  frequency  is  at  or  above 
the  family  level  in  birds  (Arnold  and  Owens 

2002) ;  thus,  comparative  analyses  (Bennett 


1 Dept,  of  Avian  Sciences,  Univ.  of  California,  Da- 
vis, CA  95616,  USA. 

2 Dept,  of  Biology,  Queens  Univ.,  Kingston,  ON 
K7L  3N6,  Canada. 

3 Current  address:  Dept,  of  Biology,  Wake  Forest 
Univ.,  Winston-Salem,  NC  27109,  USA. 

4 Corresponding  author;  e-mail:  da@wfu.edu 


and  Owens  2002,  Westneat  and  Stewart  2003) 
require  data  from  as  many  higher-order  taxa 
as  possible. 

Here,  we  present  parentage  data  from  Naz- 
ca Boobies  {Sula  granti ),  a socially  monoga- 
mous seabird  in  the  family  Sulidae,  for  which 
published  data  on  EPP  frequency  is  lacking. 
While  Nazca  Boobies  exhibit  life-history 
characteristics  associated  with  a low  EPP  rate 
(Bennett  and  Owens  2002) — such  as  long  life, 
extended  parental  care,  and  small  broods  (An- 
derson 1993,  Anderson  and  Apanius  2003) — 
they  nest  colonially  in  the  presence  of  many 
potential  copulatory  partners  (Nelson  1978), 
females  spend  extended  periods  unattended  in 
the  colony  while  their  mate  forages  at  sea,  and 
they  have  unusually  low  hatching  success 
(60%)  due  to  infertility  or  early  embryo  death 
(Anderson  1990).  The  low  hatching  success 
could  be  due  to  low  sperm  quality  in  some 
males,  which  might  induce  females  to  select 
for  insurance  sperm  outside  the  pair  bond,  al- 
though other  aspects  of  their  life  history  sug- 
gest that  EPP  should  be  rare. 

In  1990,  we  studied  Nazca  Boobies  breed- 
ing at  the  large  colony  at  Punta  Cevallos,  Isla 
Espanola,  Galapagos  Islands,  Ecuador  (1°20' 
S,  89°  40'  W).  Huyvaert  and  Anderson  (2004) 
give  details  of  the  study  site.  We  collected 
blood  samples  from  10  single-chick  broods 
(January  1990,  with  unknown  initial  clutch 
and  brood  sizes)  and  13  two-chick  broods 
(December  1990)  and  their  social  parents 
(adults  that  brooded  the  young).  In  this  pop- 
ulation, clutch  size  is  either  one  or  two  (An- 
derson 1990);  hence,  the  December  sample  al- 
most certainly  represents  complete  families, 
but  we  are  not  certain  about  initial  clutch  or 
brood  size  of  the  January  sample.  Single- 
chick broods  in  the  first  sampling  effort  were 
the  products  of  single  chicks  from  one-  or 
two-egg  clutches,  or  the  survivor  of  obligate 
siblicide  (almost  always  the  first-hatched 


244 


SHORT  COMMUNICATIONS 


245 


chick)  in  a two-chick  brood  (Humphries  et  al. 
2006).  Two-chick  broods  were  targeted  in  the 
second  sampling  effort  to  determine  whether 
siblicide  masked  a high  EPP  rate  in  second- 
hatched  chicks.  Families  were  chosen  ran- 
domly from  across  the  colony  and  had  typical 
distances  to  neighboring  sites  (mean  = 2.7  m 
± 1.55  SD;  Anderson  1993).  Using  syringes, 
we  drew  blood  samples  from  the  brachial  vein 
and  transferred  them  to  vacutainers.  Blood 
was  stored  in  Queens  lysis  buffer  (Seutin  et 
al.  1991)  at  ambient  temperature  in  the  field 
and  later  at  4°  C.  DNA  was  extracted  from  the 
blood  samples  following  the  procedures  of 
Seutin  et  al.  (1991).  After  testing  various  com- 
binations of  restriction  enzymes  and  multilo- 
cus probes  for  quality  and  quantity  of  bands, 
all  booby  DNA  was  cut  with  Mbo  I and  hy- 
bridized with  radioactively  labeled  minisatel- 
lite probes  33.15  (Jeffreys  et  al.  1985)  and  per 
(Shin  et  al.  1985).  Electrophoresis,  Southern 
blotting,  and  prehybridization  followed  Smith 
et  al.  (1991),  except  that  we  used  5 pg  of 
DNA  per  sample,  and  Immobilon  (Millipore) 
transfer  membranes.  Transfer  membranes 
were  hybridized,  washed,  and  autoradio- 
graphed  following  Smith  et  al.  (1991),  except 
that  the  membranes  were  washed  in  2X  SSC, 
0.1%  SDS.  After  probing  with  both  minisa- 
tellites, the  membranes  were  probed  a third 
time  with  lambda  DNA  to  reveal  lambda  size 
markers  in  each  lane  to  facilitate  scoring  of 
homologous  fragments  in  different  lanes. 

We  assessed  parentage  of  nestlings  by  com- 
paring bands  in  the  2-  to  12-kb  range  of  nest- 
lings with  those  of  their  putative  parents  on  the 
autoradiographs.  Bands  were  scored  by  mark- 
ing acetate  sheets,  using  different  colors  for 
maternally  and  paternally  derived  bands.  Bands 
were  considered  identical  if  their  centers  were 
less  than  1 mm  apart  and  they  did  not  differ 
greatly  in  density.  We  calculated  the  degree  of 
band-sharing  between  putative  parents  and  off- 
spring to  determine  whether  we  could  exclude 
a parent  and,  if  so,  which  one.  Band-sharing 
(D)  was  calculated  as  D = 2 {nAB)/{nA  + «B)> 
where  nAB  is  the  number  of  bands  shared  by 
birds  A and  B,  and  nA  and  nB  are  the  number 
of  bands  in  birds  A and  B,  respectively  (Wetton 
et  al.  1987).  Both  Jeffreys  33.15  and  per  probes 
produced  DNA  fingerprints  similar  to  those  de- 
scribed for  other  bird  species.  Band-sharing 
was  calculated  as  the  mean  of  the  D values  of 


the  two  probes.  We  used  the  band-sharing  of 
mates  as  an  estimate  of  the  band-sharing  of 
unrelated  birds,  and  used  that  estimate  to  eval- 
uate the  relationship  of  putative  parents  and 
their  offspring.  Because  the  probability  of  a 
brood  having  two  chicks  increased  with  in- 
creasing band-sharing  values  (logistic  regres- 
sion, x2  = 4.34,  df  = 1,  P = 0.037),  we  eval- 
uated the  January  1990  and  December  1990 
groups  separately;  generally,  band-sharing  was 
greater  in  the  December  sample,  for  all  pair- 
wise analyses  of  family  members.  Because 
there  was  no  a priori  difference  in  how  fami- 
lies were  selected  in  the  two  sampling  periods, 
and  given  that  the  two  sets  of  DNA  fingerprints 
were  prepared  by  different  lab  workers  at  dif- 
ferent times,  the  difference  in  average  similar- 
ity values  may  have  resulted  from  methodolog- 
ical differences,  not  biological  differences  in 
relatedness. 

In  one-chick  broods,  band-sharing  of  mated 
pairs  averaged  0.330  ± 0.115  SD;  band-shar- 
ing between  offspring  and  mothers  averaged 
0.617  ± 0.076  and  that  between  offspring  and 
fathers  averaged  0.625  ± 0.066.  The  smallest 
band-sharing  value  between  an  offspring  and 
a parent  (0.533)  exceeded  the  largest  value  of 
band-sharing  in  mated  pairs  (0.527;  Fig.  1). 
These  non-overlapping  distributions  provide 
no  indication  of  extra-pair  parentage  in  one- 
chick  broods. 

In  two-chick  broods,  band-sharing  of  mated 
pairs  averaged  0.418  ± 0.106;  band-sharing 
between  offspring  and  mothers  averaged 
0.699  ± 0.073,  and  between  offspring  and  fa- 
thers it  averaged  0.680  ± 0.117.  Although 
four  band-sharing  values  between  an  offspring 
and  a parent  were  less  than  the  largest  value 
of  band-sharing  in  mated  pairs  (0.564;  Fig.  1), 
they  do  not  provide  reliable  evidence  of  extra- 
pair parentage.  In  three  cases  (indicated  by  pa- 
rentheses in  Fig.  1),  DNA  degradation  in  pa- 
rental samples  caused  low  band-sharing  val- 
ues between  all  members  of  the  family,  in- 
cluding between  the  mated  adults.  The 
offspring  in  this  family  displayed  no  unattrib- 
utable  bands,  indicating  that  the  putative  par- 
ents were  in  fact  the  genetic  parents.  This  was 
the  lone  instance  of  poor  quality  DNA  among 
our  samples.  In  the  remaining  two  cases  (in- 
dicated by  the  rotated  parentheses  in  Fig.  1), 
the  identity  of  the  social  father  was  questioned 
on  behavioral  grounds  after  blood  samples 


246 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


FIG.  1 . Distribution  of  band-sharing  values  in 
Nazca  Booby  broods,  expressed  as  cumulative  per- 
centages from  lowest  to  highest  values.  Dotted  vertical 
lines  show  the  maximum  band-sharing  value  for  un- 
related adults  (mated  pairs).  Values  in  parentheses  are 
the  result  of  either  poor  quality  DNA  (•)  or  uncertain 
parentage  S.  Blood  samples  were  collected  in  January 
and  December  1990  at  Punta  Cevallos,  Isla  Espanola, 
Galapagos,  Ecuador. 

had  been  taken  from  the  two  offspring  and 
two  adults  present  at  the  nest  site.  At  the  time 
of  sampling,  we  had  observed  a male  standing 
near  the  female  and  offspring,  and  assumed 
that  he  was  the  social  father;  on  subsequent 
days,  however,  another  male  consistently  at- 
tended this  brood  and  the  original  male  in- 
stead appeared  to  be  a neighbor.  We  were  not 
able  to  obtain  a blood  sample  from  the  other 
putative  father.  This  was  the  one  instance  in 
which  family  membership  was  uncertain. 
Omitting  these  two  families  from  consider- 
ation, all  band-sharing  values  of  offspring  and 
putative  parents  exceeded  the  largest  band- 
sharing value  of  mated  adults  in  two-chick 
broods  (0.564;  Fig.  1).  Excluding  the  four 
chicks  of  the  two  questionable  broods,  our  es- 
timate of  EPP  frequency  in  the  32  chicks  was 
0 (95%  Cl  = 0-0.109),  and  in  the  21  broods 
it  was  also  0 (95%  Cl  = 0-0.162). 

This  low  EPP  frequency  of  Nazca  Boobies 
conforms  to  the  expectations  based  on  empir- 
ical data  from  other  long-lived  seabirds  (Grif- 
fith et  al.  2002)  and  theoretical  considerations 
of  the  likely  selection  forces  acting  on  such 
species  (Mauck  et  al.  1999,  Bennett  and 


Owens  2002,  Westneat  and  Stewart  2003).  It 
also  matches  behavioral  data  showing  that  fe- 
male boobies  cooperate  with  extra-pair  males 
in  permitting  extra-pair  copulation;  during  the 
8 days  preceding  egg-laying,  however,  they 
engage  almost  exclusively  in  within-pair  cop- 
ulations (DJA  unpubl.  data).  Thus,  while  ex- 
tra-pair copulation  (EPC)  is  common  in  both 
the  Nazca  Booby  (61%  of  females  had  >1 
EPC;  DJA  unpubl.  data)  and  the  related  Blue- 
footed Booby  ( S . nebouxir,  Osorio-Beristain 
and  Drummond  1998),  EPP  is  not  (see  also 
Hunter  et  al.  1992,  Schwartz  et  al.  1999).  The 
benefits,  if  any,  of  EPC  to  females  appear  un- 
related to  any  genetic  benefits,  such  as  fertil- 
ization insurance  that  could  result  from  ob- 
taining extra-pair  sperm.  This  intriguing  dis- 
parity between  a high  frequency  of  EPC  and 
a low  rate  of  EPP  places  the  Nazca  Booby  in 
an  unusual  position  in  the  spectrum  of  avian 
mating  systems  that  merits  further  study. 

ACKNOWLEDGMENTS 

We  thank  the  Servicio  Parque  Nacional  Galapagos 
for  permission  to  conduct  this  study,  the  Charles  Dar- 
win Research  Station  and  TAME  airline  for  logistical 
support,  I.  von  Lippke  and  P.  R.  Sievert  for  field  as- 
sistance, and  D.  F.  Westneat  and  two  anonymous  re- 
viewers for  helpful  comments.  The  University  of  Cal- 
ifornia Agricultural  Experiment  Station  and  the  Natu- 
ral Science  and  Engineering  Research  Council  provid- 
ed financial  support. 

LITERATURE  CITED 

Anderson,  D.  J.  1990.  Evolution  of  obligate  siblicide 
in  boobies.  1.  A test  of  the  insurance-egg  hypoth- 
esis. American  Naturalist  135:334-350. 
Anderson,  D.  J.  1993.  Masked  Booby  ( Sula  dactyla- 
tra).  The  Birds  of  North  America,  no.  73. 
Anderson,  D.  J.  and  V.  Apanius.  2003.  Actuarial  and 
reproductive  senescence  in  a long-lived  seabird: 
preliminary  evidence.  Experimental  Gerontology 
38:757-760. 

Arnold,  K.  E.  and  I.  P.  F.  Owens.  2002.  Extra-pair 
paternity  and  egg  dumping  in  birds:  life  history, 
parental  care  and  the  risk  of  retaliation.  Proceed- 
ings of  the  Royal  Society  of  London,  Series  B 
269:1263-1269. 

Bennett,  P.  M.  and  I.  P.  F.  Owens.  2002.  Evolutionary 
ecology  of  birds:  life  histories,  mating  systems  and 
extinction.  Oxford  University  Press,  New  York. 
Dickinson,  J.,  J.  Haydock,  W.  Koenig,  M.  Stanback, 
and  F.  Pitelka.  1995.  Genetic  monogamy  in  sin- 
gle-male groups  of  Acorn  Woodpeckers,  Melaner- 
pes  formicivorus.  Molecular  Ecology  4:765—769. 
Double,  M.  and  A.  Cockburn.  2000.  Pre-dawn  infi- 
delity: females  control  extra-pair  mating  in  Superb 


SHORT  COMMUNICATIONS 


247 


Fairy-wrens.  Proceedings  of  the  Royal  Society  of 
London,  Series  B 267:465-470. 

Griffith,  S.  C.,  I.  P.  F.  Owens,  and  K.  A.  Thurman. 
2002.  Extra  pair  paternity  in  birds:  a review  of 
interspecific  variation  and  adaptive  function.  Mo- 
lecular Ecology  1 1:2195-2212. 

Haydock,  J.,  W.  D.  Koenig,  and  M.  T.  Stanback. 
2001.  Shared  parentage  and  incest  avoidance  in 
the  cooperatively  breeding  Acorn  Woodpecker. 
Molecular  Ecology  10:1515-1525. 

Humphries,  C.  A.,  V.  D.  Arevalo,  K.  N.  Fischer,  and 
D.  J.  Anderson.  2006.  Contributions  of  marginal 
offspring  to  reproductive  success  of  Nazca  Booby 
(Sula  granti)  parents:  tests  of  multiple  hypotheses. 
Oecologia  147:379-390. 

Hunter,  F.  M.,  T.  Burke,  and  S.  E.  Watts.  1992.  Fre- 
quent copulation  as  a method  of  paternity  assur- 
ance in  the  Northern  Fulmar.  Animal  Behaviour 
44:149-156. 

Huyvaert,  K.  P.  and  D.  J.  Anderson.  2004.  Limited 
dispersal  in  the  Nazca  Booby.  Journal  of  Avian 
Biology  35:46-53. 

Jeffreys,  A.  J.,  V.  Wilson,  and  S.  L.  Thein.  1985. 
Hypervariable  ‘minisatellite’  regions  in  human 
DNA.  Nature  314:67-73. 

Mauck,  R.  A.,  E.  A.  Marschall,  and  P.  G.  Parker. 
1999.  Adult  survival  and  imperfect  assessment  of 
parentage:  effects  on  male  parenting  decisions. 
American  Naturalist  154:99-109. 

Moreno,  J.,  L.  Boto,  J.  A.  Fargallo,  A.  de  Leon, 
and  J.  Potti.  2000.  Absence  of  extra-pair  fertil- 
isations in  the  Chinstrap  Penguin  Pygoscelis  ant- 
arctica.  Journal  of  Avian  Biology  31:580-583. 

Mulder,  R.  A.,  P.  O.  Dunn,  R.  A.  Cockburn,  K.  A. 
Lazenby-Cohen,  and  M.  J.  Howell.  1994.  Help- 
ers liberate  female  fairy-wrens  from  constraints  on 


extra-pair  mate  choice.  Proceedings  of  the  Royal 
Society  of  London,  Series  B 255:223-229. 

Nelson,  J.  B.  1978.  The  Sulidae:  gannets  and  boobies. 
Oxford  University  Press,  Oxford,  United  Kingdom. 

Osorio-Beristain,  M.  and  H.  Drummond.  1998.  Non- 
aggressive  mate  guarding  by  the  Blue-footed  Boo- 
by: a balance  of  female  and  male  control.  Behav- 
ioral Ecology  and  Sociobiology  43:307-315. 

Schwartz,  M.  K.,  D.  J.  Boness,  C.  M.  Schaeff,  P. 
Majluf,  E.  A.  Perry,  and  R.  C.  Fleischer.  1999. 
Female-solicited  extrapair  matings  in  Humboldt 
Penguins  fail  to  produce  extrapair  fertilizations. 
Behavioral  Ecology  10:242-250. 

Seutin,  G.,  B.  N.  White,  and  P.  T.  Boag.  1991.  Pres- 
ervation of  avian  blood  and  tissue  samples  for 
DNA  analyses.  Canadian  Journal  of  Zoology  69: 
82-90. 

Shin,  H.-S.,  T.  A.  Bargiello,  B.  T.  Clark,  F.  R.  Jack- 
son,  and  M.  W.  Young.  1985.  An  unusual  coding 
sequence  from  a Drosophila  clock  gene  is  con- 
served in  vertebrates.  Nature  317:445-448. 

Smith,  H.  G.,  R.  Montgomerie,  T.  Poldmaa,  B.  N. 
White,  and  P.  T.  Boag.  1991.  DNA  fingerprinting 
reveals  relation  between  tail  ornaments  and  cuck- 
oldry  in  Barn  Swallows,  Hirundo  rustica.  Behav- 
ioral Ecology  2:90-98. 

Spottiswoode,  C.  and  A.  P.  M0ller.  2004.  Extrapair 
paternity,  migration,  and  breeding  synchrony  in 
birds.  Behavioral  Ecology  15:41-57. 

Westneat,  D.  F.  and  I.  R.  K.  Stewart.  2003.  Extra- 
pair paternity  in  birds:  causes,  correlates,  and  con- 
flict. Annual  Review  of  Ecology  and  Systematics 
34:365-396. 

Wetton,  J.  H.,  R.  E.  Carter,  D.  T.  Paricin,  and  D. 
Walters.  1987.  Demographic  study  of  a wild 
House  Sparrow  population  by  DNA  fingerprint- 
ing. Nature  327:147-149. 


The  Wilson  Journal  of  Ornithology  1 18(2):247-251,  2006 

Golden-cheeked  Warbler  Males  Participate  in  Nest-site  Selection 

Allen  E.  Graber,134  Craig  A.  Davis,1  and  David  M.  Leslie,  Jr.2 3 4 


ABSTRACT. — Nest-site  selection  behaviors  have 
rarely  been  described  for  songbirds.  Furthermore,  male 
involvement  in  nest-site  selection  is  generally  assumed 
to  be  minimal  among  most  species,  especially  those 


1 Dept,  of  Zoology,  Oklahoma  State  Univ.,  430  Life 
Sciences  West,  Stillwater,  OK  74078,  USA. 

2 Oklahoma  Coop.  Fish  and  Wildlife  Research  Unit, 
U.S.  Geological  Survey,  Oklahoma  State  Univ.,  404 
Life  Sciences  West,  Stillwater,  OK  74078,  USA. 

3 Current  address:  Arizona  Game  and  Fish  Dept., 
Research  Branch,  2221  W.  Greenway  Rd.,  Phoenix, 
AZ  85023,  USA. 

4 Corresponding  author;  e-mail:  agraber@azgfd.gov 


predominantly  exhibiting  female  nest  building.  This 
assumption  has  held  true  for  the  federally  endangered 
Golden-cheeked  Warbler  ( Dendroica  chrysoparia ),  a 
breeding  resident  of  central  Texas.  We  observed  Gold- 
en-cheeked Warbler  males  and  females  searching  for 
nest  sites  together  on  three  separate  occasions,  2001- 
2003.  Although  rare,  such  observations  add  to  our 
knowledge  of  the  life  history  of  songbirds.  Received 
20  April  2005,  accepted  11  January  2006. 


For  a breeding  pair  of  birds,  the  nest-site 
selection  process  can  be  a critical  step  in  es- 


248 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


tablishing  a pair  bond;  certainly,  the  site  se- 
lected will  often  affect  the  pair’s  reproductive 
success  (Martin  1998).  The  final  choice  of 
nest  placement,  whether  made  by  the  male, 
female,  or  both,  will  likely  be  influenced  by 
several  factors  (e.g.,  local  resource  availability 
[presence  of  nesting  materials,  food],  inter- 
and  intraspecific  competition,  and  habitat  fea- 
tures influencing  microclimate,  brood  parasit- 
ism or  predation)  that  contribute  to  the  quality 
and  quantity  of  fledglings  reared  (Martin  and 
Roper  1988).  Recent  literature  on  this  topic 
has  focused  on  gaining  a better  understanding 
of  the  relationship  between  nest  placement 
and  predation  (e.g.,  Wilson  and  Cooper  1998, 
Siepielski  et  al.  2001,  Boulton  et  al.  2003,  Da- 
vis 2005) — the  leading  cause  of  reproductive 
failure  in  birds  and  a significant  selective 
force  on  avian  breeding  behaviors  (Ricklefs 
1969,  Martin  1992).  Far  less  attention  has 
been  given  to  how  birds  actually  select  a site. 

Information  on  the  behavioral  processes  in- 
volved in  nest-site  selection  for  wood  war- 
blers, including  the  federally  endangered 
Golden-cheeked  Warbler  ( Dendroica  chryso- 
paria),  is  generally  lacking  (Morse  1989, 
Ladd  and  Gass  1999).  In  a review  of  The 
Birds  of  North  America  series,  we  found  that 
information  on  the  nest-site  selection  process 
is  well  described  for  only  15  of  the  51  wood 
warblers  (families  Parulidae  and  Peucedrami- 
dae).  Furthermore,  among  species  predomi- 
nantly exhibiting  female  nest  building,  the 
role  of  the  male  in  nest-site  selection  is  often 
assumed  to  be  minor  (Kaufman  1996,  Ladd 
and  Gass  1999).  With  few  exceptions  (see 
Ficken  1964,  Meanley  1971,  Nolan  1978), 
males  have  been  observed  only  mate-guarding 
and  singing  subdued,  infrequent  songs,  while 
females  actively  engage  in  nest-site  selection 
activities  (Pulich  1976,  Guzy  and  Lowther 
1997,  Wright  et  al.  1998). 

Few  data  exist  on  the  nest-site  selection 
processes  of  Golden-cheeked  Warblers  (Ladd 
and  Gass  1999),  although  aspects  of  their 
breeding  biology  and  nesting  characteristics 
have  been  described  in  detail  (Bent  1953,  Pul- 
ich 1976,  Ladd  and  Gass  1999).  The  Golden- 
cheeked Warbler  is  a habitat  specialist  with  a 
limited  range.  Its  nesting  habitats  are  closed- 
canopy,  low-growing  woodlands  dominated 
by  mature  Ashe  juniper  ( Juniperus  ashei)  and 
oaks  ( Quercus  spp.;  Ladd  and  Gass  1999). 


Such  habitats  are  restricted  to  limestone 
slopes,  canyons,  and  adjacent  uplands  in  the 
Edwards  Plateau  and  Llano  Uplift  of  central 
Texas  (Pulich  1976,  Kier  et  al.  1977).  Nests 
are  constructed  by  the  female  with  strips  of 
mature  Ashe  juniper  bark  and  are  typically 
placed  in  Ashe  junipers,  but  sometimes  in 
oaks  or  other  hardwoods.  Nests  are  usually 
located  in  the  upper  two-thirds  of  a tree,  av- 
eraging 5-7  m above  ground  (Pulich  1976). 

In  the  only  comprehensive  study  of  Golden- 
cheeked Warblers,  Pulich  (1976)  wrote  that 
the  male  might  accompany  the  female  in  her 
search  for  a nesting  site.  He  described  an  ob- 
servation made  on  1 April  1961,  in  which  a 
female — paying  no  attention  to  her  mate — 
flew  to  the  ground  and  picked  at  unidentified 
objects,  briefly  investigated  an  old  nest  in  a 
juniper,  and  flew  across  a ravine  to  another 
tree;  the  male  guarded  his  mate,  actively 
chased  an  approaching  satellite  male,  and  sang 
infrequently.  Pulich  (1976)  concluded  that  the 
female  chooses  the  nest  site,  but  he  gave  no 
description  of  the  behavioral  repertoire  in- 
volved in  her  selection  of  the  site.  Pulich 
(1976:82)  did  acknowledge  that  he  had  likely 
missed  some  sexual  displays  that  play  a role 
in  establishing  the  pair  bond  because  “the 
courtship  of  the  Golden-cheeked  Warbler 
seems  to  be  carried  on  in  utmost  secrecy.”  In 
another  study,  Golden-cheeked  Warbler  males 
were  observed  presenting  strips  of  juniper 
bark  to  their  mates,  but  courtship  displays 
were  not  observed  prior  to  nest  building 
(Lockwood  1996). 

Here,  we  document  male  and  female  Gold- 
en-cheeked Warblers  actively  searching  for 
nest  sites  together  on  Fort  Hood,  an  active 
U.S.  Army  installation  in  Bell  and  Coryell 
counties,  Texas  (31°  10'  N,  97°  45'  W).  We  re- 
corded these  events  during  a 3-year  study  in- 
volving detailed  behavioral  observations  of 
color-banded  Golden-cheeked  Warbler  males. 

On  2 April  2003  at  13:15  CST,  a Golden- 
cheeked Warbler  male  was  heard  singing  the 
“A-song,”  a song-type  associated  with  male- 
female  interactions  (Bolsinger  2000).  The  pair 
was  observed  displaying  nest-site  trying  be- 
haviors (Ficken  1964)  in  several  tree  forks 
within  a cluster  of  shin  oaks  ( Q . sinuata ).  Try- 
ing behaviors  were  characterized  by  both  the 
male  and  the  female  squatting  simultaneously 
or  alternately  in  potential  nearby  sites  while 


SHORT  COMMUNICATIONS 


249 


vigorously  pivoting  clockwise  and  counter- 
clockwise. Pivots,  consisting  of  half-rotations 
(180°)  and  up  to  two  full  rotations  (720°),  in- 
cluded outward  and  downward  extension  of 
wings  and  upward  elevation  of  the  tail.  Ex- 
tension of  the  limbs  may  have  provided  tactile 
information  about  the  suitability  of  the  site 
(Nolan  1978). 

Interruptions  to  trying  pivots  included 
pressing  the  breast,  belly,  and  sides  against 
limbs  as  if  “nest-shaping,”  attentively  exam- 
ining the  site,  or  hopping  to  other  prospective 
sites  (all  within  the  same  shin  oak  cluster).  At 
times,  the  female  appeared  to  gather  infor- 
mation from  her  “advertising”  mate  and  re- 
sponded to  his  trying  behavior  by  approaching 
the  potential  nest  site  as  soon  as  he  left.  In 
general,  female  nest-site  inspection  behaviors 
seemed  to  be  more  persistent  than  those  of  her 
mate,  who  infrequently  sang  a muted  A-song, 
exhibited  mate-guarding  behavior,  and  paused 
more  often.  These  activities  lasted  —180  sec. 

On  4 April  2003  at  1 1:05,  we  observed  the 
same  female  collecting  juniper  strips  and  then 
flying  to  a nest  under  construction,  24  m away 
from  the  previously  observed  trying  location 
and  4.5  m above  ground  in  the  outer  branch 
of  an  Ashe  juniper.  The  female  appeared  to  be 
in  her  1st  day  of  nest  construction,  as  a nest 
platform  was  beginning  to  take  shape.  There 
was  no  sign  of  her  mate  at  that  time. 

Similar  nest-site  trying  behaviors  were  re- 
corded on  two  separate  occasions — one  in 
2001  (1  April  at  13:12)  and  one  in  2002  (29 
March  at  10:47).  In  each  case,  we  observed 
females  in  the  initial  phases  of  nest-building 
3 days  following  our  observations  of  trying 
behaviors.  These  nest-site  selection  activities 
differed  somewhat  from  those  observed  in 
2003  with  respect  to  the  observation  duration 
(estimated  mean  for  both  observations  = 70 
sec),  the  degree  of  male  participation  (less  in 
2001;  fewer  pivot  maneuvers  in  2002,  but  a 
similar  proportion  of  time  spent  hopping  to 
prospective  sites),  the  tree  species  in  which 
nest-site  trying  took  place  (Ashe  juniper  in 
2001  and  2002),  and  the  distance  between  the 
nest-site  trying  site  and  the  actual  nest  site 
(mean  distance  for  both  observations  = 23  m). 

Although  detailed  information  on  the  be- 
havioral processes  of  nest-site  selection  is 
rare,  trying  or  sizing  prospective  nest  sites — 
by  examining  the  site,  squatting,  depressing 


the  sternal  region,  nest-shaping,  pivoting,  el- 
evating the  tail,  and  extending  the  feet  and 
wings — is  common  among  several  warbler 
species  (e.g.,  American  Redstart  [Setophaga 
ruticilla ; Ficken  1964],  Cerulean  Warbler 
[Dendroica  cerulea\  Oliarnyk  and  Robertson 
1996],  Prairie  Warbler  [D.  discolor,  Nolan 
1978],  and  Swainson’s  Warbler  [ Limnothlypis 
swainsonii ; Meanley  1971]).  Reports  of  war- 
bler males  participating  in  these  activities, 
however,  are  highly  unusual  (Morse  1989). 

Our  three  observations  of  trying  behaviors 
constitute  the  only  such  behaviors  we  wit- 
nessed during  our  study,  and  we  did  not  ob- 
serve males  or  females  performing  trying  ac- 
tivities on  their  own.  In  another  study,  Nolan 
(1978)  found  that  male  Prairie  Warblers  be- 
haved very  much  like  their  mates  in  10%  of 
about  300  observations.  In  the  other  90%  of 
Nolan’s  observations,  the  male  followed  and 
watched  the  female,  performed  display  flights, 
and  sang  irregular,  muted  songs.  Similarly, 
American  Redstart  males  have  been  observed 
only  occasionally  trying  sites  while  their  mates 
also  perform  trying  activities  (Ficken  1964). 
Meanley  (1971),  Robinson  (1990),  and  Oliar- 
nyk and  Robertson  (1996)  reported  male 
Swainson’s  Warblers,  Louisiana  Waterthrushes, 
and  Cerulean  Warblers  (respectively)  engaged 
in  similar  nest-site  trying  behaviors  with  their 
mates,  but  they  did  not  specify  the  frequency 
at  which  these  behaviors  occurred.  Meanley 
(1971)  also  reported  that  male  Swainson’s  War- 
blers might  examine  nest  sites  alone. 

Interestingly,  the  males  of  species  considered 
most  closely  related  to  Golden-cheeked  War- 
blers (e.g..  Black-throated  Green  Warbler  [D. 
virens],  Hermit  Warbler  [ D . occidentalism, 
Townsend’s  Warbler  [. D . townsendi],  and  Black- 
throated  Gray  Warbler  [ D . nigrescens])  do  not 
appear  to  participate  in  nest-site  selection.  The 
females  either  “size”  or  “examine”  prospective 
sites  (Black- throated  Green  and  Townsend’s 
warblers;  Morse  1993  and  Wright  et.  al  1998, 
respectively)  or  settle  into  a fork  and  flit  around 
for  5-15  sec  (Black-throated  Gray  Warbler; 
Guzy  and  Lowther  1997),  while  the  male  fol- 
lows closely  and  infrequently  utters  soft  songs. 
This  apparent  difference  in  nest-site  selection 
strategy  and  display  may  be  a function  of  the 
secretive  behavior  exhibited  by  these  species 
during  pair  formation;  it  is  certainly  plausible 
that  active  male  participation  occurs  in  these 


250 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


species,  but  has  simply  not  yet  been  observed. 
Based  on  well-studied  warbler  species,  Morse 
(1989:169)  reasoned  that  species’  repertoires  are 
extensive,  making  explicit  comparisons  among 
species  difficult  to  derive:  . major  differenc- 

es may  lie  in  the  frequency  with  which  a display 
is  performed,  rather  than  the  ability  to  perform 
it.”  Alternatively,  males  of  species  closely  re- 
lated to  the  Golden-cheeked  Warbler  may  not 
exhibit  similar  nest-site  selection  activities  be- 
cause visual  displays  in  wood  warblers  are  not 
necessarily  correlated  with  phylogeny  (Morse 
1989).  Lovette  and  Bermingham  (1999)  suggest 
that  adaptive  differences  in  behavioral  charac- 
ters exhibited  by  DencLroica  species  may  have 
developed  long  after  their  explosive  speciation. 

Male  birds  may  exhibit  varying  degrees  of 
participation  in  nest-site  selection  by  (1)  se- 
lecting the  site  alone,  (2)  mate-guarding  to 
protect  their  genetic  investment,  (3)  perform- 
ing displays  to  synchronize  the  pair’s  repro- 
ductive cycle,  and/or  (4)  performing  displays 
to  determine  an  actual  location  that  shows  the 
most  promise  for  successfully  fledging  young. 
Hansell  (2000)  suggests  that  increased  paren- 
tal care  by  both  parents  can  be  found  among 
species  in  which  both  sexes  build  the  nest  to- 
gether; perhaps  the  same  holds  true  for  species 
exhibiting  joint  male-female  nest-site  selec- 
tion. In  a review  of  The  Birds  of  North  Amer- 
ica species  accounts,  we  identified  96  species 
from  1 1 orders  and  35  families  in  which  both 
sexes  actively  engage  in  nest-site  selection. 
Among  these  species,  both  sexes  participate 
in  feeding  young  in  81  (98%)  of  the  83  spe- 
cies where  at  least  one  sex  feeds  young.  Close 
relatives  of  the  Golden-cheeked  Warbler,  how- 
ever, all  exhibit  biparental  feeding  (as  is  ex- 
pected in  nidicolous  species),  but  do  not  ap- 
pear to  show  biparental  nest-site  selection 
(Morse  1993,  Guzy  and  Lowther  1997,  Wright 
et  al.  1998,  Ladd  and  Gass  1999).  The  life- 
history  traits  (e.g.,  long-term  pair  bonds,  role 
of  sexes  in  parental  investment)  common  to 
avian  species  that  engage  in  joint  male-female 
nest-site  selection  deserve  additional  study. 

ACKNOWLEDGMENTS 

We  thank  J.  L.  Sterling  Graber  for  co-observing  the 
2003  nest-site  selection  behavior  and  subsequently  lo- 
cating the  nest.  We  also  thank  numerous  field  techni- 
cians and  volunteers,  including  M.  Atkinson,  V.  Bump, 
B.  C.  Cook,  K.  Cutrera,  J.  L.  Granger,  K.  J.  Moore,  H. 


Oswald,  H.  Schreiber,  and  R.  K.  Williams,  who  tire- 
lessly spent  hours  in  the  field  for  our  concurrent  study 
dealing  with  the  tolerance  of  Golden-cheeked  Warblers 
to  nonconsumptive  recreation.  C.  M.  Abbruzzesse,  A. 
D.  Anders,  P.  M.  Cavanagh,  J.  D.  Cornelius,  D.  C.  Dear- 
born, T.  J.  Hayden,  D.  M.  Herbert,  S.  Jester,  D.  L.  Koeh- 
ler, R.  I.  Leyva,  B.  Peak,  C.  E.  Pekins,  and  M.  M.  Stake 
provided  assistance  with  project  planning,  initiation,  and 
logistical  support.  We  thank  R.  T.  Churchwell,  S.  Davis, 
C.  G.  Ladd,  S.  A.  McClure,  T.  J.  O’Connell,  J.  H.  Rap- 
pole,  and  an  anonymous  reviewer  for  commenting  on 
earlier  versions  of  this  manuscript.  Funding  for  our  re- 
search was  provided  by  the  Department  of  the  Army 
and  was  administered  by  the  U.S.  Army  Construction 
Engineering  Research  Laboratory  and  the  Oklahoma 
Cooperative  Fish  and  Wildlife  Research  Unit 
(Oklahoma  State  University,  Oklahoma  Department  of 
Wildlife  Conservation,  U.S.  Geological  Survey,  and 
Wildlife  Management  Institute  cooperating).  Informa- 
tion contained  in  this  manuscript  does  not  necessarily 
reflect  the  position  or  policy  of  the  government,  and  no 
official  endorsement  should  be  inferred. 

LITERATURE  CITED 

Bent,  A.  C.  1953.  Life  histories  of  North  American 
wood  warblers.  U.S.  National  Museum  Bulletin, 
no.  203.  [Reprinted  1963,  Dover  Publications, 
New  York] 

Bolsinger,  J.  S.  2000.  Use  of  two  song  categories  by 
Golden-cheeked  Warblers.  Condor  102:539-552. 
Boulton,  R.  L.,  P.  Cassey,  C.  Schipper,  and  M.  F. 
Clarke.  2003.  Nest  site  selection  by  Yellow-faced 
Honeyeaters  Lichenostomus  chrysops.  Journal  of 
Avian  Biology  34:267-274. 

Davis,  S.  K.  2005.  Nest-site  selection  patterns  and  the 
influence  of  vegetation  on  nest  survival  of  mixed- 
grass  prairie  passerines.  Condor  107:605-616. 
Ficken,  M.  S.  1964.  Nest-site  selection  in  the  Ameri- 
can Redstart.  Wilson  Bulletin  76:189. 

Guzy,  M.  J.  and  P.  E.  Lowther.  1997.  Black-throated 
Gray  Warbler  ( Dendroica  nigrescens ).  The  Birds 
of  North  America,  no.  319. 

Hansell,  M.  H.  2000.  Bird  nests  and  construction  be- 
havior. Cambridge  University  Press,  Cambridge, 
United  Kingdom. 

Kaufman,  K.  1996.  Lives  of  North  American  birds. 

Houghton  Mifflin,  Boston,  Massachusetts. 

Kier,  R.  S.,  L.  E.  Garner,  and  L.  F.  Brown,  Jr.  1977. 
Land  resources  of  Texas.  Bureau  of  Economic  Ge- 
ology, University  of  Texas,  Austin. 

Ladd,  C.  and  L.  Gass.  1999.  Golden-cheeked  Warbler 
( Dendroica  chrysoparia).  The  Birds  of  North 
America,  no.  420. 

Lockwood,  M.  W.  1996.  Courtship  behavior  of  Golden- 
cheeked Warblers.  Wilson  Bulletin  108:591-592. 
Lovette,  I.  J.  and  E.  Bermingham.  1999.  Explosive 
speciation  in  the  New  World  Dendroica  warblers. 
Proceedings  of  the  Royal  Society  of  London,  Se- 
ries B 266:1629-1636. 

Martin,  T.  E.  1992.  Interaction  of  nest  predation  and 


SHORT  COMMUNICATIONS 


251 


food  limitation  in  reproductive  strategies.  Current 
Ornithology  9:163-197. 

Martin,  T.  E.  1998.  Are  microhabitat  preferences  of 
coexisting  species  under  selection  and  adaptive? 
Ecology  79:656-670. 

Martin,  T.  E.  and  J.  J.  Roper.  1988.  Nest  predation 
and  nest-site  selection  of  a western  population  of 
the  Hermit  Thrush.  Condor  90:51-57. 

Meanley,  B.  1971.  Additional  notes  on  prenesting  and 
nesting  behavior  of  the  Swainson’s  Warbler.  Wil- 
son Bulletin  83:194. 

Morse,  D.  H.  1989.  American  warblers:  an  ecological 
and  behavioral  perspective.  Harvard  University 
Press,  Cambridge,  Massachusetts. 

Morse,  D.  H.  1993.  Black-throated  Green  Warbler 
{Dendroica  virens).  The  Birds  of  North  America, 
no.  55. 

Nolan,  V.,  Jr.  1978.  The  ecology  and  behavior  of  the 
Prairie  Warbler  ( Dendroica  discolor).  Ornitholog- 
ical Monographs,  no.  26. 

Oliarnyk,  C.  J.  and  R.  J.  Robertson.  1996.  Breeding 
behavior  and  reproductive  success  of  Cerulean 


Warblers  in  southeastern  Ontario.  Wilson  Bulletin 
108:673-684. 

Pulich,  W.  M.  1976.  The  Golden-cheeked  Warbler:  a 
bioecological  study.  Texas  Parks  and  Wildlife  De- 
partment, Austin,  Texas. 

Ricklefs,  R.  E.  1969.  An  analysis  of  nesting  mortality 
in  birds.  Smithsonian  Contributions  in  Zoology  9: 
1-48. 

Robinson,  W.  D.  1990.  Louisiana  Waterthrush  forag- 
ing behavior  and  microhabitat  selection  in  south- 
ern Illinois.  M.Sc.  thesis.  Southern  Illinois  Uni- 
versity, Carbondale. 

Siepielski,  A.  M.,  A.  D.  Rodewald,  and  R.  H.  Yah- 
ner.  2001.  Nest-site  selection  and  nesting  success 
of  the  Red-eyed  Vireo  in  central  Pennsylvania. 
Wilson  Bulletin  113:302-307. 

Wilson,  R.  R.  and  R.  J.  Cooper.  1998.  Acadian  Fly- 
catcher nest  placement:  does  placement  influence 
reproductive  success?  Condor  100:673-679. 

Wright,  A.  L.,  G.  D.  Hayward,  S.  M.  Matsuoka,  and 
P.  H.  Hayward.  1998.  Townsend’s  Warbler  ( Den- 
droica townsendi).  The  Birds  of  North  America, 
no.  333. 


The  Wilson  Journal  of  Ornithology  1 18(2):25 1-254,  2006 


Provisioning  of  Magellanic  Woodpecker  ( Campephilus  magellanicus) 
Nestlings  with  Vertebrate  Prey 

Valeria  S.  Ojeda12  and  M.  Laura  Chazarreta1 2 


ABSTRACT.— During  the  2003-2004  and  2004- 
2005  nesting  seasons,  we  studied  parental  behavior  at 
seven  Magellanic  Woodpecker  ( Campephilus  magel- 
lanicus) nests  in  Argentine  Patagonia.  Food  items  de- 
livered to  nestlings  included  wood-boring  larvae 
(57.6%),  arachnids  (13.1%),  and  vertebrates  (4.6%,  in- 
cluding a bat,  lizards,  and  avian  eggs  and  nestlings). 
Less  frequent  items  were  adult  insects,  caterpillars,  and 
pupae.  Small,  unidentified  invertebrate  prey  made  up 
19.8%  of  the  observations.  Males  delivered  most  of 
the  large  prey  (wood-boring  larvae  and  vertebrates; 
61.7%),  while  females  brought  most  of  the  small  prey 
(arachnids  and  small,  unidentified  invertebrates; 
79.6%),  suggesting  differences  in  foraging  strategies 
between  sexes.  This  is  the  first  published  account  of 
Magellanic  Woodpeckers  provisioning  nestlings  with 
vertebrates.  The  frequency  of  Magellanic  Woodpecker 
predation  on  vertebrates  outside  of  the  breeding  sea- 
sons is  unknown.  Received  26  January  2005,  accepted 
5 December  2005. 


1 Univ.  Nacional  del  Comahue,  Depto.  de  Zoologia 
y Ecologia,  8400  Bariloche,  Argentina. 

2 Corresponding  author;  e-mail: 
campephilus  @bariloche.com.ar 


Although  several  woodpecker  species  (es- 
pecially melanerpine  species)  regularly  prey 
on  the  nestlings  and  eggs  of  other  birds,  and 
a small  number  of  species  occasionally  cap- 
ture lizards  or  even  mice,  picids  are  generally 
not  considered  to  be  important  predators  of 
vertebrates  (Short  1982,  del  Hoyo  et  al.  2002). 
The  diet  of  the  Magellanic  Woodpecker  ( Cam- 
pephilus magellanicus ),  the  largest  Neotropi- 
cal picid,  remains  largely  unstudied;  the  spe- 
cies is  considered  a specialist  predator  of 
large,  wood-boring  larvae  (Short  1970,  1982). 
There  is  only  one  record  of  a Magellanic 
Woodpecker  capturing  vertebrate  prey  (a  liz- 
ard, Liolaemus  sp.;  Ojeda  2003),  and,  based 
on  what  was  known  about  the  species’  diet, 
the  event  was  reported  as  opportunistic.  Re- 
cent observations,  however,  suggest  that  ver- 
tebrate predation  by  the  Magellanic  Wood- 
pecker may  be  more  common  than  previously 
believed.  Here,  we  present  data  on  food 


252 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


items — including  vertebrates — delivered  to 
nestlings. 

METHODS 

From  November  to  January  (2003-2004 
and  2004-2005  nesting  seasons),  we  studied 
parental  behavior  of  Magellanic  Woodpeckers 
in  native  lenga  ( Nothofagus  pumilio ) forests 
near  Bariloche  (41°  08'  S,  71°  12'  W)  in  Na- 
huel  Huapi  National  Park,  Argentine  Patagon- 
ia. The  area  is  characterized  by  lakes,  glacial 
valleys,  and  mountain  slopes  covered  by  for- 
ests dominated  by  southern  beech  {Nothofa- 
gus spp.).  Elevations  range  from  400  to  3,480 
m,  the  mean  annual  temperature  is  8°  C,  and 
winds  are  predominantly  westerly.  Annual 
rainfall  ranges  from  500  to  2,000  mm  and  oc- 
curs primarily  in  winter  (Paruelo  et  al.  1998). 

The  study  was  carried  out  at  two  forested 
sites  (Challhuaco  Valley  and  Otto  Mount)  lo- 
cated 15  km  apart.  Forest  composition  was 
similar  between  the  two  sites,  but  Otto  Mount 
was  being  intensively  logged  at  the  time  of 
our  observations.  Throughout  the  nesting  sea- 
son, we  observed  the  woodpeckers’  daily  rou- 
tine at  seven  nests  once  per  week,  from  dawn 
to  dusk  (—06:00-21:00  UTC-3).  We  found 
one  nest  at  the  Otto  Mount  site  and  six  at  the 
Challhuaco  Valley  site.  We  made  our  obser- 
vations from  ground  blinds  10-20  m from 
nest  trees,  and  observed  woodpeckers  with  8X 
binoculars  and  a 25  X spotting  scope.  Nests 
were  watched  for  a total  of  654  hr  (41  days; 
5-9  days/nest). 

Because  of  marked  sexual  dimorphism 
(Short  1970)  and  strong  territoriality  (VSO 
pers.  obs.),  adults  did  not  need  to  be  marked. 
Magellanic  Woodpeckers  normally  made  one 
or  more  stops  before  going  to  the  nest  en- 
trance, and  once  there,  they  perched  for  a few 
seconds  before  feeding  nestlings.  This  per- 
mitted identification  of  the  more  conspicuous 
prey  items  to  at  least  the  level  of  class.  Iden- 
tification of  prey  to  the  species  level  was 
made  via  direct  observation  of  predation 
events  or  during  laboratory  analysis  of  prey 
items  found  at  the  bottom  of  nest  cavities  (in- 
spected every  5—10  days). 

During  the  first  3 weeks  of  the  nestling  pe- 
riod, the  adults  normally  entered  the  nest  cav- 
ity either  without  prey  or  with  items  too  small 
to  be  detected  (Ojeda  2004).  Because  we  saw 
no  vertebrate  prey  delivered  during  this  time, 


we  assumed  that  vertebrate  prey  were  not  de- 
livered to  nestlings  until  they  were  older. 
Hence,  the  provisioning  data  analyzed  in  this 
paper  correspond  to  the  middle  and  last  parts 
of  the  nestling  period  (nestlings  20-48  days 
of  age,  on  average),  when  prey  were  large 
enough  to  be  detected. 

RESULTS  AND  DISCUSSION 

We  recorded  852  deliveries  of  conspicuous 
prey  at  seven  nests.  Total  deliveries  per  nest 
ranged  from  72  to  180.  Males  made  52.6% 
(range  = 38.0-74.6%)  of  all  prey  deliveries, 
while  females  delivered  47.4%  (range  = 
23.4-62.0%). 

Most  identified  prey  were  wood-boring  lar- 
vae, arachnids,  and  vertebrates  (Table  1).  Ver- 
tebrate prey  was  delivered  to  all  nests,  pri- 
marily by  males;  most  “vertebrates”  deliv- 
ered by  females  were  birds’  eggs  (n  = 4).  Al- 
though small  sample  sizes  precluded  statistical 
testing  for  differences  in  feeding  behavior 
among  pairs  or  sexes,  large  prey  (wood-boring 
larvae  and  vertebrates)  were  mostly  (61.7%) 
brought  by  males,  while  small  prey  (arachnids 
and  unidentified  small  invertebrates)  were 
mostly  (79.6%)  brought  by  females,  suggest- 
ing potential  differences  in  foraging  strategies 
between  sexes.  Short  (1970)  proposed  such 
differences  in  foraging  behavior  based  on  the 
species’  sexual  dimorphism  in  bill  size. 

Based  on  their  slender  shape  and  dark  col- 
oration, the  lizard  prey  we  observed  were 
most  likely  Liolaemus  sp.  (N.  Ibargiiengoytia 
pers.  comm.).  The  eggs  delivered  to  the  nests 
varied  in  coloration  from  white,  to  pink,  to 
Niagara-green  and  were  small-  to  medium- 
sized. Although  we  did  not  identify  many  of 
the  nestling  prey  items  (n  = 14)  delivered  to 
woodpecker  nestlings,  at  least  one  individual 
of  seven  species  (mostly  passerines)  was  iden- 
tified: Patagonian  Sierra-Finch  ( Phrygilus  pa- 
tagonicus ),  Austral  Thrush  ( Turdus  falcklan- 
dii ),  House  Wren  {Troglodytes  aedon),  Thom- 
tailed  Rayadito  {Aphrastura  spinicauda ), 
Striped  Woodpecker  {Picoides  lignarius). 
White-throated  Treerunner  {Pygarrhichas  al- 
bogularis),  and  Fire-eyed  Diucon  {Xolmis  py- 
rope).  On  several  occasions,  lizards  and  nest- 
lings brought  by  adults  were  so  large  that  they 
could  not  be  swallowed  by  the  woodpecker 
nestlings.  In  such  cases,  after  several  failed 
feeding  attempts,  the  prey  was  left  at  the  bot- 


SHORT  COMMUNICATIONS 


253 


TABLE  1.  Percentages  of  852  prey  items  delivered  by  male  ( n = 
Woodpeckers  ( Campephilus  magellanicus ) to  nestlings  in  seven  nests 
2004  and  2004-2005  nesting  seasons. 

= 448)  and  female  ( n = 404)  Magellanic 
in  Argentine  Patagonia  during  the  2003- 

Prey  type  (n) 

Both  sexes 

Male 

Female 

Invertebrates 

Wood-boring  larvae  (491) 

57.6 

65.6 

48.8 

Arachnids  (112) 

13.1 

9.6 

17.1 

Adult  insects  (31) 

3.6 

3.3 

4.0 

Caterpillars  (4) 

0.5 

0.7 

0.2 

Pupae  (6) 

0.7 

1.3 

0.0 

Unidentified  invertebrates  (169) 

19.8 

12.0 

28.5 

All  invertebrates  (813) 

95.4 

92.5 

98.6 

Vertebrates 

Lizards  (13) 

1.5 

2.7 

0.2 

Nestlings  (14) 

1.6 

3.1 

0.0 

Avian  eggs  (8) 

0.9 

0.9 

1.0 

Bats  (1) 

0.1 

0.2 

0.0 

Unidentified  vertebrates  (3) 

0.4 

0.4 

0.2 

All  vertebrates  (39) 

4.6 

7.4 

1.5 

tom  of  the  nest  cavity;  on  one  occasion,  how- 
ever, an  attending  male  flew  to  a nearby  tree 
with  the  prey  and  ate  it  (a  lizard). 

The  identity  of  avian  prey  or  potential  avian 
prey  also  was  determined  in  several  additional 
ways.  In  one  case,  a woodpecker  provisioned 
its  nestling  with  four  similar,  small  nestlings, 
each  brought  individually.  Between  these  de- 
liveries, the  woodpecker  flew  away  from,  and 
returned  to,  its  nest  from  the  same  direction. 
On  the  last  three  return  trips,  the  woodpecker 
was  followed  by  a pair  of  Thorn-tailed  Ray- 
aditos  that  were  vigorously  harassing  it,  but 
with  no  effect.  We  interpreted  this  event  as 
woodpecker  predation  on  a brood  of  rayaditos. 
On  another  occasion,  we  witnessed  a male 
woodpecker  vigorously  pecking  on,  and  chis- 
eling out,  the  bark  wall  that  protected  a House 
Wren  nest  in  a natural  crevice;  however,  the 
woodpecker  was  suddenly  interrupted  by  his 
mate’s  arrival  and  he  discontinued  his  peck- 
ing. When  we  examined  the  half-opened  wren 
cavity,  we  found  three  small  hatchlings.  The 
adult  wrens  were  not  present  during  the  pre- 
dation attempt. 

We  also  recorded  the  characteristic  foraging 
signs  of  Magellanic  Woodpeckers  at  several 
(n  = 11)  small  woodpecker  cavities  that  had 
been  partially  destroyed.  Below  some  cavities, 
we  observed  a row  of  Magellanic  Woodpecker 
feeding  holes  that  descended  from  the  lower 
lip  of  the  cavity  entrance  to  the  floor  level  of 


the  nest  chamber.  In  other  cases,  it  appeared 
that  Magellanic  Woodpeckers  had  pecked 
only  at  the  level  of  the  nest  chamber’s  floor, 
where  a hole  about  the  size  of  the  nest  en- 
trance had  been  drilled.  Originally,  these  small 
cavities  had  been  excavated  by  Striped  Wood- 
peckers or  White-throated  Treerunners,  and 
some  contained  the  cup  nests  of  secondary 
cavity  nesters.  Due  to  differences  in  body  size 
and  feeding  habits  between  the  Magellanic 
Woodpecker  and  these  much  smaller  species 
(Short  1970,  1982),  competition  is  not  a likely 
explanation  for  the  destructive  behavior  ob- 
served. It  appears  that  such  cavities  were  de- 
stroyed to  reach  the  nest  chamber  at  the  bot- 
tom of  the  cavity. 

This  is  the  first  published  account  of  Ma- 
gellanic Woodpeckers  provisioning  their  nest- 
lings with  vertebrates.  Though  wood-boring 
larvae  may  be  the  primary  food  of  this  wood- 
pecker throughout  its  range,  there  is  increas- 
ing evidence  that  Magellanic  Woodpeckers 
are  opportunistic  foragers  that  will  take  a wide 
variety  of  prey.  In  addition  to  insects,  verte- 
brates, and  eggs,  they  have  also  been  recorded 
feeding  on  vegetable  matter  (including  sap)  at 
locations  throughout  much  of  their  range 
(Ojeda  2003,  Schlatter  and  Vergara  2005). 

ACKNOWLEDGMENTS 

We  are  indebted  to  J.  M.  Karlanian,  M.  Gelain,  A. 
Ortiz,  P.  Taccari,  Y.  Sasal,  and  C.  Bacchi  for  their  in- 


254 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


dispensable  assistance  during  the  fieldwork  and  to  N. 
Ibargiiengoytia  for  her  assessment  concerning  Pata- 
gonian reptiles.  We  are  grateful  to  IDEA  WILD  (Fort 
Collins,  CO,  USA),  Birders’  Exchange  (Colorado 
Springs,  CO,  USA),  and  J.  Masello  for  their  donations 
of  equipment  fundamental  to  this  study.  This  research 
was  supported  by  CANON  USA  through  its  National 
Parks  Science  Scholars  Program  for  the  Americas.  We 
thank  R.  N.  Conner,  L.  L.  Short,  and  K.  Franzreb  for 
constructive  comments  that  helped  improve  this  man- 
uscript. 

LITERATURE  CITED 

del  Hoyo,  J.,  A.  Elliott,  and  J.  Sargatal  (Eds.). 
2002.  Handbook  of  the  birds  of  the  world,  vol.  7: 
jacamars  to  woodpeckers.  Lynx  Edicions,  Barce- 
lona, Spain. 

Ojeda,  V.  2003.  Magellanic  Woodpecker  frugivory  and 
predation  on  a lizard.  Wilson  Bulletin  115:208- 
210. 


Ojeda,  V.  2004.  Breeding  biology  and  social  behaviour 
of  Magellanic  Woodpeckers  ( Campephilus  magel- 
lanicus ) in  Argentine  Patagonia.  European  Journal 
of  Wildlife  Research  50:18-24. 

Paruelo,  J.  M.,  A.  Beltran,  E.  Jobbagy,  O.  E.  Sala, 
and  R.  A.  Golluscio.  1998.  The  climate  of  Pa- 
tagonia: general  patterns  and  controls  on  biotic 
processes.  Ecologia  Austral  8:85-101. 

Schlatter,  R.  P.  and  P.  Vergara.  2005.  Magellanic 
Woodpecker  ( Campephilus  magellanicus ) sap 
feeding  and  its  role  in  the  Tierra  del  Fuego  forest 
bird  assemblage.  Journal  of  Ornithology  146:188- 
190. 

Short,  L.  L.  1970.  The  habits  and  relationships  of  the 
Magellanic  Woodpecker.  Wilson  Bulletin  82:115- 
129. 

Short,  L.  L.  1982.  Woodpeckers  of  the  world.  Dela- 
ware Museum  of  Natural  History  Monographs, 
no.  4. 


The  Wilson  Journal  of  Ornithology  1 18(2):254— 256,  2006 


Reverse  Mounting  and  Copulation  Behavior  in  Polyandrous  Bearded 
Vulture  ( Gypaetus  barbatus)  Trios 

Joan  Bertran1 2  and  Antoni  Margalida12 


ABSTRACT. — We  present  the  first  report  of  reverse 
mounting  in  the  Bearded  Vulture  ( Gypaetus  barbatus). 
The  reverse  mounting,  which  occurred  in  the  Pyrenees 
of  northeastern  Spain,  took  place  between  the  female 
and  the  alpha  male  in  a polyandrous  trio.  The  function 
of  reverse  mountings  is  discussed  in  relation  to  the 
previously  reported  high  frequency  of  male-male 
mountings  in  this  raptor  species.  Received  25  April 
2005,  accepted  17  January  2006. 


Reverse  mounting,  in  which  the  female 
mounts  the  male,  has  been  described  in  a 
number  of  bird  species  (see  James  1983, 
Nuechterlein  and  Storer  1989).  This  behavior 
has  been  rarely  documented  in  raptors,  how- 
ever, except  for  a few  isolated  cases  in  species 
such  as  American  Kestrel  ( Falco  sparverius'. 
Bowman  and  Curley  1986)  and  Egyptian  Vul- 
ture {Neophron  percnopterus\  Donazar  1993). 


1 Bearded  Vulture  Study  and  Protection  Group, 
Apdo.  43,  E-25520  El  Pont  de  Suert  (Lleida)  Spain. 

2 Corresponding  author;  e-mail: 
margalida@gauss.entorno.es 


We  describe  a case  of  reverse  mounting  in 
a polyandrous  trio  of  Bearded  Vultures  {Gy- 
paetus barbatus).  Bearded  Vultures  are  terri- 
torial and  socially  monogamous  (Hiraldo  et  al. 
1979);  however,  in  the  Pyrenees  (in  both 
Spain  and  France),  where  the  species’  largest 
European  population  occurs,  polyandrous  co- 
alitions are  relatively  common  (Heredia  and 
Donazar  1990).  The  birds  in  this  population 
maintained  104  breeding  territories  (R.  Here- 
dia and  M.  Razin  pers.  comm.),  18  of  which 
were  occupied  by  polyandrous  trios.  Before 
egg-laying.  Bearded  Vultures  in  the  Pyrenees 
engage  in  their  copulations  for  an  average  of 
67  days  (range  = 50-90;  Bertran  and  Mar- 
galida  1999),  during  which  male-male  mount- 
ings in  trios  occasionally  occur  (Bertran  and 
Margalida  2003). 

Between  2004  and  2005,  we  monitored  a 
polyandrous  trio  of  Bearded  Vultures  in  the 
central  Pre-Pyrenees  mountains  in  Catalonia, 
northeastern  Spain,  during  their  courtship  pe- 
riod (200  hr  of  observation).  We  sexed  and 
identified  the  individuals  by  observing  their 


SHORT  COMMUNICATIONS 


255 


TABLE  1.  Number  of  male-female,  male-male,  and  reverse  mounting  copulation  attempts  observed  in  mo- 
nogamous pairs  ( n = 8)  and  polyandrous  trios  ( n = 5)  of  Bearded  Vultures  in  the  Pyrenees,  northeastern  Spain, 


2004-2005. 


Male-Female 

Male-Male 

Female-Male 

Source 

Pairs 

189 

— 

0 

Bertran  and  Margalida  (1999) 

Trios 

356 

39 

1 

This  study 

copulatory  activities  and  specific  plumage  pat- 
terns. On  30  October  2004  at  12:19  UTC+1 
(84  days  before  egg-laying),  the  female 
mounted  the  alpha  male  after  she  had  been 
mounted  unsuccessfully  by  the  beta  male.  Fol- 
lowing the  female’s  mount,  the  alpha  male 
drove  the  beta  male  off  the  perching  site.  The 
duration  of  the  reverse  mounting  (8  sec)  was 
similar  to  that  of  behaviorally  successful 
male-female  copulations  recorded  in  other 
polyandrous  groups  (mean  = 10.49  sec  ± 
1.30  SD,  range  = 8-14,  n = 37;  Bertran  and 
Margalida  2004). 

Previously,  researchers  have  studied  reverse 
mounting  in  the  context  of  pair  formation,  de- 
gree of  sexual  motivation,  or  reversal  of  sex- 
ual dominance  (Nuechterlein  and  Storer  1989, 
Bowen  et  al.  1991,  Ortega-Ruano  and  Graves 
1991).  Due  to  their  physical  and  behavioral 
characteristics,  it  has  been  suggested  that  fe- 
male Bearded  Vultures  can  dominate  males 
(see  Negro  et  al.  1999);  in  the  Cattle  Egret 
C Bubulcus  ibis),  reverse  mounting  has  been  as- 
sociated with  establishing  dominance  (Fujioka 
and  Yamagishi  1981).  However,  if  reverse 
mounting  were  of  adaptive  value  (e.g.,  to 
maintain  female  dominance  or  to  strengthen 
heterosexual  couplings),  it  likely  would  be 
more  common.  On  the  other  hand,  sexual  in- 
teractions outside  the  context  of  fertilization 
appear  to  be  relatively  common  in  polyan- 
drous trios  (Table  1),  and  reverse  mounting 
might  simply  be  a side  effect  of  male-male 
mountings.  That  is,  the  function  of  reverse 
mounting  may  be  to  regulate  socio-sexual  ten- 
sions— similar  to  the  function  of  male-male 
mountings  (Bertran  and  Margalida  2003,  see 
also  Heg  and  van  Treuren  1990,  Cockbum 
2004).  Further  research  is  needed  to  determine 
whether  reverse  mounting  is  the  result  of  con- 
frontational situations  or  helps  to  regulate 
them. 


ACKNOWLEDGMENTS 

We  thank  A.  Bonada,  X.  Macia,  P.  Romero,  and  E. 
Vega  for  their  help  during  fieldwork,  and  D.  Heg  and 
two  anonymous  referees  for  their  comments.  This 
study  was  supported  by  Departament  de  Medi  Ambient 
i Habitatge  de  la  Generalitat  de  Catalunya  and  Min- 
isterio  de  Medio  Ambiente. 

LITERATURE  CITED 

Bertran,  J.  and  A.  Margalida.  1999.  Copulatory  be- 
havior of  the  Bearded  Vulture.  Condor  101:161  — 
164. 

Bertran,  J.  and  A.  Margalida.  2003.  Male-male  cop- 
ulations in  polyandrous  Bearded  Vultures  ( Gypae - 
tus  barbatus):  an  unusual  mating  system  in  rap- 
tors. Journal  of  Avian  Biology  34:334-338. 
Bertran,  J.  and  A.  Margalida.  2004.  Do  females 
control  matings  in  polyandrous  Bearded  Vulture 
( Gypaetus  barbatus ) trios?  Ethology,  Ecology  and 
Evolution  16:181-186. 

Bowen,  B.  S.,  R.  R.  Koford,  and  S.  L.  Vehrencamp. 
1991.  Seasonal  pattern  of  reverse  mounting  in  the 
Groove-Billed  Ani  ( Crotophaga  sulcirostris ). 
Condor  93:159-163. 

Bowman,  R.  and  E.  M.  Curley.  1986.  Reverse  mount- 
ing in  the  American  Kestrel.  Wilson  Bulletin  98: 
472-473. 

Cockburn,  A.  2004.  Mating  systems  and  sexual  con- 
flict. Pages  81-191  in  Ecology  and  evolution  of 
cooperative  breeding  in  birds  (W.  D.  Koenig  and 
J.  L.  Dickinson,  Eds.).  Cambridge  University 
Press,  Cambridge,  United  Kingdom. 

Donazar,  J.  A.  1993.  Los  buitres  ibericos:  biologia  y 
conservacion.  J.  M.  Reyero  (Ed.),  Madrid,  Spain. 
Fujioka,  M.  and  S.  Yamagishi.  1981.  Extramarital  and 
pair  copulations  in  the  Cattle  Egret.  Auk  98:134- 
144. 

Heg,  D.  and  R.  van  Treuren.  1998.  Female-female 
cooperation  in  polygynous  Oystercatchers.  Nature 
391:687-691. 

Heredia,  R.  and  J.  A.  Donazar.  1990.  High  frequency 
of  polyandrous  trios  in  an  endangered  population 
of  Lammergeiers  Gypaetus  barbatus  in  northern 
Spain.  Biological  Conservation  53:163-171. 
Hiraldo,  F,  M.  Delibes,  and  J.  Calderon.  1979.  El 
Quebrantahuesos  Gypaetus  barbatus  (L.).  Mono- 
grafias,  no.  22,  Instituto  para  la  Conservacion  de 
la  Naturaleza,  Madrid,  Spain. 

James,  P.  C.  1983.  Reverse  mounting  in  the  North- 


256 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


western  Crow.  Journal  of  Field  Ornithology  54: 
418-419. 

Negro,  J.  J.,  A.  Margalida,  F.  Hiraldo,  and  R.  He- 
redia. 1999.  The  function  of  the  cosmetic  colour- 
ation of  Bearded  Vultures:  when  art  imitates  life. 
Animal  Behaviour  58:F14-F17. 


Nuechterlein,  G.  L.  and  R.  W.  Storer.  1989.  Reverse 
mounting  in  grebes.  Condor  91:341-346. 

Ortega-Ruano,  J.  and  J.  A.  Graves.  1991.  Reverse 
mounting  during  the  courtship  of  the  European 
Shag  Phalacrocorax  aristotelis.  Condor  93:859- 
863. 


The  Wilson  Journal  of  Ornithology  1 18(2):256— 259,  2006 


Natural  Occurrence  of  Crowing  in  a Free-living  Female  Galliform,  the 

California  Quail 

Jennifer  M.  Gee12 


ABSTRACT. — The  vocalizations  of  galliform  spe- 
cies are  typically  sexually  dimorphic  in  that  only  the 
males  crow.  I observed  crowing  by  a female  California 
Quail  ( Callipepla  californica),  a galliform  species  that 
ranges  along  the  Pacific  coast  of  North  America.  I re- 
corded the  female  crowing  during  a period  of  the 
breeding  season  when  many  other  females  were  paired. 
The  female’s  crow  was  similar  in  frequency  to  a typ- 
ical male  crow,  though  it  was  slightly  shorter  in  du- 
ration. I discuss  possible  mechanisms  and  conditions 
that  could  result  in  female  crowing.  Received  28  Feb- 
ruary 2005,  accepted  21  December  2005. 


California  Quail  {Callipepla  californica) 
show  pronounced  sexual  dimorphism  in  call- 
ing behavior:  males  crow,  whereas  females  do 
not.  The  California  Quail’s  crow  is  commonly 
called  the  Male  Advertisement  or  cow  call 
(Sumner  1935,  Williams  1969).  Males  usually 
crow  early  in  the  breeding  season  (Williams 
1969),  or  when  their  mates  are  incubating  or 
die  (JMG  pers.  obs.).  Crowing  males  often 
perch  in  conspicuous  locations  and  counter- 
call to  each  other.  To  my  knowledge,  there 
have  been  no  previous  reports  of  female  Cal- 
ifornia Quail  crowing  under  natural  condi- 
tions, although  Genelly  (1955)  observed  an 
instance  of  crowing  in  female  California  Quail 
that  were  held  under  captive  conditions. 

I observed  a female  California  Quail  crow- 
ing in  the  foothills  of  the  Santa  Rosa  Moun- 


1 Dept,  of  Ecology  and  Evolutionary  Biology, 
Princeton  Univ.,  Princeton,  NJ  08544-1003,  USA. 

2 Current  address:  W139  Mudd  Hall.  Dept,  of  Neu- 
robiology and  Behavior,  Cornell  Univ..  Ithaca,  NY 
14853,  USA;  e-mail:  jmg233@cornell.edu 


tains,  California  (33°  22'  N,  1 16°  15'  W),  dur- 
ing the  breeding  season  when  many  males 
were  crowing  (March  2000).  In  that  region, 
the  ranges  of  California  and  Gambel’s  (C. 
gambelii ) quail  overlap  and  hybrids  or  back- 
crosses  compose  approximately  60%  of  the 
population  (Gee  2003).  From  14  to  16  March, 
while  conducting  daily  observations  (>7  hr/ 
day)  at  this  site  with  a spotting  scope,  I ob- 
served a female  California  Quail  crowing  for 
1-  to  2-hr  periods.  This  female  approached  to 
approximately  5 m in  response  to  calls  that  I 
made  with  a quail  call,  and  she  continued 
crowing  from  that  distance  for  more  than  10 
min.  Both  California  and  Gambel’s  quail  are 
sexually  dimorphic;  thus,  I used  field  mark- 
ings to  identify  the  sex  and  species  of  the 
crowing  bird.  I identified  the  individual  as  fe- 
male by  her  lack  of  secondary  sex  traits  (e.g., 
brown  cap,  black  face  with  white  margin),  and 
as  a California  Quail  by  the  presence  of  scaled 
breast  feathers,  forward-pointing  crest,  and 
overall  blue-gray  body  plumage  (not  buff). 
However,  backcrosses  may  look  very  similar 
to  pure  parental  types  (Gee  2003).  I was  un- 
able to  trap  the  bird,  so  I could  use  neither 
genotyping  to  confirm  the  sex  or  species  des- 
ignation nor  laparotomy  to  examine  the  inter- 
nal anatomical  sex.  Despite  plumage  traits, 
there  is  potentially  some  ambiguity  as  to  the 
“true”  sex  and  species  of  this  individual. 

I used  Canary  1 .2.4  (www.birds. Cornell. 
edu/brp/SoundSoftware.html)  and  Syrinx 
(Burt  2005)  sound  analysis  programs  to  digi- 
tize recordings  and  prepare  spectrograms  from 
which  frequency  and  sound  duration  were 


SHORT  COMMUNICATIONS 


257 


0.0 


05 


TO 


Time  (sec) 


FIG.  1.  Spectrograms  (kHz/sec)  of  a female  Cali- 
fornia Quail  crow  (A)  compared  with  typical  male  ad- 
vertisement calls  of  California  (B),  hybrid  (C),  and 
Gambel’s  (D)  quail  in  a region  of  range  overlap  (de- 
scribed in  detail  in  Gee  2003),  in  the  foothills  of  the 
Santa  Rosa  Mountains,  California.  Recordings  were 
made  between  1998  and  2001. 


measured.  Many  low-frequency  noises  ob- 
scured the  first  harmonic  (fundamental  fre- 
quency) of  the  call;  therefore,  I measured  the 
peak  frequency  of  the  harmonic  nearest  the 
fundamental  frequency  because  it  was  clearly 
visible  in  all  spectrograms  (Fig.  1).  The  fe- 
male’s crow  was  approximately  the  same  fre- 


quency as,  but  slightly  shorter  in  duration 
than,  that  of  an  average  male  California  Quail 
(Table  1).  The  female  exhibited  male-typical 
crowing  posture  and  behavior,  calling  from  a 
conspicuous  rock  outcrop  to  males  that  were 
crowing  in  the  distance.  Though  it  was  a year 
of  moderate  reproductive  success,  the  female 
did  not  appear  to  have  a mate,  nor  was  she  a 
local  resident  based  on  detailed  observations 
of  color-banded  individuals  at  this  location 
(for  methods  see  Gee  2003). 

Conditions  and  mechanisms  that  could  have 
caused  this  female’s  unusual  behavior  include 
(1)  elevated  testosterone  due  to  increased  fe- 
male competition,  or  (2)  elevated  testosterone 
coupled  with  age-dependent  decrease  in  ovar- 
ian function  and  estrogen  production.  Note 
that  in  both  cases,  I suggest  a role  for  testos- 
terone, but  without  examination  of  the  gonad, 
there  is  no  way  to  verify  the  anatomical  and 
physiological  sex  of  the  crowing,  apparently 
female  individual.  Thus,  a reproductive,  pos- 
sibly endocrine,  pathology  may  have  contrib- 
uted to  this  behavior. 

Intense  competition  may  affect  testosterone 
levels  and  crowing  behavior.  In  males,  in- 
creased testosterone  occurs  when  males  are 
competing  for  mates,  and  it  is  a normal  con- 
sequence of  reaching  breeding  condition.  Sim- 
ilarly in  females,  intense  competition  for  scarce 
resources,  such  as  food  or  mates,  could  elevate 
testosterone  levels  or  its  rate  of  conversion  to 
other  steroids.  When  California  Quail  were 
kept  in  female-biased  pens,  females  became 
more  aggressive  and  began  crowing,  possibly 
due  to  intense  competition  (reported  in  Calkins 
et  al.  1999).  Although  the  sex  ratio  from  Jan- 
uary to  June  at  my  study  site  was  not  signifi- 
cantly skewed  (52:48,  n = 130),  local  move- 
ments could  have  created  periods  of  unusually 
intense  female  competition.  The  crowing  fe- 
male was  unpaired  and  part  of  a wave  of  tran- 
sient residents,  many  of  which  appeared  to  be 
in  small  groups  of  4-6  individuals. 

Testosterone  has  been  shown  to  play  a role 
in  the  crowing  behavior  of  male  Gambel’s 
Quail  and  female  Japanese  Quail  ( Cotumix  ja- 
ponica ).  In  Gambel’s  Quail,  testosterone  injec- 
tions administered  during  July  (late  breeding 
season)  caused  normal  adult  males,  but  not  fe- 
males, to  call  more  frequently  and  behave  more 
aggressively  (Williams  1969).  However,  when 
female  Japanese  Quail  were  both  ovariecto- 


258 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


TABLE  1.  Call  duration  and  peak  frequency  (mean  ± SD)  of  California,  Gambel’s,  and  hybrid  quail  in  a 
region  of  range  overlap  (described  in  detail  in  Gee  2003),  in  the  foothills  of  the  Santa  Rosa  Mountains,  California. 
Recordings  were  made  between  1998  and  2001.  Sample  sizes  for  call  duration  and  frequency  of  different  males 
are  as  follows:  Callipepla  calif ornica  (8,  8),  hybrid  (16,  8),  C.  gambelii  (11,  8).  Multiple  recordings  were  made 
of  the  calling  female.  Only  the  clearest  recording  was  measured,  although  her  crows  appeared  very  similar  to 
one  another. 


Female 

Male 

C.  califomica 

C.  califomica 

Hybrid 

C.  gambelii 

Duration  (sec) 

0.36 

0.38  (0.03) 

0.45  (0.02) 

0.53  (0.03) 

Frequency  (kHz) 

1.98 

2.03  (0.16) 

1.85  (0.17) 

1.86  (0.90) 

mized  and  treated  with  testosterone,  they 
crowed  and  strutted  similar  to  males  (Adkins 
and  Adler  1972;  Adkins  1975;  Balthazart  et  al. 
1983,  1996).  Thus,  two  factors  may  cause  fe- 
male crowing:  increased  levels  of  testosterone 
and  decreased  ovarian  function.  Ovarian  func- 
tion appears  to  diminish  with  age  in  Gambel’s 
and  California  quail,  as  evidenced  by  the  ac- 
quisition of  partial  male  plumage  among  some 
older  females  (Hagelin  and  Kimball  1997)  and 
the  finding  that  sexually  dimorphic  plumage  is 
estrogen-dependent  in  many  galliforms  (Domm 
1939,  Owens  and  Short  1995).  In  the  case  re- 
ported here,  the  age  of  the  crowing  female  was 
unknown,  and  she  showed  no  evidence  of  par- 
tial male  plumage.  Although  both  vocalizations 
and  plumage  could  be  affected  by  ovarian 
function,  female  crowing  and  partial  male 
plumage  are  not  coupled  and  are  likely  regu- 
lated by  different  mechanisms.  Vocalizations 
appear  to  be  governed,  in  part,  by  increased 
numbers  of  androgen  receptors  in  the  vocal 
control  regions  or  by  steroid-converting  en- 
zymes. For  example,  administering  the  aro- 
matase  inhibitor,  fadrazole,  results  in  crowing 
by  female  Japanese  Quail  (Marx  et  al.  2004). 
Similarly,  the  crowing  and  strutting  of  male 
Japanese  Quail  largely  depend  on  the  conver- 
sion of  testosterone  to  dihydrotestosterone  and 
on  androgen  receptors  (Adkins-Regan  2005). 

In  the  sympatric  population  I studied,  male- 
typical  plumage  is  infrequent  (but  consistently 
present)  in  female  California,  Gambel’s,  and 
hybrid  quail,  while  male-typical  vocalizations 
are  not.  Approximately  1%  of  the  banded  fe- 
male California,  Gambel’s,  and  hybrid  quail 
have  partial  male  plumage,  and  they  may  pair 
and  breed  normally  (JMG  pers.  obs.).  In  con- 
trast, I observed  only  one  female  with  male- 
typical  calling  patterns.  This  difference  sug- 


gests that  separate  mechanisms  govern  sexu- 
ally dimorphic  plumage  compared  to  sexually 
dimorphic  vocalizations,  but  it  also  suggests 
that  different  selective  pressures  may  act  on 
plumage  and  crowing.  The  consequences  of 
female  crowing  may  be  severe,  particularly  if 
crowing  is  associated  with  other  aggressive 
and  territorial  behaviors,  as  it  is  in  both  New 
World  quail  (Johnsgard  1988)  and  Japanese 
Quail  (Balaban  1997).  Thus,  female  crowing 
may  occur  only  under  the  rare  circumstances 
when  it  and  other  aggressive  behaviors — 
which  are  typical  among  reproductive 
males — do  not  decrease  the  reproductive  fit- 
ness of  female  quail. 

ACKNOWLEDGMENTS 

Financial  support  was  provided  by  a National  Sci- 
ence Foundation  Predoctoral  Fellowship;  a National 
Science  Foundation  Doctoral  Dissertation  Improve- 
ment Grant  DEB-0073271;  an  Environmental  Protec- 
tion Agency  Science-to-Achieve-Results  Graduate  Fel- 
lowship U-9 157290 1-0;  Sigma  Xi  Grants-in-Aid  of 
Research;  the  American  Ornithologists’  Union  Betty 
Carnes  Memorial  Award;  and  The  Reserve  Commu- 
nity Association  of  The  Reserve,  Palm  Desert,  Cali- 
fornia. Carolyn  Stillwell  helped  to  draw  figures.  For 
comments,  I thank  S.  M.  Correa,  M.  L.  Tomaszycki, 
J.  C.  Hagelin,  and  three  anonymous  referees. 

LITERATURE  CITED 

Adkins,  E.  K.  1975.  Hormonal  basis  of  sexual  differ- 
entiation in  the  Japanese  Quail.  Journal  of  Com- 
parative Physiological  Psychology  89:61-71. 
Adkins,  E.  K.  and  N.  T.  Adler.  1972.  Hormonal  con- 
trol of  behavior  in  the  Japanese  Quail.  Journal  of 
Comparative  Physiological  Psychology  81:27-36. 
Adkins-Regan,  E.  K.  2005.  Hormones  and  animal  social 
behavior  (Monographs  in  behavior  and  ecology). 
Princeton  University  Press,  Princeton,  New  Jersey. 
Balaban,  E.  1997.  Changes  in  multiple  brain  regions 
underlie  species  differences  in  a complex,  con- 


SHORT  COMMUNICATIONS 


259 


genital  behavior.  Proceedings  of  the  National 
Academy  of  Sciences,  USA  94:2001-2006. 

Balthazart,  J.,  M.  Schumacher,  and  M.  A.  Ottin- 
ger.  1983.  Sexual  differences  in  the  Japanese 
Quail:  behavior,  morphology  and  intracellular  me- 
tabolism of  testosterone.  General  and  Compara- 
tive Endocrinology  51:191-207. 

Balthazart,  J.,  O.  Tlemcani,  and  G.  E Ball.  1996. 
Do  sex  differences  in  the  brain  explain  sex  differ- 
ences in  the  hormonal  induction  of  reproductive 
behavior?  What  25  years  of  research  on  the  Jap- 
anese Quail  tells  us.  Hormones  and  Behavior  30: 
627-661. 

Burt,  J.  2005.  Syrinx-PC:  a Windows  program  for 
spectral  analysis,  editing,  and  playback  of  acoustic 
signals,  ver.  2.4.  www.syrinxpc.com  (accessed  14 
July  2005). 

Calkins,  J.  D.,  J.  C.  Hagelin,  and  D.  F.  Lott.  1999. 
California  Quail  ( Callipepla  californica).  The 
Birds  of  North  America,  no.  473. 

Domm,  L.  V.  1939.  Modifications  in  sex  and  secondary 
sexual  characters  in  birds.  Pages  227-327  in  Sex  and 
internal  secretions:  a survey  of  recent  research  (E. 
Allen,  Ed.).  Williams  and  Wilkins,  Baltimore,  Mary- 
land. 


Gee,  J.  M.  2003.  How  a hybrid  zone  is  maintained: 
behavioral  mechanisms  of  interbreeding  between 
California  and  Gambel’s  quail  ( Callipepla  califor- 
nica and  C.  gambelii).  Evolution  57:2407-2415. 

Genelly,  R.  E.  1955.  Annual  cycle  in  a population  of 
California  Quail.  Condor  57:263-285. 

Hagelin,  J.  C.  and  R.  T.  Kimball.  1997.  A female 
Gambel’s  Quail  with  partial  male  plumage.  Wil- 
son Bulletin  109:544-546. 

Johnsgard,  P.  A.  1988.  The  quails,  partridges,  and 
francolins  of  the  world.  University  of  Nebraska 
Press,  Lincoln. 

Marx,  G.,  A.  Jurkevich,  and  R.  Grossmann.  2004. 
Effects  of  estrogens  during  embryonic  develop- 
ment on  crowing  in  the  domestic  fowl.  Physiology 
and  Behavior  82:637-645. 

Owens,  I.  P.  F.  and  R.  V.  Short.  1995.  Hormonal  basis 
of  sexual  dimorphism  in  birds:  implications  for 
new  theories  of  sexual  selection.  Trends  in  Ecol- 
ogy and  Evolution  10:44-47. 

Sumner,  E.  L.,  Jr.  1935.  A life  history  study  of  the 
California  Quail,  with  recommendations  for  its 
conservation  and  management.  California  Fish 
and  Game  21:167-253,  275-342. 

Williams,  H.  W.  1969.  Vocal  behavior  of  adult  Cali- 
fornia Quail.  Auk  86:631-659. 


The  Wilson  Journal  of  Ornithology  1 18(2):259— 261,  2006 


Poult  Adoption  and  Nest  Abandonment  by  a Female  Rio  Grande 

Wild  Turkey  in  Texas 

Steve  T.  Metz,1  Kyle  B.  Melton,1  Ray  Aguirre,2  Bret  A.  Collier,14 
T.  Wayne  Schwertner,3 4  Markus  J.  Peterson,1  and  Nova  J.  Silvy1 


ABSTRACT. — While  evaluating  reproductive  pa- 
rameters in  Rio  Grande  Wild  Turkeys  ( Meleagris  gal- 
lopavo  intermedia ) in  the  Edwards  Plateau  region  of 
Texas,  we  observed  a case  of  poult  adoption  and  aban- 
donment of  an  active  nest.  In  wild  turkeys,  adoption 
of  poults  has  been  described  previously,  but  during  our 
observation  the  hen  also  abandoned  her  nest  at  a late 
stage  of  incubation.  Most  research  discussing  adoption 
in  gallinaceous  birds  has  focused  on  brood  abandon- 
ment after  hatch.  Although  poult  adoption  in  conjunc- 
tion with  nest  abandonment  is  probably  rare,  our  ob- 
servations indicate  that  it  can  occur,  at  least  in  Rio 


1 Dept,  of  Wildlife  and  Fisheries  Sciences,  Texas 
A&M  Univ.,  College  Station,  TX  77843-2258,  USA. 

2 Texas  Parks  and  Wildlife  Dept.,  Comfort,  TX 
78013,  USA. 

3 Texas  Parks  and  Wildlife  Dept.,  Mason,  TX  76856, 
USA. 

4 Corresponding  author;  e-mail:  bret@tamu.edu 


Grande  Wild  Turkeys.  Received  7 June  2005,  accepted 
16  February  2006. 


Species  such  as  gulls  ( Lams  spp.),  terns 
{Sterna  spp.),  and  geese  ( Branta  spp.)  readily 
adopt  offspring  (Pierottie  and  Murphy  1987, 
Saino  et  al.  1994,  Larsson  et  al.  1995).  North- 
ern Bobwhites  {Colinus  virginianus ) utilize 
brood  abandonment  and  adoption  as  a strategy 
for  increasing  nesting  opportunities  (Burger  et 
al.  1995,  DeMaso  et  al.  1997),  but  document- 
ed cases  of  gallinaceous  birds  adopting  off- 
spring are  rare  (Martin  1989,  Mills  and  Rum- 
ble 1991).  Adoption  of  poults  by  Merriam’s 
Wild  Turkeys  ( Meleagris  gallopavo  merriami) 
has  been  described  (Mills  and  Rumble  1991), 
and  Healy  (1992)  reported  nest  abandonment 


260 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


by  a captive  hen  that  was  attracted  to  the  calls 
of  another  brood.  In  May  2005,  we  observed 
a Rio  Grande  Wild  Turkey  (A/.  g.  intermedia ) 
hen  adopt  a poult  and  then  abandon  her  own 
nest  in  Kerr  County.  Texas.  To  our  knowledge, 
adoption  in  conjunction  with  nest  abandon- 
ment has  not  been  documented  before  in  the 
wild. 

As  part  of  a study  to  evaluate  the  reproduc- 
tive ecology  of  Rio  Grande  Wild  Turkeys  in 
Texas,  we  tracked  a radio-tagged  juvenile  hen 
through  two  nesting  attempts  on  the  Kerr 
Wildlife  Management  Area  in  Kerr  County 
(30°  04'  N,  99°  20'  W),  Texas.  On  11  April 
2005,  we  found  her  first  nest,  which  contained 
13  eggs,  and  we  estimated  nest  age  at  3 days. 
On  15  April,  the  nest  was  depredated,  and  the 
hen  subsequently  renested  on  28  April.  After 
28  April,  we  checked  the  hen’s  nesting  status 
>5  times  per  week.  On  7 May,  the  second  nest 
contained  12  eggs  and  nearby  we  set  up  an 
infrared  trail  camera  (Moultrie  Game  Spy®) 
to  monitor  the  nest.  From  8 to  21  May,  we 
never  observed  the  hen  off  the  nest,  and. 
based  on  our  intensive  tracking  of  the  hen, 
there  was  no  possibility  that  she  hatched  this 
poult  several  days  early. 

At  16:00  CST  on  21  May,  we  found  the  hen 
incubating  her  second  nest.  On  the  following 
day  at  1 1 :00,  we  located  the  hen  about  600  m 
from  the  nest.  We  approached  to  —15  m of 
the  hen  and  observed  her  bedded  down  in  a 
grassy  area  dominated  by  little  bluestem 
(Schizachyrium  scoparium ).  Upon  further  ap- 
proach. she  flushed.  Within  about  1 min,  a 
poult,  estimated  to  be  4 days  old,  ran  from  the 
grassy  area  where  the  hen  had  been  bedded. 
We  then  examined  the  hen’s  nest  and  found 
all  12  eggs  present  and  intact.  We  also  floated 
the  eggs  and  estimated  that  they  were  at  day 
23  of  incubation  (Healy  1992). 

On  23  May,  we  relocated  the  radio-tagged 
hen  in  an  effort  to  catch  and  radio-tag  the 
poult;  however,  the  hen  was  moving  and  we 
were  unable  to  locate  the  poult.  On  the  fol- 
lowing day,  the  hen  was  relocated  again,  this 
time  with  the  poult.  On  26  May,  we  captured 
the  poult,  estimated  its  age  as  9 days,  radio- 
tagged  it  with  a 1.2-g  poult  transmitter  (Bow- 
man et  al.  2002;  Advanced  Telemetry  Sys- 
tems, Isanti.  Minnesota),  and  released  it. 

Other  than  anecdotal  evidence  and  the  ar- 
ticle by  Mills  and  Rumble  (1991),  there  is  lit- 


tle available  information  on  the  frequency  of 
adoption  in  wild  turkeys.  Whereas  Mills  and 
Rumble  (1991)  reported  poult  adoption  by  tur- 
key hens  both  with  and  without  existing 
broods,  the  hen  we  observed  had  abandoned 
her  clutch  of  12  eggs  after  considerable  in- 
vestment (>20  days  of  incubation)  to  care  for 
a single  poult.  While  such  cases  of  abandon- 
ment and  adoption  are  probably  rare,  our  ob- 
servations indicate  that  it  can  occur  in  Rio 
Grande  Wild  Turkeys.  Possible  causes  might 
include  hen  physiological  condition  or  chang- 
es in  photoperiod  (Scanes  et  al.  1979,  Youn- 
gren  et  al.  1993.  Bedecarrats  et  al.  1997,  Sharp 
et  al.  1998).  The  hen  that  we  observed  was  in 
the  latter  stages  of  incubation  on  a second  nest 
when  the  adoption  event  occurred;  thus,  her 
levels  of  luteinizing  hormone  and  prolactin 
may  have  changed  sufficiently  to  promote  be- 
havioral changes  (i.e.,  poult-rearing  behavior 
in  preference  to  continued  incubation).  Addi- 
tional research  is  needed  to  clarify  what  might 
trigger  simultaneous  poult  adoption  and  nest 
abandonment  in  turkeys. 

ACKNOWLEDGMENTS 

Funding  for  this  project  was  provided  by  the  Texas 
Turkey  Stamp  Fund  through  the  Texas  Parks  and  Wild- 
life Department,  the  National  Wild  Turkey  Federation, 
and  the  Department  of  Wildlife  and  Fisheries  Sciences, 
Texas  A&M  University.  This  research  was  conducted 
under  Texas  A&M  University  Animal  Use  Permit  No. 
2005-005.  We  appreciate  the  comments  from  three 
anonymous  reviewers  that  improved  this  manuscript. 

LITERATURE  CITED 

Bedecarrats,  G..  D.  Guemene,  and  M.  A.  Richard- 
Yris.  1997.  Effects  of  environmental  and  social 
factors  on  incubation  behavior,  endocrinological 
parameters,  and  production  traits  in  turkey  hens 
( Meleagris  gallopavo).  Poultry  Science  76:1307- 
1314. 

Bowman,  K..  M.  C.  Wallace,  W.  B.  Ballard,  J.  H. 
Brunjes,  IV.  M.  S.  Miller,  and  J.  M.  Hellman. 
2002.  Evaluation  of  two  techniques  for  attaching 
radio  transmitters  to  turkey  poults.  Journal  of 
Field  Ornithology  73:276-280. 

Burger,  L.  W,  Jr.,  M.  R.  Ryan,  T.  V.  Dailey,  and  E. 
W.  Kurzejeski.  1995.  Reproductive  strategies, 
success,  and  mating  systems  of  Northern  Bob- 
whites  in  Missouri.  Journal  of  Wildlife  Manage- 
ment 59:417-426. 

DeMaso,  S.  J..  A.  D.  Peoples,  S.  A.  Cox,  and  E.  S. 
Parry.  1997.  Survival  of  Northern  Bobwhite 
chicks  in  western  Oklahoma.  Journal  of  Wildlife 
Management  61:846-853. 


SHORT  COMMUNICATIONS 


261 


Healy,  W.  M.  1992.  Behavior.  Pages  46-65  in  The 
Wild  Turkey:  biology  and  management  (J.  G. 
Dickson,  Ed.).  Stackpole  Books,  Mechanicsburg, 
Pennsylvania. 

Larsson,  K.,  H.  Tegelstrom,  and  P.  Forslund.  1995. 
Intraspecific  nest  parasitism  and  adoption  of 
young  in  the  Barnacle  Goose:  effects  on  survival 
and  reproductive  performance.  Animal  Behavior 
50:1349-1360. 

Martin,  K.  M.  1989.  Pairing  and  adoption  of  offspring 
by  replacement  male  Willow  Ptarmigan:  behavior, 
costs,  and  consequences.  Animal  Behavior  37: 
569-578. 

Mills,  T.  R.  and  M.  A.  Rumble.  1991.  Poult  adoption 
in  Merriam’s  Wild  Turkeys.  Wilson  Bulletin  103: 
137-138. 

Pierotti,  R.  and  E.  C.  Murphy.  1987.  Intergenera- 
tional  conflicts  in  gulls.  Animal  Behavior  35:435- 
444. 

Saino,  N.,  M.  Fasola,  and  E.  Crocicchia.  1994. 


Adoption  behavior  in  Little  and  Common  terns 
(Aves:  Sternidae):  chick  benefits  and  parent’s  fit- 
ness costs.  Ethology  97:294-309. 

Scanes,  C.  G.,  P.  J.  Sharp,  S.  Harvey,  P.  M.  M.  God- 
den,  A.  Chadwick,  and  W.  S.  Newcomer.  1979. 
Variations  in  plasma  prolactin,  thyroid  hormones, 
gonadal  steroids  and  growth  hormone  in  turkeys 
during  the  induction  of  egg  laying  and  molt  by 
different  photoperiods.  British  Poultry  Science  20: 
143-148. 

Sharp,  P.  J.,  A.  Dawson,  and  R.  W.  Lea.  1998.  Control 
of  luteinizing  hormone  and  prolactin  secretion  in 
birds.  Comparative  Biochemistry  and  Physiology 
C:  Pharmacology,  Toxicology  and  Endocrinology 
119:275-282. 

Youngren,  O.  M.,  M.  E.  Elhalawani,  J.  L.  Silsby, 
and  R.  E.  Phillips.  1993.  Effect  of  reproductive 
condition  on  luteinizing  hormone  and  prolactin  re- 
lease induced  by  electrical  stimulation  of  the  tur- 
key hypothalamus.  General  and  Comparative  En- 
docrinology 89:220-228. 


The  Wilson  Journal  of  Ornithology  1 1 8(2):26 1—263,  2006 


Predation  by  a Blue-crowned  Motmot  ( Momotus  momota) 
on  a Hummingbird 


J.  Mauricio  Garcfa-C.12  and  Rakan  A.  Zahawi1 2 


ABSTRACT. — We  describe  predation  of  a Green- 
crowned  Brilliant  ( Heliodoxa  jacula)  by  a Blue- 
crowned  Motmot  ( Momotus  momota)  in  southern  Cos- 
ta Rica.  We  did  not  witness  the  capture  of  the  hum- 
mingbird, but  did  observe  the  motmot  swallow  the 
prey  whole.  Although  the  diet  of  the  Blue-crowned 
Motmot  is  highly  variable  and  can  include  birds,  this 
is  the  first  report  of  predation  on  an  adult  humming- 
bird. Received  27  January  2005,  accepted  4 December 
2005. 


Members  of  the  family  Momotidae  have 
been  observed  eating  a wide  range  of  fruits, 
arthropods,  and  small  vertebrates  (Meyer  de 
Schauensee  1964,  Ridgely  and  Gwynne  1989, 
Stiles  and  Skutch  1989,  Karr  et  al.  1990,  Rem- 
sen  et  al.  1993).  Although  Remsen  et  al. 
(1993)  indicate  that  arthropods,  supplemented 
by  fruits,  are  the  more  important  component 
of  motmot  diets,  vertebrates  have  also  been 


1 Organization  for  Tropical  Studies,  P.O.  Box  676- 
2050,  San  Pedro,  San  Jose,  Costa  Rica. 

2 Corresponding  author;  e-mail:  mgarcia@ots.ac.cr 


found  in  the  stomachs  of  some  Momotidae 
species  (Wetmore  1968,  Stiles  and  Skutch 
1989).  Specifically,  motmots  have  been  ob- 
served eating  poison  dart  frogs  (Master  1999), 
snakes  (Stiles  and  Skutch  1989),  mice  (Del- 
gado-V.  and  Brooks  2003),  and  bats  (Chacon- 
Madrigal  and  Barrantes  2004). 

The  Blue-crowned  Motmot  ( Momotus 
momota),  found  throughout  the  lowlands  and 
middle  elevations  (to  —1,500  m)  of  Costa 
Rica  (Stiles  and  Skutch  1989),  forages  on 
large  spiders,  earthworms,  insects,  nestling 
birds,  and  small  snakes  and  lizards  (Stiles  and 
Skutch  1989,  Henderson  2002).  There  are, 
however,  no  known  accounts  of  motmots  eat- 
ing adult  birds.  Here,  we  describe  predation 
on  an  adult  hummingbird  by  a Blue-crowned 
Motmot. 

The  incident  occurred  on  the  morning  of  27 
February  2004  at  the  Las  Cruces  Biological 
Field  Station  (8°  47'  N,  82°  57'  W)  of  the  Or- 
ganization for  Tropical  Studies  in  San  Vito, 
Coto  Brus,  Puntarenas,  Costa  Rica  (elevation 


262 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


= 1,100  m,  annual  rainfall  = 3,988  mm).  (For 
a full  description  of  the  site,  see  Mintken  and 
Gunther  1991  and  Spencer  1991).  At  07:30 
CST,  we  observed  a motmot — perched  on  the 
cement  stairs  in  front  of  a station  building — 
with  a Green-crowned  Brilliant  ( Heliodoxa  ja- 
cula ) in  its  bill.  The  motmot  held  the  hum- 
mingbird by  its  body  and  repeatedly  beat  it 
against  the  cement.  The  hummingbird  ap- 
peared freshly  dead  and  was  easily  identifi- 
able. As  we  did  not  witness  the  capture,  the 
hummingbird  may  have  been  dead  or  injured 
prior  to  capture,  although  there  are  no  ac- 
counts of  motmots  eating  prey  they  did  not 
kill. 

At  07:35,  the  motmot  flew  to  the  ground 
~7  m away  and  continued  to  beat  the  hum- 
mingbird against  the  ground.  At  07:40,  it 
moved  under  a building  and  beat  the  hum- 
mingbird against  a rock  for  almost  1 min.  As 
a result,  most  of  the  hummingbird’s  feathers 
were  lost  and  its  bill  was  broken.  At  07:43, 
the  motmot  moved  out  from  under  the  build- 
ing to  a grassy  area  with  some  tree  cover  and 
continued  to  beat  the  hummingbird  against  the 
ground.  At  this  point,  the  motmot  was  7 m 
from  its  mate,  which  was  perched  on  a tree 
branch  2 m high  and  present  for  the  entire 
period;  it  did  not  make  any  attempt  to  move 
closer  to  the  motmot  with  the  hummingbird. 
The  motmot  never  used  its  feet  to  manipulate 
or  hold  the  prey;  the  entire  time  it  held, 
turned,  and  manipulated  the  hummingbird 
only  with  its  bill. 

At  07:54,  the  motmot  attempted,  but  failed, 
to  swallow  the  hummingbird  whole.  The  mot- 
mot threw  the  hummingbird  on  the  ground, 
picked  it  up  again  with  its  bill,  and  continued 
to  beat  it  against  the  ground.  At  07:56,  the 
motmot  again  tried  to  swallow  the  humming- 
bird and  was  successful.  It  held  the  humming- 
bird by  the  back  and  swallowed  it  back  end 
first.  The  motmot  then  flew  to  a tree  branch 
and  perched  near  its  mate. 

Reported  sources  of  adult  hummingbird 
mortality  include  arthropods  (e.g.,  Butler 
1949,  Hildebrand  1949,  Carignan  1988,  Gra- 
ham 1997),  frogs  (Monroe  1957),  and  several 
avian  taxa:  small  raptors  (e.g.,  Lowery  1938, 
Mayr  1966,  Stiles  1978),  Great  Shrike  Tyrants 
( Agriornis  livida;  Martinez  del  Rio  1992), 
Baltimore  Orioles  {Icterus  galbula ; Wright 
1962),  and  Dusky-green  Oropendolas  {Psar- 


ocolius  atrouirens;  Graves  1978).  Ours  is  the 
first  report  of  a Blue-crowned  Motmot  eating 
an  adult  bird  of  any  kind.  Our  observation  is 
best  explained  as  an  opportunistic  event  and 
broadens  the  range  of  predators  that  kill  and 
eat  hummingbirds. 

ACKNOWLEDGMENTS 

We  would  like  to  thank  K.  G.  Murray  for  helpful 
comments  on  earlier  drafts  of  the  manuscript.  T.  L. 
Master,  J.  V.  Remsen,  and  an  anonymous  reviewer 
made  invaluable  comments.  The  Organization  for 
Tropical  Studies  provided  logistical  support  at  Las 
Cruces  Biological  Station. 

LITERATURE  CITED 

Butler,  C.  1949.  Hummingbird  killed  by  praying 
mantis.  Auk  66:286. 

Carignan,  J.  M.  1988.  Predation  on  Rufous  Hum- 
mingbird by  praying  mantid.  Texas  Journal  of  Sci- 
ence 40:1 1 1. 

Chacon-Madrigal,  G.  and  G.  Barrantes.  2004. 
Blue-crowned  Motmot  {Momotus  momota ) pre- 
dation on  a long-tongued  bat  (Glossophaginae). 
Wilson  Bulletin  116:108-110. 

Delgado- V.,  C.  A.  and  D.  M.  Brooks.  2003.  Unusual 
vertebrate  prey  taken  by  Neotropical  birds.  Omi- 
tologia  Colombiana  1:63—65. 

Graham,  D.  L.  1997.  Spider  webs  and  windows  as 
potentially  important  sources  of  hummingbird 
mortality.  Journal  of  Field  Ornithology  68:98- 
101. 

Graves,  G.  R.  1978.  Predation  on  hummingbird  by 
oropendola.  Condor  80:251. 

Henderson,  C.  L.  2002.  Field  guide  to  the  wildlife  of 
Costa  Rica.  University  of  Texas  Press,  Austin. 
Hildebrand,  E.  M.  1949.  Hummingbird  captured  by 
praying  mantis.  Auk  66:286. 

Karr,  J.  R.,  S.  K.  Robinson,  J.  G.  Blake,  and  R.  O. 
Bierregaard,  Jr.  1990.  Birds  of  four  Neotropical 
forests.  Pages  237-269  in  Four  Neotropical  rain- 
forests (A.  H.  Gentry,  Ed.).  Yale  University  Press, 
New  Haven,  Connecticut. 

Lowery,  G.  H.,  Jr.  1938.  Hummingbird  in  a Pigeon 
Hawk’s  stomach.  Auk  55:280. 

Martinez  del  Rio,  C.  1992.  Great  Shrike-Tyrant  pre- 
dation on  a Green-backed  Firecrown.  Wilson  Bul- 
letin 104:368-369. 

Master,  T.  1999.  Predation  by  Rufous  Motmot  on 
black-and-green  poison  dart  frog.  Wilson  Bulletin 
111:439-440. 

Mayr,  E.  1966.  Hummingbird  caught  by  Sparrow 
Hawk.  Auk  83:644. 

Meyer  de  Schauensee,  R.  1964.  The  birds  of  Colom- 
bia, and  adjacent  areas  of  South  and  Central 
America.  Livingston  Publishers,  Narberth,  Penn- 
sylvania. 

Mintken,  J.  and  B.  Gunther.  1991.  The  Wilson  Bo- 
tanical Garden.  Principes  35:124-126. 


SHORT  COMMUNICATIONS 


263 


Monroe,  M.  1957.  Hummingbird  killed  by  frog.  Con- 
dor 59:69. 

Remsen,  J.  V.,  M.  A.  Hyde,  and  A.  Chapman.  1993. 
The  diets  of  Neotropical  trogons,  motmots,  bar- 
bets  and  toucans.  Condor  95:178-192. 

Ridgely,  R.  S.  and  J.  A.  Gwynne,  Jr.  1989.  A guide 
to  the  birds  of  Panama  with  Costa  Rica,  Nicara- 
gua, and  Honduras,  2nd  ed.  Princeton  University 
Press,  Princeton,  New  Jersey. 

Spencer,  D.  1991.  Tropical  temptation  in  Costa  Rica. 
Cornell  Plantations  46:12-15. 


Stiles,  F.  G.  1978.  Possible  specialization  for  hum- 
mingbird-hunting in  the  Tiny  Hawk.  Auk  95:550- 
553. 

Stiles,  F.  G.  and  A.  F.  Skutch.  1989.  A guide  to  the 
birds  of  Costa  Rica.  Cornell  University  Press,  Ith- 
aca, New  York. 

Wetmore,  A.  1968.  The  birds  of  the  Republic  of  Pan- 
ama, part  2.  Columbidae  (pigeons)  to  Picidae 
(woodpeckers).  Smithsonian  Institution  Press, 
Washington,  D.C. 

Wright,  B.  S.  1962.  Baltimore  Oriole  kills  humming- 
bird. Auk  79:112. 


The  Wilson  Journal  of  Ornithology  1 18(2):264-266,  2006 


Once  Upon  a ‘Time  in  American  Ornithology 


Alexander  Wilson,  namesake  of  The  Wilson 
Journal  of  Ornithology,  was  bom  on  6 July 
1766  in  Scotland.  There,  he  trained  and  worked 
as  a weaver  (and  a poet).  In  1794,  he  emigrated 
to  the  U.S.,  and  for  9 years  he  worked  as  a 
teacher.  His  own  education  had  been  sketchy, 
however,  and  he  had  to  study  to  teach.  Even- 
tually, North  America’s  birds  and  wilderness 
held  him — there  were  more  birds  and  species 
than  in  his  native  Scotland.  Eager  to  begin  a 
life  work,  Wilson  set  out  on  1 June  1803  to 
draw  “all  the  finest  birds  of  America.”  For  the 
next  10  years,  he  wrote  and  illustrated  his  sem- 
inal work,  American  Ornithology,  the  first  sci- 
entific treatment  of  American  birds  and  the  first 
to  stress  natural  history  and  field  biology.  He 
had  an  untaught  skill  in  painting,  but  William 
Bartram,  America’s  foremost  naturalist  and  a 
neighbor  in  his  hometown  of  Philadelphia, 
taught  him  how  to  draw.  Bartram  also  an- 
swered Wilson’s  natural  history  queries,  and  in- 
spired and  instructed  him  in  ornithology,  bot- 
any, and  bird  illustration. 

Wilson  was  still  teaching  in  1805,  but  art 
and  science  dominated  his  thoughts  as  his 
drawing  improved.  He  took  a job  as  an  assis- 
tant editor  with  a publishing  house  in  1806 
and  ultimately  convinced  the  publisher  to  sup- 
port his  developing  work,  but  only  if  Wilson 
could  get  commitments  from  250  subscribers 
at  $120  each.  On  7 April  1807,  a brochure  for 
American  Ornithology  was  sent  to  2,500  of 
the  most  eminent  people  in  the  U.S. 

As  time  allowed,  Wilson  traveled  exten- 
sively, widening  his  knowledge  of  birdlife  and 
gathering  information  on  the  distribution, 
nesting  habits,  and  movements  of  North 
American  birds.  He  often  traveled  on  foot  or 
by  horseback,  and  while  accumulating  bird 
lore,  always  equipped  himself  with  a shotgun, 
paint,  paper,  sketching  materials,  and  a note- 
book. Wilson  made  four  great  adventures — 
through  dense  forests  and  swamps,  across  In- 
dian territory,  and  in  all  seasons — traversing 
every  state  in  the  Union,  often  alone,  in  search 
of  birds  and  subscribers.  On  his  first  trip,  Oc- 
tober to  December  1804,  he  traveled  1,300 


miles  from  Philadelphia  to  Niagara  Falls  and 
back,  mostly  on  foot,  but  also  by  stagecoach, 
skiff,  and  sloop.  Then,  in  September  1808,  he 
was  off  to  New  England  in  search  of  birds  and 
subscribers  willing  to  commit  $120  for  his 
American  Ornithology.  During  the  winter  of 
1808-1809,  he  continued  his  fieldwork  and 
search  for  subscribers,  traveling  south  by 
horseback  to  Maryland,  New  Jersey,  Virginia, 
North  and  South  Carolina,  and  Georgia.  Wil- 
son’s longest  expedition  began  in  January  of 
1810,  when  he  went  from  Philadelphia  to 
Pittsburgh,  then  south  on  the  Ohio  River  to 
Louisville  in  a skiff  (that  Wilson  christened 
“Ornithologist”),  then  overland  to  Natchez, 
through  hostile  Chickasaw  Indian  territory, 
and  finally  on  to  New  Orleans. 

The  list  of  subscribers  to  Wilson’s  American 
Ornithology  included  some  of  the  greatest  per- 
sonalities of  his  time:  President  Thomas  Jef- 
ferson. Robert  Fulton  (inventor  of  the  first 
commercial  steamship),  and  Thomas  Paine. 
Wilson  also  enlisted  the  assistance  of  Meri- 
wether Lewis,  who  provided  bird  specimens — 
collected  during  his  remarkable  1804-1806  ex- 
pedition with  William  Clark — from  which  Wil- 
son could  draw  birds  of  western  origin. 

Wilson  died  of  dysentery  at  the  age  of  47 
in  1813,  just  before  publication  of  the  8th  vol- 
ume of  American  Ornithology.  The  9th  and 
last  volume  was  compiled  by  George  Ord 
from  Wilson’s  notes  and  drawings. 

It  was  on  a trip  through  the  southern  coastal 
states  that  Wilson  recorded  the  following  orni- 
thological observation.  On  2 February  1809,  12 
miles  outside  Wilmington,  North  Carolina,  he 
collected  two  Ivory-billed  Woodpeckers  ( Cam - 
pephilus  principalis ),  and  slightly  wounded  a 
third  (a  male).  Wilson’s  illustrations  of  the  Ivo- 
ry-billed Woodpecker  (Fig.  1)  in  his  American 
Ornithology  came  from  drawings  he  made  of 
the  injured  bird  while  in  his  Wilmington  hotel 
room.  The  original  reference  is:  Brewer,  T.  M. 
1840.  Wilson’s  American  Ornithology,  with 
notes  by  Jardine.  Otis,  Broaders,  and  Co.,  Bos- 
ton, Massachusetts. — JAMES  A.  SEDGWICK; 
e-mail:  jim_sedgwick@usgs.gov 


264 


ONCE  UPON  A TIME  IN  AMERICAN  ORNITHOLOGY 


265 


FIG.  1.  Wilson’s  Ivory-billed  Woodpecker  (top  right,  bottom  center),  Pileated  Woodpecker  (top  and  bottom 
left),  and  Red-headed  Woodpecker  (bottom  right).  Illustrations  of  the  Ivory-billed  Woodpecker  were  drawn  from 
a live  bird  that  Wilson  took  to  his  hotel  room  in  Wilmington,  North  Carolina,  in  1809.  Color  plate  from:  Wilson, 
A.  1829.  American  ornithology;  or.  The  natural  history  of  the  birds  of  the  United  States.  Collins  & Co.,  New 
York.  Image  courtesy  of  the  Josselyn  Van  Tyne  Memorial  Library,  University  of  Michigan,  Ann  Arbor. 


266 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 . No.  2.  June  2006 


The  first  place  I observed  this  bird  at,  when  on  my  way  to  the  south,  was  about 
twelve  miles  north  of  Wilmington  in  North  Carolina.  There  I found  the  bird  from 
which  the  drawing  of  Fig.  131  was  taken.  This  bird  was  only  wounded  slightly  in 
the  wing,  and.  on  being  caught,  uttered  a loudly  reiterated  and  most  piteous  note, 
exactly  resembling  the  violent  crying  of  a young  child;  which  terrified  my  horse 
so,  as  nearly  to  have  cost  me  my  life.  It  was  distressing  to  hear  it.  I carried  it  with 
me  in  the  chair,  under  cover,  to  Wilmington.  In  passing  through  the  streets,  its 
affecting  cries  surprised  every  one  within  hearing,  particularly  the  females,  who 
hurried  to  the  doors  and  windows  with  looks  of  alarm  and  anxiety.  I drove  on,  and, 
on  arriving  at  the  piazza  of  the  hotel,  where  I intended  to  put  up.  the  landlord  came 
forward,  and  a number  of  other  persons  who  happened  to  be  there,  all  equally 
alarmed  at  what  they  heard;  this  was  greatly  increased  by  my  asking,  whether  he 
could  furnish  me  with  accommodations  for  myself  and  my  baby.  The  man  looked 
blank  and  foolish,  while  the  others  stared  with  still  greater  astonishment.  After 
diverting  myself  for  a minute  or  two  at  their  expense,  I drew  my  Woodpecker  from 
under  the  cover,  and  a general  laugh  took  place.  I took  him  up  stairs,  and  locked 
him  up  in  my  room,  while  I went  to  see  my  horse  taken  care  of.  In  less  than  an 
hour,  I returned,  and,  on  opening  the  door,  he  set  up  the  same  distressing  shout, 
which  now  appeared  to  proceed  from  grief  that  he  had  been  discovered  in  his 
attempts  at  escape.  He  had  mounted  along  the  side  of  the  window,  nearly  as  high 
as  the  ceiling,  a little  below  which  he  had  begun  to  break  through.  The  bed  was 
covered  with  large  pieces  of  plaster;  the  lath  was  exposed  for  at  least  fifteen  inches 
square,  and  a hole,  large  enough  to  admit  the  fist,  opened  to  the  weather-boards; 
so  that,  in  less  than  another  hour,  he  would  certainly  have  succeeded  in  making  his 
way  through.  I now  tied  a string  round  his  leg,  and,  fastening  it  to  the  table,  again 
left  him.  I wished  to  preserve  his  life,  and  had  gone  off  in  search  of  suitable  food 
for  him.  As  I reascended  the  stairs,  I heard  him  again  hard  at  work,  and  on  entering 
had  the  mortification  to  perceive  that  he  had  almost  entirely  ruined  the  mahogany 
table  to  which  he  was  fastened,  and  on  which  he  had  wreaked  his  whole  vengeance. 
While  engaged  in  taking  the  drawing,  he  cut  me  severely  in  several  places,  and, 
on  the  whole,  displayed  such  a noble  and  unconquerable  spirit,  that  I was  frequently 
tempted  to  restore  him  to  his  native  woods.  He  lived  with  me  nearly  three  days, 
but  refused  all  sustenance,  and  I witnessed  his  death  with  regret. 


The  Wilson  Journal  of  Ornithology  1 1 8(2):267-276,  2006 


Ornithological  Literature 

Compiled  by  Mary  Gustafson 


BIRDS  OF  BELIZE.  By  H.  Lee  Jones,  illus- 
trated by  Dana  Gardner.  University  of  Texas 
Press,  Austin.  2003:  317  pp.,  56  color  plates 
with  facing-page  figure  captions,  234  range 
maps,  28  numbered  figures.  ISBN:  0292740662, 
$60.00  (cloth).  ISBN:  0292701640,  $34.95  (pa- 
per).— Being  a country  where  English  is  spo- 
ken, and  which  still  retains  70%  of  its  native 
habitat,  it  is  no  surprise  that  Belize  is  an  increas- 
ingly popular  destination  for  ornithologists  and 
birders  alike.  In  fact,  hundreds  of  birders  visit 
this  tiny  country  annually  to  enjoy  its  rich  avi- 
fauna, natural  beauty,  and  amazingly  friendly 
residents.  For  the  past  decade,  ornithologists 
and  birders  visiting  Belize  were  served  quite 
well  by  Howell  and  Webb’s  A Guide  to  the 
Birds  of  Mexico  and  Northern  Central  Amer- 
ica (Oxford  University  Press,  1995).  As  mas- 
terful as  that  work  is,  however,  the  85 1 -page 
tome  weighs  in  at  a hefty  3.4  lbs.,  a bit  much 
to  carry  in  the  field.  A much  more  portable, 
but  clearly  outdated,  option  is  Peterson  and 
Chalif’s  A Field  Guide  to  Mexican  Birds 
(Houghton  Mifflin,  1973).  Now  there  is  a third 
option:  Birds  of  Belize  is  the  first  guide  to 
comprehensively  cover  all  574  species  known 
to  occur  in  this  birder-friendly  country.  All 
regularly  occurring  species  are  illustrated,  in- 
cluding North  American  migrants  that  spend 
only  part  of  the  year  in  Belize.  Neither  Peter- 
son and  Chalif  nor  Howell  and  Webb  illustrate 
North  American  migrants,  and  both  guides  in- 
clude many  Mexican  species  that  do  not  occur 
in  Belize.  For  anyone  who  is  not  thoroughly 
familiar  with  bird  distribution  in  Central 
America,  the  convenience  of  having  only  Be- 
lizean birds  in  one  volume  is  difficult  to  over- 
state. Birds  of  Belize  is  also  two-thirds  the 
weight  of  Howell  and  Webb,  though  still  a bit 
large  to  easily  tote  in  the  field.  The  guide’s 
format  is  traditional  and  easy  to  use,  with 
plates  and  brief,  facing-page  text  in  the  front; 
more-comprehensive  text  and  detailed  maps 
are  in  the  back.  The  facing-page  text  covers 
not  only  identification  notes,  but  also  the  spe- 
cies’ status,  distribution,  and  habitat — incred- 
ibly useful  information  found  in  few  other 


guides.  Probably  the  greatest  strength  of  this 
guide  is  its  superb,  authoritative  text.  Lee 
Jones’s  knowledge  about  the  birds  of  Belize 
is  unsurpassed.  He  gives  excellent  descrip- 
tions of  status,  distribution,  and  general  iden- 
tification features  for  each  species.  His  notes 
on  habitat  are  particularly  helpful  for  anyone 
seeking  a particular  species,  and  his  descrip- 
tions of  vocalizations  are  unusually  complete, 
accurate,  and  helpful.  For  those  seeking  more 
in-depth  information  on  a particular  subject,  a 
comprehensive  bibliography  is  available. 

For  better  or  worse,  the  quality  of  a field 
guide  depends,  to  a large  degree,  on  the  qual- 
ity of  its  illustrations.  The  illustrations  in  this 
guide  are  attractive  and,  in  most  cases,  more 
than  adequate  to  convey  the  important  iden- 
tifying characters.  In  general.  Neotropical  res- 
ident species  are  better  illustrated  than  North 
American  migrants;  plates  of  antbirds,  wrens, 
becards,  and  tanagers  are  particularly  lovely 
and  accurate.  For  some  of  the  more  difficult 
ID  questions,  however,  the  illustrations  fall 
short  and  other  sources  may  need  to  be  con- 
sulted. For  example,  all  the  raptors  in  flight 
are  misshapen  and  the  plumage  markings  of 
many  are  incorrect.  Those  in  Howell  and 
Webb  are  far  superior.  Likewise,  the  Leptotila 
doves,  which  are  best  identified  by  general 
color  pattern,  look  too  similar;  Gray-fronted 
Dove  (L.  rufaxilla ) should  be  more  rufous- 
brown  above  with  contrasting  gray  nape  and 
head;  White-tipped  Dove  (L.  verreauxi ) 
should  be  more  gray-brown;  and  Gray-chested 
Dove  (L.  cassini ) should  have  a more  con- 
trasting gray  breast.  Again,  those  in  Howell 
and  Webb  are  much  better. 

For  many  of  the  North  American  migrants, 
such  as  shorebirds,  gulls,  and  terns,  one  would 
be  much  better  served  by  consulting  some  of 
the  better  North  American  references  such  as 
The  Sibley  Guide  (Alfred  A.  Knopf,  2000).  In- 
deed, the  juvenile  Red-footed  Booby  ( Sula 
sula ),  the  small  Calidris  sandpipers,  the  Com- 
mon ( Sterna  hirundo)  and  Roseate  (S.  doug- 
allii)  terns,  the  Empidonax  flycatchers,  the  im- 
mature Cape  May  Warbler  ( Dendroica  tigri- 


267 


268 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


na ),  and  the  basic-plumaged  Palm  Warbler  ( D . 
palmarum ) are  probably  not  identifiable  from 
their  illustrations  in  Birds  of  Belize ; the  ju- 
venile Yellow-crowned  Night-Heron  (Nyctan- 
assa  violacea ) should  show  pale-edged  wing 
coverts  and  a black  bill;  the  juvenile  Black- 
crowned  Night- Heron  ( Nycticorax  nycticorax ) 
has  shorter  legs  than  those  portrayed;  the 
Myiarchus  flycatchers  are  too  dark  and  small- 
headed with  incorrect  wing  patterns  (see  Sib- 
ley or  Howell  and  Webb  for  better  illustra- 
tions). The  shapes  are  a bit  off  on  many  spe- 
cies; note  especially  that  the  Clay-colored 
(Spizella  pallida ),  Chipping  {S.  passerina ), 
Lincoln’s  ( Melospiza  lincolnii ),  and  Savannah 
(Passerculus  sandwichensis)  sparrows  are  all 
shown  with  similar  proportions,  including 
identical  tail  lengths.  In  life,  these  species  dif- 
fer markedly  in  proportions  (see  Sibley). 
Countless  other  small  mistakes  make  some  of 
the  illustrations  less  useful  than  they  could  be. 

The  text  has  a few  minor  shortcomings. 
Whereas  habitats  are  nicely  described,  there  is 
little  or  nothing  about  habits  of  birds;  how 
they  move,  how  they  feed,  what  they  eat, 
whether  they  are  easy  or  hard  to  see,  how  they 
nest.  Although  such  information  may  be 
somewhat  limited  for  many  Neotropical  spe- 
cies, what  is  known  for  any  one  species  could 
have  been  included  in  a few  short  lines  with- 
out making  the  book  much  larger — particular- 
ly since  the  line  spacing  was  larger  than  it 
needed  to  be.  It  is  also  unfortunate  that  the 
text  was  printed  on  heavy,  glossy  paper,  which 
added  unnecessary  weight  and  thickness  to 
the  book. 

Several  aspects  of  the  guide’s  layout  could 
have  been  improved.  Most  notably,  bird  sizes 
should  have  been  indicated  on  the  plates.  Size, 
after  all,  is  a critical  starting  point  in  the  iden- 
tification process.  Also,  it  is  impossible  to  go 
quickly  from  the  plates  to  the  maps.  One  must 
go  from  the  plates  to  the  text  to  find  out  what 
page  the  map  is  on.  The  maps  themselves  are 
a bit  confusing.  Supposedly,  range  maps  for 
species  that  occur  throughout  Belize  are  not 
included,  which  undoubtedly  saves  space  but 
may  be  confusing  for  someone  not  familiar 
with  the  birds  of  the  region.  Plus,  some  maps 
for  colonial  waterbirds  are  misleading.  For  ex- 
ample, the  Great  Egret  ( Ardea  alba ) map  is 
illustrated  with  four  dots  indicating  the  loca- 
tions of  breeding  colonies,  yet  there  is  no  in- 


dication of  where  foraging  birds  occur  outside 
(or  during,  for  that  matter)  the  breeding  sea- 
son. Other  species,  such  as  Red-footed  Booby, 
which  clearly  has  a more  limited  nonbreeding 
distribution  than  Great  Egret,  are  mapped  in  a 
similar  way.  Rarities  for  which  there  are  few 
records  are  not  mapped,  which  is  also  under- 
standable; however,  a number  of  species  that 
occur  regularly  in  parts  of  Belize,  such  as 
Black-crested  Coquette  ( Lophornis  helenae), 
are  not  mapped.  The  migration  distribution  is 
not  mapped  for  any  species,  though  it  certain- 
ly would  have  been  helpful. 

Although  less  than  perfect.  Birds  of  Belize 
is  still  an  attractive,  authoritative,  and  very 
useful  guide.  Its  positive  attributes  far  out- 
weigh its  shortcomings.  It  serves  as  a handy 
reference  for  Belizean  birds  and  is  recom- 
mended as  the  guide  of  choice  to  most  birders 
visiting  this  splendid  country. — MICHAEL 
O BRIEN,  WINGS,  Inc.,  West  Cape  May, 
New  Jersey;  e-mail;  tsweet@comcast.net 


ARIZONA  BREEDING  BIRD  ATLAS. 
Edited  by  Troy  E.  Corman  and  Cathryn  Wise- 
Gervais.  University  of  New  Mexico  Press,  Al- 
buquerque. 2005;  646  pp.,  5 figures,  12  tables, 
336  photographs  (53  habitat  photos,  281  bird 
photos),  281  maps,  270  habitat  charts,  194 
phenology  graphs.  ISBN;  0826333796. 
$45.00  (cloth). — State  breeding  bird  atlases 
get  better  and  better.  Arizona’s  raises  the  stan- 
dard once  again.  Authoritative  species  ac- 
counts, illustrated  with  generous  use  of  color, 
make  presentation  of  data  thorough,  clear,  and 
vivid.  Atlas  workers  (atlasers)  recorded  283 
breeding  species,  plus  19  potential  breeders. 
The  270  main  species  accounts  brim  with  at- 
las-derived information,  more  than  many  state 
atlases  provide. 

Each  2-page  account  features  the  usual  state 
map,  with  easy-to-discem  color-coding  to  de- 
pict the  three  confidence  levels  portraying  the 
likelihood  of  breeding  within  a given  atlas 
block.  The  block  statistics  summarize  the 
number  of  priority  blocks  and  topographic 
quads  (1  ;74,000-scale  maps)  in  which  field 
workers  recorded  the  species.  Color  photo- 
graphs supply  the  obligatory  depiction  of 
birds  in  the  species  accounts.  Each  account 
also  includes  two  informative  charts;  a breed- 


ORNITHOLOGICAL  LITERATURE 


269 


ing  phenology  chart  (for  species  with  ade- 
quate data)  and  a graph  depicting  habitat  use. 

Arizona’s  atlas  project  specified  40  habitat 
types  within  seven  habitat  landscapes  (tundra, 
forests  and  woodlands,  scrublands,  grasslands, 
desert  lands,  wetlands,  and  urban/agricultur- 
al). Illustrated  with  color  photographs,  a pre- 
liminary chapter  on  habitat  describes  each  of 
the  40  habitats  and  reports  on  status  and  dis- 
tribution. Many  habitats,  especially  those  in 
desert  systems,  suffer  declines  attributable  to 
human  activities  and  exacerbated  by  Arizona’s 
burgeoning  population. 

Unlike  some  other  atlases  in  which  phenol- 
ogy charts  report  the  range  of  dates  in  which 
atlasers  recorded  each  stage  of  breeding  (i.e., 
atlas  breeding  phenology  codes),  phenology 
charts  in  the  Arizona  Atlas  simply  report  over- 
all breeding  activity.  The  atlas  also  highlights 
an  interesting  facet  of  Arizona  bird  life — the 
summer  “monsoon”  season  in  July  and  Au- 
gust. Monsoons  stimulate  second  nestings  by 
such  species  as  Canyon  Towhee  ( Pipilo  fus- 
cus ),  Rufous-crowned  Sparrow  ( Aimophila  ruf- 
iceps ),  Eastern  Meadowlark  ( Sturnella  mag- 
na),  and  maybe  Common  Yellowthroat 
( Geothlypis  trichas ),  as  well  as  the  first  and 
only  nestings  by  Cassin’s  ( Aimophila  cassi- 
nii ),  Botteri’s  ( Aimophila  botterii ) and  Grass- 
hopper ( Ammodramus  savannarum)  sparrows. 
Varied  Bunting  ( Passerina  versicolor ),  and 
possibly  Lazuli  Bunting  ( Passerina  amoena). 

Three  topics  organize  the  species  accounts: 
Habitat,  Breeding,  and  Distribution  and  Sta- 
tus. Under  Habitat,  authors,  referring  briefly 
to  habitat  preferences  reported  by  previous  au- 
thors, analyze  the  principal  habitats  in  which 
atlasers  found  the  species.  The  Breeding  sec- 
tion leads  with  short  expositions  about  breed- 
ing biology,  often  derived  from  the  Birds  of 
North  America  series,  and  compares  these  pre- 
cepts with  atlas  observations. 

The  Distribution  and  Status  section  reports 
on  the  species’  seasonal  status,  and  then  com- 
pares atlas  findings  with  previous  works  on 
Arizona,  particularly  the  seminal  work  by  Al- 
lan Phillips,  Joe  Marshall,  and  Gale  Monson, 
The  Birds  of  Arizona  (University  of  Arizona 
Press,  1964)  and — for  species  occurring  pri- 
marily in  Mexico — The  Birds  of  Sonora  by 
Steve  Russell  and  Gale  Monson  (University  of 
Arizona  Press,  1990).  The  discussion  details 
where  and  with  what  frequency  field  workers 


detected  the  species,  provides  comments  on  its 
detectability,  and  concludes  with  an  analysis 
of  the  species  status  and  conservation  stand- 
ing. 

One  slightly  distracting  theme  in  many  spe- 
cies accounts  involves  a small  section  of  Ar- 
izona where  the  Apache  Nation  refused  atlas- 
ers access  to  tribal  lands.  Their  section  of  the 
White  Mountains  (east-central  Arizona)  con- 
tains one  of  Arizona’s  few  areas  of  high-ele- 
vation habitat.  Authors  of  species  accounts 
frequently  lament  the  lack  of  coverage  in  the 
missing  priority  blocks  (30  out  of  1,834)  and 
often  project  species’  likely  ranges  in  the 
missing  blocks. 

Some  species  accounts  contain  a unique 
feature:  measurements  of  nest-site  character- 
istics. Field  workers  measured  or  described 
characteristics  of  3,507  nests  of  184  species, 
including  nest  height  and  nest  tree.  For  ex- 
ample, atlasers  found  121  Phainopepla 
( Phainopepla  nitrens ) nests  in  17  tree  species 
(almost  half  in  palo  verde,  Parkinsonia  sp.)  at 
a median  height  of  2.4  m (range  = 1—10  m). 

Nineteen  authors  contributed  species  ac- 
counts, although  the  editors  wrote  most  of 
them.  They  follow  an  admirably  consistent 
style  with  comparable  contents,  although  an 
editor’s  eye  might  pick  out  a few  grammatical 
goofs  (e.g.,  hanging  participial  phrases  that 
most  readers  will  not  notice)  and  a few  typos. 

The  first  part  of  the  book  discusses  the  de- 
tails of  atlas  organization,  methods,  limita- 
tions and  biases,  and  summarizes  the  results. 
One  chapter  covers  geography,  climate,  and 
habitats,  and  another  offers  a brief  history  of 
Arizona  ornithologists.  (The  first  recorded 
bird  observations  came  from  Coronado’s  ex- 
pedition in  1540-1542,  although  we  do  not 
learn  what  he  claimed  to  see.)  It  concludes  by 
quoting  Elliot  Coues’  sharing  “a  sort  of  char- 
itable pity  for  the  rest  of  the  poor  world,  who 
are  not  ornithologists,  and  have  not  the  chance 
of  pursuing  the  science  in  Arizona.” 

The  summary  of  results  appropriately  starts 
by  recognizing  the  710  field  workers  (those 
who  surveyed  one  or  more  atlas  blocks)  and 
422  block  helpers,  who  put  in  51,737  hr  of 
field  work  (plus  18,119  hr  of  travel  time). 
Blocks  with  the  most  species  are  distributed 
along  a northwesterly  line  from  the  south- 
eastern corner  of  Arizona  to  the  center  of  the 
state,  from  the  Chiricahua  and  Huachuca 


270 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


mountains  north  to  the  White  Mountains,  and 
west  along  the  Mogollon  Rim  as  far  as  Pres- 
cott. Mourning  Dove  ( Zenaida  macroura ) 
heads  the  list  of  species  reported  in  the  most 
blocks,  followed  by  Ash-throated  Flycatcher 
( Myiarchus  cinerascens),  House  Finch  ( Car - 
podacus  mexicanus).  Common  Raven  ( Corvus 
corax).  Red-tailed  Hawk  ( Buteo  jamaicensis ), 
Northern  Mockingbird  ( Mimus  polyglottos). 
Black-throated  Sparrow  ( Amphispiza  bilinea- 
ta ),  and  Brown-headed  Cowbird  ( Molothrus 
ater).  Arizona  specialties  in  the  top  21  include 
Cactus  Wren  ( Campylorhynchus  brunneicap- 
illus),  Phainopepla,  and  Verdin  {Auriparus 
flaviceps). 

In  many  states,  atlas  field  workers  have  suc- 
ceeded in  surveying  remote  and  rugged  re- 
gions that  avian  researchers  ordinarily  do  not 
study.  In  Arizona,  their  efforts  have  expanded, 
or  filled  in,  the  known  ranges  of  many  species. 
In  contrast,  they  have  also  identified  several 
declining  species,  including  Buff-breasted 
Flycatcher  ( Empidonax  fulvifrons),  American 
Dipper  ( Cinclus  mexicanus),  and  Evening 
Grosbeak  ( Coccothraustes  vespertinus).  Many 
species  accounts  detail  declines  due  to  habitat 
destruction — especially  the  loss  of  saguaros 
( Carnegiea  gigantea)  felled  by  wildfires  and 
urbanization.  Atlas  results  show  that  a sur- 
prising number  of  species  have  a limited  range 
in  Arizona — aside  from  the  Mexican  species 
that  occasionally  wander  northward  into  the 
southeastern  mountains.  Overall,  this  atlas 
provides  fascinating,  thorough,  accessible  in- 
formation about  Arizona’s  unique  breeding 
avifauna— HUGH  E.  KINGERY,  Franktown, 
Colorado;  e-mail:  ouzels@juno.com 


NESTING  BIRDS  OF  A TROPICAL 
FRONTIER:  THE  LOWER  RIO  GRANDE 
VALLEY  OF  TEXAS.  By  Timothy  Brush. 
Texas  A&M  University  Press,  College  Sta- 
tion. 2005:  245  + xiv  pp.,  31  color  photo- 
graphs, 1 1 color  illustrations,  2 tables,  5 maps. 
ISBN:  1585444367,  $50.00  (cloth).  ISBN: 
1585444901,  $24.95  (paper).— The  Lower 
Rio  Grande  Valley  of  Texas  is  well  known  to 
ornithologists  and  birders  alike  who  have  an 
interest  in  the  avifauna  of  the  United  States. 
Many  species  of  birds  with  a more  tropical 
distribution  reach  the  northern  portion  of  their 


range  in  southern  Texas,  and  the  Valley,  as  it 
is  often  referred  to,  offers  easy  accessibility  to 
the  habitats  that  these  birds  occupy.  The  geo- 
graphic area  covered  includes  the  four  south- 
ern-most counties  in  Texas:  Cameron,  Hidal- 
go, Starr,  and  Willacy.  The  two  eastern  coun- 
ties— Cameron  and  Willacy — and  southern 
Hidalgo  County  are  part  of  the  recently 
formed  delta  of  the  Rio  Grande;  thus,  the  land 
use  is  largely  devoted  to  row-crops.  For  a va- 
riety of  reasons,  the  remainder  of  the  Valley 
is  less  conducive  to  agriculture  and,  histori- 
cally, ranching  has  been  the  primary  industry. 
During  the  past  2 decades,  the  human  popu- 
lation in  these  four  counties  has  steadily  in- 
creased and  subsequent  urbanization  is  readily 
apparent.  Conservation  agencies,  both  public 
and  private,  have  made  great  efforts  to  protect 
remaining  patches  of  native  vegetation,  partic- 
ularly in  the  eastern  half  of  the  Valley.  These 
four  counties  cover  approximately  1 .2  million 
ha  and  can  boast  an  avifauna  of  just  over  500 
documented  species. 

As  the  title  states,  this  book  focuses  on  the 
breeding  avifauna  of  the  Lower  Rio  Grande 
Valley  of  Texas.  The  majority  of  the  book 
takes  a narrative  format  that  is  easy  to  read 
and  discusses  all  species  that  either  breed  reg- 
ularly or  occasionally  within  the  area.  At  the 
beginning  of  the  book,  there  is  a short  section 
that  includes  color  photos  of  selected  species 
as  well  as  several  habitat  shots.  Compelling 
among  these  are  aerial  photos  of  Santa  Ana 
National  Wildlife  Refuge  taken  prior  to  the 
construction  of  Falcon  Dam  (in  1953)  and  in 
1981  to  compare  changes  in  the  condition  of 
the  Rio  Grande  and  surrounding  land  use.  The 
remainder  of  the  color  section  includes  several 
paintings  by  Gerald  Sneed  depicting  various 
nesting  birds  of  the  Valley.  These  paintings 
provide  something  that  photos  can’t  convey, 
the  feeling  of  being  in  the  Valley’s  natural 
habitats. 

The  introductory  chapters  provide  a base- 
line understanding  of  the  Lower  Rio  Grande 
Valley.  There  are  overviews  of  topography 
and  climate,  as  well  as  an  interesting  historical 
perspective  of  land  use  and  its  effect  on  eco- 
logical diversity.  Two  of  the  remaining  chap- 
ters in  the  introductory  section  include  a brief 
discussion  of  the  basic  habitats  found  in  the 
study  area  and  seasonal  changes  in  the  avifau- 
na. A highlight  of  the  book  is  the  extensive 


ORNITHOLOGICAL  LITERATURE 


271 


References  section,  which  will  be  a great  help 
to  anyone  working  on  the  avifauna  of  South 
Texas.  The  bulk  of  the  book  is  composed  of 
species  accounts. 

The  accounts  include  all  species  (171)  for 
which  there  is  at  least  one  acceptable  breeding 
record.  At  the  time  of  writing,  Eurasian  Col- 
lared-Doves  ( Streptopelia  decaocto ) were  just 
beginning  to  arrive  in  the  Valley,  but  have 
now  taken  hold  and  can  be  added  to  that  list. 
As  might  be  expected,  the  lengths  of  species 
accounts  vary  greatly.  Brush  gives  extended 
coverage  to  species  that  are  South  Texas  spe- 
cialties and  other  species  that  may  be  of  par- 
ticular interest  due  to  their  behavior,  ecology, 
or  changes  in  relative  abundance.  The  longer 
species  accounts  form  the  heart  of  the  book 
and  contain  fairly  detailed  information  about 
the  natural  history  of  those  species  in  the  Val- 
ley. Accounts  of  the  remaining  species  vary 
in  length,  with  most  including  mention  of  the 
habitats  used  by  the  specific  species.  Brush 
specifically  mentions  that  the  style  of  the  spe- 
cies accounts  is  a hybrid  between  standard  re- 
gional works  and  other  natural  history  writing 
that  relies  heavily  on  personal  experiences  and 
field  notes.  In  many  ways,  this  adds  interest- 
ing aspects  to  the  species  accounts  in  which 
Brush  has  particular  interest,  such  as  Green 
Parakeet  {Aratinga  holochlora ),  Northern 
Beardless-Tyrannulet  ( Camptostoma  imber- 
be ),  Tropical  Parula  ( Parula  pitiayumi ),  and 
Altamira  Oriole  ( Icterus  gularis). 

My  main  quibble  with  the  book  is  that  some 
species  that  are  irregular  breeders  in  the  Val- 
ley are  covered  very  briefly,  sometimes  with 
only  a couple  of  lines.  I would  have  liked  to 
see  more  detailed  information  on  these  occur- 
rences. I also  question  the  inclusion  of  Yel- 
low-faced Grassquit  ( Tiaris  olivaceus ) as  hav- 
ing a breeding  record  in  Texas.  In  my  mind, 
a single  male  building  a “nest”  does  not  qual- 
ify as  a nesting  attempt,  but  this  is  a minor 
point.  In  the  introductory  section  of  the  book. 
Brush  does  point  out  that  there  are  five  sub- 
species endemic,  or  nearly  endemic,  to  the  Ta- 
maulipan  Biotic  Province;  in  the  species  ac- 
counts, however,  more  detailed  information  is 
not  included  for  all  these  taxa.  He  does  dis- 
cuss the  “Brownsville”  Common  Yellow- 
throat  ( Geothlypis  trichas  insperata ) and  other 
subspecies  that  occur  in  the  Valley,  although 
I would  have  liked  a more  in-depth  treatment 


of  these  taxa,  such  as  that  given  to  the  Valley 
specialties.  If  more  research  is  needed  on 
these  taxa,  this  would  have  been  a good  op- 
portunity to  point  out  major  gaps  in  the  cur- 
rent knowledge.  As  mentioned  previously. 
Brush  relies  heavily  on  his  own  field  experi- 
ence in  the  Valley,  thereby  adding  a nice  di- 
mension to  the  book  for  those  taxa  with  which 
he  has  personal  experience.  For  other  species, 
however,  his  brief  notes  don’t  always  add  to 
the  account.  Overall,  I found  the  book  to  be 
very  informative  and  would  recommend  it  to 
ornithologists  and  birders  alike  who  are  inter- 
ested in  the  avifauna  of  Texas. — MARK  W. 
LOCKWOOD,  Texas  Parks  and  Wildlife  De- 
partment, Fort  Davis,  Texas;  e-mail: 
mark.  lockwood@tpwd.  state,  tx.  us 


BIRDS  OF  WASHINGTON:  STATUS 
AND  DISTRIBUTION.  Edited  by  Terence  R. 
Wahl,  Bill  Tweit,  and  Steven  G.  Mlodinow. 
Oregon  State  University  Press,  Corvallis. 
2005:  x + 436  pp.,  285  maps.  ISBN: 
0870720494.  $65.00  (cloth). — State  avifaunal 
works  used  to  be  the  province  of  professional 
ornithologists  working  for  the  U.S.  Biological 
Survey,  other  government  agencies,  or  uni- 
versities. In  recent  years,  as  ornithological  re- 
search has  moved  into  physiological,  molec- 
ular, and  evolutionary  hypothesis  testing,  fau- 
nal investigation  and  summarization  increas- 
ingly have  been  delegated  to  dedicated  and 
talented  nonprofessionals.  This  volume  was 
developed  by  a team  of  46  authors,  including 
the  3 editors.  Many  have  biological  training 
and  employment,  but  I doubt  that  working  on 
this  book  fit  into  any  of  their  job  descriptions. 

Birds  of  Washington  includes  a short  intro- 
ductory section,  followed  by  species  accounts 
for  about  482  accepted  species.  The  wrap-up 
includes  a brief  discussion  of  non-established 
introduced  species,  including  accounts  for 
Mute  Swan  ( Cygnus  olor ),  Mandarin  Duck 
(Aix  galericulata ),  American  Black  Duck 
( Anas  rubripes ),  Monk  Parakeet  ( Myiopsitta 
monachus),  and  eight  species  of  hypothetical 
occurrence.  Appendices  include  a table  of  oc- 
currences by  habitat  and  brief  biographies  of 
the  46  authors. 

This  book  serves  an  important  function  as 


272 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  2,  June  2006 


an  up-to-date  status  check  on  bird  occurrence, 
distribution,  abundance,  and  changes  therein. 
The  included  material  appears  reliable  and  au- 
thoritative, but  I am  frustrated  by  what  is  not 
included.  This  is  a bare-bones  treatment  with 
minimal  analysis  presented.  The  introductory 
material  includes  an  explanation  of  the  species 
account  format,  a full  page  of  abbreviations 
used,  a chapter  by  Christopher  Chappell  on 
Bird  Habitats  of  Washington,  one  on  Avian 
Conservation  by  Joseph  Buchanan,  a brief  dis- 
cussion of  the  history  of  field  ornithology  in 
Washington,  a description  of  the  recent  infor- 
mation sources  used,  and  slightly  more  than  a 
page  on  changes  in  status  and  distribution 
over  the  past  half-century.  The  habitat  chapter 
provides  a listing  and  descriptions  of  30  hab- 
itats and  a lucid  explanation  of  the  basis  for 
their  delimitation.  All  of  the  other  sections  left 
me  wishing  for  more  detail.  The  history  chap- 
ter essentially  begins  with  Jewett  et  al.’s  Birds 
of  Washington  State  (University  of  Washing- 
ton Press,  Seattle,  1953)  and  does  not  even 
mention  W.  L.  Dawson,  who  wrote  the  mon- 
umental first  state  bird  books  for  Washington 
and  California.  The  Changes  in  Status  and 
Distribution  section  lacks  a discussion  of  the 
number  of  species  occurring  in  Washington, 
or  the  rate  of  addition  of  new  species.  The 
treatment  of  increases  and  decreases  in  range 
and  abundance  describes  general  classes  of 
causes  and  gives  examples,  but  without 
enough  detail  to  really  give  a reader  much 
sense  of  the  magnitude  or  prevalence  of  these 
changes. 

The  species  accounts  for  regularly  occur- 
ring species  begin  with  a brief  statement  of 
status  in  Washington.  Abundance  categories 
are  based  on  likelihood  of  encounter  rather 
than  estimates  of  actual  numbers  present.  A 
graphic  illustrating  seasonal  occurrence  and 
relative  abundance  follows,  then  a listing  of 
subspecies  in  Washington,  if  more  than  one, 
and  a listing  of  the  habitats  used.  A section 
titled  Occurrence  provides  detail  on  distribu- 
tion, abundance,  and  changes  thereof.  An  op- 
tional Remarks  section  is  followed  by  Note- 
worthy Records,  which  includes  high  counts 
and  unusual  dates.  Authorship  is  acknowl- 
edged for  the  accounts  of  accepted  species  but 
not  for  those  of  introduced  and  hypothetical 
species.  Very  detailed  distribution  maps — 
based  mainly  on  the  distribution  of  suitable 


habitat — accompany  283  of  the  accounts.  Sea- 
sonal changes  in  distribution  are  indicated 
with  different  shades  of  gray. 

Vagrants  receive  much  shorter  accounts, 
which  list  their  occurrences  in  Washington 
and  sometimes  a little  information  on  the  spe- 
cies' normal  distribution  and  abundance.  The 
term  “casual  vagrant”  is  used  in  place  of  the 
traditional  “accidental”  for  the  species  with 
the  fewest  records.  Inclusion  as  an  accepted 
species  is  based  on  acceptance  by  the  Wash- 
ington Bird  Records  Committee.  Corroborat- 
ing evidence  is  usually  mentioned,  but  up  to 
30  species  appear  to  be  accepted  based  only 
on  observer  descriptions  (the  text  is  not  al- 
ways clear  on  this).  I imagine  that  most  of 
these  records  were  accurate,  but  several  (e.g.. 
Little  Curlew,  Numenius  minutus ; Ruby- 
throated  Hummingbird.  Archilochus  colubris; 
Mourning  Warbler.  Oporornis  Philadelphia ; 
Nelson’s  Sharp-tailed  Sparrow,  Ammodramus 
nelsoni ) present  non-trivial  identification  is- 
sues. Citations  for  many  of  the  records  of  rar- 
ities refer  to  the  Records  Committee  reports 
rather  than  the  original  sources.  The  locality 
information  for  Washington  records  often 
lacks  county  or  other  regional  reference,  so 
someone  not  familiar  with  Washington  geog- 
raphy will  need  a good  gazetteer  to  locate 
Asotin,  Crockett  L.,  Stanwood,  Twisp,  Wal- 
lula,  and  so  on. 

This  book  will  be  useful  to  Washington 
birders  interested  in  the  status  of  the  birds 
they  see.  It  will  also  be  of  interest  to  scholars 
interested  in  dynamics  of  biogeography,  range 
expansion,  range  contraction,  and  vagrancy. 
Unfortunately,  the  editors  apparently  did  not 
recognize  this  latter  audience,  and  have  not 
made  the  information  of  interest  to  scholars  as 
accessible  as  they  could  have. — WAYNE 
HOFFMAN,  Newport.  Oregon;  e-mail: 
whoffman  @ peak.org 


PEREGRINE  FALCON;  STORIES  OF  THE 
BLUE  MEANIE.  By  James  Enderson.  original 
art  by  Robert  Katona.  University  of  Texas  Press, 
Austin.  2005:  266  pp.,  18  photographs,  23  line 
drawings.  ISBN:  0292705905.  $65.00  (cloth). 
ISBN:  0292706243,  $22.95  (paper).— Professor 
Emeritus  James  Enderson  of  Colorado  College 
has  written  an  engaging  and  very  readable 


ORNITHOLOGICAL  LITERATURE 


273 


memoir  that  centers  on  the  decline  and  recovery 
of  the  Peregrine  Falcon  ( Falco  peregrinus),  a 
now-revered  raptor  that  suffered  near  extinction 
in  much  of  its  range  beginning  in  the  mid-20th 
century.  The  dramatic  and  remarkable  recovery 
of  this  species  in  North  America,  following  the 
banning  of  DDT,  is  certainly  one  of  the  most 
significant  conservation  victories  of  the  last  cen- 
tury, and  Jim  Enderson  was  a major  player  on 
a team  that  won  the  game.  The  book’s  illustra- 
tions include  well-chosen  black  and  white  pho- 
tographs, as  well  as  many  original  drawings  by 
artist  Robert  Katona,  whose  contributions  add 
significantly  to  the  book’s  success. 

Enderson’s  account  might  well  be  required 
reading  for  young  ornithology  students;  cer- 
tainly, it  must  be  that  for  graduate  students 
and  established  professionals.  Enderson  tells 
his  story  well,  and  much  of  the  ground  he 
covers  in  this  book  is  now  covered  with  actual 
or  allegorical  asphalt,  no  longer  accessible  to 
students  currently  embarking  on  careers.  The 
stars  that  crossed  for  Enderson  were  falconry 
and  science.  He  clearly  has  a passion  for  both, 
and  he  was  able  to  weave  threads  from  each 
to  build  a career  full  of  adventure,  scientific 
puzzle-solving,  and  a cast  of  characters  that 
might  have  come  from  a novel. 

The  introductory  chapter  is  a splendid  de- 
scription of  the  Peregrine  Falcon,  certainly 
“one  of  the  best-studied  wild  animals  on  the 
planet.’’  ( The  Birds  of  North  America  species 
account  lists  over  300  references.)  Enderson 
provides  us  with  an  excellent  summary  of  this 
remarkable  species’  speed,  biology,  sexual  di- 
morphism, coloration,  distribution,  hunting 
techniques,  and  other  critical  life-history  com- 
ponents. The  nickname  “Blue  Meanie” — 
used  throughout  the  book — is  credited  to  En- 
derson’s fellow  peregrine  researcher  and  good 
friend,  W.  Grainger  Hunt. 

Enderson’s  story  begins  in  the  early  1960s 
with  his  searches  for  falcons.  He  focused  at 
first  on  Prairie  Falcons  ( Falco  mexicanus), 
then  (like  all  falconers  of  that  era)  dreamed  of 
peregrines.  His  chance  to  engage  with  pere- 
grines was  finally  realized  when  two  falconers 
invited  him  to  visit  the  Queen  Charlotte  Is- 
lands off  the  British  Columbia  coast,  which  at 
the  time  was  the  site  of  the  densest  population 
of  nesting  peregrines  anywhere  in  their  cos- 
mopolitan range.  One  suspects  that  Enderson’s 
rappelling  skills  might  have  had  something  to 


do  with  the  invitation,  but  the  story  of  the 
expedition  is  wonderful  autobiography  and 
adventure.  Even  in  his  quest  for  falconry 
birds,  Enderson’s  scientific  orientation  shines 
through.  For  example,  the  expedition  guide 
shot  a harbor  seal  ( Phoca  vitulina)  to  feed 
nestling  peregrines  recently  taken  from  the 
wild  and,  in  describing  the  butchering,  Ender- 
son cannot  resist  the  temptation  (or  obliga- 
tion) to  tell  us  why  the  seal’s  flesh  was  so  dark 
(it  relates  to  storing  high  levels  of  oxygen 
when  diving).  He  went  home  with  his  first 
peregrine — the  most  highly  valued  species  in 
the  world  of  American  falconry. 

By  that  time,  Enderson  was  a graduate  stu- 
dent, and  soon  thereafter  landed  a job  at  Col- 
orado College,  in  part  because  of  his  connec- 
tion with  Robert  Stabler,  a professor  at  the 
college  and  a famous  pioneer  falconer.  Pere- 
grines had  been  declining  for  a decade,  but 
the  picture  was  blurred  in  part  by  the  secrecy 
that  surrounded  nest  sites — those  who  knew 
the  bird  were  not  eager  to  tell  their  stories, 
and  most  attributed  local  declines  to  egg  col- 
lectors or  falconers.  (In  California,  a few  of 
us  who  watched  nesting  peregrines  knew  that 
eggs  had  been  laid,  but  when  we  later  returned 
to  cliffs,  the  adults  defended  weakly  or  not  at 
all,  and  the  nest  ledges  were  empty.  At  one 
site  where  a friend  had  lavished  a landowner 
with  canned  hams  and  whiskey  in  an  attempt 
to  exclude  the  reviled  “eggers,’’  we  conclud- 
ed that  the  egg-collectors  had  come  in  from 
the  sea!)  Surveys  were  then  initiated  (which 
turned  out  to  be  post-decline  surveys),  and 
Enderson  was  one  of  the  first  surveyors.  He 
checked  some  50  historical  nesting  sites, 
largely  in  the  intercontinental  West,  and  found 
only  13  pairs.  The  picture  would  worsen  be- 
fore it  was  over. 

The  watershed  event  was  the  Peregrine 
Conference  of  1965,  where  peregrine  scien- 
tists and  falconers  assembled  in  Madison, 
Wisconsin,  to  assess  the  extent  of  the  decline 
and  speculate  on  the  reasons  for  it.  Enderson 
was  there  and  he  was  much  impressed  by 
what  he  saw.  Hypotheses  explored  as  possible 
causes  of  decline  included:  the  peregrine  is  a 
“wilderness  species,’’  egg  collecting,  falcon- 
ers, drought,  and  pesticides.  Enderson  omits 
mention  of  Rachel  Carson  ( Silent  Spring  was 
published  in  1962),  and  fails  to  point  out  that 
there  was  significant  resistance  to  even  dis- 


274 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  2,  June  2006 


cussing  pesticides  at  the  conference.  Roger 
Tory  Peterson  was  sitting  not  far  from  me.  and 
at  one  frustrating  moment  when  the  inevitable 
discussion  about  pesticides  was  sidestepped 
by  a U.S.  Fish  and  Wildlife  Service  represen- 
tative, a frustrated  and  angry  Peterson  stabbed 
the  table  with  his  wooden  pencil! 

But  Enderson  tells  the  DDE  story  well.  It 
was  indeed  like  a mystery  novel,  with  the  pri- 
mary culprit  being  DDE,  not  DDT.  or  dieldrin, 
or  any  other  of  the  array  of  biocides  that 
Carson  had  described  in  Silent  Spring.  As  En- 
derson describes  it,  there  were  many  dead-end 
roads  traveled,  in  part  because  DDE  was  not 
toxic  to  insects;  therefore  it  had  been  little  in- 
vestigated. It  turned  out  to  be  the  primary  cul- 
prit that  caused  eggshell  thinning  and  was  re- 
sponsible for  most  of  the  population  decline. 
The  proof  would  not  emerge  until  the  parent 
compound  DDT  was  banned,  but  along  the 
way,  experimental  science  provided  strong  ev- 
idence. Enderson  rightly  credits  David  Peakall 
with  discovering  DDE  in  peregrine  eggshells 
that  were  collected  in  1948,  only  a couple  of 
years  after  the  “wonder  insecticide”  had  been 
introduced  into  general  agricultural  use. 

Some  of  the  best  parts  of  Enderson’s  book 
are  his  stories  of  peregrine  surveys  in  North 
America,  and  eventually  in  other  parts  of  the 
world.  Enderson  participated  in  many,  if  not 
most,  of  these  surveys,  and  his  tales  of  ca- 
noeing Arctic  rivers,  dangling  from  ropes  on 
500-foot  cliffs,  and  interacting  with  sundry 
bureaucrats  make  good  reading.  He  tells  won- 
derful tales  of  remarkable  characters,  often 
with  a little  spice  and  always  with  excellent 
descriptions  of  character.  For  example,  put  in 
the  care  of  a “surly  sergeant”  on  a Texas 
beach  when  he  is  trapping  migrating  pere- 
grines, Enderson  wins  the  day  by  trapping, 
banding,  and  releasing  four  birds.  The  surly 
sergeant  had  been  assigned  to  drive  Enderson 
(and  Clayton  White,  another  giant  in  the  per- 
egrine story)  as  punishment,  but  ended  up  an 
enthusiastic  trapper.  There  are  tales  of  many 
others. 

Enderson  was  also  part  of  the  group  that 
managed,  at  long  last,  to  breed  falcons  in  cap- 
tivity. In  a chapter  titled  Timely  Invention  of 
Peregrine  Husbandry,  Enderson  describes  this 
technology  in  detail  (I  could  have  done  with- 
out the  illustrated  description  of  collecting 
peregrine  semen  in  the  seam  of  a rubber  gas- 


ket placed  on  the  head  of  the  collector!).  Then 
(remarkably),  he  describes  his  theft  of  nestling 
peregrines  from  some  of  the  last  productive 
eyries  in  the  United  States  and  Canada.  It  was 
a matter  of  the  means  justifying  the  ends,  one 
supposes,  but  it  may  raise  some  eyebrows. 

Enderson — such  an  intimate  part  of  the  cap- 
tive breeding  and  release  program  that  is 
widely  credited  with  “saving  the  pere- 
grine”— points  out  that  what  really  saved  the 
species  was  the  (then)  controversial  decision 
in  1972  by  William  Ruckelshaus,  head  of  the 
new  Environmental  Protection  Agency,  to  ban 
most  uses  of  DDT.  He  also  asserts  that  the 
Peregrine  Falcon  would  have  recovered  on  its 
own  from  the  small  reservoir  of  functional 
breeding  pairs  left  here  and  there,  but  it  would 
have  taken  much  longer,  especially  in  those 
areas  from  which  it  had  disappeared  entirely. 
The  release  program  was  very  popular,  and 
resulted  in  the  elevation  of  the  peregrine  to 
the  status  of  absolute  charisma.  It  had  gone 
from  the  reviled,  often  shot  *'Duck  Hawk”  of 
the  mid-20th  century  to  one  of  the  best-loved, 
wild  vertebrates  in  the  world. 

Falconry  is  a major  topic  in  this  book,  and 
Enderson  does  it  justice.  He  describes  the 
sport’s  early  days  in  North  America,  the  col- 
orful cast  of  characters,  and  the  discovery  of 
the  Arctic  Peregrine’s  Atlantic  and  Gulf  of 
Mexico  coastal  migration  paths.  He  even  in- 
cludes a primer  on  falconry,  which  gives  the 
reader  a sense  of  what  that  passion  is  all 
about.  I was  especially  pleased  to  see  that  En- 
derson favors  the  correct  pronunciation  of  fal- 
con: these  birds  are  not  “phal- cons,”  but  “the 
historically  correct  fall- cons,'  as  in  the  word 
falling .”  A “phal- con”  is  a car  or  a football 
team;  a “fall- con”  is  a bird. 

In  the  latter  passages,  Enderson  brings  fal- 
conry up  to  date  and  describes — with  appro- 
priate bitterness — “Operation  Falcon,”  a fed- 
eral sting  operation  that,  between  1981  and 
1984,  entrapped  some  52  falconers  and  con- 
fiscated 106  raptors.  It  was  an  unfortunate 
chapter  in  the  peregrine  story. 

One  serious  omission  hangs  over  Ender- 
son’s book — a fuller  coverage  of  those  who 
sought  to  obfuscate  the  developing  truth  about 
DDE,  including  pesticide  company  employ- 
ees. Mention,  perhaps,  should  also  have  been 
made  of  the  false  claims  that  the  peregrine’s 
decline  was  faked  by  scientists  who  stood  to 


ORNITHOLOGICAL  LITERATURE 


275 


benefit  (in  terms  of  professional  fame  and  re- 
search money)  by  reporting  the  precipitous 
decline  in  numbers  of  peregrines. 

This  excellent  book  ends  with  a nicely  writ- 
ten memory  of  peregrines  having  returned  to 
two  historical  nesting  sites  from  which  they 
had  been  missing  for  decades.  The  writing 
here  is  a splendid  description  of  emotional  en- 
counters with  nature.  The  reader  is  put  at  the 
spot  and  in  the  experience,  and  when  blue 
meanies  appear  after  seemingly  fruitless 
searches,  one  shares  in  the  relief  and  exulta- 
tion. In  an  era  when  radiotelemetry  has  par- 
tially replaced  old-fashioned  fieldwork  and 
modeling  is  thought  to  be  a substitute  for 
much  of  what  has  occupied  biologists  for 
ages,  Enderson’s  book  reminds  us  of  why 
most  of  us  enter  wildlife-related  work  in  the 
first  place.  In  most  cases,  we  love  the  wild 
things  we  study,  we  admire  their  beauty,  and 
we  do  all  we  can  to  guarantee  that  succeeding 
generations  will  be  able  to  do  the  same. — 
STEVEN  G.  HERMAN,  The  Evergreen  State 
College,  Olympia,  Washington;  e-mail: 
hermans@evergreen.edu 


HAWKS  FROM  EVERY  ANGLE.  By  Jer- 
ry Liguori.  Princeton  University  Press,  Prince- 
ton, New  Jersey.  2005:  133  pp.,  68  plates,  371 
photos,  2 maps.  ISBN:  0691118248,  $55.00 
(cloth).  ISBN:  0691118256,  $19.95  (paper).— 
In  his  new  book,  Hawks  from  Every  Angle, 
Jerry  Liguori  uses  a new  and  different  ap- 
proach to  identify  19  migratory  hawk  species 
in  flight.  In  the  introduction,  Jerry  writes, 
“This  is  primarily  a visual  guide;  the  photos 
and  captions  are  the  crux  of  the  book  and  are 
meant  to  stand  on  their  own.”  Unlike  previous 
photo  guides  that  offer  images  showing  every 
field  mark  in  perfect  lighting  at  point-blank 
range,  Jerry  has  selected  high-quality  images 
that  more  accurately  reflect  true  conditions  of 
field  observation.  He  used  his  extensive  ex- 
perience studying  hawks  throughout  North 
America  to  select  images  that  reflect  realistic 
flight  profiles  and  structures  for  each  species. 
Through  these  images,  Jerry  represents  the 
full  range  of  varying  postures  the  birds  can 
show  in  flight  when  viewed  from  differing  an- 
gles and  under  varying  environmental  condi- 
tions. 


Multiple  images  are  often  stitched  together 
and  presented  side  by  side,  with  as  many  as 
six  images  per  page  representing  a single 
“plate.”  These  stitched  images  typically  show 
comparative  views  of  similar  species  at  the 
same  angle,  differing  plumages  (age,  sex, 
morph,  race)  of  a given  species,  or  different 
flight  profiles  that  illustrate  the  range  of  var- 
iation for  a given  species  under  varying  con- 
ditions. The  accompanying  captions  smartly 
describe  these  comparative  differences.  The 
author  uses  a holistic  approach  to  identifica- 
tion similar  to  that  seen  in  Hawks  in  Flight  by 
Pete  Dunne,  David  Sibley,  and  Clay  Sutton 
(Houghton  Mifflin,  1988),  except  Jerry  opts 
for  images  over  written  descriptions  as  the  pri- 
mary focus  of  the  guide.  As  such,  this  guide 
is  more  useful  in  the  field  than  its  predecessor, 
which  was  meant  to  be  read  at  home. 

The  book  is  clearly  designed  for  use  by 
hawkwatching  enthusiasts  at  hawkwatching 
sites.  In  the  introduction,  Jerry  summarizes  a 
number  of  sites  across  North  America,  graphs 
peak  migration  times  by  species,  and  adds  ta- 
bles that  summarize  high  counts,  by  species. 
The  images  generally  cover  the  entire  range 
of  expected  “looks”  each  species  may  offer 
as  it  flies  by  any  hawkwatch  site.  However, 
when  pictures  alone  won’t  suffice,  Jerry  uses 
intuitive  descriptions  of  behaviors,  such  as 
comparative  differences  in  wing  flapping  and 
flight  characteristics.  For  example,  “Sharp- 
shinned  Hawks  beat  their  wings  in  a shallow, 
snappy,  powerless  manner,  similar  to  a Robin. 

. . . Cooper’s  Hawks  almost  always  soar  with 
a slight  dihedral.  ...  In  moderate  to  high 
winds.  Sharp-shinned  Hawks  appear  hyper- 
active, unstable,  and  hesitant,  making  constant 
wing  adjustments.”  These  subjective  differ- 
ences are  used  by  seasoned  hawkwatchers  on 
a daily  basis  to  identify  distant  raptors,  but 
they  are  gleaned  from  thousands  of  hours  of 
experience  and  are  not  included  in  typical 
guides.  The  author  generally  excludes  fine  de- 
tails not  easily  seen  in  the  field  such  as  eye 
color,  and  descriptions  of  individual  feathers 
that  are  notable  only  at  very  close  range. 

The  text  is  organized  by  species  and  pre- 
sented in  a consistent  format  (the  three  accip- 
iter  species:  A.  striatus , A.  cooperii,  and  A. 
gentilis  are  treated  as  one  group  with  com- 
parative differences  highlighted  throughout). 
Each  species  account  begins  with  a general 


276 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  2,  June  2006 


overview  of  the  species  (or  species  group), 
followed  by  sections  on  migration  and  plum- 
age, respectively.  The  remainder  (and  bulk)  of 
the  account  is  dedicated  to  Flight  Style  with 
subsections  including  wing  beat,  soaring, 
head-on,  gliding  overhead,  and  wing-on/going 
away.  Portions  of  these  accounts  can  be  dif- 
ficult to  follow  at  times,  particularly  in  the  ac- 
cipiter  section,  which  continually  bounces  be- 
tween the  three  species;  however,  the  author 
addresses  this,  to  a degree,  by  using  bold  print 
to  accentuate  key  points  and  distinctive  char- 
acteristics found  throughout. 

As  anyone  familiar  with  raptors  might  ex- 
pect, maximum  coverage  was  given  to  the 
highly  variable  Red-tailed  Hawk  ( Buteo  ja- 
maicensis).  Jerry  uses  a full  14  pages  of  text 
and  images  to  thoroughly  cover  a wide  range 
of  recognized  subspecies,  races,  forms,  and 
color  morphs  in  each  age  class.  Jerry  also  cov- 
ers the  varying  age  classes  of  Bald  Eagle 
( Haliaeetus  leucocephalus ) and  Golden  Eagle 
( Aquila  chrysaetos ) with  detailed  descriptions 
of  molt  sequences  and  other  plumage  char- 
acteristics. 

This  guide  offers  as  much  insightful  com- 
mentary on  flight  characteristics  of  raptors  as 
any  guide  ever  has.  It  also  offers  a greater 
range  of  differing  perspectives  and  flight  pro- 


files than  any  previous  guide.  Unfortunately, 
despite  the  all-encompassing  title,  there  are 
some  “angles”  not  covered.  For  example, 
there  is  no  mention  or  images  of  perched 
birds,  and  there  is  no  coverage  of  general  nat- 
ural history  other  than  that  pertaining  to  mi- 
gration. The  exclusion  of  some  field  marks  def- 
initely limits  the  scope  and  usefulness  of  this 
book  as  well. 

Whereas  this  book  is  clearly  an  indispens- 
able resource  for  anyone  interested  in  hawk- 
watching, away  from  the  hawkwatch  site  it  of- 
fers little  assistance  for  the  observer  wanting 
to  identify  the  hawk  perched  in  their  back- 
yard. It  also  offers  little  for  those  curious 
about  nesting  habits  or  breeding  range  of  a 
given  species.  For  answers  to  these  questions, 
readers  will  have  to  turn  to  another  guide. 
However,  if  the  backyard  hawk  flies  from  the 
tree  and  you  are  able  to  observe  it  as  it  flaps 
straight  away,  then  Hawks  from  Every  Angle 
is  likely  just  the  ticket.  Given  the  reasonable 
price,  slim  profile,  and  the  wealth  of  personal 
wisdom  packed  into  the  pages  of  this  book,  it 
deserves  a spot  on  every  birder’s  bookshelf. 
There  is  no  one  who  can’t  learn  something 
from  this  work!— JEFFREY  BOUTON,  Leica 
Sport  Optics,  Port  Charlotte,  Florida;  e-mail: 
jbouton2@earthlink.net 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 


Editor  JAMES  A.  SEDGWICK 
U.S.  Geological  Survey 
Fort  Collins  Science  Center 
2150  Centre  Ave..  Bldg.  C. 

Fort  Collins,  CO  80256-81 18,  USA 
E-mail:  wjo@usgs.gov 


Editorial  Board  KATHY  G.  BEAL 
CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 
KARL  E.  MILLER 


Review  Editor 


Managing  Editor 
Copy  Editor 
Editorial  Assistant 


M.  BETH  DILLON 
CYNTHIA  P.  MELCHER 
ALISON  R.  GOFFREDI 


MARY  GUSTAFSON 
Texas  Parks  and  Wildlife  Dept. 

2800  S.  Bentsen  Palm  Dr. 

Mission,  TX  78572,  USA 
E-mail:  WilsonBookReview@aol.com 


GUIDELINES  FOR  AUTHORS 

Consult  the  detailed  “Guidelines  for  Authors”  found  on  the  Wilson  Ornithological  Society  Web  site  (http:// 
www.ummz.lsa.umich.edu/birds/wilsonbull.html).  Beginning  in  2007,  Clait  E.  Brain  will  become  the  new  editor 
of  The  Wilson  Journal  of  Ornithology.  As  of  1 July  2006,  all  manuscript  submissions  and  revisions  should  be 
sent  to  Clait  E.  Brain,  Editor,  The  Wilson  Journal  of  Ornithology,  5572  North  Ventana  Vista  Rd.,  Tucson,  AZ 
85750-7204,  USA.  The  New  Wilson  Journal  of  Ornithology  office  and  fax  telephone  number  will  be  (520)  529- 
0365,  and  the  E-mail  address  will  be  TWilsonJO@comcast.net. 

NOTICE  OF  CHANGE  OF  ADDRESS 

If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornithological 
Societies  of  North  America,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  %The  Museum  of  Zoology,  The 
Univ.  of  Michigan,  Ann  Arbor,  MI  48109.  Persons  having  business  with  any  of  the  officers  may  address  them 
at  their  various  addresses  given  on  the  inside  of  the  front  cover,  and  all  matters  pertaining  to  the  journal  should 
be  sent  directly  to  the  Editor. 


MEMBERSHIP  INQUIRIES 

Membership  inquiries  should  be  sent  to  James  L.  Ingold,  Dept,  of  Biological  Sciences,  Louisiana  State  Univ., 
Shreveport,  LA  71115;  e-mail:  jingold@pilot.lsus.edu 

THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 

The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  in  the  Univ.  of 
Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  Univ.  of  Michigan  in  1930.  Until  1947 
the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  ornithological  magazines  from 
members  and  friends  of  the  Society.  Two  members  have  generously  established  a fund  for  the  purchase  of  new 
books;  members  and  friends  are  invited  to  maintain  the  fund  by  regular  contribution.  The  fund  will  be  admin- 
istered by  the  Library  Committee.  Terry  L.  Root,  Univ.  of  Michigan,  is  Chairman  of  the  Committee.  The  Library 
currently  receives  over  200  periodicals  as  gifts  and  in  exchange  for  The  Wilson  Journal  of  Ornithology.  For 
information  on  the  Library  and  our  holdings,  see  the  Society's  web  page  at  http://www.ummz.lsa.umich.edu/ 
birds/wos.html.  With  the  usual  exception  of  rare  books,  any  item  in  the  Library  may  be  borrowed  by  members 
of  the  Society  and  will  be  sent  prepaid  (by  the  Univ.  of  Michigan)  to  any  address  in  the  United  States,  its 
possessions,  or  Canada.  Return  postage  is  paid  by  the  borrower.  Inquiries  and  requests  by  borrowers,  as  well  as 
gifts  of  books,  pamphlets,  reprints,  and  magazines,  should  be  addressed  to:  Josselyn  Van  Tyne  Memorial  Library, 
Museum  of  Zoology,  The  Univ.  of  Michigan,  1109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079,  USA.  Contri- 
butions to  the  New  Book  Fund  should  be  sent  to  the  Treasurer. 


This  issue  of  The  Wilson  Journal  of  Ornithology  was  published  on  5 June  2006. 


280 


Continued from  outside  back  cover 


2 47  Golden-cheeked  Warbler  males  participate  in  nest-site  selection 
Allen  E.  Graber,  Craig  A.  Davis , and  David  M.  Leslie,  Jr. 

251  Provisioning  of  Magellanic  Woodpecker  ( Campephilus  magellanicus)  nestlings  with  vertebrate  prey 
Valeria  S.  Ojeda  and  M.  Laura  Chazarreta 

254  Reverse  mounting  and  copulation  behavior  in  polyandrous  Bearded  Vulture  ( Gypaetus  barbatus)  trios 
Joan  Bertran  and  Antoni  Margalida 

25 6 Natural  occurrence  of  crowing  in  a free-living  female  galliform,  the  California  Quail 
Jennifer  M.  Gee 

259  Poult  adoption  and  nest  abandonment  by  a female  Rio  Grande  Wild  Turkey  in  Texas 

Steve  T.  Metz,  Kyle  B.  Melton,  Ray  Aguirre,  Bret  A.  Collier,  T.  Wayne  Schwertner,  Markus  J.  Peterson,  and 
Nova  J.  Silvy 

261  Predation  by  a Blue-crowned  Motmot  ( Momotus  momota)  on  a hummingbird 
J.  Mauricio  Garcla-C.  and  Rakan  A.  Zahawi 

264  Once  Upon  a Time  in  American  Ornithology 

2 67  Ornithological  Literature 


The  Wilson  Journal  of  Ornithology 

(formerly  The  Wilson  Bulletin) 


Volume  118,  Number  2 CONTENTS  June  2006 


Major  Articles 

131  Breeding  productivity  of  Bachman’s  Sparrows  in  fire-managed  longleaf  pine  forests 
James  W.  Tucker,  Jr.,  W Douglas  Robinson,  and  James  B.  Grand 

138  Variation  in  Bachman’s  Sparrow  home-range  size  at  the  Savannah  River  Site,  South  Carolina 
Jonathan  M.  Stober  and  David  G.  Krementz 

145  Nesting  success  and  breeding  biology  of  Cerulean  Warblers  in  Michigan 
Christopher  M.  Rogers 

152  Migrant  shorebird  predation  on  benthic  invertebrates  along  the  Illinois  River,  Illinois 
Gabriel  L.  Hamer,  Edward  J.  Heske,  Jeffrey  D.  Brawn,  and  Patrick  W Brown 

1 64  Composition  and  timing  of  postbreeding  multispecies  feeding  flocks  of  boreal  forest  passerines  in 
western  Canada 

Keith  A.  Hobson  and  Steve  Van  Wilgenburg 

173  Variation  in  size  and  composition  of  Bufflehead  {Bucephala  albeola)  and  Barrow’s  Goldeneye 
(. Bucephala  islandica)  eggs 

Jennifer  L.  Lavers,  Jonathan  E.  Thompson,  Cynthia  A.  Paszkowski,  and  C.  Davison  Ankney 

178  Site-specific  survival  of  Black-headed  Grosbeaks  and  Spotted  Towhees  at  four  sites  within  the 
Sacramento  Valley,  California 
Thomas  Gardali  and  Nadav  Nur 

187  Pre-migratory  fattening  and  mass  gain  in  Flammulated  Owls  in  central  New  Mexico 
John  P DeLong 

194  Morphological  variation  and  genetic  structure  of  Galapagos  Dove  {Zenaida  galapagoensis)  populations: 
issues  in  conservation  for  the  Galapagos  bird  fauna 
Diego  Santiago-Alarcon,  Susan  M.  Tanksley,  and  Patricia  G.  Parker 

208  Breeding  ecology  of  American  and  Caribbean  coots  at  Southgate  Pond,  St.  Croix:  use  of  woody 
vegetation 

Douglas  B.  McNair  and  Carol  Cramer-Burke 

218  Insular  and  migrant  species,  longevity  records,  and  new  species  records  on  Guana  Island,  British  Virgin 
Islands 

Clint  W.  Boal,  Fred  C.  Sibley,  Tracy  S.  Estabrook,  and  James  Lazell 

225  Reproductive  behavior  of  the  Yellow-crowned  Parrot  {Amazona  ochrocephala)  in  western  Panama 
Angelica  M.  Rodriguez  Castillo  and  Jessica  R.  Eberhard 

Tbl  Gregarious  nesting  behavior  of  Thick-billed  Parrots  ( Rhynchopsitta  pachyrhyncha)  in  aspen  stands 
Tiberio  C.  Monterrubio-Rico,  Javier  Cruz-Nieto,  Ernesto  Enkerlin-Hoeflich,  Diana  Venegas-Holguin, 
Lorena  Tellez-Garcia,  and  Consuelo  Marin-Togo 

Short  Communications 

244  No  extra-pair  fertilization  observed  in  Nazca  Booby  ( Sula  granti)  broods 
David  J.  Anderson  and  Peter  T.  Boag 


Continued  on  inside  back  cover 


8H34 

Y&  Wilson  Journal 

of  Ornithology 

Volume  118,  Number  3,  September  2006 

MCZ 

LIBRARY 

DEC  09  2010 
harvarp 

UNIVERSITY 


Published  by  the 
Wilson  Ornithological  Society 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  ornithologist. 


President — Doris  J.  Watt,  Dept,  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  IN  46556,  USA;  e-mail: 
dwatt@saintmarys.edu 

First  Vice-President — James  D.  Rising,  Dept,  of  Zoology,  Univ.  of  Toronto,  Toronto,  ON  M5S  3G5, 
Canada;  e-mail:  rising@zoo.utoronto.ca 

Second  Vice-President — E.  Dale  Kennedy,  Biology  Dept.,  Albion  College,  Albion,  MI  49224,  USA; 
e-mail:  dkennedy@albion.edu 

Editor — James  A.  Sedgwick,  U.S.  Geological  Survey,  Fort  Collins  Science  Center,  2150  Centre  Ave., 
Bldg.  C,  Fort  Collins,  CO  80526,  USA;  e-mail:  wjo@usgs.gov 

Secretary — Sara  R.  Morris,  Dept,  of  Biology,  Canisius  College,  Buffalo,  NY  14208,  USA;  e-mail: 
morriss@canisius.edu 

Treasurer — Melinda  M.  Clark,  52684  Highland  Dr.,  South  Bend,  IN  46635,  USA;  e-mail:  MClark@tcservices.biz 

Elected  Council  Members — Robert  C.  Beason,  Mary  Gustafson,  and  Timothy  O’Connell  (terms  expire 
2006);  Mary  Bomberger  Brown,  Robert  L.  Curry,  and  James  R.  Hill,  III  (terms  expire  2007);  Kathy  G. 
Beal,  Daniel  Klem,  Jr.,  and  Douglas  W.  White  (terms  expire  2008). 

Membership  dues  per  calendar  year  are:  Active,  $21.00;  Student,  $15.00;  Family,  $25.00;  Sustaining, 
$30.00;  Life  memberships,  $500  (payable  in  four  installments). 

The  Wilson  Journal  of  Ornithology  is  sent  to  all  members  not  in  arrears  for  dues. 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 
(formerly  The  Wilson  Bulletin ) 

THE  WILSON  JOURNAL  OF  ORNITHOLOGY  (ISSN  1559-4491)  is  published  quarterly  in  March,  June, 
September,  and  December  by  the  Wilson  Ornithological  Society,  810  East  10th  St.,  Lawrence,  KS  66044-8897.  The 
subscription  price,  both  in  the  United  States  and  elsewhere,  is  $40.00  per  year.  Periodicals  postage  paid  at  Lawrence,  KS. 
POSTMASTER:  Send  address  changes  to  OSNA,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

All  articles  and  communications  for  publications  should  be  addressed  to  the  Editor.  Exchanges  should  be  addressed 
to  The  Josselyn  Van  Tyne  Memorial  Library,  Museum  of  Zoology,  Ann  Arbor,  Michigan  48109. 

Subscriptions,  changes  of  address,  and  claims  for  undelivered  copies  should  be  sent  to  OSNA,  5400  Bosque  Blvd., 
Ste.  680,  Waco,  TX  76710.  Phone:  (254)  399-9636;  e-mail:  business@osnabirds.org.  Back  issues  or  single  copies  are 
available  for  $12.00  each.  Most  back  issues  of  the  journal  are  available  and  may  be  ordered  from  OSNA.  Special  prices 
will  be  quoted  for  quantity  orders.  All  issues  of  the  journal  published  before  2000  are  accessible  on  a free  Web  site  at  the 
Univ.  of  New  Mexico  library  (http://elibrary.unm.edu/sora/).  The  site  is  fully  searchable,  and  full-text  reproductions  of  all 
papers  (including  illustrations)  are  available  as  either  PDF  or  DjVu  files. 


© Copyright  2006  by  the  Wilson  Ornithological  Society 
Printed  by  Allen  Press,  Inc.,  Lawrence,  Kansas  66044,  U.S. A. 


COVER:  Wilson’s  Phalaropes  ( Phalaropus  tricolor ).  Illustration  by  Robin  Corcoran. 


® This  paper  meets  the  requirements  of  ANSI/NISO  Z39.48-1992  (Permanence  of  Paper). 


FRONTISPIECE.  American  Dippers  ( Cinclus  mexicanus ) nesting  in  the  Oregon  Coast  Range  exhibit  flexibility 
with  respect  to  selecting  nest  sites.  By  constructing  nesting  substrates  (nest  boxes,  ledges  on  cliffs)  to  augment 
the  availability  of  natural  sites,  Loegering  and  Anthony  (p.  281)  increased  the  number  of  actively  used  nesting 
sites  from  three  to  eight  along  a 10-km  reach  of  stream.  Original  painting  (mixed  media:  gouache  water  color 
and  acrylic)  by  Barry  Kent  MacKay. 


TIk  Wilson  Journal 

of  Ornithology 


Published  by  the  Wilson  Ornithological  Society 


VOL.  118,  NO.  3 September  2006  PAGES  281-438 

The  Wilson  Journal  of  Ornithology  1 1 8(3):28 1—294,  2006 


NEST-SITE  SELECTION  AND  PRODUCTIVITY  OF  AMERICAN 
DIPPERS  IN  THE  OREGON  COAST  RANGE 

JOHN  P.  LOEGERING1’2’3  AND  ROBERT  G.  ANTHONY1 


ABSTRACT. — Availability  of  high-quality  nest  sites  is  thought  to  limit  breeding  populations  of  American 
Dippers  ( Cinclus  mexicanus).  To  examine  this  hypothesis,  we  characterized  dipper  nest  sites,  nest-site  habitat, 
and  productivity  in  the  central  Oregon  Coast  Range.  We  also  made  additional  nest  sites  (“created”  nest  sites  = 
nest  boxes,  cliff  ledges,  hollowed  logs  that  we  constructed  or  created)  along  one  of  two  creeks.  Suitable  nest 
sites  (1)  provided  a physical  space  to  place  the  nest,  (2)  were  above  the  upper  reach  of  flooding  and  inaccessible 
to  ground  predators,  and  (3)  were  very  near  to,  or  extended  over,  the  stream’s  edge.  Given  these  requirements, 
and  within  the  context  of  swift,  unpolluted  mountain  streams,  dippers  exhibited  flexibility  in  their  nest-site 
selection  patterns  and  used  a variety  of  nesting  substrates.  Streamside  features  associated  with  dipper  nest  sites 
included  geomorphically  constrained  valleys  (i.e.,  narrow  valley  floors),  the  presence  of  trees  in  the  riparian 
zone  (not  tested  statistically,  but  nearly  universal  to  all  nest  sites),  stream  shading  from  overhead  vegetation, 
and  locations  that  were  farther  from  areas  frequented  by  humans  (e.g.,  roads).  Dippers  readily  used  nesting 
substrates  that  we  created,  more  than  doubling  the  breeding  population  on  a 10-km  reach  of  stream  (8  versus  3 
nests/  10-km  reach).  Reproductive  success  was  high  and  not  associated  with  any  habitat  feature  we  measured. 
The  factors  influencing  recruitment  in  the  Oregon  Coast  Range  remain  unknown.  Received  6 October  2004, 
accepted  5 May  2006. 


Habitat  associations  of  many  terrestrial  spe- 
cies associated  with  streams  in  the  Pacific 
Northwest  are  lacking  (Anthony  et  al.  1987; 
McGarigal  and  McComb  unpubl.  data),  but  are 
essential  for  ecologically  sound  management. 
The  American  Dipper  ( Cinclus  mexicanus)  is 


1 Oregon  Coop.  Fish  and  Wildlife  Research  Unit, 
Dept,  of  Fisheries  and  Wildlife,  Oregon  State  Univ., 
Corvallis,  OR  97331-3803,  USA. 

2 Current  address:  Natural  Resources  Dept.,  Univ.  of 
Minnesota-Crookston,  2900  University  Ave.,  Crooks- 
ton,  MN  56716-5001,  USA;  and  Dept,  of  Fisheries, 
Wildlife  and  Conservation  Biology,  Univ.  of  Minne- 
sota, St.  Paul,  MN,  USA. 

3 Corresponding  author;  e-mail:  jloegeri@umn.edu 


the  most  abundant  resident,  riparian-obligate 
bird  species  in  managed  forests  of  the  central 
Oregon  Coast  Range  (Loegering  and  Anthony 
1999).  From  Alaska  to  Panama,  dippers  are 
widely  distributed  in  mountainous  regions  of 
western  North  America  and  Central  America 
(Bent  1948,  Kingery  1996).  Generally,  nest  sites 
are  located  over,  or  near  the  edges  of,  streams, 
where  they  are  inaccessible  to  predators  and  of- 
ten sheltered  from  the  weather  (Hann  1950, 
Price  and  Bock  1983,  Kingery  1996).  More  spe- 
cifically, the  nests — constructed  with  moss  and 
enclosed  with  a domed  roof  (15-25  cm  in  di- 
ameter)— typically  are  placed  on  cliff  ledges;  on 
ledges  of  mid-stream  boulders;  in  crevices  be- 


281 


282 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


tween  boulders;  in  cavities  of  horizontal,  hollow 
logs  extending  over  streams;  under  or  within  the 
support  structure  of  bridges  (Kingery  1996,  Os- 
born 1999,  Morrissey  2004);  or  in  nest  boxes 
(Hawthorne  1979). 

Price  and  Bock  (1983)  suggested  that  dipper 
reproductive  success  may  vary  with  nest-site 
quality,  but  this  possibility  remained  untested. 
We  characterized  and  evaluated  dipper  nest-site 
selection  in  the  central  Oregon  Coast  Range  at 
three  spatial  scales  (Johnson  1980):  (1)  micro- 
habitat (approximately  0.25-1.0  m2  around  the 
nest),  (2)  macrohabitat  (approximately  1-10  m2 
around  the  nest),  and  (3)  streamside  habitat 
(>100  m2  around  the  nest).  Specifically,  we 
characterized  dipper  microhabitat  and  macro- 
habitat and  tested  the  null  hypotheses  that  (1) 
streamside  habitat  at  dipper  nests  was  not  dif- 
ferent from  that  of  randomly  selected  locations; 
(2)  reproductive  success  was  not  correlated  with 
any  features  of  nest-site  habitat  at  the  microhab- 
itat, macrohabitat,  or  streamside  scales;  and  (3) 
increased  availability  of  nest  sites  would  not  af- 
fect the  number  of  breeding  pairs.  Because  nest- 
site  availability  has  been  suggested  as  a factor 
limiting  dipper  populations  (Price  and  Bock 
1983,  Kingery  1996),  we  also  experimentally 
increased  the  number  of  available  nest  sites 
along  one  stream  and  monitored  nest  densities 
there  and  along  an  unaltered  stream  for  5 years. 

METHODS 

Study  area. — During  the  1993-1998  breeding 
seasons,  we  studied  dippers  along  Drift  (44°  25' 
N,  123°  50' W)  and  Lobster  (44°  15' N,  123° 
40'  W)  creeks  in  the  central  Oregon  Coast 
Range,  Oregon,  and,  in  1994,  along  23  addi- 
tional streams  in  6 basins  within  a 10-km  radius 
of  Drift  and  Lobster  creeks.  During  1994  we 
searched  181  km  of  streams  to  locate  nest  sites 
and  collect  microhabitat,  macrohabitat,  and 
streamside  habitat  data.  During  1993-1995,  we 
studied  reproductive  success  only  on  Drift  and 
Lobster  creeks.  During  1993—1998,  we  censused 
the  abundance  of  nests,  and  we  studied  occu- 
pancy of  natural  nest  sites  and  those  that  we 
made  only  on  Drift  and  Lobster  creeks.  These 
basins  were  located  in  Benton,  Lane,  and  Lin- 
coln counties  and  drained  into  the  Alsea  and 
Siuslaw  rivers  6 to  23  km  east  of  the  Pacific 
Ocean.  Streambed  elevations  ranged  from  3 to 
365  m,  and  the  topography  was  characterized 
by  steep  terrain  interspersed  with  moderately  flat 


valleys.  Stream  gradient  averaged  <4%,  (range 
= 0.5-1 1%),  generally  increasing  in  the  smaller, 
fourth-order  streams  (“stream  order”  is  a stream 
classification  system:  first-order  streams  are 
small,  unbranched  tributaries;  two  first-order 
streams  join  to  make  a second-order  stream,  and 
so  on;  Strahler  1957,  Everest  et  al.  1985:201). 
We  surveyed  91.4  km  of  fourth-order,  50.6  km 
of  fifth-order,  and  39.0  km  of  sixth-order 
streams,  the  mean  widths  of  which  were  4.2  m 
(range  = 1-30  m,  n = 203  randomly  selected 
points  along  Drift  and  Lobster  creeks),  10.1  m 
(range  = 2-25  m,  n = 203),  and  16.2  m (range 
= 3-38  m,  n — 100),  respectively.  The  maritime 
climate  was  characterized  by  mild,  wet  winters 
and  cool,  dry  summers.  Annual  precipitation 
was  180-300  cm,  75-85%  of  which  fell  during 
October-March.  Mean  temperature  seldom  fell 
below  0°  C in  the  winter,  and  summer  temper- 
atures rarely  exceeded  27°  C (Franklin  and  Dyr- 
ness  1973). 

Vegetation  upslope  of  riparian  areas  in  the 
Coast  Range  was  characteristic  of  the  western 
hemlock  ( Tsuga  heterophylla ) zone  (Franklin 
and  Dymess  1973)  and  was  dominated  by  sub- 
climax Douglas-fir  ( Pseudotsuga  menziesii), 
western  hemlock,  western  red  cedar  ( Thuja  pli- 
cata ),  and  red  alder  (Alnus  rubra).  Upslope  serai 
stages  ranged  from  recently  harvested  to  mature 
forests  (trees  >200  years  old).  Riparian  areas 
were  typically  forested  by  red  alder,  Douglas-fir, 
bigleaf  maple  ( Acer  macrophyllum),  and  west- 
ern red  cedar. 

Microhabitat  and  macrohabitat. — We 
searched  for  active  and  old  dipper  nests  in  1994 
{n  = 51)  along  Drift  and  Lobster  creeks  and 
along  the  23  additional  streams  to  characterize 
microhabitat,  macrohabitat,  and  streamside  hab- 
itat characteristics.  We  surveyed  all  streams  on 
foot  and  searched  within  5 m of  the  water’s  edge 
for  all  sites  capable  of  supporting  a nest  (here- 
after, nest  site),  including  mid-stream  boulders, 
debris  jams,  rootwads,  logs  >30  cm  in  diameter, 
bridges,  cliffs,  and  steep  banks.  We  collected 
microhabitat,  macrohabitat,  and  streamside  hab- 
itat data  at  every  site.  Microhabitat  variables 
measured  at  nest  sites  on  cliffs  included  height, 
width,  and  depth  of  the  supporting  ledge;  the 
average  thickness  of  moss  on  the  ledge  or  cliff; 
indicators  of  shelter  from  the  weather  (typically 
overhanging  vegetation  or  a rock  overhang),  ter- 
restrial predator  access,  and  whether  the  site  had 
a near-horizontal  ledge  or  platform  >10  X 10 


Loegering  and  Anthony  • NEST-SITE  SELECTION  OF  AMERICAN  DIPPERS 


283 


FIG.  1.  Illustration  of  a typical  mountain  stream 
bridge  in  the  Oregon  Coast  Range,  showing  the  sup- 
port beams  and  cross  member.  Typical  American  Dip- 
per nest  location  (*)  and  cross-member  angle  (0,  <90°) 
also  are  shown. 


cm.  We  considered  a nest  inaccessible  to  terres- 
trial predators  if  the  nest  ledge  did  not  extend 
horizontally  to  the  surrounding  upslope,  was  >1 
m high  or  perched  out  over  the  stream,  and  the 
cliff  was  smooth  enough  to  thwart  climbing 
predators,  such  as  American  mink  ( Mustela  vi- 
son ).  We  defined  macrohabitat  variables  as  cliff 
height  and  length,  cliff  slope  or  verticality  (90° 
was  exactly  vertical,  cliffs  <90°  sloped  away 
from  the  stream,  and  cliffs  >90°  sloped  out  over 
the  stream’s  edge),  and  cliff  vertical  area  (area 
of  cliff  face  that  was  >90°).  We  also  recorded 
height  of  the  ledge  or  nest  above  the  ground  or 
streambed,  the  height  from  nest  to  an  overhang 
above  (if  present),  and  the  horizontal  distance 
from  the  nest  to  stream  edge  at  base  winter  flow 
(hereafter,  setback  distance).  Setback  distance 
was  zero  for  nests  placed  directly  above  the 
edge  of  the  stream,  positive  for  nests  placed 
over  dry  land,  and  negative  for  nests  positioned 
over  the  stream.  We  used  winter  base  flow  be- 
cause dippers  selected  breeding  sites  in  Febru- 
ary and  March  (JPL  pers.  obs.)  when  streams 
were  at  this  level.  For  nests  in  logs  or  log  cav- 
ities, macrohabitat  variables  also  included  the 
diameter  of  the  log  and  whether  or  not  the  log 
was  coniferous.  For  nests  at  bridges  (n  = 11 
with  nests,  n = 11  without),  we  also  recorded 
the  cross-member  angle,  which  was  the  acute 
angle  (i.e.,  <90°  in  a horizontal  plane)  formed 
by  the  cross  member  and  one  of  the  load-bear- 
ing beams  (0  in  Fig.  1).  In  our  study,  dipper 


nests  on  bridges  typically  were  placed  in  this 
acute  angle  formed  by  the  load-bearing  support 
beam  and  the  cross  member. 

To  assess  the  availability  of  nest  sites  that 
were  suitable  but  not  used  by  dippers,  we  iden- 
tified every  site  in  our  study  basins  that  ap- 
peared suitable — based  on  sites  described  in  the 
literature  (Price  and  Bock  1983,  Kingery  1996) 
and  from  our  own  experience — but  did  not  cur- 
rently hold  a nest  ( n — 42).  We  erred  on  the 
side  of  possibly  including  unsuitable  sites  rather 
than  conservatively  excluding  sites  that  might 
have  served  as  nest  sites.  We  characterized  the 
microhabitat  and  macrohabitat  at  these  sites,  but 
did  not  compare  them  statistically  to  known  nest 
sites. 

Streamside  habitat. — In  1994,  we  measured 
seven  variables  (stream  shading,  distance  to  hu- 
man activity,  valley  form,  adjacent  land  use, 
canopy  cover,  stream  bank  vegetation,  and  ri- 
parian zone  vegetation)  to  characterize  and  com- 
pare streamside  and  riparian  zone  habitat  at  all 
active  and  old  nests  ( n = 22)  and  at  506  ran- 
domly selected  locations  along  Drift  and  Lob- 
ster creeks.  Streamside  habitat  was  not  assessed 
at  nests  in  other  basins.  None  of  the  randomly 
selected  locations  had  a dipper  nest  present  or 
the  microhabitat  and  macrohabitat  suitable  for  a 
dipper  nest.  We  visually  estimated  stream  shad- 
ing as  the  percentage  of  a transect  across  the 
stream  that  was  shaded  from  directly  overhead. 
Distance  to  human  activity  was  estimated  as  the 
straight-line  distance  (m)  to  areas  frequented  by 
humans  (e.g.,  roads,  dwellings,  etc.).  We  defined 
valley  form  as  either  constrained  (valley  floor 
<2  X the  width  of  the  active  channel)  or  un- 
constrained (valley  floor  >2  X the  width  of  the 
active  channel).  Adjacent  land  use  was  classified 
as  either  managed  forest  or  other  (e.g.,  residen- 
tial, agriculture,  pasture,  wilderness  area,  or 
campground).  We  visually  estimated  canopy 
cover  (nearest  percent)  in  a 5-m-diameter  plot 
25  m from  the  stream.  We  characterized  stream- 
bank  (immediately  adjacent  to  the  stream)  and 
riparian  zone  (25  m from  the  stream)  vegetation 
according  to  the  dominant  overstory  species, 
whether  the  dominant  vegetation  was  composed 
of  mature  trees  (woody  vegetation  >5  m tall 
versus  structurally  simpler,  non-tree  vegetation), 
and  whether  the  vegetation  was  coniferous. 
Thus,  there  were  four  categories  of  dominant 
vegetation:  conifer  trees  (e.g.,  Douglas-fir,  all 
size  classes  >5  m tall),  non-conifer  trees  (e.g., 


284 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


red  alder),  non-tree  conifers  (e.g.,  Douglas-fir, 
0-15  years,  <5  m tall),  and  non-tree,  non-co- 
nifers (e.g.,  shrubs,  grasses,  and  forbs).  For  the 
analyses,  we  used  two  binary  variables  (tree 
versus  non-tree;  conifer  versus  non-conifer)  to 
simplify  this  vegetation  assessment.  We  referred 
to  Hitchcock  and  Cronquist  (1973)  to  identify 
vegetation. 

Productivity. — During  1993-1995,  we 

searched  for  and  monitored  active  dipper  nests 
along  Drift  and  Lobster  creeks  to  assess  repro- 
ductive success  (n  = 16  nest  sites  and  48  nest- 
ing attempts  over  the  3 years).  We  examined  all 
nests  of  both  first  and  second  broods  at  least 
weekly,  noting  the  number  of  eggs,  chicks,  or 
fledged  young,  and  often  checked  nests  more 
frequently  near  the  estimated  fledging  date,  as 
recommended  by  Stanley  (2004).  Chicks  were 
uniquely  color  banded  at  10-14  days  of  age, 
and  hatching  dates  were  based  on  nest-initiation 
dates  and  growth  characteristics  indicative  of 
chick  age  (Sullivan  1973).  We  considered  a nest 
or  brood  successful  if  at  least  one  egg  hatched 
or  at  least  one  chick  fledged,  respectively.  The 
number  of  eggs  hatched  was  determined  during 
the  first  visit  to  the  nest  following  hatching,  and 
we  estimated  the  number  of  chicks  fledged  by 
counting  the  number  of  recently  fledged  young 
near  the  nest  during  or  after  fledging.  If  no 
fledged  young  were  observed,  we  assumed 
number  fledged  to  be  equal  to  the  number  of 
young  present  at  the  previous  nest  check  as  long 
as  the  previous  nest  check  was  >20  days  after 
hatching,  and  there  were  no  signs  of  nest  dis- 
turbance. We  also  identified  sources  of  nest  fail- 
ure whenever  possible. 

Created  nest  sites. — In  August  1993  and 
1994,  we  constructed  nine  nest  structures  (five 
nest  boxes,  two  log  cavities,  and  two  cliff  ledg- 
es; hereafter  referred  to  as  “created”  nest  sites) 
along  a segment  of  Drift  Creek  (9,480  m long) 
and  compared  dipper  nest  abundance  to  that 
along  a comparable  portion  of  Lobster  Creek 
(7,800  m long) — an  unaltered  control — to  assess 
nest  site  availability  and  saturation.  Both  reaches 
were  similar  in  size,  gradient,  geomorphology, 
and  adjacent  land  use.  We  constructed  nest  box- 
es (Loegering  1997)  similar  to  those  used  by 
Hawthorne  (1979)  in  California  and  Jost  (1970) 
in  Europe.  The  open  cavities  were  made  by  us- 
ing a brace  and  bit  in  the  ends  of  two,  nearly 
horizontal  logs  extending  over  the  stream  (min- 
imum dimensions  were  15  X 19  X 15  cm).  We 


used  a hammer  and  chisel  to  construct  two  ledg- 
es on  sandstone  cliffs  lacking  a mossy  covering. 
Two  of  the  five  nest  boxes  were  glued  to  the 
underside  of  flat-bottomed,  concrete  bridges 
(1994);  one  was  glued  to  the  wall  of  a fish  lad- 
der; one  was  screwed  to  the  inside  top  of  a 3- 
m-diameter  culvert;  and  one  was  screwed  to  the 
bottom  of  a stream- spanning  log.  All  structures 
were  >500  m from  known  nest  sites.  We  re- 
corded nest- site  use  as  we  monitored  nests  dur- 
ing 1993-1995;  during  1996-1998,  we  searched 
these  two  reaches  at  least  twice  each  year  and 
noted  only  whether  the  nest  sites  were  in  use. 
We  used  Analyses  of  Covariance  (PROC  GLM; 
SAS  Institute,  Inc.  1989)  to  compare  number  of 
active  nest  sites  between  Drift  and  Lobster 
creeks  for  the  1993-1998  breeding  seasons. 

Statistical  analyses. — We  categorized  nest 
sites  into  five  types,  based  on  their  substrate 
(hereafter  referred  to  as  nest  type):  nest  boxes, 
rock  or  moss-covered  cliff  ledges,  bridges,  cav- 
ities or  hollows  in  logs  (log  cavities),  and 
streambank  roots  or  rootwads.  We  considered 
multiple  nests  in  close  proximity  (<5  m)  as  rep- 
resentative of  one  breeding  attempt  and  one  ac- 
tive nesting  area;  within  and  across  years,  dip- 
pers may  build  more  than  one  nest  at  slightly 
different  locations,  but  will  only  use  one  nest 
(Kingery  1996).  We  observed  no  simultaneous- 
ly active  nesting  attempts  that  were  closer  than 
400  m,  although  others  have  reported  closer 
nesting  (S.  A.  H.  Osborn  pers.  comm.). 

We  used  logistic  regression  analysis  (PROC 
LOGISTIC  and  PROC  GENMOD;  SAS  Insti- 
tute, Inc.  1989)  with  a forward  variable-selec- 
tion routine  to  build  models  for  assessing  nest- 
site  selection — specifically  (1)  to  distinguish  be- 
tween bridges  used  and  not  used  by  dippers,  and 
(2)  to  compare  streamside  habitat  at  dipper  nests 
with  randomly  selected  streamside  habitat.  We 
used  a binary  response  variable  in  each  model 
to  indicate  dipper  use  (1)  versus  no  use  (0).  At 
bridges,  the  explanatory  variables  we  considered 
were  the  length,  width,  and  height  of  the  ledge; 
the  vertical  distance  to  streambed;  and  the  set- 
back distance  of  the  nest.  For  streamside  habitat, 
we  evaluated  stream  shading,  the  distance  to  hu- 
man activity,  valley  form,  adjacent  land  use, 
canopy  cover,  stream  bank  vegetation,  and  ri- 
parian zone  vegetation.  At  each  step,  all  vari- 
ables under  consideration  were  evaluated,  and 
the  variable  with  the  greatest  explanatory  power 
(greatest  reduction  in  model  deviance)  was  add- 


Loegering  and  Anthony  • NEST-SITE  SELECTION  OF  AMERICAN  DIPPERS 


285 


ed  to  the  model  (i.e.,  we  ran  each  model  chang- 
ing only  the  variable  of  interest  and  manually 
calculating  the  reduction  in  deviance).  We  ter- 
minated model-building  when  the  additional 
variable  did  not  improve  the  model’s  explana- 
tory power  by  a drop  in  deviance  (P  ^ 0.10). 
We  used  a liberal  significance  level  for  variable 
entry  because  more  conservative  levels  often 
fail  to  identify  important  variables  (Hosmer  and 
Lemeshow  2000:95).  All  models  met  the  Hos- 
mer and  Lemeshow  goodness-of-fit  test  (P  > 
0.050,  Hosmer  and  Lemeshow  2000).  No  two 
variables  were  highly  correlated  (all  r < 0.60, 
no  multicollinearity;  Neter  et  al.  1989);  models 
also  met  the  assumption  of  linearity  (Neter  et 
al.  1989).  We  tested  all  first-order  interaction 
combinations  (i.e.,  crossed  effects)  of  the  sig- 
nificant variables  for  each  model  after  the  initial 
variable  selection  (Neter  et  al.  1989,  Hosmer 
and  Lemeshow  2000).  We  identified  variables 
that  distinguished  between  (1)  microhabitat  and 
macrohabitat  at  bridges  used  by  dippers  versus 
those  not  used  and  (2)  streamside  habitats  where 
we  located  nests  versus  locations  that  we  se- 
lected at  random.  We  included  three  indicator 
(dummy)  variables  in  all  regression  models,  one 
for  basin  and  two  for  stream  order,  because  our 
objective  was  to  examine  habitat  selection  pat- 
terns after  accounting  for  any  effects  of  the  two 
stream  basins  and  three  stream  orders  (Strahler 
1957).  All  odds  ratios  (Hosmer  and  Lemeshow 
2000:50)  from  logistic  regression  analyses  are 
reported  relative  to  a base  comparison  (i.e.,  odds 
ratio  = 1).  An  odds  ratio  is  the  multiplicative 
likelihood  of  use  given  a one-unit  increase  in 
the  value  of  a given  variable.  Odds  <1  indicate 
that  an  increase  in  the  value  of  that  variable 
decreases  the  likelihood  of  use,  whereas  odds 
>1  indicate  a greater  likelihood  of  use  with  an 
incremental  increase  in  the  value  of  that  vari- 
able. 

We  used  the  Mayfield  method  (Mayfield 
1961,  1975)  to  determine  nest  survival  in  each 
stage  of  nesting,  and  program  MICROMORT 
(Heisey  and  Fuller  1985)  to  calculate  daily  sur- 
vival probabilities  and  95%  confidence  intervals 
(Cl).  We  report  bias-adjusted  interval  survival 
rates  (Heisey  and  Fuller  1985).  Estimates  were 
based  on  a 44-day  nesting  period  ( 1 9 egg-laying 
and  incubation  days,  and  25  brood-rearing 
days);  survival  was  calculated  for  each  stage 
and  the  overall  period.  We  calculated  survival 
based  on  exposure  days  (total  number  of  days 


observed).  When  we  observed  a nest  or  brood 
failure,  we  used  the  mid-date  between  the  last 
visit  and  the  previous  visit  as  the  date  of  failure. 
Although  we  did  observe  nests  with  unknown 
causes  of  failure,  we  were  certain  about  the  fate 
of  each  nest  (Manolis  et  al.  2000).  We  used  a 
Z-test  (Hensler  1985)  to  compare  observed  daily 
nest  survival  between  the  two  stream  basins, 
among  the  five  nest  types,  and  between  natural 
and  created  nest  sites.  Each  nest  site  hosted  one, 
two,  or  (rarely)  three  breeding  attempts  each 
season.  For  each  nest  site,  we  calculated  the 
mean  and  total  number  of  chicks  that  fledged. 
We  used  nonparametric  Wilcoxon’s  rank  sum 
(normal  approximation)  and  Kruskal-Wallis 
tests  (chi-square  approximation;  Sokal  and 
Rohlf  1981),  both  conducted  with  PROC 
NPAR1WAY  (SAS  Institute,  Inc.  1989)  to  com- 
pare the  mean  number  of  chicks  fledged  be- 
tween basins  and  among  nest  types,  respective- 
ly. We  used  Spearman’s  rank  correlation  to  re- 
late the  mean  number  of  chicks  produced  at 
each  site  to  19  measures  of  microhabitat,  ma- 
crohabitat, and  streamside  habitat  characteris- 
tics: mean  and  maximum  moss  thickness  on 
cliffs;  length,  width,  and  depth  of  the  nest  ledge; 
length,  height,  and  area  of  the  cliff’s  vertical 
surface;  height  and  length  of  the  nest-site  cliff; 
vertical  height  above  and  below  the  nest;  ver- 
ticality  of  the  cliff;  diameter  of  the  log  associ- 
ated with  log  nests;  setback  distance  of  the  nest; 
bridge  cross-member  angle;  stream  shading;  dis- 
tance to  humans;  and  streamside  canopy  cover. 
To  control  Type  I error  rates  during  these  si- 
multaneous multiple  comparisons,  we  used  the 
Bonferroni  method  (Bart  and  Notz  2005)  be- 
cause of  its  simplicity  and  few  assumptions. 
This  method  guarantees  a significance  level,  a, 
for  M comparisons  by  adjusting  the  critical  val- 
ue for  each  comparison  to  oJM.  We  used  a 
paired  r-test  to  remove  the  potential  confound- 
ing effect  of  nest-site  quality  when  comparing 
the  number  of  young  fledged  from  first  broods 
versus  second  broods.  To  do  this,  we  calculated 
the  difference  in  number  of  young  fledged  (first 
brood  - second  brood)  at  each  site  that  raised 
two  broods  and  ran  a /-test  on  the  difference 
(Hq:  difference  = 0).  We  used  SAS  (ver.  6. 1 and 
9.1;  SAS  Institute,  Inc.  1989)  to  complete  all 
statistical  analyses.  Values  reported  are  means 
± 1 SE. 


286 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  1 . Microhabitat  and  macrohabitat  characteristics  at  American  Dipper  nest  sites  on  cliff  ledges,  under 
bridges,  in  log  cavities,  and  on  roots  and  rootwads  in  the  Oregon  Coast  Range,  1994.  Uneven  sample  sizes 
indicate  variables  that  could  not  be  safely  evaluated  (e.g.,  a cliff  ledge  too  high  to  reach  or  rootwads  in  large, 
unstable  debris  piles)  or  would  be  nonsensical  (e.g.,  ledge  dimensions  for  either  enclosed  log  cavities  or  nests 
placed  in  a tangle  of  roots)  for  one  or  more  sites. 

Cliff  ledges  ( n = 

20) 

Bridges  (n  = 11) 

n 

Mean 

SE 

Range 

n 

Mean 

SE 

Range 

Microhabitat 

Ledge  length  (cm) 

19 

185.7 

129.1 

20-2,500 

li 

363.0 

148.5 

10-1,220 

Ledge  width  (cm) 

19 

22.1 

1.8 

10-35 

li 

17.9 

1.7 

10-31 

Ledge  to  overhang  (cm) 

19 

113.3 

39.8 

18-0= 

n 

55.6 

8.9 

24-110 

Macrohabitat 

Cliff  height  (m) 

19 

3.8 

0.4 

2. 1-9.0 

b 

— 

— 

— 

Cliff  length  (m) 

19 

20.5 

3.7 

3-50 

— 

— 

— 

— 

Cliff  verticality3 

19 

94.6 

2.3 

78-120 

— 

— 

— 

— 

Cliff  vertical  area  (m2) 

19 

44.1 

11.6 

2-225 

— 

— 

— 

— 

Height  below  nest  (m) 

20 

2.4 

0.2 

1.2-4.4 

11 

2.7 

0.2 

1.7-3. 8 

Setback  distance  of  the 

nest  (m) 

20 

-0.1 

0.1 

-1.0-0.3 

11 

-2.0 

0.5 

-4.4-0 

a 90°  is  exactly  vertical,  cliffs  <90°  slope  away  from  the  stream,  and  cliffs  >90°  slope  out  over  the  stream  edge. 
b Parameter  not  applicable  to  the  substrate  type. 


RESULTS 

We  searched  1 8 1 km  of  stream  in  eight  basins 
in  the  central  Coast  Range  in  1994  and  found 
51  active  and  old  nests.  Nest  densities  in  indi- 
vidual streams  ranged  from  1.9  to  3.4  nests/10 
km  (Loegering  1997).  We  found  20  nests  on 
cliff  ledges,  11  nests  under  bridges,  17  nests  in 
logs,  and  3 nests  associated  with  rootwads. 
Nests  on  cliffs  were  typically  placed  on  rock 
ledges;  however,  in  three  instances,  dippers  cre- 
ated ledges  by  selecting  a cliff  with  a thick, 
mossy  mat,  slipping  behind  the  moss  and  push- 
ing it  away  from  the  cliff  face,  thereby  creating 
a space  to  place  a nest.  This  method  of  ledge 
creation  has  not  been  described  previously  and 
may  be  limited  to  areas  where  moss-covered 
cliffs  are  relatively  common,  such  as  in  the  Pa- 
cific Northwest.  Nests  on  bridges  were  placed 
on  horizontal  beams  or,  in  many  instances,  on 
beams  with  ledges  that  sloped  downward  at  a 
45°  angle,  often  adjacent  to  a vertical  cross 
member.  Logs  that  hosted  dipper  nests  generally 
were  within  45°  of  horizontal,  were  damaged  by 
flood  events,  and  often  had  a shattered  end  or 
heart-rot  that  provided  a cavity  or  platform  on 
which  a nest  could  be  placed.  Roots  and  root- 
wads  used  by  dippers  as  nest  sites  were  either 
created  or  exposed  by  erosion  during  flood 
events.  We  also  found  42  sites  that  were  unoc- 


cupied but  had  the  best  potential  for  serving  as 
future  nest  sites. 

Microhabitat  and  macrohabitat. — Dipper 
nests  in  the  central  Oregon  Coast  Range  were 
typically  sheltered  from  the  weather  from  above 
(>85%  for  all  nest  types,  n — 51),  and  100% 
were  placed  on  a ledge  or  root.  On  cliffs,  dip- 
pers selected  ledges  that  were  >20  cm  long  X 
10  cm  wide  (Table  1).  We  recorded  one  nest 
that  was  placed  on  a ledge  with  only  18  cm  of 
overhead  clearance  between  the  ledge  and  a 
rock  overhang  (ledge  to  overhang;  Table  1),  but 
most  had  considerably  more  clearance.  Cliffs 
ranged  considerably  in  size;  however,  those  used 
by  dippers  were  vertical  or,  more  often,  leaned 
out  over  the  stream  (mean  cliff  verticality  = 95°; 
Table  1).  All  nests  were  safe  from  terrestrial 
predators  by  virtue  of  height  above  the  stream- 
bed  or  ground  and  setback  distance.  Nests  in  log 
cavities  tended  to  be  closer  (lower)  to  the 
streambed  than  other  nest  types,  and  all  were 
placed  over  the  stream  (Table  1).  Most  cavity 
nests  were  placed  in  coniferous  logs  (13/16;  tree 
species  was  not  recorded  for  one  log). 

Bridges  used  by  dippers  ( n = 11)  were  dis- 
tinguished from  unused  bridges  (n  = 11)  by 
their  height  and  the  angle  of  the  cross  member 
in  their  support  structure  (logistic  regression,  x2 
= 13.2,  df  = 2,  P = 0.001,  r2  = 0.70;  Table  2). 


Loegering  and  Anthony  • NEST-SITE  SELECTION  OF  AMERICAN  DIPPERS 


287 


TABLE  1 . Extended. 


Log  cavities  (n  = 17) 

Rootwads  ( n = 3) 

n 

Mean 

SE 

Range 

n 

Mean 

SE 

Range 

7 

36.4 

19.5 

10-114 

i 

29.0 

7 

29.8 

3.7 

24-44 

i 

22.0 

7 

115.8 

96.1 

15-500 

2 

38.0 

10.0 

28-48 

_ 

_ 

_ 

_ 

3 

3.1 

0.9 

2. 1-4.0 

— 

— 

— 

— 

2 

16.4 

8.7 

7.7-25.0 

— 

— 

— 

— 

2 

97.5 

7.5 

90-105 

— 

— 

— 

— 

2 

52.9 

47.1 

5.9-100 

17 

1.7 

0.1 

0. 9-3.0 

3 

1.8 

0.1 

1. 7-2.1 

17 

-2.3 

0.3 

-3.8— 0.5 

3 

-0.1 

0.2 

— 0.4-0.2 

Sites  on  bridges  used  by  dippers  were  lower 
(closer  to  the  stream;  range  = 1.71-3.81  m,  n 
= 11)  than  bridges  not  used  (range  = 2.70-8.47 
m,  n = 10)  by  dippers  (odds  ratio  = 0.01,  90% 
Cl  = 0-0.44;  Tables  2,  3),  and  the  probability 
of  dipper  use  decreased  as  bridge  cross-member 
angle  increased  to  90°  (odds  ratio  = 0.83,  90% 
Cl  = 0.69-1.0).  Overall,  bridge  cross  members 
supporting  nests  were  set  at  sharper  angles 
(79.4°  ± 5.05,  n = 10)  than  those  without  nests 
(85.6°  ± 2.08,  n = 8);  this  use  pattern  was  most 
pronounced  on  bridges  supported  by  concrete  I- 
beams  (bridges  supporting  nests:  56.7°  ± 2.8,  n 
= 3;  bridges  not  used:  84.4°  ± 3.1,  n = 5). 

Streamside  habitat  selection. — Stream  shad- 
ing, valley  form,  and  the  distance  to  human  fea- 
tures distinguished  dipper  nest  sites  from  other 
available  (unused)  locations  (logistic  regression, 
X2  = 34.4,  df  = 7,  P < 0.0001,  r2  = 0.22). 
Streams  at  dipper  nests  were  more  shaded  than 


the  available  habitat  (58%  versus  34%,  respec- 
tively; Table  4);  for  each  10%  increase  in  stream 
shading,  the  likelihood  of  dippers  selecting  that 
area  increased  by  1.6  times  (90%  Cl  on  odds 
ratio  = 1.29-1.85;  Table  5).  Streams  near  dipper 
nests  also  were  constrained  by  a steep  valley 
wall  more  often  than  they  were  at  randomly  se- 
lected available  sites  on  at  least  one  (91%  versus 
65%  of  the  observations,  respectively)  or  both 
(50%  versus  20%)  sides  of  the  stream.  Dipper 
nests  were  3.2  and  9.1  times  more  likely  to  oc- 
cur where  the  valley  walls  constrained  the 
stream  on  one  (90%  Cl  on  odds  ratio  = 0.9- 
1 1 .9)  or  both  sides  (90%  Cl  on  odds  ratio  = 
2.5-33.9)  than  in  unconstrained  reaches  (odds 
ratio  = 1;  Table  5).  Dipper  nests  were  located 
where  trees  dominated  both  sides  of  the  stream 
(91%  of  nest  locations  versus  68%  of  randomly 
selected  locations;  Table  4);  however,  we  were 
not  able  to  statistically  evaluate  the  importance 


TABLE  2.  Microhabitat  and  macrohabitat  characteristics  distinguishing  bridges  with  (n  = 11)  and  without 
(n  = 11)  American  Dipper  nests  in  the  Oregon  Coast  Range,  1994.  Odds  ratio  is  a multiplicative  likelihood  of 
use  given  a 1-unit  increase  in  the  value  of  a given  variable.  Odds  <1  indicate  that  an  increase  in  the  value  of 
that  variable  decreases  the  likelihood  of  use,  whereas  odds  >1  indicate  a greater  likelihood  of  use  with  an 
incremental  increase  in  the  value  of  that  variable. 


Variable 

Parameter  estimate 

SE 

Odds  ratio  (lower,  upper 
90%  Cl) 

Intercept 

29.468 

15.124 

Height  above  streambed 

-4.468 

2.220 

0.01  (0.00,  0.44) 

Cross-member  angle 

-0.183 

0.111 

0.83  (0.69,  1 .00) 

288 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  3.  Microhabitat  and  macrohabitat  characteristics  of  bridges  with  and  without  American  Dipper  nests 
in  the  Oregon  Coast  Range,  1994.  The  equal  sample  sizes  of  used  and  unused  nest  sites  was  coincidental  (i.e., 
these  were  not  paired  analyses). 


Used 

Unused 

n 

Mean  ± SE 

n 

Mean  ± SE 

Microhabitat 

Ledge  length  (cm) 

11 

363.0  ± 148.5 

11 

486.5  ± 155.1 

Ledge  width  (cm) 

11 

17.9  ± 1.7 

11 

14.8  ± 3.3 

Ledge  to  overhang  (cm) 

11 

55.6  ± 8.9 

11 

47.7  ± 12.6 

Macrohabitat 

Vertical  distance  to  streambed  (m) 

11 

2.7  ± 0.2 

10 

4.6  ± 0.7 

Setback  distance  of  the  nest  (m) 

11 

-2.0  ± 0.5 

8 

-2.1  ± 0.6 

of  riparian  forests.  Lastly,  dipper  nests  were  lo- 
cated farther  from  human  activity  (e.g.,  roads) 
than  unused  sites  (474  m versus  310  m,  respec- 
tively). We  were  2.5  times  more  likely  to  find 
dipper  nests  for  each  additional  km  away  from 
human  activity  (90%  Cl  on  odds  ratio  = 1.1— 
5.7;  Table  5),  although  our  farthest  nest  was 
only  2.6  km  from  a road  (Table  5). 


Productivity. — Reproductive  success  was 
markedly  high  during  each  nesting  stage  ( n = 
48  nesting  attempts  at  1 6 nest  sites  over  3 years 
along  Drift  and  Lobster  creeks).  Overall  daily 
Mayfield  survival  of  dipper  nests  was  0.991 
(1,219  exposure  days,  10  losses,  44-day  interval 
survival  - 0.692,  95%  Cl  = 0.556-0.871).  Dai- 
ly Mayfield  nest  survival  during  egg  laying  and 


TABLE  4.  Streamside  habitat  (mean  ± SE)  at  American  Dipper  nest  sites  and  randomly  selected  locations 
in  Drift  and  Lobster  creeks  in  the  Oregon  Coast  Range,  1994. 


Known  nest  sites  (n  = 22) 

Randomly  selected  locations 
(n  = 506) 

Stream  shading  (%) 

57.7  ± 5.6 

34.2  ± 1.3 

Distance  to  human  activity  (m) 

474.0  ± 131.9 

309.9  ± 21.2 

Riparian  zone  canopy  cover  (%) 

55.6  ± 3.4 

47.3  ± 0.9 

Riparian  zone  trees  (%)a 

One  bank 

4.6  ± 4.4 

27.3  ± 2.0 

Both  banks 

90.0  ± 6.1 

67.8  ± 2.1 

Riparian  zone  conifers  (%) 

One  bank 

36.4  ± 10.3 

32.0  ± 2.1 

Both  banks 

18.2  ± 8.2 

10.3  ± 1.4 

Stream  bank  trees  (%)a 

One  bank 

4.6  ± 4.4 

5.9  ± 1.1 

Both  banks 

4.6  ± 4.4 

1.4  ± 0.5 

Stream  bank  conifers  (%)a 

One  bank 

4.6  ± 4.4 

0.2  ± 0.2 

Both  banks 

b 

b 

Valley  form  (%  constrained) 

One  bank 

40.9  ± 10.5 

44.5  ± 2.2 

Both  banks 

50.5  ± 10.7 

20.2  ± 1.8 

Land  use  (%  managed  forests) 

One  bank 

0.0  ± 0.0 

17.2  ± 1.7 

Both  banks 

81.8  ± 8.2 

67.2  ± 2.1 

a Not  included  in  logistic  regression  analyses  because  too  few  nests  were  in  the  response  category  (i.e.,  ^2  nests  did  not  have  these  features). 
b All  values  were  zero. 


Loegering  and  Anthony  • NEST-SITE  SELECTION  OF  AMERICAN  DIPPERS 


289 


TABLE  5.  Riparian  habitat  variables  distinguishing  nest  sites  of  American  Dippers  (n  = 22)  and  randomly 
located  points  (n  = 506)  in  the  Oregon  Coast  Range,  1994.  We  entered  indicator  variables  for  basin  and  stream 
order  into  all  logistic  regression  models.  Odds  ratio  is  a multiplicative  likelihood  of  use  given  a 1-unit  increase 
in  the  value  of  a given  variable.  Odds  <1  indicate  that  an  increase  in  the  value  of  that  variable  decreases  the 
likelihood  of  use,  whereas  odds  >1  indicate  a greater  likelihood  of  use  with  an  incremental  increase  in  the  value 
of  that  variable. 


Variable 

Parameter  estimate 

SE 

Odds  ratio  (lower, 
upper  90%  Cl) 

Intercept 

-5.332 

1.052 

Basin  1 (design  variable) 

-0.470 

0.572 

0.63  (0.24,  1 .60) 

Order  4 (design  variable) 

-2.006 

0.928 

0.14  (0.03,  0.62) 

Order  5 (design  variable) 

-0.956 

0.841 

0.39  (0.10,  1.54) 

Stream  shading  (10%  increments) 

0.435 

0.109 

1.55  (1.29,  1.85) 

Constrained  valley  form 

One  bank 

1.159 

0.800 

3.19  (0.86,  11.87) 

Both  banks 

2.210 

0.799 

9.11  (2.45,  33.89) 

Distance  to  human  activity  (km) 

0.930 

0.487 

2.54  (1.14,  5.65) 

incubation  was  >0.988  (494.5  exposure  days,  6 
losses,  19-day  interval  survival  = 0.792,  95% 
Cl  = 0.658-0.954).  Furthermore,  daily  nest  sur- 
vival did  not  differ  between  Drift  (0.990)  and 
Lobster  (0.983)  creeks  (P  = 0.52),  among  nest 
types  (all  >0.981,  all  P > 0.05),  or  between 
created  (1.0)  and  natural  (0.981)  nest  sites  along 
Drift  Creek  (P  = 0.080).  Daily  Mayfield  sur- 
vival of  chicks  was  0.994  (724.5  exposure  days, 
4 losses,  25-day  interval  survival  = 0.869,  95% 
Cl  = 0.760-0.997),  and  did  not  differ  between 
Drift  (0.996)  and  Lobster  (0.992)  creeks  (P  = 
0.56),  among  nest  types  (all  >0.833,  all  P > 
0.16),  or  between  created  (1.0)  and  natural 
(0.991)  nest  sites  along  Drift  Creek  (P  = 0.16). 
In  1 1 attempts,  there  were  no  nest  or  brood  fail- 
ures at  created  nest  sites.  Of  the  48  nesting  at- 
tempts for  which  we  had  complete  histories,  > 1 
young  hatched  in  42  attempts  (87.5%)  and  >1 
young  fledged  in  37  attempts  (77%).  There  were 
no  obvious  sources  of  loss  for  eggs  or  chicks. 
All  six  nests  that  lost  their  entire  clutch  were 
found  empty  and  undisturbed,  and  three  of  the 
five  nests  where  all  chicks  were  lost  showed  no 
signs  of  disturbance;  one  nest  was  disturbed  and 
had  a new  male  in  the  territory,  and  one  nest 
contained  dead  chicks.  Neither  the  mean  num- 
ber of  chicks  fledged  per  nesting  attempt  per  site 
(2.3  ± 0.3,  range  = 0-4,  n — 16  sites;  Table 
6)  nor  the  total  number  of  chicks  fledged  per 
site  (mean  = 6.75  ± 1.1,  range  = 0-16,  n — 
16)  was  correlated  with  any  of  the  19  measure- 
ments of  microhabitat,  macrohabitat,  or  stream- 
side  habitat  characteristics  (Spearman’s  rank 


correlation,  all  P > 0.05;  Bonferroni-adjusted 
critical  value  for  experiment-wise  a = 0.05:  P 
— 0.003).  Mean  number  of  chicks  fledged  per 
attempt  per  site  also  did  not  differ  between  Drift 
and  Lobster  creeks  (Wilcoxon  rank  sum,  Z = 
—0.55,  df  = 15,  P = 0.58),  among  nest  types 
(Kruskal-Wallis,  X2  = 2.5,  df  = 4,  P = 0.64), 
or  between  first  and  second  broods  (paired  t - 
-0.52,  n = 14,  P = 0.61).  Overall  abundance 
of  nests  was  2.7  ± 0.7  nest  sites/linear  10  km 
of  stream  in  181  km  of  streams  in  the  Oregon 
Coast  Range  (n  = 39  streams,  mean  length  = 
4.6  ± 1.0  km). 

Nest  sites  in  our  study  were  used  repeatedly. 
Between  nesting  attempts,  dippers  typically  re- 
moved and  replaced  the  nest  cup  but  reused  the 
external  mossy  shell.  Of  the  12  nest  sites  we 
identified  in  1993,  8 were  used  every  year  for 
4 years  (otherwise  a nearby  site  within  the  same 
territory  was  active),  three  sites  were  idle  once 
during  1993-1996,  and  one  site  hosted  only  a 
single,  failed  nesting  attempt. 

Created  nest  sites. — By  1996,  all  created  nest 
sites  ( n = 9)  had  been  used  at  least  once  except 
for  one  nest  box  destroyed  by  flooding  in  early 
1996.  In  the  year  after  these  sites  were  created, 
the  number  of  active  nest  sites  on  the  experi- 
mental reach  doubled  from  three  nests  to  six 
nests,  and  the  number  remained  higher  on  Drift 
Creek  than  on  Lobster  Creek  (ANCOVA:  Fh]0 
= 6.6,  P = 0.029;  Fig.  2).  This  increase  repre- 
sents an  increase  in  the  number  of  dipper  breed- 
ing pairs,  not  additional  alternate  nest  sites,  be- 
cause we  could  uniquely  identify  one  or  both 


290 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  6.  Apparent  reproductive  success,  total  and  mean  (±  SE)  number  of  American  Dipper  young  fledged 
per  attempt  per  nest  site  along  Drift  and  Lobster  creeks  for  different  nest  substrate  types,  and  for  first  and  second 
broods  in  the  Oregon  Coast  Range,  1993-1995. 


Category 

No.  nesting 
attempts 
observed 

No.  nests 
hatching 
^1  egg 

No.  nests 
fledging 
si  chick 

Total  young 
fledged 

No.  sites 

No.  fledged 
per  attempt 
per  site 

p 

Overall 

Basin 

48 

42 

37 

108 

16 

2.3  ± 0.32 

0.58a 

Drift  Creek 

29 

26 

23 

68 

10 

2.5  ± 0.41 

Lobster  Creek 
Nest  substrate  type 

19 

16 

14 

40 

6 

2.1  ± 0.53 

0.64b 

Nest  box 

3 

3 

3 

9 

2 

2.9  ± 0.89 

Cliff  ledge 

20 

18 

15 

42 

5 

2.2  ± 0.57 

Bridge 

18 

15 

14 

42 

4 

2.5  ± 0.63 

Log  cavity 

6 

5 

5 

15 

4 

2.6  ± 0.63 

Rootwad 

Brood 

1 

1 

0 

0 

1 

0.0  ± 0.00 

0.61c 

Lirst 

28 

26 

22 

67 

14 

2.5  ± 0.39 

Second 

20 

16 

15 

41 

14 

2.7  ± 0.34 

a Wilcoxon  rank  sum  test  (normal  approximation). 
b Kruskal- Wallis  (chi-square  approximation). 
c Paired  r-test. 


mates  at  all  nests.  Nearly  all  adult  birds  (14  of 
16)  and  their  young  (17  of  23)  were  uniquely 
color  banded  at  the  nest  in  the  1st  year  of  the 
study  and  19  more  birds  were  banded  after  the 
breeding  season;  each  year  thereafter,  new  birds 
were  banded  as  they  arrived  in  the  study  area 
(140  birds  during  1993-1995).  Created  nest 
sites  were  colonized  both  by  new,  unbanded  im- 
migrants as  well  as  by  birds  that  had  previously 
bred  elsewhere  within  the  study  basins.  Assum- 
ing populations  were  similar  in  the  treated  and 


control  reaches  prior  to  the  treatment  (Drift 
Creek  actually  had  fewer  nests  prior  to  treat- 
ment), the  increased  number  of  nesting  dippers 
along  Drift  Creek  during  all  five  post-treatment 
years  is  indicative  of  a population  response. 

DISCUSSION 

Habitat  selection. — American  Dipper  nest- 
site  selection  was  disproportionately  influenced 
by  factors  at  the  largest  and  smallest  spatial 
scales.  Given  their  geographic  affinity  for  un- 


LIG.  2.  Number  of  active  American  Dipper  nests  along  a portion  of  Drift  and  Lobster  creeks,  Oregon  Coast 
Range,  1993-1998.  Seven  (A)  and  two  (B)  experimental  nest  sites  were  constructed  after  the  1993  and  1994 
breeding  seasons,  respectively,  along  Drift  Creek  (9,480-m  reach);  Lobster  Creek  (7,800-m  reach)  was  our 
reference  stream.  One  site  was  destroyed  by  flooding  prior  to  the  1995  (C)  and  1997  (D)  breeding  seasons. 


Loegering  and  Anthony  • NEST-SITE  SELECTION  OF  AMERICAN  DIPPERS 


291 


polluted,  swift  mountain  streams  in  western 
North  America  (Price  and  Bock  1983,  Kingery 
1996),  dippers  appear  to  require  structures  that 
are  large  enough  to  hold  a nest,  close  to  the 
stream,  and  high  enough  to  reduce  destruction 
from  predation  or  spring  flooding.  We  found  no 
nests  that  were  accessible  to  terrestrial  predators 
and  did  not  record  any  nest  loss  to  predators.  In 
contrast,  predation  was  the  most  important  fac- 
tor in  reducing  nest  success  of  American  Dip- 
pers in  British  Columbia  (Morrissey  2004)  and 
White-throated  Dippers  ( Cinclus  cine lus ) in 
Norway  (Efteland  and  Kyllingstad  1984),  sug- 
gesting that  predation  has  influenced  and  con- 
tinues to  influence  the  evolution  of  nest-site  se- 
lection. Moreover,  no  nests  we  found  were  >0.3 
m from  the  stream’s  edge.  Previously,  American 
Dipper  nests  have  been  found  in  trees  and 
shrubs,  and  farther  from  the  water  (Sullivan 
1966)  than  we  noted  for  American  Dippers  or 
which  Moon  (1923),  Robson  (1956),  Balat 
(1964),  and  Trochet  (1967)  noted  for  White- 
throated  Dippers;  however,  these  are  rare  occur- 
rences (Price  and  Bock  1983,  this  study).  Be- 
yond these  general  requirements,  dippers  exhib- 
ited great  flexibility  in  nest-site  selection.  Dip- 
pers will  nest  on  a diversity  of  stream  sizes  and 
substrates,  including  on  cliff  ledges,  under 
bridges,  on  midstream  boulders  (Sullivan  1973, 
Price  and  Bock  1983),  in  boxes  and  log  cavities, 
around  rootwads  (Hawthorne  1979,  Morrissey 
2004,  this  study),  and  occasionally  in  gaps  in 
rock  walls  and  bridge  drainpipes  (Everett  and 
Marti  1979).  A comprehensive  study  of  nest 
characteristics  of  White-throated  Dippers  in  Eu- 
rope yielded  similar  results  (Shaw  1978). 

Streamside  habitat  at  dipper  nests  differed 
from  that  available,  which  may  reflect  micro- 
habitat and  macrohabitat  selection.  Geomorphi- 
cally  constrained  valleys  have  steeper  slopes, 
more  cliffs,  and  a greater  potential  for  micro- 
habitats that  are  suitable  for  nesting  (e.g.,  ledg- 
es) than  unconstrained  valleys.  Most  dipper  nest 
sites  were  located  where  trees  dominated  the  ad- 
jacent riparian  zone  on  both  sides  of  the  stream, 
and  nest  sites  were  located  where  the  stream 
was  more  heavily  shaded  by  those  trees.  The 
importance  of  streamside  forests  extends  be- 
yond the  observed  pattern.  Riparian-zone  trees 
contribute  large  logs  to  the  stream  and  stream 
bank  (Meehan  et  al.  1977,  Swanson  and  Lien- 
kaemper  1978,  Keller  and  Swanson  1979,  Se- 
dell  et  al.  1988).  Logs  from  mature  hardwoods 


and  conifers  not  only  add  wood  to  the  stream 
and  increase  its  structural  complexity,  but  also 
may  provide  nest  sites.  To  a large  extent,  future 
nest-site  availability  may  depend  on  there  being 
a conifer  component  in  riparian  areas;  >80%  of 
the  nests  that  we  found  in  log  cavities  were  in 
coniferous  logs,  yet  in  only  42%  of  the  basins 
were  conifers  the  dominant  trees  on  either  side 
of  the  stream.  Moreover,  32%  of  the  nests  we 
found  were  placed  on  or  in  large  dead  wood; 
however,  nests  associated  with  large  wood  or 
logs  are  listed  as  only  occasional  (0-5%  of  ob- 
served nests)  or  are  not  mentioned  at  all  in  pre- 
vious reviews  (Ealey  1977,  Kingery  1996). 

We  have  revealed  at  least  three  lines  of  evi- 
dence that  suitable  nest  sites  for  dippers  may  be 
in  short  supply.  First,  during  our  surveys,  only 
1 of  42  sites  that  we  identified  as  possible  nest 
sites — but  with  no  evidence  of  past  use — met 
minimal  criteria  that  we  derived  from  the  liter- 
ature (Price  and  Bock  1983,  Kingery  1996). 
Twenty-eight  (67%,  n = 42)  of  these  sites  failed 
to  meet  the  minimal  requirements  of  a platform 
or  ledge  ^10  X 10  cm,  inaccessibility  to  ter- 
restrial predators,  being  > 1 m above  the  stream- 
bed,  and  (for  bridges)  having  a cross-member 
angle  of  <80°.  Eight  suitable  nest  sites  (19%) 
were  <500  m from  a dipper  nest  and  likely 
within  the  same  territory  (see  Ealey  1977,  Price 
and  Bock  1983,  this  study).  Of  the  remaining 
six  (14%)  unused  sites,  three  were  >1.2  m from 
the  water,  and  two  sites  were  subjectively  clas- 
sified as  “poor”  sites.  Second,  sites  that  were 
used  were  occupied  nearly  every  year.  Third,  the 
use  of  created  nest  sites  further  corroborated  the 
possibility  that  nest  sites  are  limited.  All  created 
structures  were  colonized  within  2 years  of  their 
creation,  except  for  one  that  was  destroyed  by 
flooding  before  it  could  be  used.  This  more  than 
doubled  the  number  of  active  nests  on  Drift 
Creek  from  1993  to  1996.  Overall,  the  lack  of 
suitable  but  unused  sites,  the  high  re-occupancy 
rate,  and  the  rapid  colonization  of  created  sites 
indicates  that  suitable  nest  sites  may  have  lim- 
ited the  abundance  of  dippers  in  our  study  ba- 
sins. However,  there  may  be  regional  variation, 
as  Feck  and  Hall  (2004)  found  several  unoc- 
cupied sites  in  Wyoming  and  concluded  that 
macroinvertebrate  prey  strongly  affected  dipper 
breeding  presence  in  their  study  area. 

Productivity. — Reproductive  success  was  not 
correlated  with  any  feature  of  nest-site  habitat 
at  the  microhabitat,  macrohabitat,  or  streamside 


292 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


scales.  Productivity  and  survival  were  high  and 
did  not  differ  among  nest  sites  or  nest  types. 
Similarly,  Feck  and  Hall  (2004)  found  that  pro- 
ductivity was  unrelated  to  any  of  the  physical 
or  biological  variables  they  considered.  Con- 
versely, Price  and  Bock  (1983)  found  nest  and 
fledging  success  positively  correlated  with  an 
index  of  nest-site  quality.  We  did  not  detect  any 
reproductive  advantage  attributable  to  differenc- 
es in  nest-site  characteristics  for  the  16  nest  sites 
in  our  study;  once  the  minimum  criteria  for  suit- 
ability were  met,  nests  were  generally  success- 
ful. The  number  of  chicks  fledged  per  attempt 
in  Oregon  was  greater  and  less  variable  than  that 
reported  in  Colorado  (Price  and  Bock  1983), 
Montana  (Bakus  1959,  Sullivan  1973),  Wyo- 
ming (Feck  and  Hall  2004),  Alberta  (Ealey 
1977),  or  Europe  (Balat  1964),  but  lower  than 
what  was  reported  in  British  Columbia  (Morris- 
sey 2004).  In  our  study,  the  abundance  of  breed- 
ing American  Dippers  was  lower  than  it  was  in 
British  Columbia  (Morrissey  2004),  Colorado 
(Price  and  Bock  1983),  or  Alberta  (Ealey  1977); 
however,  both  Sullivan  (1973)  and  Osborn 
(1999)  found  nest  densities  in  Montana  (0.3 
nests/km)  that  were  comparable  to  ours  (0.3 
nests/km)  and  noted  that  the  majority  of  nests 
found  were  under  bridges.  Suitable  nest  sites  ap- 
pear to  limit  the  breeding  population  elsewhere 
(e.g..  Price  and  Bock  1983,  Fite  1984,  Kingery 
1996,  Osborn  1999). 

Minimum  nest-site  requirements. — Based  on 
the  relatively  high  levels  of  productivity,  dippers 
appeared  to  select  nest  sites  that  met  minimal 
requirements  for  a site  to  be  suitable  for  suc- 
cessful reproduction,  specifically  (1)  an  ade- 
quate ledge  or  physical  space  for  a nest,  (2) 
close  proximity  to  the  stream’s  edge,  (3)  safety 
from  terrestrial  predators,  and  (4)  a low  chance 
of  spring  flooding.  The  presence  of  an  adequate 
ledge  seems  obvious,  but  not  all  cliffs  offer  suit- 
able nesting  space.  The  smallest  log  cavity  used 
was  13  cm  in  height,  width,  and  depth.  Suitable 
ledges  also  should  exceed  13  cm  and  be  larger 
if  the  ledge  is  not  horizontal.  Nests  can  be  in- 
accessible to  ground  predators  because  of  the 
elevation  of  the  nest  ledge  and/or  the  distance 
from  the  stream’s  edge.  Every  dipper  nest  we 
located  ( n = 51)  was  over  the  stream  or  its  edge 
(>77%  had  <0  m setback  distance),  or  was 
within  0.3  m of  the  stream’s  edge  (<23%);  this 
was  also  the  case  in  the  Oregon  Cascades  (n  = 
30,  Loegering  1997).  Further  inaccessibility 


may  be  afforded  on  high  ledges  associated  with 
near-vertical  cliffs.  Bridges  have  this  obvious 
advantage;  however,  not  all  have  suitable  ledges 
for  nest  placement.  Bridges  constructed  of  used 
railroad  flatcars  provide  excellent  nest  sites  if 
the  ledges  do  not  extend  to  the  abutments,  thus 
allowing  mammalian  access.  Bridges  with  con- 
crete, I-style  beams  provide  suitable  nest  sites, 
but  only  if  the  central  cross  member  between 
the  parallel  supports  is  placed  at  an  acute  angle 
(<60°),  permitting  dippers  to  wedge  their  nest 
against  the  walls  of  the  support  and  cross  mem- 
ber. Interestingly,  dippers  in  Utah  and  Montana 
nested  successfully  by  nesting  under  bridges 
without  cross  members  (R.  E.  Donnelly  and  S. 
A.  H.  Osborn  pers.  comm.).  By  virtue  of  their 
position  over  the  stream,  log  cavities  are  even 
more  protected  from  predators.  Sufficient  di- 
ameter is  needed  for  logs  to  develop  cavities 
large  enough  to  hold  a nest.  We  found  nests  in 
logs  that  were  40-150  cm  in  diameter;  however, 
a 3 1-cm-diameter  branch  overhanging  the 
stream  contained  a nest  cavity  that  was  created 
when  the  majority  of  the  branch  was  ripped  off 
by  a windstorm  or  spring  flood. 

Management  implications. — Breeding  dipper 
populations  in  the  Oregon  Coast  Range  appear 
to  be  limited  by  the  availability  of  suitable  nest- 
ing substrates.  Suitable  dipper  nest  sites,  and 
consequently  recruitment  from  those  sites,  are 
dependent  on  the  physical  characteristics  of  the 
nest-site.  However,  suitable  sites  are  not  abun- 
dant and  are  mostly  products  of  geomorphology 
and  human  development  (i.e.,  bridges).  If  war- 
ranted, effective  options  to  increase  breeding 
abundance  include  providing  nest  boxes,  creat- 
ing ledges  and  cavities,  and  modifying  existing 
structures  (e.g.,  bridges)  to  provide  suitable  nest 
sites.  A long-term,  natural  alternative  for  nest- 
site  recruitment  may  be  the  conservation  of 
large  coniferous  logs  in  riparian  systems.  Tim- 
ber harvest  operations  that  reduce  the  amount 
of  large  wood  along  streams  should  be  avoided, 
and  managers  should  protect  and  encourage  co- 
nifer-dominated riparian  areas.  Large  logs  that 
fall  into  the  stream  channel  and  along  the  stream 
bank  from  riparian  areas  or  the  upslopes  (Swan- 
son et  al.  1976,  Van  Sickle  and  Gregory  1990, 
Fetherston  et  al.  1995)  can  contribute  to  hetero- 
geneity in  the  channel  and  riparian  zone  (Keller 
and  Swanson  1979,  Bilby  1988,  Gregory  et  al. 
1991)  and  potentially  serve  as  nest  or  foraging 
sites  (S.  A.  H.  Osborn  pers.  comm.)  for  dippers. 


Loegering  and  Anthony  • NEST-SITE  SELECTION  OF  AMERICAN  DIPPERS 


293 


Coniferous  logs  also  have  greater  longevity  than 
comparably  sized  red  alder  logs  (Swanson  and 
Lienkaemper  1978),  and  they  have  the  potential 
to  reach  a larger  diameter,  further  increasing 
their  persistence  as  nest  sites  or  structural  com- 
ponents of  riparian  systems.  Current  guidelines 
for  private  and  state  forests  (Oregon  Forest 
Practices  Act  1994)  that  require  maintenance  of 
the  community  structure  and  specific  conifer 
stocking  levels  in  riparian  areas  appear  to  be 
adequate.  Overall,  resources  needed  by  dippers 
should  be  adequately  protected  by  the  guidelines 
for  federal  forests  (Forest  Ecosystem  Manage- 
ment Assessment  Team  1993),  which  limit  dis- 
turbance in  riparian  areas.  Unfortunately,  our 
sampling  of  riparian  habitat  extended  only  25  m 
from  the  stream’s  edge  and  did  not  allow  us  to 
evaluate  differing  buffer  widths  in  riparian 
zones.  Subsequent  research  should  address  this 
concern. 

ACKNOWLEDGMENTS 

The  Coastal  Oregon  Productivity  Enhancement  Pro- 
gram and  the  U.S.  Bureau  of  Land  Management  funded 
this  study.  The  Oregon  Cooperative  Fish  and  Wildlife 
Research  Unit  provided  financial  and  technical  support 
throughout  the  entire  project.  We  are  grateful  to  N.  V. 
Marr,  L.  L.  Loegering,  and  K.  Popper  for  field  assistance. 
Earlier  drafts  of  this  manuscript  were  greatly  improved 
by  comments  from  K.  G.  Beal,  R.  E.  Donnelly,  H.  E. 
Kingery,  S.  A.  H.  Osbom,  and  two  anonymous  reviewers. 
Methods  and  techniques  used  to  capture  and  handle  dip- 
pers were  approved  by  Oregon  State  University’s  Insti- 
tutional Animal  Care  and  Use  Committee. 

LITERATURE  CITED 

Anthony,  R.  G„  E.  C.  Meslow,  and  D.  S.  deCalesta. 
1987.  The  role  of  riparian  zones  for  wildlife  in 
westside  Oregon  forests:  what  we  know  and  don’t 
know.  Pages  5-12  in  Managing  Oregon’s  riparian 
zone  for  timber,  fish,  and  wildlife.  National  Council 
for  Air  and  Stream  Improvement  Technical  Bulle- 
tin, no.  514,  Research  Triangle  Park,  North  Caro- 
lina. 

Bakus,  G.  J.  1959.  Observations  on  the  life  history  of 
the  Dipper  in  Montana.  Auk  76:190-207. 

Balat,  F.  1964.  Breeding  biology  and  population  dy- 
namics in  the  Dipper.  Zoologicke  Listy  13:305- 
320. 

Bart,  J.  R.  and  W.  I.  Notz.  2005.  Analysis  of  data  in 
wildlife  biology.  Pages  72-105  in  Techniques  for 
wildlife  investigations  and  management,  6th  ed.  (C. 
E.  Braun,  Ed.).  The  Wildlife  Society,  Bethesda, 
Maryland. 

Bent,  A.  C.  1948.  Pages  96-113  in  Life  histories  of 
North  American  nuthatches,  wrens,  thrashers,  and 


their  allies.  U.S.  National  Museum  Bulletin,  no. 
195. 

Bilby,  R.  E.  1988.  Interactions  between  aquatic  and  ter- 
restrial systems.  Pages  13-29  in  Streamside  man- 
agement: riparian  wildlife  and  forestry  interactions 
(K.  J.  Raedeke,  Ed.).  Institute  of  Forest  Resources, 
Contribution  no.  59,  University  of  Washington, 
Seattle. 

Ealey,  D.  M.  1977.  Aspects  of  the  ecology  and  be- 
haviour of  a breeding  population  of  dippers  (Cin- 
clus  mexicanus : Passeriformes)  in  southern  Alber- 
ta. M.Sc.  thesis.  University  of  Alberta,  Edmonton. 

Efteland,  S.  and  K.  Kyllingstad.  1984.  Nesting  suc- 
cess in  a southwestern  Norwegian  dipper  Cinclus 
cinclus  population.  Fauna  Norvegica,  Series  C Cin- 
clus 7:7-11. 

Everest,  F.  H.,  N.  B.  Armantrout,  S.  M.  Keller,  W. 
D.  Parante,  J.  R.  Sedell,  T.  E.  Nickelson,  J.  M. 
Johnston,  and  G.  N.  Haugen.  1985.  Salmonids. 
Pages  199-230  in  Management  of  wildlife  and 
fish  habitats  in  forests  of  western  Oregon  and 
Washington  (E.  R.  Brown,  Ed.).  USDA  Forest 
Service,  Pacific  Northwest  Research  Station,  Port- 
land, Oregon. 

Everett,  S.  W.  and  C.  D.  Marti.  1979.  Unusual  dip- 
per nests  found  in  Utah.  North  American  Bird 
Bander  4:58-59. 

Feck,  J.  and  R.  O.  Hall,  Jr.  2004.  Response  of  Amer- 
ican Dippers  ( Cinclus  mexicanus ) to  variation  in 
stream  water  quality.  Freshwater  Biology  49: 
1123-1137. 

Fetherston,  K.  L.,  R.  J.  Naiman,  and  R.  E.  Bilby. 
1995.  Large  woody  debris,  physical  process,  and 
riparian  forest  development  in  montane  river  net- 
works of  the  Pacific  Northwest.  Geomorphology 
13:133-144. 

Fite,  M.  K.  1984.  Vocal  behavior  and  interactions 
among  parents  and  offspring  in  the  American  Dip- 
per. M.Sc.  thesis,  Utah  State  University,  Logan. 

Forest  Ecosystem  Management  Assessment  Team. 
1993.  Forest  ecosystem  management:  an  ecologi- 
cal, economic,  and  social  assessment.  USDA  For- 
est Service,  US  National  Oceanic  and  Atmospher- 
ic Administration,  USDI  Bureau  of  Land  Man- 
agement, USDI  Fish  and  Wildlife  Service,  Na- 
tional Park  Service,  and  Environmental  Protection 
Agency,  Washington,  D.C. 

Franklin,  J.  F.  and  C.  T.  Dyrness.  1973.  Natural  veg- 
etation of  Oregon  and  Washington.  Oregon  State 
University  Press,  Corvallis. 

Gregory,  S.  V.,  F.  J.  Swanson,  W.  A.  McKee,  and  K. 
W.  Cummins.  1991.  An  ecosystem  perspective  of 
riparian  zones.  Bioscience  41:540-551. 

Hann,  H.  W.  1950.  Nesting  behavior  of  American  Dip- 
per in  Colorado.  Condor  52:49-62. 

Hawthorne,  V.  M.  1979.  Use  of  nest  boxes  by  dippers 
on  Sagehen  Creek,  California.  Western  Birds  10: 
215-216. 

Heisey,  D.  M.  and  T.  K.  Fuller.  1985.  Estimation  of 
survival  and  cause-specific  mortality  rates  using  te- 


294 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


lemetry  data.  Journal  of  Wildlife  Management  49: 
668-674. 

Hensler,  G.  L.  1985.  Estimation  and  comparison  of 
functions  of  daily  nest  survival  probabilities  using 
the  Mayfield  method.  Pages  289-301  in  Statistics 
in  ornithology  (B.  J.  T.  Morgan  and  P.  M.  North, 
Eds.).  Springer- Verlag,  New  York. 

Hitchcock,  C.  L.  and  A.  Cronquist.  1973.  Flora  of 
the  Pacific  Northwest.  University  of  Washington 
Press,  Seattle. 

Hosmer,  D.  W.,  Jr.,  and  S.  Lemeshow.  2000.  Applied 
logistic  regression,  2nd  ed.  John  Wiley  and  Sons, 
New  York. 

Johnson,  D.  H.  1980.  Comparison  of  usage  and  avail- 
ability measurements  for  evaluating  resource  pref- 
erence. Ecology  61:65-71. 

Jost,  O.  1970.  Erfolgreiche  Schutzmassnahmen  in  den 
Brutevieren  der  Wassermsel  ( Cinclus  cinclus). 
Angewandte  Ornithologie  3:101-108.  [In  Ger- 
man] 

Keller,  E.  A.  and  F.  J.  Swanson.  1979.  Effects  of 
large  organic  material  on  channel  form  and  fluvial 
processes.  Earth  Surface  Processes  4:361-380. 

Kingery,  H.  E.  1996.  American  Dipper  ( Cinclus  mex- 
icanus).  The  Birds  of  North  America,  no.  229. 

Loegering,  J.  P.  1997.  Abundance,  habitat  association, 
and  foraging  ecology  of  American  Dippers  and 
other  riparian-associated  wildlife  in  the  Oregon 
Coast  Range.  Ph.D.  dissertation,  Oregon  State 
University,  Corvallis. 

Loegering,  J.  P.  and  R.  G.  Anthony.  1999.  Distribu- 
tion, abundance,  and  habitat  association  of  ripar- 
ian-obligate and  -associated  birds  in  the  Oregon 
Coast  Range.  Northwest  Science  73:168-185. 

Manolis,  J.  C.,  D.  E.  Anderson,  and  F.  J.  Cuthbert. 
2000.  Uncertain  nest  fates  in  songbird  studies  and 
variation  in  Mayfield  estimation.  Auk  117:15- 
626. 

Mayfield,  H.  F.  1961.  Nesting  success  calculated  from 
exposure.  Wilson  Bulletin  73:255-261. 

Mayfield,  H.  F.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

Meehan,  W.  R.,  F.  J.  Swanson,  and  J.  R.  Sedell. 
1977.  Influences  of  riparian  vegetation  on  aquatic 
ecosystems  with  particular  reference  to  salmonid 
fishes  and  their  food  supply.  Pages  137-145  in 
Importance,  preservation,  and  management  of  ri- 
parian habitat:  a symposium  (R.  R.  Johnson  and 
D.  A.  Jones,  Tech.  Coords.).  General  Technical 
Report  RM-43,  USD  A Forest  Service,  Rocky 
Mountain  Research  Station,  Fort  Collins,  Colora- 
do. 

Moon,  H.  J.  1923.  Dipper  nesting  away  from  water. 
British  Birds  17:59. 

Morrissey,  C.  A.  2004.  Effect  of  altitudinal  migration 
within  a watershed  on  the  reproductive  success  of 
American  Dippers.  Canadian  Journal  of  Zoology 
82:800-807. 

Neter,  J.,  W.  Wasserman,  and  M.  H.  Knutner.  1989. 


Applied  linear  regression  models,  2nd  ed.  Irwin, 
Inc.,  Homewood,  Illinois. 

Oregon  Forest  Practices  Act.  1994.  Water  protec- 
tion rules,  scientific  and  policy  considerations. 
Oregon  Department  of  Forestry,  Salem. 

Osborn,  S.  A.  H.  1999.  Factors  affecting  the  distri- 
bution and  productivity  of  the  American  Dipper 
{Cinclus  mexicanus ) in  western  Montana:  does 
streamside  development  play  a role?  M.Sc.  thesis, 
University  of  Montana,  Missoula. 

Price,  F.  E.  and  C.  E.  Bock.  1983.  Population  ecology 
of  the  dipper  {Cinclus  mexicanus ) in  the  Front 
Range  of  Colorado.  Studies  in  Avian  Biology, 
no.  7. 

Robson,  R.  W.  1956.  The  breeding  of  the  dipper  in 
north  Westmorland.  Bird  Study  3:170-180. 

SAS  Institute,  Inc.  1989.  SAS/STAT  user’s  guide, 
ver.  6,  4th  ed.  SAS  Institute  Inc.,  Cary,  North  Car- 
olina. 

Sedell,  J.  R.,  P.  A.  Bisson,  F.  J.  Swanson,  and  S.  V. 
Gregory.  1988.  What  we  know  about  large  trees 
that  fall  into  streams  and  rivers.  Pages  47-81  in 
From  the  forest  to  the  sea:  a story  of  fallen  trees 
(C.  Maser,  R.  F.  Tarrant,  J.  M.  Tappe,  and  J.  F. 
Franklin,  Eds.).  General  Technical  Report  PNW- 
229,  USDA  Forest  Service,  Pacific  Northwest  Re- 
search Station,  Portland,  Oregon. 

Shaw,  G.  1978.  Breeding  biology  of  the  dipper.  Bird 
Study  25:149-160. 

Sokal,  R.  R.  and  F.  J.  Rohlf.  1981.  Biometry:  the 
principles  and  practice  of  statistics  in  biological 
research,  2nd  ed.  W.  H.  Freeman  and  Company, 
San  Francisco,  California. 

Stanley,  T.  R.  2004.  When  should  Mayfield  model 
data  be  discarded?  Wilson  Bulletin  116:267-269. 

Strahler,  A.  N.  1957.  Quantitative  analysis  of  water- 
shed geomorphology.  Transactions  of  the  Ameri- 
can Geophysical  Union  38:913-920. 

Sullivan,  J.  O.  1966.  A dipper  nest  away  from  water. 
Condor  68:107. 

Sullivan,  J.  O.  1973.  Ecology  and  behavior  of  the 
dipper,  adaptations  of  a passerine  to  an  aquatic 
environment.  Ph.D.  dissertation.  University  of 
Montana,  Missoula. 

Swanson,  F.  J.  and  G.  W.  Lienkaemper.  1978.  Physi- 
cal consequences  of  large  organic  debris  in  Pacific 
Northwest  streams.  General  Technical  Report 
PNW-69,  USDA  Forest  Service,  Pacific  North- 
west Research  Station,  Portland,  Oregon. 

Swanson,  F.  J..  G.  W.  Lienkaemper,  and  J.  R.  Sedell. 
1976.  History,  physical  effects,  and  management 
implications  of  large  organic  debris  in  western 
Oregon  streams.  General  Technical  Report  PNW- 
56,  USDA  Forest  Service,  Pacific  Northwest  Re- 
search Station,  Portland,  Oregon. 

Trochet,  B.  1967.  Un  nid  de  cincle  sur  un  arbre.  Jean 
le  Blanc:  100-101.  [In  French] 

Van  Sickle,  J.  and  S.  V.  Gregory.  1990.  Modeling 
inputs  of  large  woody  debris  to  streams  from  fall- 
ing trees.  Canadian  Journal  of  Forest  Research  20: 
1593-1601. 


The  Wilson  Journal  of  Ornithology  1 18(3):295— 308,  2006 


UPLAND  BIRD  COMMUNITIES  ON  SANTO,  VANUATU, 
SOUTHWEST  PACIFIC 

ANDREW  W.  KRATTER,1 4 JEREMY  J.  KIRCHMAN,23  AND 
DAVID  W.  STEADMAN1 


ABSTRACT. — We  surveyed  indigenous  landbirds  at  two  upland,  mostly  forested  sites  in  southwestern  Santo, 
Vanuatu.  One  site  (Wunarohaehare,  600-1,250  m elevation)  lies  on  the  western,  rain-shadowed  slope  of  Mt. 
Tabwemasana.  The  other  (Tsaraepae,  500-700  m elevation)  is  16  km  to  the  south,  on  the  southeastern,  very  wet 
slope  of  Peak  Santo.  These  are  the  richest  single-site  bird  communities  yet  surveyed  in  Vanuatu,  with  30  species 
of  resident  birds  recorded  at  each  site,  27  of  which  were  common  to  both  sites,  including  6 species  endemic  to 
Vanuatu.  We  judged  that  12  of  the  shared  species  were  common  at  both  sites.  The  non-overlapping  species  were 
a megapode,  a parrot,  and  four  understory  passerines.  We  present  new  data  on  vocalizations  for  four  species 
endemic  to  Vanuatu  ( Ptilinopus  tannensis,  Todiramphus  farquhari,  Neolalage  banksiana)  or  to  Vanuatu  plus 
New  Caledonia  ( Clytorhynchus  pachycephaloides ).  We  found  less  seasonality  in  breeding  than  previously  re- 
ported for  Vanuatu.  Most  human  impact  at  the  sites  today  may  be  from  non-native  mammals  (rats,  cats,  pigs, 
cows),  along  with  low  levels  of  hunting  and  forest  clearing.  Based  on  prehistoric  bones  from  elsewhere  in 
Vanuatu,  we  suspect  that  formerly  the  sites  on  Santo  may  have  supported  additional  species  of  megapode,  hawk, 
parrot,  and  starling.  Received  28  July  2005,  accepted  14  March  2006. 


The  Republic  of  Vanuatu  (12,195  km2 3 4;  Fig. 
1)  consists  of  12  islands  >270  km2  and  nearly 
100  smaller  ones  in  the  tropical  Pacific  Ocean. 
Approximately  190,000  persons  inhabit  70  is- 
lands (Lai  and  Fortune  2000)  that  range  from 
active  volcanoes  to  limestone  islands  to  older, 
geologically  composite  islands,  such  as  Santo 
(MacFarlane  et  al.  1988,  Nunn  1994).  Anal- 
yses of  avian  distributions  in  Vanuatu,  based 
largely  on  collections  made  during  the  Whit- 
ney South  Sea  Expedition  on  31  islands  in 
1926  and  1927  (e.g.,  Mayr  1934,  1941),  have 
been  important  in  the  development  of  evolu- 
tionary theory  (Mayr  1963)  and  the  fields  of 
island  biogeography  (MacArthur  and  Wilson 
1967)  and  community  ecology  (Diamond 
1975).  Aside  from  the  study  by  Scott  (1946), 
field  ornithology  in  Vanuatu  lagged  until  the 
Percy  Sladen  expedition  of  1971  focused  on 
inter-island  and  altitudinal  patterns  of  avian 
distribution  across  six  islands  in  the  archipel- 
ago (Medway  and  Marshall  1975).  Despite  the 
continued  interest  by  ecologists  in  the  results 


1 Florida  Museum  of  Natural  History,  Univ.  of  Flor- 
ida, P.O.  Box  117800,  Gainesville,  FL  32611,  USA. 

2 Dept,  of  Zoology,  Univ.  of  Florida,  Gainesville,  FL 
32611,  USA. 

3 Current  address:  New  York  State  Museum,  Room 
3023,  Cultural  Education  Center,  Albany,  NY  12230, 
USA. 

4 Corresponding  author;  e-mail: 
kratter@flmnh.ufl.edu 


of  surveys  conducted  decades  ago  (e.g.,  San- 
derson et  al.  1998,  Gotelli  and  Entsminger 
2001),  little  recent  attention  has  been  paid  to 
gathering  new  data  on  intra-  and  inter-island 
variation  in  Vanuatu’s  bird  communities  (al- 
though see  Bowen  1997).  Bregulla  (1992) 
summarized  information  on  identification, 
life-history,  and  distribution  for  each  species 
recorded  from  the  island  group,  yet  made  it 
clear  that  much  remains  to  be  learned  about 
the  basic  biology  of  Vanuatu’s  birds.  Although 
most  biogeographic  analyses  of  insular  faunas 
(or  floras)  are  based  on  lists  of  species  from 
an  entire  island,  such  lists  typically  contain 
species  that  seldom,  if  ever,  interact  because 
they  are  not  syntopic.  Especially  on  large  is- 
lands such  as  Santo,  the  sets  of  species  found 
at  single  sites  provide  fertile  grounds  for  anal- 
ysis. 

In  2002  and  2003,  we  made  two  trips  to 
Santo,  Vanuatu’s  largest  (3,900  km2)  and  high- 
est (1,879  m)  island,  home  to  eight  of  the  nine 
bird  species  endemic  to  the  archipelago  (Bre- 
gulla 1992).  We  surveyed  birds  at  two  mid- 
elevation rainforest  sites,  one  each  on  the 
southeastern  (windward)  and  western  (lee- 
ward) slopes  of  Santo’s  rugged  west-coast 
mountain  range.  Our  surveys  were  based  on 
sight/sound  records,  mist  netting,  tape-record- 
ings, and  specimens  collected:  skins  with 
wings  spread,  skeletons,  tissues,  stomach  con- 
tents, and  ectoparasites  from  the  same  indi- 


295 


296 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


0 5 10  15  20  25  km 


FIG.  1 . Map  of  Espiritu  Santo,  Vanuatu,  with  an  inset  of  Melanesia.  Islands  and  island  groups  mentioned 
in  the  text  are  named.  Sites  of  bird  surveys  conducted  from  2002-2003  by  the  authors  are  indicated  by  the 
triangle  (Wunarohaehare)  and  the  square  (Tsaraepae),  and  filled  circles  indicate  sites  surveyed  by  Bowen  (1997; 
Loru  Protected  Area)  and  Medway  and  Marshall  (1975;  Nokovula,  Apuna  River,  Hog  Harbour).  Asterisks  = 
mountain  peaks  >1,400  m;  dashed  line  = 600-m  contour. 


vidual,  along  with  data  on  habitat,  molt,  diet, 
and  reproductive  condition.  Such  information 
is  a first  step  in  the  investigation  of  ecological, 
morphological,  and  genetic  differences  among 
populations,  and  it  is  important  for  conserva- 
tion efforts  that  often  focus  on  endemic  taxa. 

In  this  paper,  we  present  the  results  of  our 
surveys  at  each  site,  focusing  on  Vanuatu’s 
endemic  and  poorly  known  species.  We  also 
present  comparisons  with  previous  surveys  at 
sites  elsewhere  on  Santo  and  in  the  Solomon 
Islands. 


METHODS 

The  island  of  Espiritu  Santo  (generally 
called  Santo;  Fig.  1)  probably  originated  in 
the  Oligocene  (ca.  25-30  mya)  through  vol- 
canism  and  tectonic  uplift,  although  most  of 
its  land  formed  during  or  since  the  Miocene 
through  these  same  processes  (Mallick  1975, 
Collot  and  Fisher  1989).  Much  of  the  island’s 
eastern  half  is  flat  or  has  rolling  hills,  with 
most  land  <300  m in  elevation  and  very  little 
of  it  above  600  m.  The  western  half  of  Santo 
is  dominated  by  a north-south  trending  moun- 


Knitter  et  al.  • SANTO,  VANUATU  BIRD  COMMUNITIES 


297 


TABLE  1.  Study  sites  and  mist-netting  effort  on  Santo,  Vanuatu,  2002-2003. 


Site  (latitude,  longitude) 

Major  habitats 

Netting  dates 

Elevation  (m) 

No.  nets 

Net-hr 

Wusi  village  (15°  22.7'  S, 
166°  39.7'  E) 

Dry  lowland  forest, 
secondary  scrub 

22-27  Oct  2002, 
4-5  Nov  2002 

0-50 

8 

165 

Wunarohaehare1*  (15°  20.5'  S, 
166°  40.5'  E) 

Humid  premontane  forest, 
forest  patches,  grassy 
ridge 

29  Oct-2  Nov 
2002 

600-1200 

18 

337 

Kerevalissy  village  (15°  35.7'  S, 
166°  50.0'  E) 

Secondary  lowland  forest 
patches 

3-6  and  14  Jun 
2003 

200 

5 

14 

Tsaraepae3  (15°  32.7'  S, 
166°  48.4'  E) 

Wet,  primary,  premontane 
forest 

7-14  Jun  2003 

500-700 

15 

575 

a Primary  study  sites. 


tain  range  that  reaches  its  greatest  height  at 
Mt.  Tabwemasana  (1,879  m).  Prevailing 
winds  push  moist  air  off  the  Pacific  Ocean 
across  the  eastern  lowlands  and  into  the  east- 
or  southeast-facing  slopes  of  the  main  cordi- 
llera. Thus,  the  eastern  and  southern  slopes  of 
the  cordillera  are  humid  with  high  precipita- 
tion, whereas  the  western  slopes,  which 
plunge  into  the  Pacific  with  little  development 
of  a coastal  plain,  lie  in  a rain  shadow  and  are 
relatively  dry. 

From  22  October  to  5 November  2002,  we 
(AWK,  JJK)  mist-netted  and  observed  birds  in 
dry  forest  and  scrub  in  the  vicinity  of  Wusi 
(Fig.  1,  Table  1),  a village  in  the  rain  shadow 
on  the  western  coast  10  km  west  of  Mt.  Tab- 
wemasana, and  in  humid  premontane  forests 
and  grassy  ridges  from  600  to  1,250  m ele- 
vation on  the  northern  slope  of  Mt.  Wunaro- 
haehare  (denoted  by  a triangle  in  Fig.  1 ; Table 
1).  At  Wunarohaehare,  figs  ( Ficus  spp.)  and 
nutmegs  ( Myristica  spp.)  are  the  dominant 
fruiting  trees.  Tree  ferns  ( Cyathea  spp.,  Dick- 
sonia  spp.)  become  common  above  700  m in 
a transitional  habitat  between  the  “high-stat- 
ure  lowland  rain  forest”  and  the  “montane 
cloud  forest”  (described  in  Mueller-Dombois 
and  Fosberg  1998).  The  weather  at  Wunaro- 
haehare is  cool  and  moist  in  the  morning,  as 
cloud  cover  descends  below  600  m.  By  10:00 
UTC  + 11,  however,  the  clouds  dissipate  and 
the  canopy  receives  direct  sunlight.  Short  pe- 
riods (<1  hr)  of  rain  occur  most  afternoons. 

From  3 to  14  June  2003,  AWK,  JJK,  and 
DWS  worked  on  the  southern  slopes  of  Peak 
Santo  (also  called  Lairiri;  1,704  m),  —16  km 
south-southeast  of  Mt.  Tabwemasana.  This 
area  received  the  full  precipitative  effects  of 
moist  air  coming  off  the  Pacific,  and  was 


much  wetter  than  sites  in  the  rain  shadow — 
Wusi  and  Wunarohaehare.  From  3 to  7 June, 
we  surveyed  a patchy  secondary  forest  near 
Kerevalissy  village  (Fig.  1,  Table  1),  a land- 
scape dominated  by  coconut  plantations,  —4 
km  north  of  the  coastal  village  of  Ipayato. 
From  7 to  14  June,  we  mist-netted  (Table  1) 
and  observed  birds  on  the  southern  slopes  of 
Peak  Santo  at  Tsaraepae  (—500  m;  denoted  by 
the  square  in  Fig.  1)  and  on  nearby  slopes  up 
to  700  m elevation.  Ridges  in  the  lower  ele- 
vations had  a broken  canopy  and  were  cleared 
of  undergrowth,  grazed  by  feral  cattle  ( Bos 
taurus ),  and  browsed  by  feral  pigs  ( Sus  scro- 
fa ).  The  area  >700  m was  mainly  tall  (canopy 
12-25  m)  forest. 

Trees  identified  (by  DWS)  to  genus  includ- 
ed Garuga  (Burseraceae),  Calophyllum  (Clu- 
siaceae),  Elaeocarpus  (Elaeocarpaceae),  Her- 
nandia  (Hemandiaceae),  Ficus  (Moraceae;  at 
least  five  species,  some  of  them  emergent), 
and  Myristica  (Myristicaceae);  those  we  iden- 
tified to  species  included  Barringtonia  edulis 
(Barringtoniaceae)  and  Endospermum  medul- 
losum  (Euphorbiaceae;  often  emergent).  There 
also  were  a number  of  unknown  species,  in- 
cluding various  Myrtaceae  and  Rubiaceae. 
Also  present  were  Pandanus  spp.  (Pandana- 
ceae),  tree-ferns  ( Cyathea  spp.;  Cyatheaceae), 
and  Dicksonia  spp.  (Dicksoniaceae).  The  edg- 
es included  trees  and  shrubs  of  Macaranga 
spp.  (Euphorbiaceae),  lnocarpus  fagifer  (Fa- 
baceae),  Ficus  spp.,  Piper  spp.  (Piperaceae), 
Alphitonia  spp.  (Rhamnaceae),  Pipturus  spp. 
(Urticaceae),  palms  ( Cocos  spp.;  Metroxylon 
spp.  [Arecaceae]),  and  thickets  of  Hibiscus  til- 
iaceus  (Malvaceae),  bananas  (Musaceae),  and 
gingers  (Zingiberaceae). 

The  weather  at  Kerevalissy  and  Tsaraepae 


298 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


in  2003  was  extremely  wet,  with  heavy  rain- 
fall occurring  on  1 1 of  our  12  days.  On  9 days 
we  estimated  that  the  daily  rainfall  exceeded 
100  mm,  including  6 days  (5,  6,  8,  9,  12,  and 
13  June)  on  which  it  probably  exceeded  150 
mm.  The  excessive  rain  was  due  to  an  unusu- 
ally late  tropical  storm  that  paused  just  north 
of  Santo  over  the  Banks  and  Torres  islands. 
Because  avian  activity  did  not  diminish  no- 
ticeably during  rains  at  Tsaraepae,  we  con- 
ducted our  sight/sound  surveys  and  set  mist 
nets  even  during  the  very  rainy  weather.  Vo- 
calizations were  tape-recorded  on  several  days 
at  each  of  our  two  primary  sites  (Table  1),  and 
the  original  tapes  were  deposited  in  the  Flor- 
ida Museum  of  Natural  History  (UF)  Sound 
Archives.  Birds  were  collected  according  to 
the  stipulations  of  our  permits  from  the  Va- 
nuatu Ministry  of  Lands,  Environment  Unit. 
Specimens  were  prepared  as  various  combi- 
nations of  round  skins,  complete  or  partial 
skeletons,  and  with  spread  wings.  Stomach 
contents  and  two  tissue  samples  were  taken 
from  each  specimen;  one  tissue  sample  is 
housed  at  UF  and  the  other  at  the  Louisiana 
State  University  Museum  of  Natural  Science. 
All  non-tissue  material  is  housed  at  UF.  As  far 
as  we  know,  neither  tissue  nor  skeletal  spec- 
imens of  birds  had  been  collected  previously 
in  Vanuatu.  The  skeletal  specimens  of  the  Va- 
nuatu endemics  Ducula  bakeri , Ptilinopus 
tannensis,  Todiramphus  farquhari , Neolalage 
banks ianci,  Zosterops  flavifrons,  and  Glycifo- 
hia  notabilis  (see  Tables  2 and  3 for  English 
common  names)  are  the  first  in  the  world’s 
inventories. 

In  addition  to  our  work  at  the  two  primary 
sites,  JJK  and  AWK  collected  and  surveyed 
birds  in  patchy  forested  sites  near  sea  level  on 
the  eastern  coast  of  Santo  for  2 days  in  Oc- 
tober—November  2002  and  for  4 days  in  June 
2003.  In  northern  Santo,  AWK  visited  lowland 
forests  of  the  Vatte  Conservation  Area  (near 
Matantas;  Fig.  1)  from  17  to  19  November 
2002.  DWS  visited  Aore  Island  (Fig.  1)  on 
15-16  June  2003,  surveying  (sight/sound 
only)  birds  in  patches  of  tall  (canopy  15—30 
m)  lowland  rainforest. 

Although  this  was  the  first  visit  to  Santo  by 
all  three  authors.  AWK  and  especially  DWS 
have  wide  experience  with  the  avifauna  in 
western  Oceania.  They  know  the  vocalizations 
and  behaviors  of  all  but  one  of  the  genera 


found  on  Santo.  Nonetheless,  cryptic  species 
may  have  been  missed  if  they  were  not  vocal 
during  our  visits. 

RESULTS 

Diversity  and  community  composition. — We 
recorded  33  indigenous  species  of  landbirds  at 
Wunarohaehare  and  Tsaraepae,  with  27  spe- 
cies common  to  both  sites  (Table  2).  As  is  the 
case  across  most  of  Oceania  (Steadman  1997, 
2006b),  pigeons  and  doves  (Columbidae) 
composed  a large  part  of  the  avifauna;  the 
same  seven  species  of  columbids  were  found 
at  each  site.  We  also  recorded  seven  of  the 
eight  species  endemic  to  Vanuatu,  failing  to 
record  only  Aplonis  santovestris  (see  below). 
Six  of  the  endemic  species  (all  but  Megapo- 
dius  layardi ) were  recorded  at  both  sites. 

Although  three  species  of  non-native  birds 
are  widespread  on  Santo  (Red  Junglefowl, 
Gallus  gallus;  Common  Myna,  Acridotheres 
tristis;  and  Black-headed  Munia,  Lonchura 
malacca),  the  only  one  we  recorded  was  G. 
gallus , and  it  was  uncommon  (<5/day)  at  both 
sites.  All  three  species  were  common  in  plan- 
tations and  villages  at  elevations  lower  than 
those  of  Wunarohaehare  and  Tsaraepae.  Con- 
tamination of  the  bird  communities  by  non- 
native species  on  Santo  is  minor  (by  Pacific 
Island  standards);  however,  both  sites  are 
heavily  infested  with  non-native  mammals.  At 
Tsaraepae,  we  noted  feral  cats  ( Felis  catus ), 
pigs  ( Sus  scrofa),  and  cows  ( Bos  taurus );  dogs 
( Canis  familiaris ) seemed  to  be  confined  to 
villages.  Inside  our  leaf  house  at  Tsaraepae, 
DWS  snap-trapped  10  rats  (7  Rattus  rattus,  3 
R.  exulans ) in  3 nights,  using  only  two  traps. 

Although  species  richness  was  the  same  at 
our  two  primary  sites,  composition  of  the 
landbird  communities  differed  slightly.  Me- 
gapodius  layardi , Charmosyna  palmarum , 
and  Clytorhynchus  pachycephaloides  were 
found  only  at  Tsaraepae,  although  the  latter 
species  was  found  in  the  dry  forests  near  Wusi 
(lower  elevations  than  at  Wunarohaehare). 
The  mound-building  Megapodius  layardi  may 
be  absent  from  dry  forests  due  to  unsuitable 
soil  conditions.  Our  failure  to  record  Char- 
mosyna palmarum  at  Wunarohaehare  may 
have  been  a consequence  of  its  nomadic  habits 
(see  C.  palmarum  species  account,  below). 
Three  species  with  widespread  distributions  in 
Oceania — Lalage  leucopyga , Turdus  polioce- 


Kratter  et  al.  • SANTO,  VANUATU  BIRD  COMMUNITIES 


299 


phalus,  and  Petroica  multicolor — were  not  re- 
corded at  Tsaraepae.  The  four  passerine  spe- 
cies found  at  only  one  of  the  two  sites  have 
been  recorded  on  both  sides  of  the  cordillera 
(Medway  and  Marshall  1975;  Table  3),  so 
their  apparent  absence  at  one  site  may  be  re- 
lated to  inadequate  sampling.  We  note,  how- 
ever, that  our  guides  at  Tsaraepae  did  not  rec- 
ognize the  illustration  in  Bregulla  (1992)  of 
Turdus  poliocephalus,  suggesting  that  the  lo- 
cal absence  of  this  conspicuous  species  was 
genuine.  The  guides  did  not  distinguish  be- 
tween L.  maculosa  and  L.  leucopyga  (Hakei 
language  names  for  Lalage  were  “vasoimoto” 
and  “losoloso,”  which  seemed  to  apply  to  ei- 
ther species),  so  it  is  possible  that  the  latter 
species  was  present.  Our  guides  did  know  Pe- 
troica multicolor , however,  and  called  it  “pa- 
nopano.” 

We  observed  inter-site  differences  in  the  al- 
titudinal ranges  of  some  species.  Two  endemic 
species  characteristic  of  the  highlands  ( Ducula 
bakeri  and  Glycifohia  notabilis)  were  more 
common  at  Tsaraepae  than  at  Wunarohaehare, 
where  D.  bakeri  was  not  seen  below  800  m. 
At  Tsaraepae,  D.  bakeri  was  found  regularly 
as  low  as  500  m and  locally  in  forest  patches 
as  low  as  200  m along  the  trail  south  toward 
the  coast.  At  Tsaraepae,  the  fantail,  Rhipidura 
spilodera,  was  scarce  above  500  m,  but  at 
Wunarohaehare  it  was  common  up  to  800  m. 
Some  species  associated  with  less  forested 
habitats  ( Todiramphus  chloris,  Lalage  macu- 
losa, Gerygone  flavolateralis ) were  found  at 
higher  elevations  at  Wunarohaehare,  where  we 
sampled  open  habitats  up  to  1,000+  m;  at  Tsa- 
raepae, however,  we  did  not  find  these  species 
at  elevations  above  550  m,  which  were  almost 
entirely  forested. 

Seasonality  of  reproduction. — Our  visit  to 
Wunarohaehare  during  October-November 
coincided  with  the  reported  breeding  period 
for  most  species  of  birds  in  Vanuatu,  which 
generally  is  September-February  (Bregulla 
1992).  Our  visit  to  Tsaraepae  took  place  dur- 
ing June,  a month  when  Bregulla  (1992) 
found  breeding  activity  for  only  5 of  the  33 
species  we  recorded  (Table  2).  We  found  less 
evidence  of  marked  seasonality  in  breeding, 
with  signs  of  reproductive  activity  (enlarged 
gonads  in  specimens,  active  nests,  or  recently 
fledged  juveniles)  in  20  of  23  species  at  Wun- 
arohaehare and  12  of  20  species  at  Tsaraepae 


(Table  4).  We  suspect,  nevertheless,  that  the 
difference  between  the  two  sites  (87%  versus 
60%  of  species)  does  reflect  seasonal  trends 
more  than  inter-site  variation. 

Selected  Species  Accounts 

We  present  our  findings  for  species  endemic 
to  Vanuatu  and  for  some  others  that  are  poorly 
known  in  Vanuatu  or  throughout  their  range. 

Megapodius  layardi. — The  endemic  Vanu- 
atu Megapode  was  not  recorded  at  Wunaro- 
haehare, but,  at  Tsaraepae  on  1 1 and  12  June, 
three  individuals  were  heard  calling  at  an  el- 
evation of  550  m in  the  thick  undergrowth 
near  an  active  incubation  mound  in  a large 
tract  of  forest.  This  was  the  only  mound  near 
Tsaraepae  known  to  our  guides.  Another  bird 
was  observed  in  a dense  Hibiscus  tiliaceus 
thicket  at  600  m on  1 1 June.  Single  birds  also 
were  seen  twice  in  secondary  forest  patches 
on  Santo’s  eastern  coast,  and  once  near  Ma- 
tantas.  Villagers  showed  us  eggs  from  an  ac- 
tive mound  near  Matevulu  on  16  June. 

Chalcophaps  indica. — This  terrestrial  dove 
is  widespread  in  Oceania,  with  the  subspecies 
C.  i.  sandwichensis  confined  to  New  Caledon- 
ia, the  Santa  Cruz  Group,  and  Vanuatu.  Abun- 
dant in  disturbed  forest  and  forest  edge  from 
sea  level  to  400  m elevation  (lower  than  either 
study  site),  the  Emerald  Dove  was  much  less 
common  in  more  mature  forest  near  our  two 
primary  study  sites.  In  337  net-hr  at  Wuna- 
rohaehare, only  one  bird  was  netted  at  eleva- 
tions >500  m,  whereas  five  were  netted  in  165 
net-hr  at  0-50  m near  Wusi.  Because  it  sel- 
dom vocalizes  and  is  rather  furtive,  mist-net- 
ting  may  yield  better  evidence  of  the  Emerald 
Dove’s  population  density  than  auditory  or  vi- 
sual data.  The  species  is  common  in  village 
gardens,  where  it  often  is  lured  with  papaya 
{Carica  papaya)  into  traps;  stomachs  of  near- 
ly all  collected  individuals  contained  seeds  of 
this  non-native  plant.  The  four  birds  taken 
near  Wusi  village  included  two  males  with  en- 
larged testes,  an  adult  male  (no  bursa)  with 
unenlarged  testes,  and  an  adult  female  (no 
bursa,  convoluted  oviduct)  with  slightly  en- 
larged ova.  The  single  bird  from  Tsaraepae 
was  a male  with  enlarged  testes. 

Ptilinopus  tannensis. — Endemic  to  Vanua- 
tu, the  Tanna  Fruit  Dove  was  common  (up  to 
15  per  day)  at  each  site,  especially  in  montane 
forests.  This  fruit  dove  was  heard  much  more 


300 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  2.  Summary  of  native  bird  communities  at  two  sites  (Wunarohaehare,  600-1,200  m;  Tsaraepae, 
500-700  m)  on  Santo,  Vanuatu,  surveyed  in  2002-2003.  E = endemic  to  Vanuatu,  e = endemic  to  Vanuatu 
plus  New  Caledonia  and/or  the  Santa  Cruz  Group.  Relative  abundance:  c = common  (encountered  regularly  by 
all  observers),  u = uncommon  (encountered  daily  or  almost  daily  in  small  numbers),  r = rare  (encountered 
fewer  than  five  times),  — = not  recorded.  Foraging  guild  (microhabitat/prey):  A = aerial,  C = canopy,  T = 
terrestrial,  U = understory,  F = fruit,  G = granivore  (seeds),  I = insects  and  other  invertebrates,  N = nectar, 
V = vertebrates.  Avian  nomenclature  follows  Dickinson  (2003),  except  that  we  do  not  recognize  Aerodramus, 
which  has  been  used  for  some  species  in  Collocalia  (but  see  Price  et  al.  2004). 


Species 

Megapodiidae 

Megapodius  layardi,  Vanuatu  Megapode  (E) 

Accipitridae 

Circus  approximans.  Swamp  Harrier 
Columbidae 

Columba  vitiensis  leopoldi , White-throated  Pigeon 
Macropygia  m.  mackinlayi,  Mackinlay’s  Cuckoo-Dove 
Chalcophaps  indica  sandwichensis,  Emerald  Dove 
Ptilinopus  tannensis,  Tanna  Fruit  Dove  (E) 

Ptilinopus  greyii.  Red-bellied  Fruit  Dove  (e) 

Ducula  p.  pacifica.  Pacific  Imperial  Pigeon 
Ducula  bakeri,  Vanuatu  Imperial  Pigeon  (E) 

Psittacidae 

Trichoglossus  haematodus  massena.  Rainbow  Lorikeet 
Charmosyna  palmarum , Palm  Lorikeet  (e) 

Cuculidae 

Chrysococcyx  lucidus  layardi.  Shining  Bronze-Cuckoo 
Apodidae 

Collocalia  esculenta  uropygialis,  Glossy  Swiftlet 
Collocalia  v.  vanikorensis,  Uniform  Swiftlet 

Alcedinidae 

Todiramphus  farquhari.  Chestnut-bellied  Kingfisher  (E) 
Todiramphus  chloris  santoensis,  Collared  Kingfisher 

Meliphagidae 

Glycifohia  n.  notabilis.  White-bellied  Honeyeater  (E) 

Myzomela  cardinalis  tenuis.  Cardinal  Honeyeater 

Acanthizidae 

Gerygone  flavolateralis  correiae.  Fan-tailed  Gerygone 
Artamidae 

Artamus  leucorhynchus  tenuis.  White-breasted  Woodswallow 
Campephagidae 

Coracina  caledonica  thilenii,  Melanesian  Cuckoo-shrike 
Lalage  maculosa  modesta,  Polynesian  Triller 
Lalage  leucopyga  albiloris.  Long-tailed  Triller 

Pachycephalidae 

Pachycephala  [pectoralis]  caledonica  intacta.  New  Caledonian 
Whistler  (e) 

Petroicidae 

Petroica  multicolor  ambrynensis.  Pacific  Robin 
Rhipiduridae 

Rhipidura  [ fuliginosa]  albiscapa  brenchleyi.  Gray  Fantail 
Rhipidura  s.  spilodera.  Streaked  Fantail 


Relative  abundance 

Foraging 

Wunarohaehare  Tsaraepae  guild 


T/F,G,I 


A/V 


T, U,C/F,G 
U/F 
T/G,I,F 
C/F 

U, C/F 
C/F 
C/F 


C/N,F 

C/N 


C/I? 


A/I 

A/I 


U/I,V 

C/I,V 


C/N,I 
C/N, I 


U,C/I 

A/I 


U,C/F,I 

U,C/F,I 

U/F,I 


U/I 


— U,C/F,I 


U/I 

T,U/I 


Kratter  et  al.  • SANTO,  VANUATU  BIRD  COMMUNITIES 


301 


TABLE  2.  Continued. 

Relative  abundance 

Foraging 

guild 

Species 

Wunarohaehare 

Tsaraepae 

Monarchidae 

Neolalage  banksiana,  Buff-bellied  Monarch  (E) 

C 

C 

U/I 

Clytorhynchus  pachycephaloides  grisescens , Southern  Shrikebill  (e) 

— 

U 

U/I 

Myiagra  caledonica  marinae,  Melanesian  Flycatcher 

C 

C 

U,C/I 

Zosteropidae 

Zosterops  flavifrons  brevicauda,  Yellow-fronted  White-eye  (E) 

C 

C 

U,C/N,F,I 

Zosterops  lateralis  tropicus.  Silver-eye 

c 

C 

U,C/N,F,I 

Turdidae 

Turdus  poliocephalus  vanikorensis.  Island  Thrush 

c 

— 

T,U/F,I 

often  than  seen,  although  it  called  less  fre- 
quently than  the  Red-bellied  Fruit  Dove.  Con- 
trary to  Medway  and  Marshall  (1975)  and 
Bowen  (1997),  we  found  the  Tanna  Fruit 
Dove  above  500  m;  it  remained  common  up 
to  the  highest  continuous  forests  that  we 
reached  at  both  Wunarohaehare  (800  m)  and 
Tsaraepae  (700  m).  The  most  common  call 
was  a series  (~10+)  of  low,  upwardly  inflect- 
ing woot  notes,  spaced  up  to  2 sec  apart.  In- 
frequently, it  also  gave  a soft,  single  woot 
note. 

We  found  the  Tanna  Fruit  Dove  breeding  at 
both  sites.  Bregulla  (1992)  reported  its  nesting 
status  as  poorly  known,  with  previous  evi- 
dence reported  only  in  April  and  May,  a time 
of  little  breeding  activity  among  other  land- 
birds  in  Vanuatu.  At  Wunarohaehare,  a nearly 
fledged  nestling  was  found  on  the  ground  after 
a windy  evening,  and  two  males  had  enlarged 
testes  and  a female  had  enlarged  ova.  At  Tsa- 
raepae, the  one  bird  collected  was  a female 
with  enlarged  ova. 

Ptilinopus  greyii. — The  monotypic  Red- 
bellied  Fruit  Dove  is  confined  to  New  Cale- 
donia, the  Loyalty  Islands,  and  Vanuatu.  The 
species  was  abundant  (<50/day)  at  both  sites 
in  heavily  disturbed  to  mature  forests  and  at 
all  elevations.  It  vocalized  throughout  the  day. 
All  specimens  showed  evidence  of  breeding: 
at  Wunarohaehare,  these  included  a female 
with  a ruptured  follicle,  another  with  enlarged 
ova,  a male  with  enlarged  testes,  and  a re- 
cently fledged  juvenile;  at  Tsaraepae,  the  spec- 
imens included  two  males  with  enlarged  tes- 
tes, a female  with  enlarged  ova,  and  two  ju- 
veniles. 


Ducula  bakeri. — The  monotypic  Vanuatu 
Imperial  Pigeon  is  endemic  to  seven  islands 
in  northern  Vanuatu.  Although  rare  or  absent 
in  the  lowlands  of  Santo,  it  was  common  at 
Tsaraepae,  where  two  or  three  calling  individ- 
uals often  were  audible  from  many  points  on 
a forested  ridge  at  —600  m,  and  we  recorded 
as  many  as  20  on  single  days.  It  was  less  com- 
mon on  the  disturbed  slopes  below  500  m, 
although  we  heard  it  in  a forest  patch  adjacent 
to  Kerevalissy  on  14  June.  At  Wunarohaehare, 
we  found  the  Vanutau  Imperial  Pigeon  only  at 
elevations  >800  m,  where  up  to  three  indi- 
viduals called  in  heavy  forest  cover  on  most 
days.  The  birds  taken  at  Tsaraepae  were  an 
adult  female  with  enlarged  ova  and  a juvenile 
male.  They  differed  little  in  plumage,  and  both 
had  Myristica  spp.  fruits  in  their  crops  and 
stomachs. 

Charmosyna  palmarum. — The  monotypic 
Palm  Lorikeet  is  endemic  to  Vanuatu  and  the 
Santa  Cruz  Group.  We  recorded  this  species 
only  twice  (a  flock  of  six  on  8 June,  a group 
of  two  on  1 1 June),  both  times  in  a Ficus  spp. 
tree  with  large,  fleshy  fruits,  in  humid  forest 
at  650  m on  the  main  ridge  at  Tsaraepae.  Al- 
though more  characteristic  of  montane  than 
lowland  habitats,  the  Palm  Lorikeet  seems  to 
undergo  population  fluctuations  and  has  a pro- 
pensity to  wander  (Medway  and  Marshall 
1975,  Bregulla  1992).  Its  preferred  foods 
(flowers  and  fruits)  may  have  been  scarce  at 
the  time  of  our  visits. 

Collocalia  esculenta  uropygialis  and  C.  v. 
vanikorensis. — Each  of  these  widespread 
swiftlets  was  common  at  Tsaraepae.  The 
Glossy  Swiftlet  (C.  esculenta  uropygialis ; 20- 


302 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  3.  Indigenous  birds  recorded  (+  = present,  - = not  recorded)  at  six  sites  on  Santo,  Vanuatu,  2002- 
2003.  English  common  names  are  provided  for  the  species  not  included  in  Table  2.  E = endemic  to  Vanuatu,  e 
= endemic  to  Vanuatu  plus  New  Caledonia  and/or  the  Santa  Cruz  Group.  Sources  are  Bowen  (1997)  for  Loru 
Protected  Area;  Medway  and  Marshall  (1975)  for  Apuna  River,  Hog  Harbor,  and  Nokovula;  and  our  own  data 
for  Wunarohaehare  and  Tsaraepae.  For  each  site,  the  elevation  (m)  is  included. 


Species 

Loru  Protected 
Area  0-120  m 

Apuna  River 
100  m 

Hog  Harbor 
160  m 

Wunarohaehare 
600-1,250  m 

Tsaraepae 
500-700  m 

Nokovula 
1,120  m 

Megapodius  layardi  (E) 

+ 

+ 

- 

- 

+ 

- 

Falco  peregrinus.  Peregrine  Falcon 

+ 

- 

- 

- 

+ 

Circus  approximans 

+ 

+ 

- 

+ 

+ 

+ 

Gallir alius  philippensis , Banded  Rail 

+ 

- 

- 

- 

- 

- 

Columba  vitiensis 

+ 

+ 

- 

+ 

+ 

- 

Macropygia  mackinlayi 

+ 

+ 

+ 

+ 

+ 

+ 

Chalcophaps  indica 

+ 

+ 

+ 

+ 

+ 

+ 

Ptilinopus  tannensis  (E) 

+ 

+ 

- 

+ 

+ 

— 

Ptilinopus  greyii  (e) 

+ 

+ 

+ 

+ 

+ 

+ 

Ducula  pacifica 

+ 

+ 

+ 

+ 

+ 

- 

Ducula  bakeri  (E) 

- 

- 

- 

+ 

+ 

+ 

Trichoglossus  haematodus 

+ 

+ 

+ 

+ 

+ 

- 

Charmosyna  palmarum  (e) 

- 

- 

- 

- 

+ 

+ 

Chrysococcyx  lucidus 

- 

- 

- 

+ 

+ 

- 

Tyto  alba.  Barn  Owl 

+ 

- 

- 

- 

- 

- 

Collocalia  esculenta  uropygialis 

+ 

+ 

- 

+ 

+ 

+ 

Collocalia  v.  vanikorensis 

- 

+ 

- 

+ 

+ 

- 

Todiramphus  farquhari  (E) 

+ 

+ 

+ 

+ 

+ 

- 

Todiramphus  chloris 

+ 

- 

- 

+ 

+ 

- 

Glycifohia  n.  notabilis  (E) 

- 

- 

- 

+ 

+ 

+ 

Myzomela  cardinalis 

- 

+ 

+ 

+ 

+ 

+ 

Gerygone  flavolateralis 

- 

+ 

+ 

+ 

+ 

+ 

Artamus  leucorhynchus 

+ 

— 

— 

+ 

+ 

+ 

Coracina  caledonica 

+ 

+ 

+ 

+ 

+ 

+ 

Lalage  maculosa 

- 

- 

- 

+ 

+ 

- 

Lalage  leucopyga 

- 

- 

- 

+ 

- 

+ 

Pachycephala  [pectoralis ] caledonica  (e)  + 

+ 

+ 

+ 

+ 

+ 

Petroica  multicolor  ambrynensis 

- 

- 

- 

+ 

- 

+ 

Rhipidura  [fuliginosa]  albiscapa 

+ 

- 

- 

+ 

+ 

- 

Rhipidura  spilodera 

+ 

+ 

+ 

+ 

+ 

+ 

Neolalage  banksiana  (E) 

Clytorhynchus  pachycephaloides  grise- 

+ 

+ 

+ 

+ 

+ 

+ 

scens  (e) 

+ 

+ 

+ 

- 

+ 

— 

Myiagra  caledonica 

Cichlornis  whitneyi,  Melanesian  Thick- 

+ 

+ 

+ 

+ 

+ 

etbird 

- 

- 

- 

- 

- 

+ 

Zosterops  flavifrons  (E) 

+ 

+ 

+ 

+ 

+ 

+ 

Zosterops  lateralis 

Alponis  zelandica.  Rufous-winged  Star- 

+ 

+ 

+ 

+ 

ling  (e) 

- 

- 

- 

- 

— 

+ 

Turdus  poliocephalus  vanikorensis 
Erythrura  cyaneovirens.  Red-headed 

+ 

+ 

+ 

+ 

Parrotfinch 

- 

- 

- 

- 

- 

+ 

Total  species 

25 

22 

16 

30 

30 

24 

Total  endemic  species  (E  + e) 

8 

8 

6 

8 

11 

8 

50/day)  generally  flew  much  closer  to  the 
ground  than  the  Uniform  Swiftlet  (C.  v.  van- 
ikorensis;  <20/day,  except  for  loose  flocks  of 
—400  that  passed  over  on  several  mornings  at 
Tsarapae,  all  flying  west).  Both  species  were 


noted  at  all  sites  visited  on  Santo.  Despite  our 
careful  observations  of  all  swiftlets  detected 
on  Santo,  we  did  not  record  the  White-rumped 
Swiflet  ( Collocalia  spodiopygia),  which  was 
unknown  to  our  guides. 


TABLE  4.  Avian  specimen  data  from  Santo,  Vanuato,  October-November  2002  and  June  2003.  Specimens  collected  at  low  elevations  around  Wusi  and 
Kerevalissy  villages  and  montane  study  sites  are  included.  E = endemic  to  Vanuatu,  e = endemic  to  Vanuatu  plus  New  Caledona  and/or  the  Santa  Cruz  Group. 
See  Tables  2 and  3 for  English  common  names.  Juvenile  status  determined  by  presence  of  bursa  of  Fabricius,  degree  of  skull  ossification,  condition  of  reproductive 
tract,  and  plumage.  Breeding  evidence  (+  or  — ) determined  on  the  basis  of  condition  of  reproductive  tract,  active  nests,  or  recently  fledged  juveniles;  NI  = no 
information. 


Kratter  et  al.  • SANTO,  VANUATU  BIRD  COMMUNITIES 


g-a 
* s 


+ i + + + i 


I Z + 1 Z 1 + z z 


y+  + b|+  i z 


+ + + + + ZZZ  1 + + 1 + + 1 + + + + + + + + + + + 


<N 

r~- 

<n 

<N  O 
cm  r- 


I I 


On 

0 

r- 

o 

-cf 

NO 

. 00 

on 

1 1 1 

1 1 1 1 NO 

1 1 

1 ^ 

1 1 ~ 

| 

I °)  rt 

r*« 

1 1 1 

1 1 1 1 <N 

1 1 

1 CM 

1 1 — ' 

\ On  — 

CO 

CM 

CO 

cd 

00 

d 

CM 

o 

>n 

-t 

1 1 

cm"  0 

cn 

1 ^ 1 

1 1 

1 ^ 

1 1 

00  | 

| d in 

00 

Tj- 

t NO  1 

1 CM 

1 r-  1 

M O'  M 
m . — 

CM  CM 
. 00 


-8 


CM  On 
„ NO 

3 - 
o >n 

(N  NO 


IN 

I I 2 


CM  ^ 
O O 
co 


o 

no 

On  CO 
iri  On 
„ CO 
CM  . 

oo 
it  co 


NO 
CM  O 
00  — 


cm  . in  t oo 

O NO  — 1 On 

cm  oo  in  co  — 1 


■2  S3 

§ 2 

.8:1 


•3 

5a 

S 

3 

.ft 

I 

^3 

Q W 


32  ft 
ft  £ 


S.a  Si 

S oa  <0 

3 3 

ft  Cl  S3. 

g O'  5 

»2  S3  S3 
3S  3 3 
3 3 3 

U a,  a. 


<3 

I * 

ft  ^ 
. a 
ft  -a 

-S  -S 
s s 

ft  ft 
3 3 

Q Q 


't 

O) 

q 

(N 

q 

CM 

NO 

• 

cd 

1—I 

rn 

1—! 

cd 

Tt 

CM 

CM 

1— H 

t-H 

in 

ON 

„ 

„ 

co" 

„ 

00 

CM 

NO 

in 

0 

T-H 

O 

in 

0 

cd 

't 

1 ^ 

cd 

in 

d d 

in 

On 

(N 

d 

CM 

NO 

CM 

00 

1 CM 

CM 

06 

r-  —i 

CM 

— 1 

't 

■'t 

00  T+ 


O ~ On 
oo  6 t 


o on  o no  on 
Tt  ft  oo  -t  d 


on  co  co  in  co  o 


inh- 
°°.  ft  NO  ON  — 

tj-  co  m cm 


cj  2 £ £ ^ 2 

_ On  00 

oo  co  . .t  r)  o oo  (N 

(NMOOOO^-i(N^- 


co  n 
cd  cm 


ft  <3 
ft  ft 


o w 


3-  .ft 

to  to 

3 3 
-3  3 
ft  ft 


<3  « 


ft  *5 

.2  <3 
32  Ci 

S s 
•S  o 
** 
ft  c 
_i  3 ~3 

c ^ s 
J -9  « 

& § ■§ 
3 ft  C3 

£ 


w 


NJ  ri 

ft  s 


a §1.1 

PII 

^ 2S  <3  ^ 
„ ft  ft  e>o 

I ill 

§ .ft.  .ft.  "S 

ft  3 3 ^ 

i.  ft;  ft  ^ 


s 

J -S  a 

8 "N 


ft 

s 

b S3 


ft 

to  3 

3 3 

•Si  § 

ft.  > 

b ,3 

•2  "ft 

3 3 

5 ft- 

ft  £ 

ft  -2 


6? 

t | 

00  ^ 

ft  ^ ^ ^3 
“ ■ -60  C 

IN 


^ 2 
ft.  Ci- 
2 S2 
c 


G S't3 


303 


304 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Todiramphus  farquhari. — Endemic  to  San- 
to, Malo,  and  Malakula,  the  Chestnut-bellied 
Kingfisher  was  slightly  more  common  in  the 
wet  forests  near  Tsaraepae  (<5/day)  than  in 
the  dry  forests  of  the  western  slope,  although 
we  recorded  up  to  six  daily  at  Wunarohaehare. 
It  was  most  common  in  high-canopy  forests, 
but  also  persisted  in  forest  patches,  even  near 
Kerevalissy  village.  It  ranged  from  the  low- 
lands up  to  at  least  800  m,  overlapping  the 
entire  elevational  range  of  its  larger  congener, 
the  Collared  Kingfisher  (T.  chloris ),  which 
prefers  more  open  habitat.  The  Chestnut-bel- 
lied Kingfisher  was  very  vocal  at  both  sites, 
often  singing  throughout  the  day.  The  call  is 
a series  of  ascending  notes  with  decreasing 
intervals,  not  the  “monotonous  single  note” 
described  by  Bowen  (1997).  The  two  birds 
collected  at  Wunarohaehare,  both  at  600  m, 
were  adult  males,  one  in  non-reproductive 
condition  (testes  3 X 1.5  mm)  and  the  other 
with  somewhat  enlarged  testes  (6X4  mm). 
Evidence  of  reproductive  activity  at  Tsaraepae 
included  a juvenile  male  (probably  in  first  pre- 
basic  molt,  with  heavy  wing  molt  and  mod- 
erate body  molt),  and  two  adult  females  with 
convoluted  oviducts  but  unenlarged  ova. 
Stomachs  contained  the  remains  of  large  bee- 
tles (including  Cerambycidae),  large  orthop- 
terans,  spiders,  skinks,  and  geckos. 

Glycifohia  notabilis. — The  monotypic 
White-bellied  Honeyeater  is  endemic  to  Santo 
and  Malakula.  With  Dickinson’s  (2003)  place- 
ment of  this  species  in  the  genus  Glycifohia 
(previously  classified  as  Phylidonyris ),  its 
only  congener — the  Barred  Honeyeater  (G. 
undulata) — is  endemic  to  New  Caledonia. 
Previously,  both  had  been  placed  in  the  wide- 
spread Australian  genus,  Phylidonyris.  The 
White-bellied  Honeyeater  occurred  in  similar 
abundance  between  600  and  800  m at  both 
sites,  usually  in  large  tracts  of  forest.  Often, 
these  birds  congregated  at  flowering  trees  in 
noisy  groups  of  <15  individuals.  Of  four 
specimens  (two  from  each  site),  only  one  was 
reproductively  active,  a male  from  Wunaro- 
haehare with  enlarged  testes.  The  other  bird 
from  this  site,  an  adult  female  (no  bursa;  skull 
100%  ossified),  had  minute  ova,  a straight  ovi- 
duct (probably  had  not  yet  bred),  and  its 
wings,  tail,  and  body  were  molting.  An  adult 
female  from  Tsaraepae  had  these  same  char- 
acteristics. A young  male  (bursa  2X2  mm) 


from  Tsaraepae  also  was  molting,  probably  its 
first  pre-basic  molt. 

Petroica  multicolor  ambrynensis. — The 
subspecies  of  Pacific  Robin  from  Santo,  P.  m. 
ambrynensis , is  one  of  5 subspecies  from  Va- 
nuatu and  14  across  Oceania.  In  the  Solomons 
and  New  Guinea,  the  Pacific  Robin  is  restrict- 
ed to  montane  forests.  Although  apparently  re- 
stricted to  high-elevation  forests  (>500  m)  on 
Santo,  the  Pacific  Robin  may  be  found  at  low- 
er elevations  elsewhere  in  Vanuatu.  JJK  found 
it  to  be  common  in  lowland  forests  on  the 
rain-shadowed  Dillon’s  Bay  area  of  western 
Erromango.  On  Efate,  however,  DWS  found 
it  in  humid,  mid-elevation  forest  (—350  m).  In 
addition  to  not  finding  the  Pacific  Robin  at 
Tsaraepae  (although  our  guides  there  knew  of 
this  species),  no  one  has  recorded  it  from  any 
lowland  location  on  the  wet  (eastern)  side  of 
Santo.  Medway  and  Marshall  (1975)  recorded 
it  at  an  elevation  of  1,100  m on  the  eastern 
flank  of  Mt.  Tabwemasana,  but  we  recorded 
robins  (up  to  four  daily)  only  in  forest  from 
650  to  800  m near  Wunarohaehare.  The  three 
specimens  were  two  adult  males  with  enlarged 
testes  and  seminal  vesicles,  and  an  adult  fe- 
male that  probably  had  nested  recently  (ova 
not  enlarged,  but  oviduct  somewhat  thickened 
and  convoluted). 

Neolalage  banksiana. — The  Buff-bellied 
Monarch  belongs  to  a monotypic  genus  en- 
demic to  Vanuatu.  It  occurs  on  most  major 
islands  south  to  Efate  and  was  common  at 
both  of  our  primary  study  sites,  with  daily  re- 
cords of  up  to  25  at  Wunarohaehare  and  12  at 
Tsaraepae.  It  was  found  most  often  in  pairs  or 
family  groups  in  the  undergrowth  of  forest 
patches  or  large  tracts  of  forest,  especially 
where  vine  tangles  or  thickets  of  Hibiscus  til- 
iaceus  dominate  the  understory,  although 
some  birds  were  found  in  forests  with  an  open 
understory. 

The  song  of  the  Buff-bellied  Monarch  is  ap- 
parently undescribed;  Bregulla  (1992)  stated 
that,  “.  . . it  is  said  to  have  melodious  song.” 
AWK  tape-recorded  a bird  singing  in  scrubby 
dry  forest  adjacent  to  Wusi  village  on  the 
morning  of  25  October.  The  song  had  a stut- 
tering, jumbled  beginning,  then  three  rapid  se- 
ries of  reedy,  high-pitched,  whistled  notes. 
The  first  and  last  series  consisted  of  three  de- 
scending notes,  whereas  the  second  series 
consisted  of  only  two  descending  notes:  tee- 


Kratter  et  al.  • SANTO,  VANUATU  BIRD  COMMUNITIES 


305 


dee-dee — tee-dee — tee-dee-deee.  The  song, 
which  lasts  ~3  sec,  resembled  that  of  the  Fan- 
tailed Gerygone  ( Gerygone  flavolateralis ) but 
was  shorter,  and  the  tone  of  the  notes  was 
more  pure.  The  call  note  (a  drawn-out,  single 
burry  note  that  increased  in  amplitude)  was 
given  between  songs.  The  song  was  heard  (in- 
frequently) in  montane  forests  at  Wunarohae- 
hare  as  well,  but  not  at  Tsaraepae  the  follow- 
ing June.  Nevertheless,  Buff-bellied  Monarchs 
called  frequently  throughout  the  day  at  both 
sites,  especially  pairs  that  called  to  one  anoth- 
er while  foraging. 

Breeding  activity  of  this  species  was  pro- 
nounced at  Wunarohaehare,  where  a near-fin- 
ished nest  was  discovered  on  1 November,  2.5 
m above  ground  in  the  fork  of  a sapling  in 
humid  forest.  The  nest  was  similar  to  those 
described  for  the  species  by  Bregulla  (1992) 
and  Bowen  (1997).  At  least  two  pairs  of  Buff- 
bellied  Monarchs  were  found  accompanied  by 
recently  fledged  young  at  Wunarohaehare. 
Two  of  the  three  adult  males  taken  at  Wuna- 
rohaehare had  enlarged  testes;  the  other  male 
had  somewhat  enlarged  testes,  whereas  the  fe- 
male lacked  a bursa  but  had  a straight  oviduct, 
indicating  that  she  had  not  bred  previously.  At 
Tsaraepae,  one  of  the  two  adult  male  speci- 
mens had  enlarged  testes.  The  other  three 
specimens  from  Tsaraepae  were  young  birds 
with  bursae  and  incompletely  ossified  skulls. 
The  plumage  of  adult  males  is  slightly  more 
vividly  colored  than  that  of  adult  females  or 
non-adults. 

Clytorhynchus  pachycephaloides  grise- 
scens. — The  inconspicuous  Southern  Shrike- 
bill  species  is  found  only  in  New  Caledonia 
and  Vanuatu.  The  subspecies  C.  p.  grisescens 
is  endemic  to  Vanuatu.  Once  we  learned  its 
vocalizations  (see  below),  we  recorded  <4/ 
day  in  dense  forest  at  Tsaraepae  (600-650  m). 
Although  we  netted  four  (in  165  net-hr)  in  dry 
forest  near  sea  level  at  Wusi  village,  we  nei- 
ther netted  (in  337  net-hr)  nor  recorded  any  in 
the  higher-elevation  forests  at  Wunarohaehare. 
One  also  was  seen  by  AWK  at  the  Vatte  Con- 
servation Area  in  northern  Santo  in  November 
2002,  and  the  species  was  heard  often  and 
seen  occasionally  in  lowland  forests  at  the 
Loru  Protected  Area  (Bowen  1997).  Shrike- 
bills  were  netted  rarely  (0.006/net-hr)  at  two 
lowland  forest  sites  east  of  the  main  cordillera 
by  Medway  and  Marshall  (1975),  although 


none  was  found  at  their  higher-elevation  site 
(1,120  m).  The  birds  we  observed  were  slug- 
gish, perching  from  near  the  ground  to  8 m 
above  ground  in  the  humid  forest. 

Bregulla  (1992)  described  the  Southern 
Shrikebill’s  song  as  highly  variable  “drawn 
out  whistled  sounds  in  cadence.”  On  10  June 
at  Tsaraepae,  AWK  tape-recorded  a three-part 
song  made  up  of  two  evenly  spaced,  harsh 
chek  notes,  followed  by  a descending,  drawn- 
out,  burry  whistle.  The  most  commonly  re- 
corded call  was  a single,  burry  musical  note, 
similar  to  that  of  the  Buff-bellied  Monarch, 
but  less  raspy  and  dropping  in  pitch  at  the  end. 

Testes  of  the  male  collected  at  Tsaraepae 
were  somewhat  enlarged  (10  X 5 mm),  indi- 
cating recent  reproductive  activity.  The  four 
taken  near  sea  level  at  Wusi  were  adults  (no 
bursae,  skull  100%  ossified)  consisting  of  two 
reproductively  active  males  (testes  enlarged) 
and  a nonbreeding  male  and  female. 

Zosterops  flavifrons. — Endemic  to  Vanuatu, 
the  Yellow-fronted  White-eye  was  one  of  the 
most  common  forest  birds  at  both  sites,  as  it 
is  throughout  much  of  the  archipelago  (AWK, 
JJK,  DWS  pers.  obs.).  Up  to  75  were  found 
daily  from  near  sea  level  to  the  highest  ele- 
vations that  we  visited  (1,250  m at  Wunaro- 
haehare, 700  m at  Tsaraepae).  We  often  found 
White-eyes  in  fruiting  trees,  where  flocks  of 
<15  kept  up  a persistent  chatter.  It  co-oc- 
curred  at  some  forest  edges  with  a larger  con- 
gener (Z.  lateralis , the  Silver-eye),  although 
the  latter  usually  was  absent  from  the  large 
tracts  of  mature  forest  where  the  Yellow-front- 
ed White-eye  was  most  common.  At  Wuna- 
rohaehare, all  adult  specimens  were  in  repro- 
ductive condition  (three  males,  two  females). 
At  Tsaraepae,  all  five  specimens  were  young 
birds  (with  bursae  and/or  incompletely  ossi- 
fied skulls):  two  were  undergoing  wing  molt, 
three  were  in  tail  molt,  and  all  were  under- 
going body  molt. 

Turdus  poliocephalus  vanikorensis. — The 
extremely  polytypic  Island  Thrush  (51  rec- 
ognized subspecies;  Dickinson  2003)  occurs 
irregularly  from  the  Philippines  to  Samoa. 
Among  the  eight  subspecies  occurring  in  Va- 
nuatu is  T.  p.  vanikorensis , found  on  Santo, 
Malo,  and  the  Santa  Cruz  Group.  Similar  to 
the  Pacific  Robin,  today  the  Island  Thrush  is 
restricted  to  montane  forests  on  some  islands 
(e.g..  New  Guinea,  New  Ireland),  whereas  on 


306 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


others  (e.g.,  Rennell  in  the  Solomon  Islands) 
it  lives  in  the  lowlands.  Fossils  from  coastal 
sites  in  Tonga  (where  it  no  longer  occurs)  and 
New  Ireland  indicate  that  the  Island  Thrush 
has  undergone  considerable  range  contraction 
since  the  arrival  of  humans  on  the  islands 
(Steadman  1993,  2006b). 

The  Island  Thrush  was  absent  at  Tsaraepae 
but  common  at  Wunarohaehare,  where  we 
found  it  in  dry  forests  near  sea  level  (0.03/ 
net-hr),  in  montane  forests  at  600-800  m in 
elevation  (0.03/net-hr),  and  in  forest  patches 
at  1,250  m (0.08/net  hr).  Birds  collected  near 
Wusi  included  adults  of  both  sexes  with  en- 
larged gonads.  The  current  distribution  of  the 
Island  Thrush  on  Santo  resembles  that  of  the 
Pacific  Robin  in  being  present  in  dry  forest  on 
the  western  slopes  of  the  cordillera  but  absent 
(or  very  rare)  in  humid  forests  to  the  east. 
Likewise,  Bowen  (1997)  did  not  record  it  at 
the  Loru  Protected  Area.  This  may  reflect  a 
recent  change  in  its  status  east  of  the  cordi- 
llera, where  the  Island  Thrush  was  recorded 
frequently  at  two  lowland  forest  sites  in  1971 
(Hog  Harbour,  Apuna  River;  Medway  and 
Marshall  1975).  Predation  by  feral  cats  may 
be  the  cause  of  the  apparent  decline  of  the 
Island  Thrush  on  Santo. 

DISCUSSION 

Inter-site  ( intra-island ) comparisons. — Of 
the  39  species  of  landbirds  recorded  from  at 
least  one  of  the  six  surveyed  sites  on  Santo 
(Fig.  1,  Table  3),  only  17  (44%)  were  found 
at  five  or  six  sites.  These  included  5 of  the  1 1 
endemic  or  near-endemic  species.  Three  spe- 
cies known  from  Santo  ( Gallicolumba  sanc- 
taecrucis , Cacomantis  pyrrhophanus , Aplonis 
santovestris ) were  not  recorded  at  any  of  the 
sites.  That  more  species  are  not  more  wide- 
spread on  Santo  may  be  due  to  elevational 
factors;  nine  species  are  known  only  from  one 
or  more  of  the  three  highland  (>500  m)  sites 
(. Aplonis  santovestris  also  is  restricted  to  high- 
lands), and  two  species  ( Gallirallus  philippen- 
sis,  Tyto  alba ) are  recorded  only  from  lowland 
sites  (<500  m).  Of  the  remaining  species 
found  at  fewer  than  five  sites,  some  preferred 
more  open  habitats  ( Todiramphus  chloris, 
Rhipidura  albiscapa , Artamus  leucorhynchus , 
Zosterops  lateralis)  and  some  were  rare  ( Me - 
gapodius  layardi , Falco  peregrinus,  Galli- 
columba sanctaecrucis,  Charmosyna  palma- 


rum , Aplonis  zelandica );  for  unknown  rea- 
sons, others  ( Columba  vitiensis,  Ptilinopus 
tannensis,  Cacomantis  pyrrhophanus , Collo- 
calia  vanikorensis,  Turdus  poliocephalus ) oc- 
cur only  locally. 

The  inter-site  variation  in  landbird  com- 
munities on  Santo  is  noteworthy.  In  island 
biogeography,  it  has  been  common  practice  to 
assemble  lists  based  on  the  entire  fauna  or  flo- 
ra of  an  island,  even  though  many  species  may 
rarely,  if  ever,  interact  because  they  are  not 
syntopic.  Because  much  of  island  biogeogra- 
phy theory  (e.g.,  Mac  Arthur  and  Wilson  1967; 
Diamond  and  Marshall  1977;  Diamond  1980, 
1982;  Mayr  and  Diamond  2001)  is  based  on 
analyses  at  the  community  level,  it  may  be 
more  biologically  informative  to  compare  the 
avifauna  from  single  sites,  rather  than  the  en- 
tire avifuana  of  islands,  especially  on  large  is- 
lands where  strong  elevational  and  precipita- 
tion gradients  occur  (e.g.,  Santo).  Aside  from 
the  massive  island  of  New  Guinea,  there  is  no 
island  in  Melanesia  for  which  bird  survey  data 
have  been  published  for  as  many  sites  as  those 
on  Santo.  We  urge  biologists  working  on  is- 
lands to  undertake  the  surveys  needed  to  gen- 
erate data  on  presence/absence,  relative  abun- 
dance, and  habitat  preference  of  birds  from 
single  sites. 

Inter-archipelago  comparisons. — Com- 
pared with  a forested  lowland  site  on  the  sim- 
ilarly sized  island  of  Isabel  (3,995  km2;  Fig. 
1)  in  the  Solomon  Islands  (Kratter  et  al. 
2001a,  2001b),  the  species  richness  at  the  sites 
on  Santo  was  much  lower  (25-30  versus  59 
resident  species  of  forest  birds).  Pigeons  and 
doves  contributed  equally  to  richness  (seven 
species  at  sites  on  either  island),  whereas  pas- 
serine diversity  was  not  as  rich  on  Santo  but 
contributed  a higher  percentage  to  species 
richness  (15-16  species  or  50-53%  at  the 
Santo  sites,  versus  21  species  or  36%  at  Isa- 
bel). The  sites  on  Santo  also  had  markedly 
fewer  hawks  and  falcons  (one  compared  with 
five  species  on  Isabel),  parrots  (two  versus  six 
species),  and  kingfishers  (two  versus  six  spe- 
cies). In  addition,  the  sites  on  Santo  held  a 
smaller  portion  of  the  entire  forest  bird  avi- 
fauna than  that  found  along  the  Garanga  River 
on  Isabel:  the  30  species  found  at  either  Wun- 
arohaehare or  Tsaraepae  represent  7 1 % of  the 
42  species  known  from  Santo,  whereas  the  59 
species  found  along  the  Garanga  River  rep- 


Kratter  et  al.  • SANTO,  VANUATU  BIRD  COMMUNITIES 


307 


resent  84%  of  the  70  species  of  landbirds 
known  from  Isabel.  This  may  have  been  due, 
in  part,  to  our  longer  stay  at  the  Garanga  River 
site  (21  days  over  2 years  versus  6 and  7 days 
at  Wunarohaehare  and  Tsaraepae,  respective- 
ly). Another  possible  factor  is  that,  for  a given 
island  in  Oceania,  lowland  forests  tend  to  sup- 
port richer  bird  communities  than  montane 
forests  (Mayr  and  Diamond  2001). 

Species  not  recorded  at  our  sites. — At  Wun- 
arohaehare and  Tsaraepae,  we  failed  to  record 
seven  species  known  to  occur  in  forests  on 
Santo — the  Peregrine  Falcon  ( Falco  peregri- 
nus),  Santa  Cruz  Ground  Dove  ( Gallicolumba 
sanctaecrucis ),  Fan-tailed  Cuckoo  ( Caco - 
mantis  pyrrhophanus),  Rufous-winged  Star- 
ling ( Aplonis  zelandicus ),  Mountain  Starling 
{A.  santovestris),  Melanesian  Thicketbird 
{Cichlornis  whitneyi ),  and  Red-headed  Parrot- 
finch  ( Erythrura  cyaneovirens).  Our  guides 
knew  the  Peregrine  Falcon  and  called  it  “vus- 
avusa”  in  the  Hakei  language;  it  may  be  a rare 
resident  at  or  near  our  sites,  most  likely  in 
areas  with  cliffs.  The  Santa  Cruz  Ground 
Dove  is  considered  rare  in  montane  forests 
(Bregulla  1992);  our  guides,  however,  knew  it 
and  called  it  “nono.”  Perhaps  restricted  to  the 
lowlands,  the  Fan-tailed  Cuckoo  has  become 
rare  in  Vanuatu  (Bregulla  1992),  and  our 
guides  did  not  recognize  it.  The  Fan-tailed 
Cuckoo  also  was  not  recorded  at  the  other 
four  sites  surveyed  in  1971  and  1995  (Med- 
way and  Marshall  1975,  Bowen  1997),  al- 
though Bregulla  (1992)  considered  it  uncom- 
mon on  Santo. 

The  Mountain  Starling  is  known  to  occur 
only  in  cloud  forest  at  elevations  >1,150  m 
on  Santo  (Medway  and  Marshall  1975),  and, 
on  the  southern  slopes  of  Peak  Santo,  the  star- 
ling was  not  found  below  1,400  m (Bregulla 
1992).  The  Rufous-winged  Starling  is  thought 
to  be  common  in  forests  at  around  1,000  m 
on  Santo  (Bregulla  1992).  Although  it  could 
be  absent  from  the  drier  forests  on  the  western 
slope,  we  suspect  that  we  would  have  found 
it  on  the  wetter  southern  slopes  had  the  rain- 
fall diminished,  thereby  allowing  us  access  to 
higher  elevations.  Our  guides  did  not  recog- 
nize the  illustrations  (in  Bregulla  1992)  of  ei- 
ther starling  species.  The  Melanesian  Thick- 
etbird is  a streamside  specialist,  and  we  did 
not  sample  streamsides  at  either  site.  Our 
guides  knew  the  species,  however,  explaining 


that  it  lives  close  to  the  ground  along  high- 
elevation  streams;  they  called  the  male  “sisi- 
via”  and  the  female  “sisiriva.”  The  Red- 
headed Parrotfinch  {Erythrura  cyaneovirens ) 
is  an  uncommon  fig  specialist  suspected  of  be- 
ing nomadic,  which  likely  explains  its  absence 
from  seemingly  suitable  habitats  if  the  large, 
fleshy  fig  fruits  that  it  prefers  (Bregulla  1992; 
DWS  pers.  obs.  on  Efate  Island,  3 August 
1997)  are  scarce  or  absent.  Our  guides  had  no 
name  for  Red-headed  Parrotfinch. 

Finally,  bones  from  archaeological  sites 
elsewhere  in  Vanuatu  give  clues  about  which 
species  once  may  have  lived  on  Santo.  DWS 
and  JJK  have  identified  extinct  or  extirpated 
species  of  megapode  {Megapodius  unde- 
scribed sp.)  and  hawk  ( Accipiter  cf . fasciatus) 
on  Efate,  flightless  rail  ( Porzana  undescribed 
sp.)  and  parrot  {Eclectus  infectus ; Steadman 
2006a)  on  Malakula,  and  starling  {Aplonis  un- 
described sp.)  on  Erromango.  Given  that  most 
volant  species  of  Pacific  Island  landbirds  were 
more  widespread  before  the  arrival  of  humans 
on  the  islands  (Steadman  1995,  2006b),  we 
suspect  that  these  (or  similar  species  in  the 
case  of  flightless  rails)  once  lived  on  Santo 
and  many  other  islands  in  Vanuatu. 

ACKNOWLEDGMENTS 

We  kindly  thank  D.  Kalfatak  and  E.  Bani  of  the 
Vanuatu  Environment  Unit  for  permission  to  undertake 
this  research,  and  for  logistical  assistance.  We  are  very 
grateful  to  R.  Regenvanu  and  the  staff  of  the  Vanuatu 
National  Museum  for  crucial  logistical  support  and  ad- 
vice. For  assistance  at  our  field  sites  we  thank  K.  Al- 
vea.  Chief  Bua  and  the  residents  of  Wusi,  W.  Dauron, 
M.  K.  Hart,  R.  Kolomule,  Chief  P.  Leon  and  the  resi- 
dents of  Kerevalissy,  S.  Nisak,  R.  Palo,  and  P.  Tav- 
ouiruja.  Funding  (to  DWS)  was  provided  by  the  Uni- 
versity of  Florida  Division  of  Sponsored  Research  (Re- 
search Opportunity  Fund  grant  U001)  and  National 
Science  Foundation  grant  EAR-9714819.  We  thank  G. 
Dutson  and  two  anonymous  reviewers  for  constructive 
comments  on  the  manuscript. 

LITERATURE  CITED 

Bowen,  J.  1997.  The  status  of  the  avifauna  of  Loru 
Protected  Area,  Santo,  Vanuatu.  Bird  Conserva- 
tion International  7:331-344. 

Bregulla,  H.  L.  1992.  Birds  of  Vanuatu.  Anthony 
Nelson,  Shropshire,  United  Kingdom. 

Collot,  J.  Y.  and  M.  A.  Fisher.  1989.  Formation  of 
fore-arc  basins  by  collision  between  seamounts 
and  accretionary  wedges:  an  example  from  the 
New  Hebrides  subduction  zone.  Geology  17:930- 
933. 


308 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Diamond,  J.  M.  1975.  Assembly  of  species  commu- 
nities. Pages  342-444  in  Ecology  and  evolution 
of  communities  (M.  L.  Cody  and  J.  M.  Diamond, 
Eds.).  Belknap  Press,  Cambridge,  Massachusetts. 

Diamond,  J.  M.  1980.  Species  turnover  in  island  bird 
communities.  Proceedings  of  the  International  Or- 
nithological Congress  17:777-782. 

Diamond,  J.  M.  1982.  Effects  of  the  species  pool  size 
on  species  occurrence  frequencies:  musical  chairs 
on  islands.  Proceedings  of  the  National  Academy 
of  Sciences,  USA  79:2420-2424. 

Diamond,  J.  M.  and  A.  G.  Marshall.  1977.  Niche 
shifts  in  New  Hebridean  birds.  Emu  77:61-62. 

Dickinson,  E.  C.  (Ed.).  2003.  The  Howard  and  Moore 
complete  checklist  of  the  birds  of  the  world. 
Princeton  University  Press,  Princeton,  New  Jer- 
sey. 

Gotelli,  N.  J.  and  G.  L.  Entsminger.  2001.  Swap  and 
fill  algorithms  in  null  model  analysis:  rethinking 
the  knight’s  tour.  Oecologia  129:281-291. 

Kratter,  A.  W.,  D.  W.  Steadman,  C.  E.  Smith,  C.  E. 
Filardi,  and  H.  P.  Webb.  2001a.  Avifauna  of  a 
lowland  forest  site  on  Isabel,  Solomon  Islands. 
Auk  118:472-483. 

Kratter,  A.  W.,  D.  W.  Steadman,  C.  E.  Smith,  and 
C.  E.  Filardi.  2001b.  Reproductive  condition, 
molt,  and  body  mass  of  birds  from  Isabel,  Solo- 
mon Islands.  Bulletin  of  the  British  Ornitholo- 
gists’ Club  121:128-144. 

Lal,  B.  V.  and  K.  Fortune.  2000.  The  Pacific  Islands: 
an  encyclopedia.  University  of  Hawaii  Press,  Hon- 
olulu. 

MacArthur,  R.  H.  and  E.  O.  Wilson.  1967.  The  the- 
ory of  island  biogeography.  Princeton  University 
Press,  Princeton,  New  Jersey. 

MacFarlane,  A.,  J.  N.  Carney,  A.  J.  Crawford,  and 
H.  G.  Greene.  1988.  Vanuatu:  a review  of  onshore 
geology.  Pages  45-9 1 in  Geology  and  offshore  re- 
sources of  Pacific  Island  arcs:  Vanuatu  region  (H. 
G.  Greene  and  F.  L.  Wong,  Eds.).  Circum-Pacific 
Council  for  Energy  and  Mineral  Resources  Earth 
Science  Series,  vol.  8,  Houston,  Texas. 

Mallick,  D.  I.  J.  1975.  Development  of  the  New  Heb- 
rides archipelago.  Philosophical  Transactions  of 


the  Royal  Society  of  London,  Series  B 272:277- 
285. 

Mayr,  E.  1934.  Birds  collected  during  the  Whitney 
South  Sea  Expedition.  29,  Notes  on  the  genus  Pe- 
troica.  American  Museum  Novitates  no.  714. 

Mayr,  E.  1941.  Birds  collected  during  the  Whitney 
South  Sea  Expedition.  45,  Notes  on  New  Guinea 
Birds.  8.  American  Museum  Novitates  no.  1133. 

Mayr,  E.  1963.  Animal  species  and  evolution.  Har- 
vard University  Press,  Cambridge,  Massachusetts. 

Mayr,  E.  and  J.  Diamond.  2001.  Birds  of  Northern 
Melanesia.  Oxford  University  Press,  Oxford,  Unit- 
ed Kingdom. 

Medway,  L.  and  A.  G.  Marshall.  1975.  Terrestrial 
vertebrates  of  the  New  Hebrides:  origin  and  dis- 
tribution. Philosophical  Transactions  of  the  Royal 
Society  of  London,  Series  B 272:423-365. 

Mueller-Dombois,  E.  and  F.  R.  Fosberg.  1998.  Veg- 
etation of  the  tropical  Pacific  Islands.  Springer- 
Verlag,  New  York. 

Nunn,  P.  D.  1994.  Oceanic  islands.  Blackwell  Publish- 
ers, Oxford,  United  Kingdom. 

Price,  J.  J.,  K.  P.  Johnson,  and  D.  H.  Clayton.  2004. 
The  evolution  of  echolocation  in  swiftlets.  Journal 
of  Avian  Biology  35:135-143. 

Sanderson,  J.  G.,  M.  P.  Moulton,  and  R.  G.  Self- 
ridge. 1998.  Null  matrices  and  the  analysis  of  spe- 
cies co-occurrences.  Oecologia  116:275-283. 

Scott,  W.  E.  1946.  Birds  observed  on  Espiritu  Santo, 
New  Hebrides.  Auk  63:362-368. 

Steadman,  D.  W.  1993.  Biogeography  of  Tongan  birds 
before  and  after  human  impact.  Proceedings  of  the 
National  Academy  of  Sciences,  USA  90:818-822. 

Steadman,  D.  W.  1995.  Prehistoric  extinctions  of  Pa- 
cific Island  birds:  biodiversity  meets  zooarchaeol- 
ogy. Science  267:1123-1131. 

Steadman,  D.  W.  1997.  The  historic  biogeography  and 
community  ecology  of  Polynesian  pigeons  and 
doves.  Journal  of  Biogeography  24:157-173. 

Steadman,  D.  W.  2006a.  A new  species  of  extinct  par- 
rot (Psittacidae:  Eclectus ) from  Tonga  and  Vanu- 
atu, South  Pacific.  Pacific  Science  60:137-145. 

Steadman,  D.  W.  2006b.  Extinction  and  biogeography 
of  tropical  Pacific  birds.  University  of  Chicago 
Press,  Chicago,  Illinois. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):309-3 15,  2006 


A DESCRIPTION  OF  THE  FIRST  MICRONESIAN  HONEYEATER 
( MYZOMELA  RUBRATRA  SAFFORDT)  NESTS  FOUND  ON 
SAIPAN,  MARIANA  ISLANDS 

THALIA  SACHTLEBEN,134  JENNIFER  L.  REIDY,2  AND  JULIE  A.  SAVIDGE1 2 3 4 


ABSTRACT. — We  provide  the  first  descriptions  of  Micronesian  Honey  eater  ( Myzomela  rubratra  saffordi) 
nests  {n  = 7)  and  nestlings  (n  = 6)  from  Saipan  in  the  Mariana  Islands.  Measured  nests  ( n = 3)  averaged  46.7 
mm  in  inner  cup  diameter,  65.7  mm  in  outer  diameter,  41.3  mm  in  cup  height,  and  55.3  mm  in  external  nest 
height.  We  found  all  nests  in  two  species  of  native  trees,  1.47-5.1  m above  the  ground.  Nesting  materials  were 
primarily  vine  tendrils  and  Casuarina  equisetifolia  needles.  We  also  report  observations  of  parental  behavior. 
Nests,  nest  placements,  and  behaviors  appeared  broadly  similar  to  those  reported  for  this  species  prior  to  its 
extirpation  on  Guam,  and  on  other  islands  in  Micronesia.  Received  2 May  2005,  accepted  26  January  2006. 


The  Meliphagidae  family  (honeyeaters)  is 
restricted  to  the  Australo-Papuan  region 
(Mayr  1945).  Micronesian  Honeyeaters  {My- 
zomela rubratra ) occur  throughout  the  high 
islands  (i.e.,  those  of  volcanic  origin  rising 
more  than  a few  meters  above  sea  level)  of 
Micronesia,  with  subspecies  endemic  to  Palau 
(M.  r.  kobayashii),  Yap  (M.  r.  kurodai ), 
Chuuk  (M.  r.  major),  Pohnpei  {M.  r.  dichro- 
mata),  Kosrae  (M.  r.  rubratra),  and  the  Mar- 
iana Islands  (M.  r.  saffordi ; Pratt  et  al.  1987). 
Within  the  Mariana  Islands,  Baker  (1951) 
found  that  birds  from  Guam,  Rota,  Tinian,  and 
Saipan  are  similar  with  respect  to  morpho- 
metric measurements,  and  he  does  not  sepa- 
rate them  taxonomically.  Micronesian  Hon- 
eyeaters, along  with  most  other  native  forest 
birds,  were  extirpated  from  Guam  in  the  mid- 
1980s  with  the  arrival  and  range  expansion  of 
the  brown  treesnake  ( Boiga  irregularis ; Sav- 
idge  1987,  Wiles  et  al.  2003).  Surveys  on 
Rota,  Tinian,  and  Saipan  (the  inhabited  islands 
of  the  Commonwealth  of  the  Northern  Mari- 
ana Islands  [CNMI])  have  indicated  that  Mi- 
cronesian Honeyeaters  are  less  numerous  on 
Saipan  than  on  Rota  or  Tinian  (Pratt  et  al. 
1979,  Ralph  and  Sakai  1979,  Jenkins  and 
Aguon  1981,  Jenkins  1983,  Craig  1996),  al- 


1  Dept,  of  Fishery  and  Wildlife  Biology,  Colorado 
State  Univ.,  Fort  Collins,  CO  80523-1474,  USA. 

2 Dept,  of  Fisheries  and  Wildlife,  Univ.  of  Missouri, 
Columbia,  MO  65211,  USA. 

3 Current  address:  High  Desert  Ecological  Research 
Inst.,  15  SW  Colorado  Ave.,  Ste.  300.  Bend,  OR 
97702,  USA. 

4 Corresponding  author;  e-mail: 
tjskiwi@yahoo.co.uk 


though  Engbring  et  al.  (1986)  found  that  den- 
sities were  greater  on  Saipan  than  on  Tinian. 
On  Saipan,  Engbring  et  al.  (1986)  counted  549 
honeyeaters  (mean  of  2.25  birds  per  station  ± 
0.14  SE),  and  estimated  the  total  Micronesian 
Honeyeater  population  at  22,573.  In  a repeat 
survey,  the  U.S.  Fish  and  Wildlife  Service 
(1997)  counted  316  honeyeaters  (mean  of  1.30 
birds  per  station  ± 0.09  SE;  no  population 
estimate  given),  indicating  a possible  decline 
in  the  honeyeater  population  between  survey 
periods. 

Little  research  has  been  published  on  the 
avifauna  of  the  Mariana  Islands,  and  many  de- 
tailed aspects  of  life  histories  are  unknown  for 
most  native  and  endemic  species  (Rodda  et  al. 
1998,  Mosher  and  Fancy  2002).  This  lack  of 
information  hampers  the  development  and  im- 
plementation of  conservation  plans.  Despite 
interdiction  measures,  the  number  of  brown 
treesnake  sightings  on  Saipan  has  increased  in 
recent  years  (Rodda  et  al.  1998;  N.  B.  Hawley 
pers.  comm.);  although  definitive  proof  is 
lacking,  75  plausible  brown  treesnake  sight- 
ings and  1 1 hand-captured  brown  treesnakes 
on  Saipan  (Gragg  2004)  indicate  that  an  in- 
cipient population  of  snakes  is  now  estab- 
lished (Colvin  et  al.  2005).  Thus,  information 
on  the  ecology  and  breeding  biology  of  all 
avian  species  in  the  CNMI  is  urgently  needed 
so  that  captive  breeding  programs  can  be  im- 
plemented. 

We  undertook  a study  to  assess  nesting  suc- 
cess of  common  forest  passerines  in  native 
and  nonnative  forests  of  Saipan.  Micronesian 
Honeyeaters  were  not  a target  species  for  this 
study,  as  they  are  reported  to  be  more  corn- 


309 


310 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


FIG.  1.  Location  of  Saipan  within  the  Commonwealth  of  the  Northern  Mariana  Islands,  and  Saipan  study 
sites  (shaded  areas)  in  which  we  searched  for  nests  of  native  forest  birds  during  2003  and  2004  to  assess  nesting 
success;  Micronesian  Honeyeater  nests  were  found  at  Marpi,  As  Teo,  Kagman,  and  Laolao  Bay.  Marpi,  As  Teo, 
and  Kagman  study  areas  were  native  forest;  Cow  Town,  Bird  Island,  Obyan,  and  Naftan  were  nonnative  tan- 
gantangan  forest;  Laolao  Bay  was  mixed  native/agriforest.  Approximate  coordinates  (taken  at  the  nearest  open 
area,  generally  a road)  for  study  sites  were  as  follows:  As  Teo  15°11'N,  145°  45' E;  Bird  Island  15°  15' 
N,  145°  48' E;  Cow  Town  15°  16'  N,  145°  49' E;  Kagman  15°09'N,  145°  16' E;  Laolao  Bay  15°09'N,  145° 
44'  E;  Marpi  15°  16'  N,  145°  47'  E;  Naftan  15°  06'  N,  145°  44'  E;  Obyan  15°  06'  N,  145°  43'  E.  The  dotted  line 
on  the  location  map  signifies  the  division  between  the  Territory  of  Guam  and  the  Commonwealth  of  the  Northern 
Mariana  Islands. 


mon  in  coconut  plantings,  shrubbery  and  gar- 
dens of  villages,  scrub,  coastal  strand,  and  di- 
verse second-growth  forest  composed  of  both 
native  and  introduced  trees  (Seale  1901,  Saf- 
ford  1902,  Pratt  et  al.  1979,  Jenkins  1983, 
Engbring  et  al.  1986).  Over  the  course  of  our 
study,  however,  we  incidentally  found  seven 
Micronesian  Honeyeater  nests.  To  our  knowl- 
edge, these  are  the  first  nests  of  this  species 
found  on  Saipan,  although  nests  have  previ- 
ously been  found  on  Guam,  and  one  nest  has 
been  found  on  Rota.  Here,  we  describe  nests 
and  nestlings  from  Saipan  and  compare  these 
descriptions  with  those  from  Guam,  Rota,  and 
other  islands  in  Micronesia  from  which  infor- 
mation is  available. 

METHODS 

Study  area. — Saipan,  located  in  the  western 
Pacific  Ocean  (15°  10'  N,  145°  45'  E;  Fig.  1), 
encompasses  a land  area  of  123  km2,  and  is 
the  second  largest  island  in  the  Marianas.  The 
island  has  a tropical  climate  with  an  annual 
mean  temperature  of  28.3°  C and  mean  annual 
rainfall  of  200—250  cm.  The  timing  of  the  wet 
and  dry  seasons  varies  somewhat  between 
years,  but  the  wet  season  usually  extends  from 


July  to  November  and  the  dry  season  from 
December  to  June.  Typhoons  may  occur  at 
any  time,  but  are  most  frequent  between  Au- 
gust and  December  (Young  1989,  Mueller- 
Dombois  and  Fosberg  1998). 

We  focused  our  study  on  two  forest  types — 
introduced  tangantangan  ( Leucaena  leuco- 
cephala)  forest  and  native  limestone  forest. 
Most  (77%)  of  the  forest  remaining  on  Saipan 
is  nonnative  (Falanruw  et  al.  1989),  and  tan- 
gantangan forest  is  estimated  to  cover  28%  of 
the  island.  This  tree  species  grows  in  dense, 
near-monocultures  on  flat  lowlands  and  pla- 
teaus (Craig  1990).  Native  limestone  forest  is 
restricted  to  cliffs  and  less  accessible  areas  not 
easily  cultivated  (Craig  1989,  Stinson  and 
Stinson  1994),  and  is  estimated  to  cover  only 
5-19%  of  Saipan  (Engbring  et  al.  1986, 
Young  1989).  Pisonia  grandis  and  Cynometra 
ramiflora  dominate  the  canopy  of  this  forest 
type,  and  C.  ramiflora  and  Guamia  mariannae 
are  the  most  common  species  in  the  understo- 
ry (Craig  1996).  Study  sites  were  selected  in 
three  native,  four  nonnative,  and  one  mixed 
forest  (Fig.  1).  The  mixed  forest  contained 
common  native  and  agriforest  trees,  including 
coconut  ( Cocos  nucifera ) and  mango  {Man- 


Sachtleben  et  al.  • MICRONESIAN  HONEYEATER  NESTS  ON  SAIPAN 


31  1 


gifera  indica).  Study  areas  were  delineated  by 
transects  marked  with  flagging. 

Avian  surveys. — We  conducted  our  study 
from  April  to  July  2003  and  February  to  May 
2004.  Micronesian  Honeyeater  nests  were 
found  while  searching  line  transects  according 
to  distance  sampling  methodology  (Buckland 
et  al.  2001)  or  incidentally  while  moving 
through  the  forest  to  monitor  nests  of  other 
species.  When  found,  each  nest  was  flagged 
and  assigned  a unique  nest  identification  num- 
ber. Nest  contents  were  visually  checked  and 
described  at  3-day  intervals,  using  a mirror  on 
a telescoping  pole  if  necessary.  We  did  not 
handle  nest  contents  while  nests  were  still  ac- 
tive; thus,  no  egg  measurements  were  made, 
and  we  visually  estimated  nestling  character- 
istics by  using  a millimeter  ruler  for  compar- 
ison. 

After  each  nesting  attempt  was  completed, 
we  measured  the  nest’s  height,  distance  from 
trunk,  and  number  and  diameter  of  supporting 
branch(es).  Tree  species  and  tree  height  were 
also  recorded.  We  used  a clinometer  to  mea- 
sure nest  and  tree  heights  (unless  these  could 
be  measured  directly  with  a steel  measuring 
tape),  a steel  measuring  tape  to  measure  dis- 
tance from  the  trunk,  and  a millimeter  ruler  to 
measure  diameters  of  supporting  branches.  We 
also  estimated  the  distance  between  the  nest 
and  the  nearest  road  in  25 -m  categories  (<25, 
26-50,  51-75,  76-100,  and  >100  m).  Nests 
were  collected  if  possible  and  measured  with 
a millimeter  ruler,  after  which  they  were  la- 
beled and  given  to  the  CNMI  Division  of  Fish 
and  Wildlife  on  Saipan. 

RESULTS 

We  discovered  seven  honeyeater  nests  on 
31  May  2003,  and  on  17  February,  9 March, 
12  March,  7 April,  9 April,  and  26  April  2004. 
Two  nests  contained  eggs,  two  contained  nest- 
lings, and  two  were  empty  when  located.  The 
female  was  sitting  on  one  nest  and  was  not 
disturbed;  in  this  case  the  nest  contents  were 
not  determined  when  the  nest  was  discovered. 
No  adults  were  in  attendance  at  three  nests 
upon  initial  discovery.  Four  nests  failed  (three 
during  incubation  and  one  at  an  undetermined 
nesting  stage),  and  three  fledged  young.  Four 
nests  were  located  in  mixed  forest,  and  one 
nest  was  located  in  each  of  the  three  native 
sites.  All  six  nests  in  which  we  observed  con- 


tents contained  two  eggs  or  two  young.  Ini- 
tially, we  mistook  two  nests  for  Bridled 
White-eye  ( Zosterops  conspicillatus  saypani ) 
nests  due  to  their  similar  size,  structure,  and 
placement.  However,  we  noticed  that  the  nests 
of  Micronesian  Honeyeaters  tended  to  have 
thinner  walls  and  deteriorated  more  rapidly 
than  Bridled  White-eye  and  Golden  White-eye 
( Cleptornis  marchei ) nests,  which  they  oth- 
erwise closely  resembled. 

Nest  composition  and  structure. — Only 
three  nests  were  accessible  and  in  adequate 
condition  for  measurement.  Cup  heights  were 
39,  40,  and  45  mm  (mean  = 41.3  mm),  and 
nest  heights  were  41,  50,  and  75  mm  (mean 
= 55.3  mm).  Internal  diameters  were  43,  47, 
and  50  mm  (mean  = 46.7  mm),  and  external 
diameters  were  55,  69,  and  73  mm  (mean  = 
65.7  mm).  Nests  were  composed  of  vine  ten- 
drils and  Casuarina  equisetifolia  needles  (Fig. 
2),  and  part  of  a leaf  skeleton  from  a native 
Pandanus  sp.  was  entwined  around  the  outer 
base  of  one  nest. 

Nest  placement. — Micronesian  Honeyeater 
nests  were  located  at  various  distances  from 
roads  (i.e.,  from  <25  to  >100  m).  Four  nests 
were  placed  in  Guamia  mariannae  and  three 
were  placed  in  a Psychotria  (genera  compris- 
ing more  than  one  species  in  CNMI,  and 
which  we  could  not  identify  to  species  level, 
are  listed  herein  only  to  the  genus  level).  Nest 
(and  tree)  heights  in  G.  mariannae  were  1.5 
m (5.6  m),  3 m (5  m),  3.5  m (6  m),  and  5.1 
m (not  obtained),  and  in  Psychotria  they  were 
1.5  m (2  m),  1.7  m (2.3  m),  and  3.8  m (8  m). 
Nests  were  placed  83-184  cm  from  the  trunk 
in  G.  mariannae  and  0-103  cm  from  the  trunk 
in  Psychotria , generally  near  the  outer  edge  of 
the  tree  (Fig.  2).  The  number  of  nest  support 
branches  varied  from  two  to  five  in  both  tree 
species,  and  support  branch  diameter  ranged 
from  1.5  to  9.7  mm  in  G.  mariannae  and  from 
1.5  to  2.5  mm  in  Psychotria. 

Egg  description. — Although  four  monitored 
nests  each  contained  two  eggs,  we  had  a clear 
view  of  the  eggs  only  in  the  nest  found  on  26 
April  2004.  The  eggs  were  creamy  white  and 
marked  with  two  distinct  rings  of  brown 
speckles,  one  ring  near  the  broad  end  and  the 
other  near  the  narrow  end  of  the  egg. 

Nestling  description. — Of  the  three  nests 
from  which  young  fledged  successfully,  we 
found  two  during  the  nestling  stage  and  one 


312 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


FIG.  2.  Micronesian  Honeyeater  ( Myzomela  rubratra  saffordi)  nest  photographed  on  Saipan,  Mariana  Is- 
lands, 19  April  2004,  showing  its  placement  at  the  outer  end  of  the  branch. 


during  the  incubation  stage.  Micronesian  Hon- 
eyeater nestlings  are  altricial  and  closely  re- 
semble Bridled  White-eye  nestlings  until  they 
develop  red  pin  feathers.  Because  nestling  de- 
velopment was  variable,  each  nest  is  treated 
separately. 

The  2003  nest  contained  eggs  when  found, 
and  the  two  nestlings  were  first  seen  at  day 
0-1  following  hatching.  At  this  age  they  were 
estimated  to  be  approximately  2 cm  in  length, 
had  dark  pink  skin,  and  were  downy  on  their 
wings  and  backs.  On  day  3-4,  the  nestlings 
had  grown  to  3—3.5  cm  in  length,  were  still 
covered  with  down,  and  their  skin  color  was 
dark  pink/purple.  They  appeared  well  fed,  as 
they  had  large,  rounded  stomachs.  At  day  6- 
7,  when  their  eyes  were  beginning  to  open, 
the  nestlings  were  4-4.5  cm  long,  with  wing 
pins  approximately  5 mm  in  length  and  back 
pins  beginning  to  erupt.  Their  heads  were  cov- 
ered in  long  down.  On  day  7-8,  the  chicks 
were  still  4-4.5  cm  long,  their  wing  and  back 
pins  were  8 and  2 mm  (respectively)  long, 
their  bills  were  beginning  to  curve,  and  their 
head  pins  still  had  not  erupted.  Underlying 
skin  color,  which  lightened  progressively 
throughout  nestling  development,  was  pale 
pink  by  this  stage.  At  day  9-10,  the  wing  pins 
were  10  mm  in  length  and  tail  and  head  pins 
had  erupted  1 mm.  Tan  brown  feathers  had 
erupted  from  the  wing  pins,  red  feathers  were 


beginning  to  erupt  from  the  back  pins,  and  1 - 
to  2-mm  head  pins  were  visible  on  day  10- 
1 1 . Both  nestlings  fledged  prematurely  on  day 
13-14,  when  the  observer  was  1 m from  the 
nest.  One  nestling  was  captured  and  returned 
to  the  nest,  but  the  second  could  not  be  relo- 
cated and  was  left  to  the  adults  who  remained 
nearby  and  were  agitated.  At  this  time,  the 
nestlings  were  estimated  at  5.5  cm  in  length, 
but  they  were  not  yet  fully  feathered.  Red 
feathers,  1 mm  in  length,  had  erupted  on  the 
back,  gray  feathers  had  erupted  on  the  head, 
and  8-mm  tail  pins  did  not  yet  have  erupted 
feathers.  The  breast  was  bare.  On  day  14-15, 
the  remaining  nestling’s  wing  feathers  had 
turned  dark  gray,  and  it  fledged  at  day  15—16. 

The  second  nest  that  fledged  young  was 
found  on  12  March  2004.  On  that  date,  the 
two  nestlings  were  already  approximately  4 
cm  in  length,  their  eyes  were  open,  and  they 
had  2-mm  long  downy  feathers  erupting  from 
the  pins  on  their  wings,  backs,  and  heads.  On 
15  March,  only  one  nestling  remained.  This 
nestling  fledged  prematurely  on  18  March 
when  the  observer  approached  to  ~3  m from 
the  nest.  The  nestling  fluttered  away,  but  it 
could  not  fly  and  was  captured  and  returned 
to  the  nest.  We  estimated  the  nestling  to  be 
4-4.5  cm  long  and  it  did  not  appear  fully 
feathered.  The  erupted  feathers  were  mostly 
black,  with  small  red  patches  of  feathers  ap- 


Sachtleben  et  al.  • MICRONESIAN  HONEYEATER  NESTS  ON  SAIPAN 


313 


pearing  on  the  head  and  back.  By  22  March, 
when  the  final  nest  check  was  performed,  this 
nestling  had  fledged. 

On  9 April  2004,  we  found  the  last  suc- 
cessful nest  by  observing  the  female  bringing 
food  to  her  two  nestlings.  The  nestlings  were 
estimated  at  3-3.5  cm  in  length  and  were  al- 
ready developing  pin  feathers.  On  13  April, 
the  nestlings  were  ~4  cm  long,  covered  with 
long,  black  pins  from  which  feathers  had 
erupted,  and  their  eyes  were  open.  Three  days 
later,  the  nestlings  were  4-4.5  cm  long  and 
their  bills  were  visible  over  the  rim  of  the  nest. 
They  were  black  all  over  with  no  red  feathers 
visible.  By  19  April,  the  nestlings  had  fledged. 

Parental  behavior. — Only  females  were  ob- 
served incubating  (n  = 5 nest  checks)  or 
brooding  nestlings  (n  — 1 nest  check).  How- 
ever, one  or  both  members  of  the  pair  were 
often  observed  close  to  the  nest.  When  ob- 
served, the  adult(s)  were  always  very  agitated. 
Typically,  one  or  both  adults  would  feign  in- 
jury, fluttering  about  low  to  the  ground  and 
drooping  one  wing.  If  only  one  adult  was  pre- 
sent, this  behavior  was  sometimes  accompa- 
nied by  scolding;  if  both  adults  were  present, 
one  adult  would  often  feign  injury  while  the 
other  scolded.  We  observed  injury-feigning 
behavior  on  9 of  26  nest  visits  and  scolding 
during  5 of  26;  this  behavior  was  observed 
only  at  nests  containing  nestlings.  Microne- 
sian  Honeyeaters  appeared  very  intolerant  of 
disturbance  at  the  nest  during  the  incubation 
stage,  as  each  time  the  incubating  female  was 
flushed  from  the  nest  during  a nest  check  ( n 
= 3),  the  nest  had  failed  by  the  next  visit. 

DISCUSSION 

Prior  to  our  study,  nests  of  Micronesian 
Honeyeaters  had  been  found  on  Guam  (Har- 
tert  1898,  Seale  1901,  Yamashina  1932,  Jen- 
kins 1983;  N.  Drahos  pers.  comm.),  Rota  (C. 
C.  Kessler  unpubl.  data),  Kosrae  and  Pohnpei 
(Baker  1951),  Chuuk  (Baker  1951,  Brandt 
1962),  Palau  (Pratt  et  al.  1980),  and  in  the 
southwest  Pacific  region  (Mayr  1945).  The 
amount  of  information  provided  varies  by 
source.  Nest  measurements  are  variable,  with 
the  following  ranges  reported  from  Guam:  cup 
height  25-50  mm,  outer  height  50-120  mm, 
internal  diameter  25-60  mm,  and  external  di- 
ameter 35-80  mm  (Hartert  1898,  Seale  1901, 
Jenkins  1983;  N.  Drahos  pers.  comm.).  The 


measurements  of  nests  we  found  on  Saipan 
fall  within  these  ranges.  In  contrast,  the  av- 
erage outer  height  of  18  nests  found  on  Chuuk 
was  20  mm,  considerably  shorter  than  nests 
from  Guam  and  Saipan,  although  the  average 
external  diameter  was  similar  (50  mm;  Brandt 
1962).  Our  nest  heights  are  also  similar  to 
those  reported  from  other  islands,  varying 
from  1.2  to  4.6  m (Hartert  1898,  Seale  1901, 
Yamashina  1932,  Mayr  1945,  Brandt  1962, 
Jenkins  1983;  N.  Drahos  pers.  comm.,  C.  C. 
Kessler  unpubl.  data). 

Similar  to  our  descriptions  of  nests  found 
on  Saipan,  nests  from  Guam,  Rota,  Chuuk, 
and  Palau  have  been  variously  described  as 
“loosely  constructed,”  “fragile,”  “frail,” 
“not  heavily  made,”  and  having  see-through 
sides  (Brandt  1962,  Pratt  et  al.  1980,  Jenkins 
1983;  C.  C.  Kessler  unpubl.  data).  In  addition, 
they  were  found  placed  among  the  outer 
branches  of  the  trees  in  which  they  were  con- 
structed (Seale  1901,  Brandt  1962,  Pratt  et  al. 
1980,  Jenkins  1983).  Unlike  the  nests  we 
found  on  Saipan,  however,  those  on  other  is- 
lands tended  to  be  found  in  open  locations, 
such  as  the  edges  of  clearings  or  the  outer 
perimeters  of  forests  (Brandt  1962,  Pratt  et  al. 
1980;  C.  C.  Kessler  unpubl.  data).  Reported 
nesting  materials  are  diverse  and  include  fine 
roots  and  fibers,  grasses,  leaves,  ferns,  weed 
stems,  and  pieces  of  coconut  bast  (Mayr  1945, 
Baker  1951,  Brandt  1962).  As  on  Saipan,  Ca- 
suarina  equisetifolia  needles  were  included  in 
nests  found  on  Guam. 

The  chief  difference  between  our  observa- 
tions and  those  of  other  authors  in  the  Mariana 
Islands  is  the  suite  of  tree  species  used  for 
nesting.  On  Saipan,  nests  were  placed  in  Psy- 
chotria  and  Guamia  mariannae  (trees  native 
to  the  Mariana  Islands),  whereas  nests  on 
Guam  were  placed  in  Pithecellobium  dulce, 
Casuarina  equisetifolia , Delonix  regia , and 
Bruguiera  gymnorrhiza,  only  two  of  which 
(C.  equisetifolia  and  B.  gymnorrhiza ) are  in- 
digenous to  the  Mariana  Islands  (Raulerson 
and  Rinehart  1991).  On  Rota,  the  nest  was 
found  in  nonnative  Acacia  confusa.  This  dif- 
ference is  likely  a reflection  of  other  authors 
working  primarily  in  habitats  that  were  dif- 
ferent from  those  in  which  we  worked  (only 
one  of  our  study  areas  comprised  mixed  native 
and  exotic  forest),  rather  than  differences  in 
honeyeater  habitat  use  among  islands. 


314 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


All  reported  clutch  sizes  are  of  one  or  two 
eggs,  although  a nest  found  on  Palau  con- 
tained three  nestlings  (Pratt  et  al.  1980).  Two- 
to  three-egg  clutches  are  characteristic  of  the 
Meliphagidae  family  (Mayr  1945).  Microne- 
sian  Honeyeater  eggs  from  Saipan,  Guam, 
Rota,  and  Chuuk  all  had  a base  color  of  white, 
off-white,  or  cream,  generally  with  rufous- 
brown  speckling,  although  Yamashina  (1932) 
described  the  speckling  as  gray  and  dark  yel- 
low-brown. The  speckling  may  be  concentrat- 
ed at  the  broader  end  (Hartert  1898,  Seale 
1901,  Brandt  1962,  Jenkins  1983),  near  the 
narrow  end  (Yamashina  1932),  near  both  ends 
(this  study),  or  may  be  scattered  over  the 
whole  egg  (Brandt  1962). 

We  found  no  comparative  descriptions  of 
nestlings  or  data  on  their  age  at  fledging. 
However,  several  authors  have  described 
fledgling  Micronesian  Honeyeaters  from 
Guam.  Seale  (1901:57)  reported  that  “.  . . the 
young  are  olive  brown  above,  yellowish  on 
the  under  parts,  washed  with  red  on  the  sides 
of  the  fore  breast  and  back;  bill  dark,  yellow- 
ish on  the  base  of  lower  mandible;  feet  and 
iris  dark.”  N.  Drahos  (pers.  comm.)  described 
a pair  of  fledgling  Micronesian  Honeyeaters 
recently  out  of  the  nest.  The  female  was 
mouse  gray  with  a faintly  rusty-red  chin,  her 
bill  was  black  with  a yellow  stripe  on  its  edge 
and  the  top  of  her  bill  was  yellow  at  the  base, 
and  her  eyes  and  feet  were  black.  He  reported 
that  the  male  was  similar,  but  the  middle  of 
the  back,  chin,  and  lower  half  of  the  head 
were  faintly  cardinal  red.  Other  authors’  de- 
scriptions are  similar  although  less  compre- 
hensive. There  are  several  dissimilarities 
among  our  descriptions  of  nestlings  from  dif- 
ferent nests,  and  between  our  descriptions  of 
nestlings  and  those  of  other  authors.  The  for- 
mer may  be  explained  by  factors  that  could 
affect  nestling  development,  including  the 
number  of  nestlings  present  in  the  nest  (thus, 
whether  provisioning  must  be  shared),  breed- 
ing experience  or  foraging  ability  of  the 
adults,  or  food  availability  in  different  study 
areas.  The  latter  presumably  is  explained  by 
continued  plumage  development  after  fledg- 
ing. Although  our  sample  size  included  only 
two  nests,  Micronesian  Honeyeater  nestlings 
seem  apt  to  leap  from  the  nest  before  they  are 
fully  ready  to  fledge,  which,  under  undis- 
turbed conditions,  seems  to  be  at  15-16  days. 


Parental  distraction  displays  of  Micronesian 
Honeyeaters  on  Saipan  appear  to  be  the  same 
as  those  of  birds  on  Guam  and  Rota,  although 
on  Guam  and  Rota  only  females  have  been 
reported  to  feign  injury  (Stophlet  1946,  Jen- 
kins 1983;  N.  Drahos  pers.  comm.). 

Three  of  the  seven  nests  we  found  on  Sai- 
pan were  in  native  limestone  forest,  which  has 
not  previously  been  reported  as  preferred  hab- 
itat for  the  Micronesian  Honeyeater;  the  spe- 
cies has  been  considered  more  common  in  co- 
conut plantings,  shrubbery  and  gardens  of  vil- 
lages, and  diverse  second-growth  forest.  Sim- 
ilarly, Cardinal  Honeyeaters  ( Myzomela 
cardinalis ) in  Samoa  are  most  abundant  in  vil- 
lage habitats  (Freifeld  1999),  and  Orange- 
breasted Honeyeaters  ( Myzomela  jugularis ) in 
Fiji  are  most  abundant  in  coconut  plantations 
(Steadman  and  Franklin  2000).  This  under- 
scores the  importance  of  obtaining  ecological 
information  for  all  native  species  to  further  the 
development  of  conservation  plans.  Some  of 
the  habitats  in  which  Micronesian  Honeyeat- 
ers are  reportedly  common,  such  as  backyard 
gardens,  would  appear  unsuitable  as  nesting 
habitat,  given  this  species’  apparent  intoler- 
ance of  disturbance  at  the  nest  and  the  likeli- 
hood of  disturbance  in  these  areas. 

Overall,  we  found  that  Micronesian  Hon- 
eyeaters on  Saipan  have  nesting  requirements 
and  behaviors  similar  to  those  on  Guam  prior 
to  their  extirpation.  Information  on  the  nesting 
requirements  of  Micronesian  Honeyeaters  on 
Saipan  should  aid  in  the  establishment  of  ef- 
fective captive  breeding  programs  for  this  spe- 
cies, and  for  future  re-establishment  on  Guam 
and  Saipan  (if  necessary)  once  brown  tree- 
snakes  have  been  controlled  or  eradicated. 

ACKNOWLEDGMENTS 

This  study  was  funded  by  the  U.S.  Fish  and  Wildlife 
Service  (USFWS)  Region  1 Office,  Portland,  Oregon. 
We  extend  special  thanks  to  the  USFWS  Marianas 
Team,  in  particular  H.  B.  Freifeld  and  A.  P Marshall; 
the  Commonwealth  of  the  Northern  Mariana  Islands 
Division  of  Fish  and  Wildlife,  especially  J.  B.  de  Cruz, 
L.  Williams,  S.  Kremer,  and  N.  B.  Hawley;  J.  Quitano 
for  allowing  us  access  to  As  Teo;  N.  Johnson  and  S. 
Mosher  for  helpful  field  information;  and  S.  Hopken 
for  translating  the  Yamashina  (1932)  paper.  We  appre- 
ciate the  constructive  comments  made  on  earlier  drafts 
of  the  manuscript  by  J.  B.  de  Cruz,  A.  B.  Franklin,  R. 
A.  Hufbauer,  R.  J.  Craig,  and  two  anonymous  referees. 


Sachtleben  et  al.  • MICRONESIAN  HONEYEATER  NESTS  ON  SAIPAN 


315 


LITERATURE  CITED 

Baker,  R.  H.  1951.  The  avifauna  of  Micronesia:  its 
origin,  evolution,  and  distribution.  University  of 
Kansas  Museum  of  Natural  History,  Lawrence. 

Brandt,  J.  H.  1962.  Nests  and  eggs  of  the  birds  of  the 
Truk  Islands.  Condor  64:416-437. 

Buckland,  S.  T.,  D.  R.  Anderson,  K.  P.  Burnham,  J. 
L.  Laake,  D.  L.  Borchers,  and  L.  Thomas.  2001. 
Introduction  to  distance  sampling:  estimating 
abundance  of  biological  populations.  Oxford  Uni- 
versity Press,  Oxford,  United  Kingdom. 

Colvin,  B.  A.,  M.  W.  Fall,  L.  A.  Fitzgerald,  and  L. 

L.  Loop.  2005.  Review  of  brown  treesnake  prob- 
lems and  control  programs:  report  of  observations 
and  recommendations.  Report  to  the  Office  of  In- 
sular Affairs,  Honolulu,  Hawaii. 

Craig,  R.  J.  1989.  Observations  on  the  foraging  ecol- 
ogy and  social  behavior  of  the  Bridled  White-eye. 
Condor  91:187-192. 

Craig,  R.  J.  1990.  Foraging  behavior  and  microhabitat 
use  of  two  species  of  white-eyes  (Zosteropidae) 
on  Saipan,  Micronesia.  Auk  107:500-505. 

Craig,  R.  J.  1996.  Seasonal  population  surveys  and 
natural  history  of  a Micronesian  bird  community. 
Wilson  Bulletin  108:246-267. 

Engbring,  J.,  F.  L.  Ramsey,  and  V.  L.  Wildman.  1986. 
Micronesian  forest  bird  survey,  1982:  Saipan,  Ti- 
nian, Agiguan,  and  Rota.  U.S.  Fish  and  Wildlife 
Service,  Honolulu,  Hawaii. 

Falanruw,  M.  C.,  T.  G.  Cole,  and  A.  H.  Ambacher. 
1989.  Vegetation  survey  of  Rota,  Tinian,  and  Sai- 
pan, Commonwealth  of  the  Northern  Mariana  Is- 
lands. Resource  Bulletin  PSW-27,  USDA  Forest 
Service,  Pacific  Southwest  Forest  and  Range  Ex- 
periment Station,  Berkeley,  California. 

Freifeld,  H.  B.  1999.  Habitat  relationships  of  forest 
birds  on  Tutuila  Island,  American  Samoa.  Journal 
of  Biogeography  26:1191-1213. 

Gragg,  J.  E.  2004.  Rodent  reduction  for  enhanced 
control  of  brown  treesnakes  ( Boiga  irregularis). 

M. Sc.  thesis,  Colorado  State  University,  Fort  Col- 
lins. 

Hartert,  E.  1898.  On  the  birds  of  the  Marianne  Is- 
lands. Novitates  Zoologicae  V:51-69. 

Jenkins,  J.  M.  1983.  The  native  forest  birds  of  Guam. 
Ornithological  Monographs,  no.  31. 

Jenkins,  J.  M.  and  C.  Aguon.  1981.  Status  of  candi- 
date endangered  bird  species  on  Saipan,  Tinian 
and  Rota  of  the  Mariana  Islands.  Micronesica  17: 
184-186. 

Mayr,  E.  1945.  Birds  of  the  southwest  Pacific.  Mac- 
Millan Co.,  New  York. 

Mosher,  S.  M.  and  S.  G.  Fancy.  2002.  Description  of 
nests,  eggs,  and  nestlings  of  the  endangered 
Nightingale  Reed- Warbler  on  Saipan,  Micronesia. 
Wilson  Bulletin  114:1-10. 

Mueller-Dombois,  D.  and  F.  R.  Fosberg.  1998.  Veg- 


etation of  the  tropical  Pacific  Islands.  Springer- 
Verlag,  New  York. 

Pratt,  H.  D.,  P.  L.  Bruner,  and  D.  G.  Berrett.  1979. 
America’s  unknown  avifauna:  the  birds  of  the 
Mariana  Islands.  American  Birds  33:227-235. 

Pratt,  H.  D.,  P.  L.  Bruner,  and  D.  G.  Berrett.  1987. 
A field  guide  to  the  birds  of  Hawaii  and  the  trop- 
ical Pacific.  Princeton  University  Press,  Princeton, 
New  Jersey. 

Pratt,  H.  D.,  J.  Engbring,  P.  L.  Bruner,  and  D.  G. 
Berrett.  1980.  Notes  on  the  taxonomy,  natural 
history,  and  status  of  the  resident  birds  of  Palau. 
Condor  82:117-131. 

Ralph,  C.  J.  and  H.  F.  Sakai.  1979.  Forest  bird  and 
fruit  bat  populations  and  their  conservation  in  Mi- 
cronesia: notes  on  a survey.  ‘Elepaio  40:21-26. 

Raulerson,  L.  and  A.  Rinehart.  1991.  Trees  and 
shrubs  of  the  Northern  Mariana  Islands.  Coastal 
Resources  Management,  Saipan,  Northern  Mari- 
ana Islands. 

Rodda,  G.  H.,  E.  W.  Campbell,  and  S.  R.  Derrick- 
son.  1998.  Avian  conservation  research  in  the 
Mariana  Islands,  western  Pacific  Ocean.  Pages 
367-381  in  Avian  conservation:  research  and 
management  (J.  M.  Marzluff  and  R.  Sallabanks, 
Eds.).  Island  Press,  Washington,  D.C. 

Safford,  W.  E.  1902.  The  birds  of  the  Marianne  Is- 
lands and  their  vernacular  names — I.  Osprey  1: 
39-70. 

Savidge,  J.  A.  1987.  Extinction  of  an  island  forest  avi- 
fauna by  an  introduced  snake.  Ecology  68:660- 
668. 

Seale,  A.  1901.  Report  of  a mission  to  Guam,  part  I. 
Avifauna.  Bernice  P.  Bishop  Museum  Occasional 
Papers  1:17-60. 

Steadman,  D.  W.  and  J.  Franklin.  2000.  A prelimi- 
nary survey  of  landbirds  on  Lakeba,  Lau  Group, 
Fiji.  Emu  100:227-235. 

Stinson,  C.  M.  and  D.  W.  Stinson.  1994.  Nest  sites, 
clutch  size  and  incubation  behavior  in  the  Golden 
White-eye.  Journal  of  Field  Ornithology  65:65- 
69. 

Stophlet,  J.  J.  1946.  Birds  of  Guam.  Auk  65:534-540. 

U.S.  Fish  and  Wildlife  Service.  1997.  Draft  1997  for- 
est bird  survey,  Saipan,  CNMI:  Nightingale  Reed 
Warbler  ( Acrocephalus  luscinia)  assessment.  U.S. 
Fish  and  Wildlife  Service,  Honolulu,  Hawaii. 

Wiles,  G.  J.,  J.  Bart,  R.  E.  Beck,  Jr.,  and  C.  F. 
Aguon.  2003.  Impacts  of  the  brown  tree  snake: 
patterns  of  decline  and  species  persistence  in 
Guam’s  avifauna.  Conservation  Biology  17:1350- 
1360. 

Yamashina,  Y.  1932.  On  a collection  of  birds’  eggs 
from  Micronesia.  Tori  7:393-413. 

Young,  F.  J.  1989.  Soil  survey  of  the  islands  of  Agui- 
jan,  Rota,  Saipan,  and  Tinian,  Commonwealth  of 
the  Northern  Mariana  Islands.  U.S.  Department  of 
Agriculture,  Soil  Conservation  Service,  Washing- 
ton, D.C. 


The  Wilson  Journal  of  Ornithology  1 18(3):3 16-325,  2006 


WITHIN-PAIR  INTERACTIONS  AND  PARENTAL  BEHAVIOR  OF 
CERULEAN  WARBLERS  BREEDING  IN  EASTERN  ONTARIO 

JENNIFER  J.  BARG,1 2 3 4  JASON  JONES,1  24  M.  KATHARINE  GIRVAN,1 3 AND 

RALEIGH  J.  ROBERTSON1 


ABSTRACT. — The  Cerulean  Warbler  ( Dendroica  cerulea)  is  currently  the  focus  of  considerable  management 
interest;  however,  our  ability  to  develop  effective  management  strategies  is  hampered  by  a dearth  of  life  history 
and  basic  behavioral  data.  Here,  we  present  information  on  male-female  interactions  of  Cerulean  Warblers  and 
parental  nest  attentiveness  that  is,  to  our  knowledge,  among  the  first  such  rigorously  collected  data  for  this 
species.  Males  feed  females  during  nest  building  and  on  the  nest  during  incubation;  the  relative  infrequency  of 
these  events  suggests  that  they  play  more  of  a role  in  pair-bond  maintenance  than  they  do  in  enhancing  female 
energetics.  Female  incubation  rhythms  were  not  significantly  influenced  by  temperature,  time  of  day,  or  egg  age. 
Compared  with  other  Dendroica  warblers,  we  observed  relatively  infrequent  female  departures  during  incubation, 
perhaps  in  response  to  a high  risk  of  nest  predation.  As  the  nestlings  aged,  females  spent  less  time  brooding 
nestlings,  presumably  to  allow  for  more  frequent  feeding;  however,  both  males  and  females  exhibited  relatively 
low  rates  of  food  delivery  compared  with  other  Dendroica  warblers.  Despite  the  low  rates  of  food  delivery, 
feeding  trips  were  more  frequent  at  successful  nests  than  unsuccessful  nests.  Our  results  suggest  that  Cerulean 
Warblers  are  tightly  constrained  by  the  competing  pressures  of  predation  risk  and  sufficient  food  provisioning 
for  nestlings.  Received  28  February  2005,  accepted  23  February  2006. 


Birds  that  form  socially  monogamous  pairs 
during  the  breeding  season  exhibit  various 
acoustic  (Kroodsma  and  Miller  1996)  and  be- 
havioral (Birkhead  and  Mpller  1992)  within- 
pair  interactions.  These  social  behaviors  can 
have  conservation  and  management  implica- 
tions; indeed,  our  ability  to  manage  or  con- 
serve species  of  interest  is  often  unwittingly 
limited  by  our  poor  understanding  of  basic  life 
history  and  behavioral  phenomena  (Komdeur 
and  Deerenberg  1997).  Hopefully,  the  careful 
documentation  of  these  behaviors  will  assist 
us  in  identifying  species’  social  requirements, 
which  may  be  used  to  augment  management 
and  conservation  strategies  based  on  habitat 
requirements.  The  Cerulean  Warbler  ( Den- 
droica cerulea ) is  a poorly  known  species  of 
particular  concern  due  to  population  declines 
of  up  to  3%  per  year  since  1966  (North  Amer- 
ican Breeding  Bird  Survey  data;  Robbins  et 
al.  1992,  Link  and  Sauer  2002),  probably  due 
to  habitat  loss  in  both  North  America  and 
South  America.  In  the  United  States,  the  spe- 
cies has  been  variously  designated  as  threat- 


1  Dept,  of  Biology,  Queen’s  Univ.,  Kingston,  ON 
K7L  3N6,  Canada. 

2 Current  address:  Dept,  of  Biology,  Vassar  College, 
Poughkeepsie,  NY  1 2604,  USA. 

3 Current  address:  Norval  Outdoor  School,  Box  226, 
Norval,  ON  LOP  1 K0,  Canada. 

4 Corresponding  author;  e-mail:  jajones@vassar.edu 


ened,  rare,  or  of  special  concern;  in  Canada, 
it  is  a species  of  special  concern  (Robbins  et 
al.  1992,  Hamel  2000,  Committee  on  the  Sta- 
tus of  Endangered  Wildlife  in  Canada  2003); 
and  it  is  listed  as  vulnerable  by  the  Interna- 
tional Union  for  Conservation  of  Nature  and 
Natural  Resources  (2004).  However,  the  de- 
sign and  implementation  of  effective  conser- 
vation and  management  strategies  has  been 
slowed  by  limited  availability  of  life  history 
and  behavioral  data  (Hamel  et  al.  2004). 

As  a result  of  long-term  research,  beginning 
in  1994  at  the  Queen’s  University  Biological 
Station  (QUBS)  in  Ontario,  Canada,  we  have 
learned  a great  deal  about  habitat  selection  be- 
havior (Jones  et  al.  2001 ; Jones  and  Robertson 
2001;  Barg  et  al.  2005,  2006),  reproductive 
ecology  and  population  dynamics  (Oliarnyk 
and  Robertson  1996,  Jones  et  al.  2004),  and 
population  structure  (Gibb  et  al.  2005,  Jones 
et  al.  2005,  Veit  et  al.  2005)  for  the  enigmatic 
Cerulean  Warbler.  Here,  we  present  data  on 
Cerulean  Warbler  male-female  interactions 
and  parental  nest  attentiveness  that  is,  to  our 
knowledge,  among  the  first  such  rigorously 
collected  data  for  this  species.  Specifically,  we 
were  interested  in  how  males  and  females  co- 
ordinate reproductive  activities,  how  they  di- 
vide parental  responsibilities,  and  how  pat- 
terns of  nest  attendance  were  influenced  by 
weather  variables,  partner  behavior,  and  nest- 
ing stage. 


316 


Barg  et  al.  • CERULEAN  WARBLER  PARENTAL  BEHAVIOR 


317 


METHODS 

We  collected  data  during  the  breeding  sea- 
sons (May-July)  of  1999-2001,  at  QUBS, 
Lake  Opinicon,  Leeds/Frontenac  counties, 
Ontario,  Canada  (44°  30'  N,  76°  20'  W).  The 
forest  there  is  characterized  as  second  growth 
deciduous,  between  80  and  90  years  old.  The 
canopy  is  dominated  by  sugar  maple  (. Acer 
saccharum),  bitternut  hickory  {Cary a cordi- 
formis ),  and  ash  {Fraxinus  spp.);  the  mid- 
and  understories  are  primarily  hophornbeam 
(known  as  ironwood  in  Canada;  Ostrya  vir- 
giniana ) and  sugar  maple  saplings.  We  used 
microclimate  data  loggers  (Onset  HOBO®  H8 
Pro  Series  data  loggers,  Bourne,  Massachu- 
setts) to  record  temperature  and  relative  hu- 
midity hourly  at  two  separate  locations  within 
the  study  site,  which  was  a 24-ha  area  on 
QUBS  property. 

Each  year,  we  captured  territorial  males  by 
using  target-netting  techniques  (whereby  a 
mist  net  was  erected  in  a male’s  territory  and 
a conspecific  playback  and  model  presentation 
were  placed  nearby  to  attract  the  male  towards 
the  net).  We  banded  all  males  with  unique 
combinations  of  color  bands  and  a Canadian 
Wildlife  Service  band.  Females  were  more 
difficult  to  capture,  as  they  were  largely  un- 
responsive to  playbacks;  thus,  we  attempted 
other  methods,  including  chickadee  mobbing 
calls,  hoop  nets  placed  at  nests,  and  owl  calls 
with  presentations  of  owl  models,  to  capture 
females.  The  few  females  we  did  catch  (also 
banded)  were  captured  opportunistically  when 
they  were  visiting  water  sources,  feeding 
fledglings  low  in  the  canopy,  collecting  nest- 
ing material,  or  flushed  off  nests  low  in  the 
canopy. 

The  Cerulean  Warbler’s  breeding  season  in 
Ontario  is  approximately  60-75  days.  Over 
the  course  of  our  long-term  study  (1996— 
2001;  201  nests),  we  determined  that  nest 
building  takes  4-7  days,  egg  laying  <7  days, 
and  incubation  10-12  days;  the  nestling  stage 
lasts  10-11  days.  The  female  does  all  the  in- 
cubating and  brooding,  and  both  males  and 
females  feed  the  young.  Nests  were  checked 
every  2-3  days.  Mirrors  attached  to  telescop- 
ing poles  were  used  to  see  into  the  nests;  if  a 
nest  could  not  be  reached  with  the  mirrors,  we 
used  parental  activities,  such  as  departure  fre- 


quency, food  delivery,  or  fecal  sac  removal, 
to  assess  nesting  status. 

We  classified  nests  that  fledged  at  least  one 
young  as  successful.  As  the  high  location  of 
nests  made  it  difficult  to  determine  their  fates 
precisely,  we  combined  all  unsuccessful  nests 
for  analyses,  whether  they  had  succumbed  to 
predation,  exposure,  abandonment,  or  some 
unknown  cause.  We  hired  a professional  tree- 
climber  to  access  nests  during  the  nestling 
stages  in  2000-2001.  On  average,  it  took  >3 
hr  per  nest  to  access  and  process  the  nestlings. 
Mean  brood  size  in  the  nine  nests  that  we  ac- 
cessed was  3.3  nestlings  (range  = 3—4). 

To  document  parental  behavior  and  within- 
pair  interactions,  we  performed  a series  of  fo- 
cal nest  watches  in  1999-2001.  For  each 
watch,  a single  observer  monitored  activity  at 
a nest  for  30  min.  Female  presence  or  absence 
at  the  nest  was  recorded  every  minute.  The 
observer  also  kept  a running  tally  of  depar- 
ture/arrival times,  male  and  female  vocaliza- 
tions, male  visits  to  the  nest,  and  feeding  trips 
made  by  the  male  and  the  female — docu- 
menting the  food  item  whenever  possible. 
Given  our  inability  to  access  most  nests,  we 
were  not  able  to  calculate  provisioning  rates 
on  a “per  nestling”  basis,  which  would  have 
allowed  us  to  control  for  any  potential  effects 
of  brood  size  on  provisioning  rates.  Nest 
watches  were  performed  on  individual  nests 
at  2-  to  3-day  intervals  until  the  nestlings 
fledged  or  the  nest  failed;  nest  status  was 
monitored  between  watches.  Where  nest  vis- 
ibility permitted,  we  videotaped  nests  for  2-hr 
periods;  this  allowed  us  to  assess  the  bout 
length  of  incubation  and  brooding  without  the 
30-min  time  constraint  of  focal-nest  watches. 
To  increase  our  nest-watch  sample  size,  we 
included  the  first  30  min  of  each  video  re- 
cording in  our  analyses;  there  were  no  signif- 
icant differences  in  the  patterns  of  incubation 
and  brooding  between  our  focal  nest  watches 
and  the  first  30  min  of  our  video  recordings 
(all  P > 0.20).  No  nest  was  watched  or  vid- 
eotaped more  than  once  on  any  given  day. 

Analysis. — We  used  analysis  of  covariance 
(ANCOVA)  to  analyze  incubation  patterns 
based  on  130  watches  (117  direct,  13  video) 
from  39  nests  and  31  females  conducted  dur- 
ing 1999-2001;  this  included  nests  of  females 
that  renested  {n  = 7).  Fixed  effects  in  the  AN- 
COVA models  were  time  of  day  and  day  of 


318 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


incubation,  with  ambient  temperature  included 
as  a covariate.  Because  we  performed  multiple 
watches  on  each  female,  “individual”  was  in- 
cluded in  the  model  as  a random  effect.  To 
control  for  seasonal  effects  (Julian  date  was 
significantly  correlated  with  ambient  temper- 
ature; r = 0.45.  P < 0.001),  we  regressed  time 
spent  incubating  per  30-min  watch  on  Julian 
date  and  used  the  residuals  from  this  linear 
regression  as  the  response  variable  in  the  AN- 
COVA  model. 

We  used  ANCOVA  to  analyze  brooding 
patterns  based  on  135  watches  (111  direct,  24 
video)  from  40  nests  and  35  females  during 
1999-2001.  Fixed  effects  in  the  ANCOVA 
models  were  time  of  day  and  nestling  age.  As 
in  the  incubation  models,  we  included  “indi- 
vidual” as  a random  effect.  We  conducted 
separate  analyses  for  two  covariates:  ambient 
temperature  and  male  feeding  rates.  For  the 
temperature  model,  we  used  the  residuals 
from  a regression  of  time  spent  brooding  on 
Julian  date  as  our  response  variable.  For  the 
male-feeding  model,  the  response  variable 
was  the  time  spent  brooding  per  30-min  watch 
(untransformed).  In  our  analysis  of  male  feed- 
ing rates,  we  only  included  2000-2001  data 
(77  watches,  31  nests,  25  females).  We  had  to 
exclude  1999  male  feeding  rate  data  due  to 
consistent  observer  bias  detected  in  that  year; 
one  field  assistant  neglected  to  consistently  re- 
cord whether  or  not  a male  was  carrying  food 
upon  arrival  at  the  nest.  We  also  used  AN- 
COVA models  to  examine  the  effect  of  am- 
bient temperature  and  male  feeding  rate  on  the 
number  of  feeding  trips  made  by  females.  As 
in  the  incubation  and  brooding  models,  we  in- 
cluded “individual”  as  a random  effect.  Male 
feeding  rate  data  were  excluded. 

We  performed  /-tests  to  compare  time  spent 
incubating  and  brooding,  and  the  number  of 
feeding  trips  (per  30-min  watch)  at  successful 
versus  unsuccessful  nests.  There  was  no  sta- 
tistically significant  difference  between  the 
average  timing  (defined  by  incubation  day)  of 
watches  on  successful  (mean  incubation  day 
of  watches  = 7.3  ± 0.4)  and  unsuccessful 
(mean  = 7.2  ± 0.4)  nests  (/  = 0.14.  df  = 128, 
P = 0.89).  In  addition  to  nest  success  (i.e., 
whether  or  not  a nest  fledged  at  least  one 
young),  we  also  included  an  analysis  of  sur- 
vival by  nesting  stage  (i.e.,  whether  or  not  a 
nest  survived  the  incubation  period)  because 


parental  activity  during  the  incubation  phase 
is  known  to  affect  nest  success  (Martin  and 
Ghalambor  1999,  Ghalambor  and  Martin 
2002).  No  nest  watches  were  performed  on 
unsuccessful  nests  after  day  10  of  the  brood- 
ing period;  therefore,  all  watches  conducted 
after  day  10  at  successful  nests  were  excluded 
from  our  analysis  of  parental  behavior.  In  this 
restricted  data  set,  there  was  no  statistically 
significant  difference  between  the  average 
timing  (defined  by  brooding  day)  of  watches 
on  successful  (mean  brooding  day  of  watches 
= 5.4  ± 0.4)  and  unsuccessful  (mean  = 5.0 
± 0.1)  nests  (f  = 1.77,  df  = 104,  P = 0.08). 
Data  are  presented  as  untransformed  means  ± 
SE.  All  statistical  analyses  were  performed  us- 
ing JMPIN  (ver.  4.0.2;  SAS  Institute,  Inc. 
2000). 

RESULTS  AND  DISCUSSION 

Reciprocal  vocalizations. — We  documented 
136  instances  of  reciprocal  vocalizations 
(male  vocalization  followed  immediately  by 
female  call)  during  the  study  period.  In  the 
context  of  reciprocal  vocalizations,  males 
were  more  likely  to  sing  quiet  songs  (whisper 
songs)  during  nest  building  than  during  the 
other  stages  of  the  nesting  cycle  (nest  build- 
ing: 62%  of  reciprocal  vocalizations;  incuba- 
tion: 18%;  brooding:  24%;  x2  = 23.09,  df  = 
2,  P < 0.001).  When  females  are  nest  build- 
ing, males  tend  to  follow  very  closely  (often 
within  1-2  m)  and  regularly  sing  whisper 
songs  directed  at  the  female  (JJB  pers.  obs.). 
Presumably,  this  following  behavior  during 
the  fertile  period  is  a form  of  mate  guarding, 
while  the  whisper  singing  with  occasional  fe- 
male response  presumably  functions  in  pair- 
bond maintenance.  Our  observations  of  male 
whisper  singing  during  nest  building  are  sim- 
ilar to  John  and  Kermott’s  (1991)  observations 
of  the  House  Wren  ( Troglodytes  aedon );  whis- 
per singing  by  male  House  Wrens  also  may 
serve  to  stimulate  ovulation  in  the  females 
(Johnson  and  Kermott  1991).  Interestingly, 
male  Cerulean  Warblers  would  frequently 
whisper  sing  while  females  inspected  potential 
nest  sites;  males  would  usually  inspect  these 
same  sites  immediately  thereafter  (JJB  pers. 
obs.).  Males  were  rarely  heard  whisper  sing- 
ing away  from  the  female  or  the  nest  (Barg  et 
al.  2005).  Whisper  singing  by  males  in  similar 


Barg  et  al.  • CERULEAN  WARBLER  PARENTAL  BEHAVIOR 


319 


contexts  has  been  observed  in  other  parts  of 
the  breeding  range  (Rogers  2006). 

Nearly  two-thirds  (63%)  of  the  reciprocal 
observations  occurred  during  the  incubation 
stage,  although  the  function  of  reciprocal  vo- 
calizations while  the  female  is  incubating  is 
unclear.  One  possibility  was  that  male  vocal- 
izations signal  an  “all-clear”  for  females  to 
leave  the  nest;  however,  this  was  not  support- 
ed by  our  data,  despite  our  expectations  based 
on  anecdotal  observation  prior  to  data  collec- 
tion. The  frequency  of  male  whisper  songs 
versus  normal  songs  did  not  influence  whether 
or  not  a female  stayed  on  the  nest  following 
the  reciprocal  vocalization  (Fisher’s  exact  test, 
P = 0.45).  Future  research  should  be  designed 
to  test  a second  possibility,  that  a female  re- 
sponse to  a male  vocalization  may  encourage 
male  care  (Halkin  1997). 

Females  regularly  chip  (without  prompting 
by  male  song)  when  departing  the  nest  for  an 
off-bout  (approximately  50%  of  departures; 
JJB  pers.  obs.),  possibly  as  a signal  to  males 
that  the  nest  is  unprotected  (e.g.,  Barber  et  al. 
1998).  During  a survey  of  15  songbird  species 
in  which  females  gave  nest-departure  calls, 
McDonald  and  Greenberg  (1991)  reported 
that,  unlike  the  Cerulean  Warbler,  most  of  the 
species  inhabit  grassy  or  shrubby  habitats  and 
that  the  calls  appear  to  reduce  male  activity  at 
the  nest,  presumably  to  reduce  the  risk  of  pre- 
dation. Male  Cerulean  Warblers  frequently  at- 
tended the  nest  for  the  duration  of  the  female’s 
off-bout,  sitting  quietly  <2  m from  the  nest  in 
the  nest  tree;  sometimes  the  male  perched  on 
the  edge  of  the  nest  but  was  never  observed 
sitting  on  the  nest  (i.e.,  no  incubating  or 
brooding)  during  our  watches.  Apparently, 
males  of  other  species  are  also  known  to  ex- 
hibit nest  vigilance  during  female  absences 
(e.g..  Northern  Mockingbird,  Mimus  poly- 
glottos ; Breitwisch  et  al.  1989). 

Mate  feeding  and  mate  quality. — We  made 
28  observations  of  males  feeding  females  (i.e., 
courtship  feeding)  during  nest  building.  Over 
half  (n  = 15)  of  these  feeding  events  were 
followed  by  copulations.  In  all  cases,  the  food 
item  presented  was  a larval  lepidopteran. 
Thirty-five  percent  of  the  males  (16  of  46) 
also  were  observed  feeding  incubating  fe- 
males (mean  = 0.70  ± 0.06  feedings/hr). 

Originally,  mate  feeding  was  hypothesized 
to  strengthen  pair  bonds  (Lack  1940)  or  to 


serve  as  an  index  of  mate  quality — thereby 
influencing  future  mate  choice  (Nisbet  1973). 
More  recently,  researchers  have  shown  that 
mate  feeding  can  represent  an  important  nu- 
tritive and  energetic  contribution  to  the  female 
(Royama  1966;  Lyon  and  Montgomerie  1985, 
1987;  Hatchwell  et  al.  1999)  and  may  com- 
pensate for  poor-quality  territories  (Lifjeld  and 
Slagsvold  1986).  Finally,  mate  feeding  may 
serve  to  reduce  the  incidence  of  brood  para- 
sitism by  Brown-headed  Cowbirds  ( Molothrus 
ater ),  presumably  by  reducing  female  activity 
and  keeping  her  on  the  nest;  this  advantage, 
however,  may  carry  the  cost  of  increased  nest 
predation  resulting  from  greater  levels  of  male 
activity  at  the  nest  (Tewksbury  et  al.  2002). 

The  hypotheses  regarding  nutrition  and  en- 
ergetics are  unlikely  candidates  for  explaining 
mate  feeding  among  Cerulean  Warblers,  pri- 
marily because  their  relative  frequency  of 
mate  feeding  is  low  (less  than  one  visit  per 
observation  hr);  however,  it  is  not  clear  how 
frequent  mate  feeding  must  be  before  it  sig- 
nificantly affects  female  condition.  Assessing 
the  potential  selection  pressure  of  brood  par- 
asitism on  mate  feeding  requires  feeding  data 
from  nests  that  were  parasitized;  however,  de- 
spite a high  density  of  cowbirds  in  the  region 
(JJ  unpubl.  data),  we  have  never  observed  Ce- 
rulean Warbler  parents  feeding  cowbird  nest- 
lings or  fledglings.  Furthermore,  since  1994 
we  have  detected  cowbird  eggs  in  only  two 
Cerulean  Warbler  nests,  both  of  which  were 
abandoned. 

We  have  made  several  observations  that  of- 
fer indirect  support  for  the  notion  that  female 
Cerulean  Warblers  are  capable  of  assessing 
mate  quality  and  potentially  basing  their  mate- 
choice  decisions  on  those  assessments.  First, 
we  witnessed  extra-pair  copulations  by  band- 
ed individuals  and,  for  the  two  complete  fam- 
ilies for  which  we  obtained  blood  samples  (on 
a separate  project),  >50%  (4/7)  of  young  were 
sired  by  a male  other  than  the  social  mate  (JJB 
unpubl.  data).  The  criteria  female  Cerulean 
Warblers  use  to  choose  extra-pair  mates  are 
unknown,  but  presumably  they  involve  judg- 
ments of  male  quality.  Second,  we  observed 
an  instance  of  double  brooding  (i.e.,  initiation 
of  a second  nest  following  a successful  first 
nest).  Double  brooding  may  occur  more  fre- 
quently, but  our  difficulty  in  capturing  females 
limits  our  understanding  of  certain  reproduc- 


320 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  1.  Incubation  patterns  (n  = 130  focal  nest  watches)  of  female  Cerulean  Warblers  at  the  Queen’s 
University  Biological  Station,  eastern  Ontario,  1999-2001,  were  not  affected  by  time  of  day,  incubation  day,  or 
ambient  temperature.  During  the  nestling  stage  (n  = 135  focal  nest  watches),  females  spent  less  time  brooding 
as  nestlings  aged.  No  interactions  were  statistically  significant  (all  P > 0.10)  in  these  ANCOVA  models.  Boldface 
values  denote  significant  model  effects.  The  male  feeding-rate  model  is  based  on  2000-2001  data  only. 

Source  of  variation 

Mean  square 

df 

F 

p 

Incubation  patterns  ( R 2 = 0.38) 

Time  of  day 

2.03 

1 

0.24 

0.62 

Incubation  day 

9.65 

13 

1.15 

0.32 

Ambient  temperature  (covariate) 

1 1.59 

1 

1.39 

0.24 

Individual  female 

8.29 

30 

0.99 

0.49 

Error 

8.36 

84 

Brooding  patterns 

Temperature  as  covariate  ( R 2 = 0.57) 

Time  of  day 

12.18 

1 

0.30 

0.58 

Nestling  age 

160.84 

13 

4.02 

<0.001 

Ambient  temperature 

11.36 

1 

0.28 

0.60 

Individual  female 

51.63 

34 

1.29 

0.18 

Error 

40.06 

85 

Male  feeding  rate  as  covariate  ( R 2 = 0.58) 

Time  of  day 

2.53 

1 

0.07 

0.80 

Nestling  age 

113.67 

1 1 

3.00 

0.006 

Male  feeding  rate 

22.09 

1 

0.56 

0.57 

Individual  female 

35.22 

24 

0.93 

0.57 

Error 

37.92 

39 

tive  behaviors.  What  makes  this  single  obser- 
vation germane  is  that  this  female  was  the  sec- 
ondary female  of  a bigamous  male,  who  pro- 
vided very  little  parental  care  to  her  first 
brood;  once  her  fledglings  were  sufficiently 
mobile,  the  female  moved  the  brood  —800  m 
(the  width  of  four  territories)  and  re-mated 
with  a different  male  (all  birds  were  banded). 
The  female’s  choice  of  a second  mate  ap- 
peared to  be  based  on  this  male’s  willingness 
to  provide  parental  care  to  her  fledglings, 
something  not  offered  by  her  first  mate.  This 
second  male  “adopted”  her  brood  by  feeding 
the  young  while  the  female  built  a new  nest 
and  laid  a clutch  of  five  eggs  (this  second 
nesting  attempt  was  unsuccessful).  Although 
this  is  the  first  documented  case  of  brood 
adoption  in  Cerulean  Warblers,  it  has  been 
documented  occasionally  in  other  wood  war- 
blers (e.g..  Hooded  Warbler,  Wilsonia  citrina\ 
Evans  Ogden  and  Stutchbury  1994).  Interest- 
ingly, the  double-brooded  female’s  new  mate 
already  had  an  active  nest  and  his  primary  fe- 
male was  incubating  at  the  time  of  brood 
adoption.  Bigamy  is  uncommon  but  regular 


on  our  study  site  (—10%  of  breeding  males 
are  bigamous;  JJB  pers.  obs.). 

Incubation  patterns. — On  average,  females 
spent  25.7  ± 0.27  min  incubating  and  made 
1.0  ± 0.1  departures  (range  = 0-2)  per  30- 
min  watch.  For  all  females  (including  those 
recorded  on  videotape),  the  average  (contin- 
uous) duration  of  an  incubation  bout  was  32.6 
± 3.5  min.  After  removing  the  effect  of  Julian 
day,  the  duration  of  incubation  bouts  was  not 
significantly  influenced  by  time  of  day,  incu- 
bation day,  or  ambient  temperature  (Table  1). 
We  detected  no  differences  in  incubation  time 
between  successful  (i.e.,  surviving  incubation 
or  fledging  at  least  one  young)  and  unsuc- 
cessful nests  (incubation:  t — 1.19,  df  = 128, 
P = 0.24;  fledging:  t = 0.089,  df  - 128,  P = 
0.93;  Fig.  1A). 

Incubating  females  are  faced  with  two  de- 
cisions, the  outcomes  of  which  largely  define 
incubation  rhythms  (Reid  et  al.  1999).  The 
first  decision — when  to  leave — is  linked  to  fe- 
male energy  levels.  The  second — when  to  re- 
turn— is  linked  to  female  foraging  efficiency. 
In  other  words,  on-bout  duration  is  linked  to 


Barg  et  al.  • CERULEAN  WARBLER  PARENTAL  BEHAVIOR 


321 


Survived  Fledged  Fledged 

stage  young  young 


Incubation  Brooding 


FIG.  1.  Cerulean  Warbler  on-bout  duration  (A) 
and  feeding  behavior  (B)  for  successful  (filled  bars) 
and  unsuccessful  (unfilled  bars)  nests.  Queen’s  Uni- 
versity Biological  Station,  eastern  Ontario.  For  the  in- 
cubation period,  we  defined  success  in  two  ways:  first, 
whether  or  not  the  clutch  hatched,  and,  second,  wheth- 
er or  not  at  least  1 young  fledged  from  the  nest.  For 
the  brooding  period,  success  was  defined  by  whether 
or  not  at  least  1 young  fledged  from  the  nest.  Data  for 
female  on-bout  duration  and  female  feeding  trips  are 
from  1999  to  2001.  Feeding  trip  data  for  male  and 
sexes-combined  are  from  2000  to  2001.  Values  pre- 
sented are  means  ± 1 SE  with  sample  size  inside  each 
column.  Brooding  sample  size  is  higher  than  incuba- 
tion sample  size  as  we  included  nests  that  were  found 
after  the  eggs  had  hatched.  Results  of  /-tests:  NS  = 
not  significant,  * = P < 0.05. 


parental  needs  as  much  as  it  is  to  embryonic 
needs  (Conway  and  Martin  2000a,  b).  That  we 
detected  no  significant  effect  of  ambient  tem- 
perature on  incubation  patterns  implies  either 
(a)  that  the  thermal  needs  of  embryos  were 
met  by  ambient  temperatures  (Webb  1987)  on 
our  study  site,  thereby  releasing  female  be- 
havior from  this  constraint  during  the  day,  or 


(b)  that  female  behavior  was  constrained  by 
other  pressures,  such  as  female  condition, 
male  behavior,  or  predation  risk.  Compared 
with  other  Dendroica  warblers  (Conway  and 
Martin  2000b),  we  observed  relatively  infre- 
quent female  departures  during  incubation 
(Table  2).  Given  the  lack  of  a significant  re- 
lationship between  incubation  rhythms  and 
temperature,  this  low  frequency  of  nest  de- 
partures may  be  indicative  of  a high  risk  of 
predation  (Martin  and  Ghalambor  1999,  Ghal- 
ambor  and  Martin  2002).  Nest  predation  is 
likely  the  primary  cause  of  nest  failure  on  our 
study  site  (Jones  et  al.  2001),  with  Blue  Jays 
( Cyanocitta  cristata ) being  the  primary  predator 
(JJB  pers.  obs.);  however,  given  the  inaccessi- 
bility of  most  of  our  nests,  we  were  unable  to 
examine  the  contents  of  most  abandoned  nests 
to  help  confirm  the  cause  of  failure. 

Brooding  and  feeding  young. — Females 
spent  20.1  ± 7.84  min  brooding  and  made  1.6 
± 0.2  departures  (range  = 0-3)  per  30-min 
watch.  For  all  females  (including  those  re- 
corded on  videotape),  the  average  (continu- 
ous) duration  of  brooding  bouts  was  16.2  ± 
1.5  min.  In  both  brooding  models  (Table  1), 
females  tended  to  brood  less  as  nestlings  aged, 
but  time  of  year,  temperature,  and  male  feed- 
ing rate  had  no  significant  effect.  We  detected 
no  differences  in  time  spent  brooding  for  suc- 
cessful versus  unsuccessful  nests  ( t = 1.63,  df 
= 104,  P = 0.1  1;  Fig.  1A). 

Both  males  and  females  averaged  1.1  ±0.1 
feeding  trips  per  30-min  watch  (range:  fe- 
males - 0-3,  males  = 0-4).  Females  fed 
more  frequently  as  nestlings  aged  and  as  male 
feeding  rate  increased  (Table  3),  corroborating 
the  findings  in  previous  studies  (e.g.,  Nolan 
1978,  Conrad  and  Robertson  1993,  Lozano 
and  Lemon  1998,  MacColl  and  Hatchwell 
2003).  Males  (t  = 2.40,  df  = 68  P = 0.019) 
but  not  females  (/  = 0.85,  df  = 93,  P = 0.40; 
Fig.  IB)  fed  nestlings  more  often  at  successful 
nests  than  at  unsuccessful  nests.  Adults  (both 
sexes  combined)  at  successful  nests  made  ap- 
proximately twice  as  many  feeding  trips  per 
30-min  watch  as  they  did  at  unsuccessful  nests 
( t = 2.12,  df  = 68,  P = 0.038;  “Both”  in  Fig. 
IB).  While  we  have  no  direct  evidence  that 
differences  in  food-delivery  rates  were  re- 
sponsible for  differences  in  nest  success,  a dif- 
ference of  1 trip  per  30-min  watch  is  larger 
than  it  first  appears.  If  we  assume  a 15-hr  day. 


322 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  2.  Parental  behavior  of  Dendroica  wood  warblers  of  northeastern  North  America.  A dash  indicates 
behaviors  for  which  we  could  find  no  published  information.  Very  few  quantitative  estimates  of  mate  feeding 
are  available;  therefore,  we  adopted  the  qualitative  classification  of  Conway  and  Martin  (2000b). 


Species 

Nest 

location 

Incubation- 
bout  length 
(min) 

No.  incubation 
departures 
(/hr) 

Male 

incubation 

feeding 

Nestling 
provisioning 
rate  (/nest/hr) 

Source 

Bay-breasted 
Warbler  ( D . 

Tree 

18 

5.5 

Moderate 

26 

Griscom  1938, 
Williams  1996 

castanea ) 

Blackburnian 
Warbler  (D. 
fused) 

Tree 

21-22 

4.2 

Infrequent 

Kendeigh  1945, 
Lawrence  1953, 
Morse  2004 

Blackpoll  War- 
bler (D.  stria- 
ta) 

Tree 

19 

5.0 

Moderate 

3/nestling/hr 

Bent  1953,  Hunt 
and  Eliason 
1999 

Black-throated 
Blue  Warbler 
( D . caerules- 

Shrub 

20-31 

2.9 

Moderate 

7 

Kendeigh  1945, 
Holmes  et  al. 
2005 

cens) 

Black- throated 
Green  Warbler 
(D.  virens) 

Tree 

50 

1.9 

12-14 

Nice  and  Nice 
1932a,  b;  Morse 
and  Poole  2005 

Cerulean  Warbler 

Tree 

33 

2.0 

Infrequent 

3-4 

This  study 

( D . cerulea) 

Chestnut-sided 
Warbler  ( D . 
pensylvanica) 

Shrub 

23 

4.5 

Moderate 

8 

Kendeigh  1945, 
Lawrence  1948, 
Tate  1970, 
Richardson  and 
Brauning  1995, 
Hanski  et  al. 
1996 

Magnolia  Warbler 

Tree 

17 

4.9 

— 

8 

Hall  1994 

( D . magnolia) 

Yellow  Warbler 
(D.  petechia) 

Shrub 

36 

3.1 

Frequent 

Kendeigh  1945, 
Hanski  et  al. 
1996,  Goosen 
and  Sealy  1982, 
Martin  et  al. 
2000 

Yellow-rumped 
Warbler  ( D . 
coronata) 

Tree 

8-10 

Martin  et  al.  2000, 
Hunt  and  Flash- 
poler  1998 

1 caterpillar/trip,  0.1  g/caterpillar,  a 10-day 
nestling  period,  and  1 extra  trip/30  min,  par- 
ents at  successful  nests  would  have  delivered 
approximately  30  g more  food  to  nestlings 
than  unsuccessful  parents. 

Because  increased  parental  activity  late  in 
the  nestling  stage  tends  to  increase  predation 
risk  (Martin  et  al.  2000),  we  find  it  surprising 
that  parents  at  successful  nests  made  more 
feeding  trips  than  parents  at  unsuccessful 
nests;  however.  Cerulean  Warblers  feed  nest- 
lings at  relatively  low  rates  compared  to  other 
passerines  (Martin  et  al.  2000;  Table  2),  which 
might  lessen  the  predation  resulting  from  in- 


creased activity.  Taken  together,  our  observa- 
tions— male  incubation  feeding,  low  rates  of 
female  departure,  low  rates  of  food  delivery, 
and  the  possible  link  between  food  provision- 
ing and  nesting  success — suggest  that  Ceru- 
lean Warblers  are  tightly  constrained  by  the 
competing  pressures  of  predation  risk  and 
food  provisioning. 

ACKNOWLEDGMENTS 

D.  M.  Aiama,  R.  D.  DeBruyn,  S.  Harding,  B.  Risk, 
A.  J.  Stevens,  and  J.  Vargas  provided  field  assistance. 
Three  anonymous  reviewers  made  valuable  contribu- 
tions to  the  manuscript.  The  Queen’s  University  Bio- 


Barg  et  al.  • CERULEAN  WARBLER  PARENTAL  BEHAVIOR 


323 


TABLE  3.  Female  Cerulean  Warblers  (tempera- 
ture ANCOVA:  n = 135  focal  nest  watches;  male 
feeding-rate  ANCOVA:  n = 77)  at  the  Queen’s  Uni- 
versity Biological  Station,  eastern  Ontario,  1999- 
2001,  fed  nestlings  more  as  nestling  aged  and  as  their 
social  mates  fed  more.  No  interactions  were  statisti- 
cally significant  (all  P > 0.10).  Boldface  values  denote 
significant  model  effects.  The  male  feeding-rate  model 
is  based  on  2000-2001  data  only. 


Source  of  variation 

Mean 

square 

df 

F 

p 

Temperature  as  covariate  ( R 2 = 

0.30) 

Time  of  day 

0.19 

1 

0.19 

0.67 

Nestling  age 

2.77 

12 

2.77 

0.004 

Ambient  temperature 

0.05 

1 

0.05 

0.82 

Individual  female 

0.23 

34 

0.23 

0.99 

Error 

1.00 

68 

Male  feeding  rate  as  covariate  ( R 2 = 

0.43) 

Time  of  day 

0.68 

1 

0.81 

0.37 

Nestling  age 

1.68 

11 

2.01 

0.051 

Male  feeding  rate 

7.52 

1 

9.01 

0.005 

Individual  female 

0.39 

24 

0.47 

0.98 

Error 

0.84 

39 

logical  Station  provided  valuable  logistical  support. 
Wildlife  Habitat  Canada,  the  Eastern  Ontario  Model 
Forest  Program,  Natural  Science  and  Engineering  Re- 
search Council  of  Canada,  and  World  Wildlife  Fund  of 
Canada  (MacNaughton  Conservation  Scholarships, 
Endangered  Species  Recovery  Fund),  Queen’s  Univer- 
sity, the  Society  of  Canadian  Ornithologists,  and  the 
American  Ornithologists’  Union  provided  financial 
support.  This  project  is  part  of  Natural  Legacy  2000, 
a nationwide  initiative  in  Canada  to  conserve  wildlife 
in  private  and  public  habitats.  We  gratefully  acknowl- 
edge the  support  of  the  Government  of  Canada’s  Mil- 
lennium Partnership  Fund. 

LITERATURE  CITED 

Barber,  P.  M.,  T.  E.  Martin,  and  K.  G.  Smith.  1998. 
Pair  interactions  in  Red-faced  Warblers.  Condor 
100:512-518. 

Barg,  J.  J.,  D.  M.  Aiama,  J.  Jones,  and  R.  J.  Robert- 
son. 2006.  Within-territory  habitat  use  and  micro- 
habitat selection  by  male  Cerulean  Warblers  ( Den - 
droica  cerulea).  Auk  123:In  press. 

Barg,  J.  J.,  J.  Jones,  and  R.  J.  Robertson.  2005.  De- 
scribing breeding  territories  of  migratory  passer- 
ines: suggestions  for  sampling,  choice  of  estima- 
tor, and  delineation  of  core  areas.  Journal  of  An- 
imal Ecology  74:139-149. 

Bent,  A.  C.  1953.  Black-polled  Warbler.  Pages  389- 
408  in  Life  histories  of  North  American  wood 
warblers.  U.S.  National  Museum  Bulletin,  no. 
203.  [Reprinted  1963,  Dover  Publications,  New 
York.] 

Birkhead,  T.  R.  and  A.  P.  M0ller.  1992.  Sperm  com- 


petition in  birds:  evolutionary  causes  and  conse- 
quences. Academic  Press,  London,  United  King- 
dom. 

Breitwisch,  R.,  N.  Gottlieb,  and  J.  Zaias.  1989.  Be- 
havioral differences  in  nest  visits  between  male 
and  female  Northern  Mockingbirds.  Auk  106: 
659-665. 

Committee  on  the  Status  of  Endangered  Wildlife 
in  Canada.  2003.  COSEWIC  assessment  and  up- 
date status  report  on  the  Cerulean  Warbler  Den- 
droica  cerulea  in  Canada.  Committee  on  the  Sta- 
tus of  Endangered  Wildlife  in  Canada,  Ottawa, 
Ontario. 

Conrad,  K.  F.  and  R.  J.  Robertson.  1993.  Patterns  of 
parental  provisioning  by  Eastern  Phoebes.  Condor 
95:57-62. 

Conway,  C.  J.  and  T.  E.  Martin.  2000a.  Effects  of 
ambient  temperature  on  avian  incubation  behavior. 
Behavioral  Ecology  1 1:178-188. 

Conway,  C.  J.  and  T.  E.  Martin.  2000b.  Evolution  of 
passerine  incubation  behavior:  influence  of  food, 
temperature,  and  nest  predation.  Evolution  54: 
670-685. 

Evans  Ogden,  L.  J.  and  B.  J.  Stutchbury.  1994. 
Hooded  Warbler  ( Wilsonia  citrina).  The  Birds  of 
North  America,  no.  110. 

Ghalambor,  C.  K.  and  T.  E.  Martin.  2002.  Compar- 
ative manipulation  of  predation  risk  in  incubating 
birds  reveals  variability  in  the  plasticity  of  re- 
sponses. Behavioral  Ecology  13:101-109. 

Gibb,  C.  E.,  J.  Jones,  M.  K.  Girvan,  J.  J.  Barg,  and 
R.  J.  Robertson.  2005.  Geographic  variation  in 
prevalence  and  parasitemia  with  Haemoproteus 
paruli  in  the  Cerulean  Warbler  ( Dendroica  ceru- 
lea). Canadian  Journal  of  Zoology  86:626-629. 

Goosen,  J.  P.  and  S.  G.  Sealy.  1982.  Production  of 
young  in  a dense  nesting  population  of  Yellow 
Warblers,  Dendroica  petechia,  in  Manitoba.  Ca- 
nadian Field-Naturalist  96:189-199. 

Griscom,  L.  1938.  The  birds  of  the  Lake  Umbagog 
region  of  Maine:  compiled  from  the  diaries  and 
journals  of  William  Brewster.  Bulletin  of  the  Mu- 
seum of  Comparative  Zoology,  part  4,  no.  66. 
Harvard  University,  Cambridge,  Massachusetts. 

Halkin,  S.  L.  1997.  Nest-vicinity  song  exchanges  may 
coordinate  biparental  care  of  Northern  Cardinals. 
Animal  Behaviour  54:189-198. 

Hall,  G.  A.  1994.  Magnolia  Warbler  ( Dendroica  mag- 
nolia). The  Birds  of  North  America,  no.  136. 

Hamel,  P.  B.  2000.  Cerulean  Warbler  (. Dendroica  cer- 
ulea). The  Birds  of  North  America,  no.  511. 

Hamel,  P.  B.,  D.  K.  Dawson,  and  P.  D.  Keyser.  2004. 
How  we  can  learn  more  about  the  Cerulean  War- 
bler ( Dendroica  cerulea).  Auk  121:7—14. 

Hanksi,  I.  K.,  T.  J.  Fenske,  and  G.  J.  Niemi.  1996. 
Lack  of  edge  effect  in  nesting  success  of  breeding 
birds  in  managed  forest  landscapes.  Auk  1 13:578- 
585. 

Hatchwell,  B.  J.,  M.  K.  Fowlie,  D.  J.  Ross,  and  A. 
F.  Russell.  1999.  Incubation  behavior  of  Long- 


324 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


tailed  Tits:  why  do  males  provision  incubating  fe- 
males? Condor  101:681-686. 

Holmes,  R.  T.,  N.  L.  Rodenhouse,  and  T.  S.  Sillett. 
2005.  Black-throated  Blue  Warbler  ( Dendroica 
caerulescens).  The  Birds  of  North  America,  no. 
87.  http://bna.birds.cornell.edu/BNA/account/ 
Black-throated_Blue_Warbler/  (accessed  Decem- 
ber 2005). 

Hunt,  R D.  and  B.  C.  Eliason.  1999.  Blackpoll  War- 
bler (. Dendroica  striata ).  The  Birds  of  North 
America,  no.  431. 

Hunt,  P.  D.  and  D.  J.  Flaspohler.  1998.  Yellow-rum- 
ped  Warbler  ( Dendroica  coronata).  The  Birds  of 
North  America,  no.  376. 

International  Union  for  Conservation  of  Nature 
and  Natural  Resources  (IUCN).  2004.  Den- 
droica cerulea.  2004  IUCN  Red  List  of  Threat- 
ened Species,  http://www.iucnredlist.org  (ac- 
cessed December  2005). 

Johnson,  L.  S.  and  L.  H.  Kermott.  1991.  The  func- 
tions of  song  in  male  House  Wrens  ( Troglodytes 
aedon ).  Behaviour  116:190-208. 

Jones,  J.,  J.  J.  Barg,  T.  S.  Sillett,  M.  L.  Veit,  and 
R.  J.  Robertson.  2004.  Minimum  estimates  of 
survival  and  population  growth  for  Cerulean  War- 
blers breeding  in  Ontario,  Canada.  Auk  121:15— 
22. 

Jones,  J.,  R.  D.  DeBruyn,  J.  J.  Barg,  and  R.  J.  Rob- 
ertson. 2001.  Assessing  the  effects  of  natural  dis- 
turbance on  a Neotropical  migrant  songbird.  Ecol- 
ogy 82:2628-2635. 

Jones,  J.,  C.  E.  Gibb,  S.  C.  Millard,  J.  J.  Barg,  M. 
K.  Girvan,  M.  L.  Veit,  V.  L.  Friesen,  and  R.  J. 
Robertson.  2005.  Multiple  selection  pressures 
generate  adherence  to  Bergmann’s  Rule  in  a Neo- 
tropical migrant  songbird.  Journal  of  Biogeogra- 
phy 32:1827-1833. 

Jones,  J.  and  R.  J.  Robertson.  2001.  Territory  and 
nest-site  selection  of  Cerulean  Warblers  in  eastern 
Ontario.  Auk  1 18:727-735. 

Kendeigh,  S.  C.  1945.  Nesting  behavior  of  wood  war- 
blers. Wilson  Bulletin  57:145-167. 

Komdeur,  J.  and  C.  Deerenberg.  1997.  The  impor- 
tance of  social  behavior  studies  for  conservation. 
Pages  262-276  in  Behavioral  approaches  to  con- 
servation in  the  wild  (J.  R.  Clemmons  and  R. 
Buchholz,  Eds.).  Cambridge  University  Press, 
New  York. 

Kroodsma,  D.  E.  and  E.  H.  Miller  (Eds.).  1996. 
Ecology  and  evolution  of  acoustic  communication 
in  birds.  Comstock  Publishing,  Ithaca,  New  York. 

Lack,  D.  1940.  Courtship  feeding  in  birds.  Auk  54: 
169-178. 

Lawrence,  L.  de  Kiriline.  1948.  Comparative  study 
of  the  nesting  behavior  of  Chestnut-sided  and 
Nashville  warblers.  Auk  65:204-219. 

Lawrence,  L.  de  Kiriline.  1953.  Notes  on  the  nesting 
behavior  of  the  Blackburnian  Warbler.  Wilson 
Bulletin  65:135-144. 

Lifjeld,  J.  T.  and  T.  Slagsvold.  1986.  The  function 
of  courtship  feeding  during  incubation  in  the  Pied 


Flycatcher  Ficedula  hypoleuca.  Animal  Behaviour 
34:1441-1453. 

Link,  W.  A.  and  J.  R.  Sauer.  2002.  A hierarchical 
analysis  of  population  change  with  application  to 
Cerulean  Warblers.  Ecology  83:2832-2840. 

Lozano,  G.  A.  and  R.  E.  Lemon.  1998.  Parental-care 
responses  by  Yellow  Warblers  ( Dendroica  pete- 
chia) to  simultaneous  manipulations  by  food 
abundance  and  brood  size.  Canadian  Journal  of 
Zoology  76:916-924. 

Lyon,  B.  E.  and  R.  D.  Montgomerie.  1985.  Incuba- 
tion feeding  in  Snow  Buntings:  female  manipu- 
lation or  indirect  male  parental  care?  Behavioral 
Ecology  and  Sociobiology  17:279-284. 

Lyon,  B.  E.  and  R.  D.  Montgomerie.  1987.  Ecolog- 
ical correlates  of  incubation  feeding:  a compara- 
tive study  of  high  arctic  finches.  Ecology  68:713- 
722. 

MacColl,  A.  D.  C.  and  B.  J.  Hatchwell.  2003.  Shar- 
ing of  caring:  nestling  provisioning  behaviour  of 
Long-tailed  Tits,  Aegithalos  caudatus,  parents  and 
helpers.  Animal  Behaviour  66:955-964. 

Martin,  T.  E.  and  C.  K.  Ghalambor.  1999.  Males 
feeding  females  during  incubation.  I.  Required  by 
microclimate  or  constrained  by  nest  predation? 
American  Naturalist  153:131-139. 

Martin,  T.  E.,  J.  Scott,  and  C.  Menge.  2000.  Nest 
predation  increases  with  parental  activity:  sepa- 
rating nest  site  and  parental  activity  effects.  Pro- 
ceedings of  the  Royal  Society  of  London,  Series 
B 267:2287-2293. 

McDonald,  M.  V.  and  R.  Greenberg.  1991.  Nest  de- 
parture calls  in  female  songbirds.  Condor  93:365- 
373. 

Morse,  D.  H.  2004.  Blackburnian  Warbler  ( Dendroica 
fusca).  The  Birds  of  North  America,  no.  102. 
http://bna.birds.cornell.edu/BNA/account/ 
Blackburnian. Warbler  (accessed  December  2005). 

Morse,  D.  H.  and  A.  F.  Poole.  2005.  Black-throated 
Green  Warbler  ( Dendroica  virens).  The  Birds  of 
North  America,  no.  55.  http://bna/birds.cornell. 
edu/BNA/account/Black-throated_Green_  Warbler 
(accessed  December  2005). 

Nice,  M.  M.  and  L.  B.  Nice.  1932a.  A study  of  two 
nests  of  the  Black-throated  Green  Warbler,  part  I. 
Bird-Banding  3:95-105. 

Nice,  M.  M.  and  L.  B.  Nice.  1932b.  A study  of  two 
nests  of  the  Black-throated  Green  Warbler,  part  II. 
Chronicle  of  the  August  nest.  Bird-Banding  3: 
157-172. 

Nisbet,  I.  C.  T.  1973.  Courtship-feeding,  egg  size  and 
breeding  success  in  Common  Terns.  Nature  241: 
141-142. 

Nolan,  V.,  Jr.  1978.  The  ecology  and  behavior  of  the 
Prairie  Warbler  Dendroica  discolor.  Ornithologi- 
cal Monographs,  no.  26. 

Oliarnyk,  C.  J.  and  R.  J.  Robertson.  1996.  Breeding 
behavior  and  reproductive  success  of  Cerulean 
Warblers  in  southeastern  Ontario.  Wilson  Bulletin 
108:673-684. 


Barg  et  al.  • CERULEAN  WARBLER  PARENTAL  BEHAVIOR 


325 


Pitelka,  F.  1940.  Breeding  behavior  of  the  Black- 
throated  Green  Warbler.  Wilson  Bulletin  52:3-18. 

Reid,  J.  M.,  P.  Monaghan,  and  G.  D.  Ruxton.  1999. 
The  effect  of  clutch  cooling  rate  on  starling,  Stur- 
nus  vulgaris,  incubation  strategy.  Animal  Behav- 
iour 58:1161-1167. 

Richardson,  M.  and  D.  W.  Brauning.  1995.  Chest- 
nut-sided Warbler  ( Dendroica  pensylvanica).  The 
Birds  of  North  America,  no.  190. 

Robbins,  C.  S.,  J.  W.  Fitzpatrick,  and  P.  B.  Hamel. 
1992.  A warbler  in  trouble:  Dendroica  cerulea. 
Pages  549-562  in  Ecology  and  conservation  of 
Neotropical  migrant  landbirds  (J.  M.  Hagan,  III, 
and  D.  W.  Johnston,  Eds.).  Smithsonian  Institution 
Press,  Washington,  D.C. 

Rogers,  C.  M.  2006.  Nesting  success  and  breeding 
biology  of  Cerulean  Warblers  in  Michigan.  Wilson 
Journal  of  Ornithology  118:145-151. 

Royama,  T.  1966.  A re-interpretation  of  courtship 
feeding.  Bird  Study  13:116-129. 


SAS  Institute,  Inc.  2000.  JMPIN  V4.0.2.  SAS  Insti- 
tute, Inc.,  Cary,  North  Carolina. 

Tate,  J.  1970.  Nesting  and  development  of  the  Chest- 
nut-sided Warbler.  Jack-Pine  Warbler  48:57-65. 

Tewksbury,  J.  J.,  T.  E.  Martin,  S.  J.  Hejl,  M.  J. 
Kuehn,  and  J.  W.  Jenkins.  2002.  Parental  care  of 
a cowbird  host:  caught  between  the  costs  of  egg- 
removal  and  nest  predation.  Proceedings  of  the 
Royal  Society  of  London,  Series  B 269:423-429. 

Veit,  M.  L.,  R.  J.  Robertson,  P.  B.  Hamel,  and  V.  L. 
Friesen.  2005.  Population  genetic  structure  and 
dispersal  across  a fragmented  landscape  in  Ceru- 
lean Warblers  ( Dendroica  cerulea).  Conservation 
Genetics  6:159-174. 

Webb,  D.  R.  1987.  Thermal  tolerance  of  avian  embry- 
os: a review.  Condor  89:874-898. 

Williams,  J.  M.  1996.  Bay-breasted  Warbler  ( Den- 
droica castanea).  The  Birds  of  North  America,  no. 
206. 


The  Wilson  Journal  of  Ornithology  1 18(3):326-332,  2006 


COMPARATIVE  SPRING  MIGRATION  ARRIVAL  DATES  IN  THE 
TWO  MORPHS  OF  WHITE-THROATED  SPARROW 

SARAH  S.  A.  CALDWELL1 2 AND  ALEXANDER  M.  MILLS1 2 


ABSTRACT. — White-throated  Sparrows  ( Zonotrichia  albicollis ) display  a plumage  dimorphism  ( white-striped 
and  tan-striped)  with  attendant  behavioral  differences,  including  greater  aggression  levels  in  white-striped  birds 
and  negative  assortative  mating,  in  which  tan-striped  birds  pair  with  white-striped  birds.  To  determine  whether 
morph  influences  migration  timing,  which  could  influence  patterns  of  assortative  mating,  we  evaluated  the 
phenology  of  northbound  migration  among  White-throated  Sparrows  from  a long-term  banding  dataset  collected 
at  a southern  Ontario  banding  station.  White-throated  Sparrows  are  sexed  by  wing-chord  length,  but  there  is  an 
intermediate  size  for  which  sex  cannot  be  assigned.  When  all  birds  were  considered  together  (both  known  and 
unknown  sexes,  n = 6,243),  the  white-striped  birds  migrated  earlier  by  slightly  more  than  2 days.  The  sexing 
criteria,  however,  appeared  to  yield  a sample  that  was  not  representative  of  the  whole  population:  when  we 
included  only  birds  for  which  sex  was  assigned  (n  = 2,794,  45%  of  all  birds),  white-striped  birds  apparently 
migrated  earlier  by  more  than  4 days,  but  separate  analyses  of  males  ( n = 1,511)  and  females  (n  = 1,283) 
revealed  no  differences  in  migration  timing  between  morphs.  By  measuring  wing-chord  lengths  of  internally 
sexed  specimens  (from  the  Royal  Ontario  Museum)  collected  during  April  to  June  ( n = 273),  we  found  that  in 
both  sexes  the  wings  of  white-striped  birds  were  about  2%  longer  than  those  of  tan-striped  birds.  When  we  used 
these  specimen  data  to  recalibrate  the  sexing  criteria,  (a)  it  was  possible  to  assign  sex  to  1.47  times  as  many 
birds  (n  = 4,121;  66%  of  all  birds),  (b)  sex  ratios  of  the  banded  birds  more  closely  approached  what  appears 
to  be  the  natural  sex  ratio  (approximately  1:1),  and  (c)  within-sex  analyses  indicated  that  white-striped  females 
migrate  earlier  than  tan-striped  females  by  about  1.3  days,  whereas  there  was  no  statistical  difference  between 
male  morphs  in  migration  timing.  Received  25  April  2005,  accepted  2 February  2006. 


The  White-throated  Sparrow  ( Zonotrichia 
albicollis ) displays  a plumage  dimorphism 
(Lowther  1961)  produced  by  an  inversion  in 
the  second  chromosome  (Thorneycroft  1966). 
The  two  morphs  are  usually  referred  to  as 
white-striped  and  tan-striped.  The  former  has 
a gray  breast  and  a bright  white  median  crown 
stripe  and  supercilium,  while  the  latter  has  a 
brown  breast  and  a dull  or  tan-colored  crown 
stripe  and  supercilium  (Lowther  1961.  Falls 
and  Kopachena  1994).  White-striped  males 
are  slightly  heavier  than  tan-striped  males  and 
white-striped  females,  which  are  heavier  than 
tan-striped  females  (Tuttle  1993).  Thorney- 
croft (1975)  showed  that  the  nestling  sex  ratio 
was  not  significantly  different  from  1:1,  and 
both  morphs  are  represented  nearly  equally  in 
adult  populations  (Falls  and  Kopachena 
1994). 

Ecological  and  behavioral  differences  be- 
tween white-striped  and  tan-striped  morphs 
include  aggression  levels,  preferred  breeding 
habitat,  and  patterns  of  parental  care  (e.g., 
Knapton  and  Falls  1982,  1983:  Knapton  et  al. 


1 Dept,  of  Zoology,  Univ.  of  Toronto,  25  Harbord 
St..  Toronto,  ON  M5S  3G5,  Canada. 

2 Corresponding  author;  e-mail: 
sarah.caldwell@utoronto.ca 


1984;  Kopachena  and  Falls  1993;  Tuttle  1993; 
Falls  and  Kopachena  1994).  In  particular, 
white-striped  males  are  most  aggressive  and 
tan-striped  females  are  least  aggressive  (Ko- 
pachena and  Falls  1993).  Tuttle  (2003)  found 
that,  compared  to  tan-striped  males,  white- 
striped  males  exhibited  higher  rates  of  at- 
tempted polygyny  and  intrusion  into  neigh- 
boring territories,  and  lower  rates  of  parental 
care  and  mate  guarding.  Negative  assortative 
mating  occurs  such  that  >95%  of  pairs  com- 
prise one  bird  of  each  morph  (Lowther  1961, 
Falls  and  Kopachena  1994,  Houtman  and 
Falls  1994).  It  has  been  proposed  that  females 
of  both  morphs  prefer  tan-striped  males,  and 
that  the  negative  assortative  mating  is  facili- 
tated, at  least  in  part,  by  the  ability  of  white- 
striped  females  to  out-compete  tan-striped  fe- 
males for  tan-striped  males  (Houtman  and 
Falls  1994). 

Notwithstanding  the  lack  of  evidence  dem- 
onstrating ratios  that  depart  from  1:1  for  sex 
or  for  morph.  Falls  and  Kopachena  (1994) 
found  unequal  numbers  of  the  two  types  of 
breeding  pair  assortments  in  Algonquin  Park, 
Ontario,  with  nearly  70%  composed  of  white- 
striped  males  and  tan-striped  females.  How- 
ever, in  another  Algonquin  Park  study,  Knap- 


326 


Caldwell  and  Mills  • WHITE-THROATED  SPARROW  ARRIVAL  DATES 


327 


ton  and  Falls  (1982)  found  the  ecological  dis- 
tribution of  tan-striped  males  to  be  much 
broader  than  that  of  white-striped  males.  In 
addition,  there  is  a male  floater  population  that 
includes  an  unknown  proportion  of  both 
morphs  (Falls  and  Kopachena  1994). 

Typical  of  males  in  migrant  passerines, 
male  White-throated  Sparrows  migrate  earlier 
than  females  (Jenkins  and  Cristol  2002).  Con- 
sidering the  higher  aggression  levels  in  both 
the  male  and  female  white-striped  morph,  ear- 
lier arrival  times  of  white-striped  birds  at  their 
breeding  grounds  would  not  be  surprising.  If 
white-striped  males  arrive  before  tan-striped 
males,  they  would  have  first  choice  of  terri- 
tory. If  white-striped  females  arrive  before 
tan-striped  females,  they  would  have  first 
choice  of  males,  allowing  them  to  pair  with 
the  preferred  tan-striped  males.  Thus,  whether 
due  to  differences  in  latitudes  of  wintering 
ranges,  different  departure  dates,  or  different 
rates  of  migration,  timing  of  northbound 
(herein  referred  to  as  “spring”)  migration 
could  represent  one  factor  influencing  nega- 
tive assortative  mating  in  this  species. 

Knapton  et  al.  (1984)  considered  morph  and 
sex  when  comparing  arrival  times  of  White- 
throated  Sparrows  at  breeding  territories  in 
Algonquin  Park.  Their  two-year  study  re- 
vealed no  significant  timing  differences  be- 
tween male  morphs,  but  there  was  an  apparent 
difference  among  females,  whereby  white- 
striped  birds  arrived  before  tan-striped  birds. 
They  were  reluctant  to  conclude  whether 
white-striped  females  were  actually  migrating 
earlier  or  were  merely  detected  earlier  due  to 
either  their  greater  levels  of  aggression  and 
vocal  behavior  or  their  earlier  association  with 
males. 

Here,  we  report  results  of  two  independent, 
but  related,  investigations.  We  began  by  con- 
sidering the  issue  of  morph-specific  migration 
timing.  To  do  this,  we  used  banding  data  from 
a bird  observatory  in  southern  Ontario  to  an- 
alyze passage  dates  of  White-throated  Spar- 
rows during  spring  migration.  We  speculated 
that  the  apparent  earlier  arrival  of  white- 
striped  females  on  the  breeding  grounds  re- 
flects real  differences  in  migration  timing; 
thus,  we  predicted  that  white-striped  females 
pass  through  earlier  than  their  tan-striped 
counterparts.  When  our  results  suggested 
problems  with  the  sexing  criteria  (wing-chord 


length),  we  used  museum  specimens  to  inves- 
tigate size  differences  between  the  two 
morphs  to  propose  new  morph-specific  sexing 
criteria  for  the  species.  With  these  new  rules, 
we  reassigned  sex  to  the  birds  in  the  banding 
data  set  and  then  repeated  the  analyses. 

METHODS 

Banding  dataset. — We  used  White-throated 
Sparrow  banding  data  collected  at  Long  Point 
Bird  Observatory  (LPBO;  42°  35'  N,  80°  15' 
W)  on  Lake  Erie  in  southern  Ontario.  This 
species  breeds  north  of  LPBO,  so  passage 
times  there  were  used  as  a proxy  for  arrival 
times  at  the  nesting  grounds.  Observatory 
mist  nets  were  opened  on  or  near  1 April,  pri- 
or to  the  mid-April  arrival  of  the  first  White- 
throated  Sparrows.  Characteristics  recorded  at 
LPBO  included  wing-chord  length,  morph, 
weight,  sex  (by  wing  chord),  date,  and  bander 
information. 

Morph  data  were  collected  from  1981 
through  1994,  so  we  restricted  our  analysis  to 
that  period.  We  arbitrarily  required  a mini- 
mum of  25  individuals  of  each  sex  per  spring 
migration  to  include  that  year’s  records  in  the 
dataset,  which  reduced  the  dataset  to  6 years 
(1985-1986,  1991-1994).  The  White-throated 
Sparrow  is  dimorphic  at  least  during  spring 
migration  and  breeding  (Atkinson  and  Ralph 
1980,  Falls  and  Kopachena  1994),  which  per- 
mitted morph  assignment  to  85%  of  the  LPBO 
birds.  Even  though  licensed  banders  train  and 
supervise  volunteers,  non-assignment  of 
morph  probably  was  due  to  bander  uncertainty 
in  cases  where  birds  with  more  intermediate 
plumage  were  caught.  Furthermore,  there  may 
be  instances  in  the  datasets  of  incorrect  morph 
assignment,  although  we  think  such  mistakes 
would  be  unlikely  during  spring  migration, 
when  birds  are  in  fresh  plumage. 

Following  convention,  the  sexing  technique 
used  by  banders  at  LPBO  was  based  on  wing- 
chord  length  (to  the  nearest  mm)  of  the  closed, 
unflattened  wing  chord,  as  measured  from  the 
most  anterior  point  of  the  wrist  joint  to  the  tip 
of  the  longest  primary.  Birds  of  both  morphs 
were  sexed  as  male  if  the  wing  chord  was  >74 
mm  and  as  female  if  the  wing  chord  was  <68 
mm.  Birds  with  wing  chords  of  69-73  mm 
were  designated  as  unsexed.  We  used  chi- 
square  analyses  to  determine  whether  the  ratio 
of  males  to  females  in  each  morph  differed 


328 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


from  a 1:1  ratio.  Julian  dates  were  used  for 
passage  dates,  and  we  followed  convention  by 
setting  alpha  levels  at  0.05  and  reporting 
means  as  ± SE. 

Analysis  of  migration  timing. — We  con- 
ducted four  one-way  analysis  of  variance 
(ANOVA)  in  three  analyses  to  determine 
whether  white-striped  and  tan-striped  birds  ar- 
rived at  different  times  and,  if  so,  whether  sex 
was  a factor.  We  used  all  birds  in  the  first  anal- 
ysis, pooling  both  sexed  and  unsexed  birds  ( n 
= 6,243).  In  the  second  analysis,  we  used  only 
sexed  birds,  but  we  pooled  both  sexes  (n  = 
2,794).  In  the  third  analysis,  we  did  not  pool 
sexes  so  that  we  could  examine  migration 
phenology  for  males  (n  = 1,511)  and  for  fe- 
males ( n = 1,283)  separately. 

Re-calibrating  the  sexing  criteria. — Initial 
analyses  (see  below)  indicated  that  using  the 
established  sexing  criteria  would  not  allow  an 
impartial  test  of  differences  in  migration  tim- 
ing between  the  two  morphs.  We  surmised 
that  there  were  slight  size  differences  between 
the  morphs  that  might  be  confounding  the 
analyses.  If  true,  using  the  established  sexing 
criteria  would  result  in  samples  that  were  not 
representative  of  the  population.  Because  fe- 
males are  smaller  than  males,  it  seemed  likely 
that  if  tan-striped  birds  were  smaller  than 
white-striped  birds,  the  sexing  criteria  would 
bias  designations  of  tan  birds  as  female  and 
white  birds  as  male.  Accordingly,  we  inves- 
tigated the  possibility  of  devising  a more  ac- 
curate, morph-specific  sexing  system  by  re- 
calibrating the  sexing  criteria  and  then  re- 
peating the  second  and  third  analyses. 

We  obtained  White-throated  Sparrow  skins 
{n  = 273)  from  the  Royal  Ontario  Museum 
(ROM)  in  Toronto,  Ontario,  Canada,  to  cali- 
brate wing-chord  length  with  sex  and  morph. 
Only  birds  collected  during  spring  (April  to 
June  of  each  year)  were  used,  and  all  speci- 
mens had  been  assigned  sex  based  on  exam- 
ination of  gonads  rather  than  by  wing  chord. 
The  length  of  the  unflattened  wing  chord  was 
measured  three  times  for  each  bird,  resulting 
in  a mean  measurement  (to  the  nearest  mm) 
that  we  used  in  our  analysis.  We  used  ANO- 
VA to  determine  whether  there  was  a within- 
sex  difference  in  wing-chord  length  between 
white-striped  and  tan-striped  birds. 

We  plotted  wing-chord  lengths  of  males  and 
females,  by  morph,  in  a histogram  to  examine 


the  range  in  overlap.  We  assumed  a normal 
distribution  within  each  sex  of  the  ROM  spec- 
imens. By  convention,  we  accepted  a two- 
tailed  alpha  level  of  0.05,  which  allowed  error 
rates  of  2.5%  on  the  upper  end  of  the  females’ 
distribution  and  on  the  lower  end  of  the 
males’  distribution.  These  measurements  were 
used  to  set  new  morph-specific  measurements 
of  wing-chord  length  for  sexing  the  birds. 

To  determine  whether  the  morph-specific 
sexing  criteria  yielded  fewer  unsexed  birds, 
we  used  a one-sample  sign  test  to  compare  the 
tallies  of  male,  female,  and  unsexed  birds  as- 
signed via  the  new  criteria  to  those  assigned 
via  the  established  criteria.  Specifically,  we 
wished  to  see  whether  the  new  criteria  in- 
creased numbers  of  white-striped  females  and 
tan-striped  males.  Chi-square  analysis  was 
used  to  determine  whether  the  ratio  of  males 
to  females  in  each  morph  differed  from  1:1 
after  the  proposed  sexing  criteria  had  been  ap- 
plied to  the  LPBO  dataset.  Once  we  deter- 
mined that  the  morph-specific  sexing  criteria 
were  superior,  as  demonstrated  by  substantial 
increases  in  sample  sizes,  we  applied  them  to 
the  LPBO  data.  Because  we  expected  migra- 
tion passage  to  be  normally  distributed  (Mills 
2005),  we  expected  the  distribution  of  accu- 
mulated percentages  of  migrants  to  be  sig- 
moid; thus,  we  applied  a third-order  polyno- 
mial model  to  our  distributions.  Once  such 
curves  were  estimated  from  the  data,  we  com- 
pared morph  passage  times  by  comparing  re- 
spective areas  under  morph-specific  curves  by 
using  integrals. 

RESULTS 

Migration  phenology  using  the  established 
sexing  criteria. — White-striped  birds  slightly 
outnumbered  tan-striped  birds  in  the  banding 
dataset  (56%  white-striped).  Using  all  banded 
birds  for  which  morph  was  assigned  ( n = 
6,243),  there  was  a significant  difference  in 
the  arrival  times  of  the  two  morphs  (F1624 1 = 
1 19.7,  P < 0.001).  White-striped  birds  arrived 
2.15  days  earlier  than  the  tan-striped  birds 
(white-striped  Cl:  0.25  days;  tan-striped  Cl: 
0.30  days). 

Using  the  established  sexing  criteria,  only 
about  45%  of  the  birds  were  sexed,  and  there 
were  significantly  fewer  white-striped  birds 
sexed  as  females  than  as  males  (n  = 1,561, 
29%  female;  x2  = 279.9,  df  = 1 . P < 0.001) 


Caldwell  and  Mills  • WHITE-THROATED  SPARROW  ARRIVAL  DATES 


329 


TABLE  1.  Number  of  male,  female,  and  unsexed  White-throated  Sparrows  of  both  tan-striped  and  white- 
striped  color  morphs,  identified  according  to  established  and  re-calibrated  sexing  criteria.  Birds  were  captured 
and  banded  at  the  Long  Point  Bird  Observatory  (LPBO),  Long  Point,  Ontario  (6  years:  1985-1986,  1991-1994). 


White-striped  birds  Tan-striped  birds 


Established 

Proposed 

Established 

Proposed 

Sex 

n 

Percent 

n 

Percent 

n 

Percent 

n Percent 

Female 

450 

13.0 

760 

21.9 

833 

30.0 

833  30.0 

Male 

1,111 

32.0 

1,560 

45.0 

400 

14.4 

968  34.9 

Unsexed 

1,909 

55.0 

1,150 

33.1 

1,540 

55.5 

972  35.1 

Total 

3,470 

2,773 

and  significantly  more  tan-striped  birds  sexed 
as  females  than  as  males  ( n = 1,233,  68% 
female;  x2  = 152.1,  df  = 1,  P < 0.001).  Fur- 
thermore, the  apparent  migration  timing  dif- 
ferences between  morphs  were  exaggerated 
when  only  sexed  birds  were  pooled  and  ana- 
lyzed, with  white-striped  birds  apparently  mi- 
grating 4.27  days  earlier  than  the  tan-striped 
birds  {FX2i92  ~ 192.7,  P < 0.001).  Finally, 
when  separate  analyses  were  conducted  for 
males  and  females,  apparent  differences  in  mi- 
gration timing  between  morphs  were  <1  day 
in  both  cases,  and  neither  was  statistically  sig- 
nificant (males:  F,  1509  = 2.71,  P = 0.10;  fe- 
males: Ft  128i  = 3.19,  P = 0.074).  According- 
ly, we  concluded  that  the  sexed  samples  were 
neither  reliable  nor  representative  of  the  pop- 
ulation, and  we  resorted  to  museum  skins  to 
see  whether  more  reliable  sexing  criteria  could 
be  employed. 

Re-calibrating  the  sexing  criteria. — Analy- 
sis of  the  ROM  skins  showed  that  the  wing 
chords  of  white-striped  females  ( n = 46; 
68.93  mm  ± 0.63)  significantly  exceeded 
those  of  tan-striped  females  (n  = 55;  67.61 
mm  ± 0.65)  by  an  average  of  1.32  mm  (F X 99 
= 8.30,  P = 0.005).  The  difference  in  male 
wing-chord  lengths  was  also  significant  (F,  170 
= 25.8,  P < 0.001),  with  those  of  white- 
striped  birds  ( n — 99;  73.31  mm  ± 0.43)  av- 
eraging 1.48  mm  longer  than  those  of  tan- 
striped  birds  (n  = 73;  71.84  mm  ± 0.34).  In 
both  sexes,  the  average  wing-chord  length  of 
white-striped  morphs  was  —2%  greater.  Using 
the  new  sexing  criteria  and  accepting  a 2.5% 
error  rate,  we  determined  that  we  could  not 
assign  sex  to  white-striped  birds  with  wing- 
chord  lengths  of  70-72  mm,  nor  to  those  of 
tan-striped  morphs  with  wing-chord  lengths  of 
69-71  mm. 


When  we  reapplied  the  revised  sexing  cri- 
teria to  the  LPBO  data  and  conducted  a one- 
sample  sign  test  on  the  data,  1.47  times  as 
many  birds  were  sexed,  a significant  increase 
(white-striped:  n = 3,470,  df  = 1,  P < 0.001; 
tan-striped:  n = 2,773,  df  = 1,  P < 0.001). 
In  addition,  sex  ratios  were  less  skewed  for 
both  morphs:  the  percentage  of  females  in- 
creased modestly  among  white-striped  birds 
(29%  to  33%)  and  decreased  dramatically 
among  tan-striped  birds  (68%  to  46%;  Table 
1).  In  both  morphs,  however,  sex  ratios  still 
differed  from  a 1:1  ratio  (white-striped:  n = 
2,320,  x2  = 275.9,  df  = 1,  P < 0.001;  tan- 
striped:  n — 1,801,  x2  = 10.1,  df  = 1,  P = 
0.001). 

Using  the  new  sexing  criteria,  we  repeated 
the  second  ANOVA  by  pooling  males  and  fe- 
males for  both  white-  {n  = 2,320)  and  tan- 
striped  ( n = 1,801)  morphs  and  comparing 
phenologies  by  morph.  White-striped  birds 
passed  LPBO  2.06  days  earlier  than  tan- 
striped  birds  (F14119  = 67.7,  P < 0.001).  Ac- 
cordingly, we  concluded  that  the  samples 
sexed  by  using  the  new  sexing  criteria  were 
representative  of  the  whole  population,  be- 
cause 2.06  days  (calculated  using  only  sexed 
birds)  is  very  close  to  2.15  days  (calculated 
using  all  birds)  and  substantially  different 
from  the  4.27-day  difference  in  migration  tim- 
ing (calculated  using  only  birds  sexed  with  the 
established  sexing  criteria). 

Migration  phenology  using  the  re-calibrat- 
ed sexing  criteria.- — Being  satisfied  with  the 
new  sexing  criteria,  we  repeated  the  third 
analysis  by  comparing  the  within-sex  passage 
dates  for  both  morphs.  Progression  of  the 
spring  passage  for  the  four  sex-morph  classes 
of  White-throated  Sparrow  at  LPBO  is  shown 
in  Figure  1 . As  expected,  third-order  polyno- 


330 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


FIG.  1 . Progression  of  spring  (northbound)  migration  among  sexes  (females:  the  two  upper  curves;  males: 
the  two  lower  curves)  and  color  morphs  of  White-throated  Sparrows  caught  and  banded  at  Long  Point  Bird 
Observatory,  Long  Point,  Ontario.  Birds  were  identified  on  the  basis  of  sex  and  morph  using  re-calibrated  sexing 
criteria  (see  text).  The  curves  represent  3rd-order  polynomials  that  describe  the  timing  of  each  group’s  passage 
( R 2 values  range  from  0.96  to  0.99).  For  each  curve,  bar  widths  represent  the  proportion  of  birds  passing  through 
on  each  particular  Julian  date.  Compared  with  tan-striped  females,  passage  was  significantly  earlier  for  white- 
striped  females  ( n = 1,593,  F,  1591  = 13.8,  P < 0.001)  by  about  1.3  days;  there  was  no  difference  in  arrival  time 
of  male  morphs  ( n = 2,528,  F,  252 6 = 2.25,  P = 0.13).  Tan-striped  females  took  7%  longer  than  white-striped 
females  to  complete  their  migration  (see  text). 

mials  described  the  migration  timing  well, 
with  the  four  R2  values  ranging  from  0.96  to 
0.99.  Using  the  1st  day  of  female  migration 
as  time  zero  and  calculating  the  areas  under 
each  such  curve  by  using  integrals,  tan-striped 
females  took  7%  longer  than  white-striped  fe- 
males to  complete  their  migration.  On  aver- 
age, this  amounted  to  a significantly  later  ar- 
rival (1.3  days,  n = 1,593,  F,  1591  = 13.8,  P < 

0.001).  Likewise,  the  passage  of  tan-striped 
males  was  2.6%  longer  than  that  of  white- 
striped  males  ( n = 2528,  F,  2526  = 2.25,  P = 

0.13). 

DISCUSSION 

Several  studies  of  aggression  levels  among 
white-striped  and  tan-striped  morphs  in 
White-throated  Sparrows  revealed  that  both 
sexes  of  the  white-striped  morph  appear  to  be 
more  aggressive  than  their  tan-striped  coun- 
terparts (e.g..  Watt  et  al.  1984,  Kopachena  and 
Falls  1993,  Collins  and  Houtman  1999).  To 


this  body  of  knowledge  we  add  the  observa- 
tion that  white-striped  females  arrive  at  the 
breeding  grounds  earlier  than  tan-striped  fe- 
males. Our  results  are  consistent — for  both 
male  and  female  arrival  dates — with  those  of 
Knapton  et  al.  (1984),  who  detected  (a)  white- 
striped  males  slightly,  but  not  significantly, 
earlier  than  tan-striped  males,  and  (b)  white- 
striped  females  significantly  earlier  than  tan- 
striped  females.  Results  of  our  study,  however, 
point  to  real  differences  in  female  migration 
timing,  rather  than  differences  in  detections  of 
white-striped  and  tan-striped  birds. 

Since  male  arrival  dates  are  similar  for  both 
morphs,  perhaps  it  is  the  earlier  arrival  of 
white-striped  females  that  facilitates  the  neg- 
ative assortative  mating  in  this  species.  This 
is  consistent  with  the  mechanism  proposed  by 
Houtman  and  Falls  (1994),  whereby  white- 
striped  females  out-compete  tan-striped  fe- 
males for  the  tan-striped  males.  We  suggest, 
however,  that  dominance  does  not  act  alone; 


Caldwell  and  Mills  • WHITE-THROATED  SPARROW  ARRIVAL  DATES 


331 


rather,  the  morph- specific  migration  phenolo- 
gies also  give  a competitive  advantage  to 
white-striped  females.  While  the  1-  to  2-day 
difference  in  timing  that  we  report  here  is 
modest,  it  is  not  implausible  that  it  is  suffi- 
cient to  confer  on  white-striped  females  a 
competitive  advantage  over  their  tan-striped 
counterparts. 

Early  arrival  can  confer  a higher  social  sta- 
tus in  migrant  birds  (e.g.,  Red-winged  Black- 
birds, Agelaius  phoeniceus’,  Cristol  1995).  In 
White-throated  Sparrows,  Watt  et  al.  (1984) 
concluded  that  the  dominance  between  female 
morphs  is  seasonally  dependent,  whereby 
white-striped  females  are  dominant  on  the 
breeding  grounds  and  the  tan-striped  females 
are  dominant  on  the  winter  grounds.  The  ear- 
lier spring  arrival  of  white-striped  females 
may  then  represent  the  switch  in  social  status 
between  female  morphs.  Inferior  social  status 
on  the  winter  grounds  could  mean  that  the 
best  strategy  for  white-striped  females  is  to 
leave  earlier  in  spring  to  attain  a higher  social 
status  than  tan-striped  females.  Others  have 
concluded,  however,  that  morph  type  has  no 
effect  on  social  rank  in  winter  (Piper  and  Wi- 
ley 1989). 

Alternatively,  we  acknowledge  the  possi- 
bility that  the  earlier  arrival  of  white-striped 
females  demonstrated  in  our  study  is  merely 
facilitated  by  their  larger  size  and  may  have 
no  functional  significance  in  negative  assor- 
tative  mating  or  dominance  relationships.  We 
think  this  unlikely,  however,  because  white- 
striped  males  are  bigger  than  tan-striped 
males,  and  yet  their  migration  phenologies  do 
not  differ. 

Because  white-striped  females  exhibit  low- 
er levels  of  parental  care  than  tan-striped  fe- 
males in  normal,  two-parent  nests,  Knapton 
and  Falls  (1983)  questioned  the  ability  of 
white-striped  females  to  raise  broods  on  their 
own  without  a mate.  If  true,  fledging  success 
among  white-striped  females  might  be  en- 
hanced if  they  pair  with  tan-striped  males,  as 
the  latter  exhibit  parental  contributions  that 
match  those  of  white-striped  females  and  ex- 
ceed those  of  white-striped  males  (Knapton 
and  Falls  1983).  In  another  study,  however, 
Whillans  and  Falls  (1990)  found  that  both 
white-striped  and  tan-striped  females  compen- 
sate in  terms  of  parental  care  when  males  are 
removed  from  the  nest,  and  both  female 


morphs  are  able  to  successfully  fledge  young. 
Whillans  and  Falls  (1990)  suggested  that  the 
difference  in  results  between  the  two  studies 
might  be  explained  by  differences  in  study 
sites  that  supported  differing  densities  of 
white-striped  males. 

Previously,  researchers  have  suggested  that 
nearly  70%  of  all  White-throated  Sparrow 
pairs  are  composed  of  white-striped  males  and 
tan-striped  females  (Thorneycroft  1975, 
Knapton  and  Falls  1983).  This  is  perplexing, 
since  the  nestling  ratio  and  the  banding  data 
we  present  suggest  that  the  morph  ratio  is 
much  closer  to  1:1.  It  is  not  known  whether 
tan-striped  birds  are  predominant  among  pop- 
ulations of  floating  males,  or  whether  white- 
striped  birds  are  predominant  among  popula- 
tions of  non-breeding  females.  White-striped 
birds  are  more  conspicuous  compared  to  their 
tan-striped  counterparts  in  song,  territorial  be- 
havior, and  overall  brightness  in  color  (Lowth- 
er  1961,  Falls  and  Kopachena  1999),  and  this 
may  influence  apparent  proportions  of  pair-as- 
sortment types. 

With  white-striped  birds  being  larger  and 
having  significantly  longer  wing  chords,  we 
feel  it  would  be  logical  to  use  two  sexing  sys- 
tems when  wing-chord  length  is  employed. 
Rising  and  Shields  (1980)  found  that,  gener- 
ally, tan-striped  males  were  slightly  smaller 
overall  than  white-striped  males,  and  that  gen- 
erally white-striped  females  were  larger  than 
tan-striped  females  in  terms  of  most  charac- 
teristics that  they  measured.  To  assist  in  more 
comprehensive  sex  assignment  and  to  gener- 
ate samples  more  accurately  representing  nat- 
ural populations,  we  suggest  that  these  new 
sexing  criteria  be  used  whenever  morph  iden- 
tification is  possible.  Although  the  sexing  cri- 
teria proposed  here  yielded  only  slightly  dif- 
ferent wing-chord  lengths  than  those  mea- 
sured by  the  established  sexing  criteria,  im- 
plementing this  change  substantially  increased 
the  number  of  birds  to  which  we  could  assign 
sex.  When  morph  identification  is  not  possi- 
ble, the  established  wing-chord  rule,  as  sug- 
gested in  Pyle  (1997),  should  be  used. 

Previously,  it  was  known  that  there  are  sev- 
eral differences  between  white-striped  and 
tan-striped  morphs  of  White-throated  Spar- 
rows, including  size,  habitat,  aggression  lev- 
els, and  parental  care  (Rising  and  Shields 
1980,  Knapton  and  Falls  1982,  Houtman  and 


332 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Falls  1994).  Our  study  reveals  yet  another  dif- 
ference: the  timing  of  spring  migration  among 
females  differs  between  morphs.  Overall,  it 
appears  that  the  White-throated  Sparrow’s 
morph-based  systems  of  migration  timing  and 
social  structure  are  unique  among  passerine 
birds. 

ACKNOWLEDGMENTS 

We  are  grateful  to  the  many  volunteers  of  the  Long 
Point  Bird  Observatory  (LPBO),  as  well  as  J.  D. 
McCracken  (LPBO — Bird  Studies  Canada),  for  pro- 
viding us  with  the  raw  data.  We  also  thank  M.  K.  Peck 
at  the  Royal  Ontario  Museum  for  allowing  us  access 
to  White-throated  Sparrow  skins.  J.  D.  Rising  provided 
support  and  helpful  comments  throughout  the  project, 
and  J.  B.  Falls  and  two  anonymous  reviewers  provided 
invaluable  suggestions  that  improved  the  manuscript. 

LITERATURE  CITED 

Atkinson,  C.  T.  and  C.  J.  Ralph.  1980.  Acquisition 
of  plumage  polymorphism  in  White-throated 
Sparrows.  Auk  97:245-252. 

Collins,  C.  F.  and  A.  M.  Houtman.  1999.  Tan  and 
white  color  morphs  of  White-throated  Sparrows 
differ  in  their  non-song  vocal  responses  to  terri- 
torial intrusion.  Condor  101:842-845. 

Cristol,  D.  A.  1995.  Early  arrival,  initiation  of  nest- 
ing, and  social  status:  an  experimental  study  of 
breeding  female  Red-winged  Blackbirds.  Behav- 
ioral Ecology  6:87-93. 

Falls,  J.  B.  and  J.  G.  Kopachena.  1994.  White-throat- 
ed Sparrow  ( Zontrichia  albicollis).  The  Birds  of 
North  America,  no.  128. 

Houtman,  A.  M.  and  J.  B.  Falls.  1994.  Negative  as- 
sortative  mating  in  the  White-throated  Sparrow, 
Zonotrichia  albicollis : the  role  of  mate  choice  and 
intra-sexual  competition.  Animal  Behaviour  48: 
377-383. 

Jenkins,  K.  D.  and  D.  A.  Cristol.  2002.  Evidence  of 
differential  migration  by  sex  in  White-throated 
Sparrows  ( Zonotrichia  albicollis).  Auk  119:539- 
543. 

Knapton,  R.  W..  R.  V.  Carter,  and  J.  B.  Falls.  1984. 
A comparison  of  breeding  ecology  and  reproduc- 
tive success  between  morphs  of  the  White-throat- 
ed Sparrow.  Wilson  Bulletin  96:60-71. 


Knapton,  R.  W.  and  J.  B.  Falls.  1982.  Polymorphism 
in  the  White-throated  Sparrow:  habitat  occupancy 
and  nest-site  selection.  Canadian  Journal  of  Zo- 
ology 60:452-459. 

Knapton,  R.  W.  and  J.  B.  Falls.  1983.  Differences  in 
parental  contribution  among  pair  types  in  the 
polymorphic  White-throated  Sparrows.  Canadian 
Journal  of  Zoology  61:1288-1292. 

Kopachena,  J.  G.  and  J.  B.  Falls.  1993.  Aggressive 
performance  as  a behavioural  correlate  of  plum- 
age polymorphism  in  the  White-throated  Sparrow 
( Zonotrichia  albicollis ).  Behaviour  124:249-266. 

Lowther,  J.  K.  1961.  Polymorphism  in  the  White- 
throated  Sparrow,  Zonotrichia  albicollis  (Gmelin). 
Canadian  Journal  of  Zoology  39:281-292. 

Mills,  A.  M.  2005.  Changes  in  the  timing  of  spring 
and  autumn  migration  in  North  American  migrant 
passerines  during  a period  of  global  warming.  Ibis 
147:259-269. 

Piper,  W.  H.  and  R.  H.  Wiley.  1989.  Correlates  of 
dominance  in  wintering  White-throated  Sparrows: 
age,  sex  and  location.  Animal  Behaviour  37:298- 
310. 

Pyle,  P.  1997.  Identification  guide  to  North  American 
birds,  part  I.  Columbidae  to  Ploceidae.  Slate  Creek 
Press,  Bolinas,  California. 

Rising,  J.  D.  and  G.  F.  Shields.  1980.  Chromosomal 
and  morphological  correlates  in  two  New  World 
sparrows  (Emberizidae).  Evolution  34:654-662. 

Thorne ycroft,  H.  B.  1966.  Chromosomal  polymor- 
phism in  the  White-throated  Sparrow,  Zonotrichia 
albicollis  (Gmelin).  Science  154:1571-1572. 

Thorneycroft,  H.  B.  1975.  A cytogenetic  study  of  the 
White-throated  Sparrow,  Zonotrichia  albicollis 
(Gmelin).  Evolution  29:611-621. 

Tuttle,  E.  M.  1993.  Mate  choice  and  stable  polymor- 
phism in  the  White-throated  Sparrow.  Ph.D.  dis- 
sertation, State  University  of  New  York,  Albany. 

Tuttle,  E.  M.  2003.  Alternative  reproductive  strate- 
gies in  the  White-throated  Sparrow:  behavioral 
and  genetic  evidence.  Behavioral  Ecology  14: 
425-432. 

Watt,  D.  J.,  C.  J.  Ralph,  and  C.  T.  Atkinson.  1984. 
The  role  of  plumage  polymorphism  in  dominance 
relationships  of  the  White-throated  Sparrow.  Auk 
101:110-120. 

Whillans,  K.  V.  and  J.  B.  Falls.  1990.  Effects  of 
male  removal  on  parental  care  of  female  White- 
throated  Sparrows,  Zonotrichia  albicollis.  Animal 
Behaviour  39:869-878. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):333— 340,  2006 


CAN  SUPPLEMENTAL  FORAGING  PERCHES  ENHANCE  HABITAT 
FOR  ENDANGERED  SAN  CLEMENTE  LOGGERHEAD  SHRIKES? 

SUELLEN  LYNN,1  24  JOHN  A.  MARTIN,13  AND  DAVID  K.  GARCELON1 2 3 4 


ABSTRACT. — Habitat  degradation  caused  by  feral  grazers  has  been  identified  as  a possible  limiting  factor 
for  the  endangered  San  Clemente  Loggerhead  Shrike  ( Lanius  ludovicianus  mearnsi).  In  1999,  we  installed 
supplemental  foraging  perches  within  shrike  breeding  territories  on  San  Clemente  Island  and  observed  shrike 
foraging  behavior  before  and  after  perches  were  installed.  Shrike  foraging  efficiency,  determined  by  measuring 
foraging  attack  distances  and  success  rates,  was  not  improved  when  supplemental  perches  were  present;  however, 
shrikes  shifted  their  focal  foraging  sites  to  areas  where  perches  were  installed.  Shrike  home  ranges  did  not 
change  size  when  supplemental  perches  were  installed,  indicating  that  foraging  areas  made  available  by  adding 
supplemental  perches  were  not  of  higher  quality  than  those  that  were  previously  available.  However,  the  addition 
of  supplemental  perches  may  have  increased  the  total  foraging  habitat  available  to  this  endangered  subspecies. 
Received  13  May  2005,  accepted  17  February  2006. 


Habitat  deficiencies  have  been  identified  as 
possible  limiting  factors  in  populations  of 
Loggerhead  Shrikes  ( Lanius  ludovicianus', 
Yosef  1994,  Cade  and  Woods  1997).  In  the 
1980s,  Scott  and  Morrison  (1990)  studied  a 
population  of  endangered  shrikes  on  San  Cle- 
mente Island  (SCI),  the  San  Clemente  Log- 
gerhead Shrike  (L.  /.  mearnsi).  In  the  late 
1890s  and  early  1900s,  Grinnell  (1897)  had 
considered  this  subspecies  “tolerably  com- 
mon; that  is,  two  or  three  could  generally  be 
seen  during  an  hour’s  walk,”  and  Linton 
(1908)  called  the  population  “fairly  well  dis- 
tributed.” By  the  1990s,  the  population  on 
SCI  had  dropped  to  a low  of  13  individuals 
(T.  Mader  unpubl.  data).  Scott  and  Morrison 
(1990)  identified  habitat  degradation  attribut- 
ed to  overgrazing  by  feral  goats  ( Capra  hir- 
cus)  as  a likely  cause  of  this  subspecies’  de- 
cline. Common  effects  of  overgrazing  by  feral 
goats  include  depletion  of  woody  species  and 
an  increase  in  exotic  vegetation  (Coblentz 
1980). 

Because  shrikes  use  elevated  substrates  as 
foraging  perches,  from  which  they  can  readily 
see  prey  and  attack  with  flights  to  the  ground 
(Bent  1950),  perches  are  an  important  com- 
ponent of  shrike  territories  (Esely  and  Bollin- 


1  Inst,  for  Wildlife  Studies,  RO.  Box  1104,  Areata, 
CA  95518,  USA. 

2 Current  address:  PRBO  Conservation  Science, 
4990  Shoreline  Hwy.  1.  Stinson  Beach,  CA  94970, 
USA. 

3 Current  address:  2144  Froude  St.,  San  Diego,  CA 
92107,  USA. 

4 Corresponding  author;  e-mail:  slynn@prbo.org 


ger  2001).  If  elevated  perches  are  lacking, 
shrikes  may  not  be  able  to  use  all  potential 
foraging  habitat  and  may,  therefore,  increase 
their  home-range  size  to  encompass  an  ade- 
quate area  of  usable  habitat.  Having  to  move 
about  larger  home  ranges  and  defend  larger 
territories  requires  that  shrikes  expend  greater 
amounts  of  energy;  this  may  result  in  a de- 
crease in  their  nutritional  status  (Yosef  and 
Grubb  1992).  The  establishment  of  larger  ter- 
ritories also  decreases  the  shrike  carrying  ca- 
pacity of  SCI’s  limited  area.  Yosef  and  Grubb 
(1994)  found  that  adding  fence  posts  to  shrike 
territories  in  Florida  resulted  in  smaller  aver- 
age territory  sizes  and  greater  breeding  den- 
sities of  shrikes.  Artificial  perches  have  also 
been  shown  to  attract  raptors,  especially  kes- 
trels ( Falco  sp.),  to  areas  that  were  otherwise 
devoid  of  appropriate  perches  (Kay  et  al. 
1994,  Wolff  et  al.  1999,  Kim  et  al.  2003). 

Optimal  foraging  theory  suggests  that  an 
animal  will  optimize  the  capture  and  con- 
sumption of  prey,  maximizing  energy  intake 
while  minimizing  energy  expenditure  (Schoe- 
ner  1971,  Mills  1979).  Therefore,  an  increase 
in  foraging  efficiency  should  be  reflected  by 
shorter  attack  distances  (less  energy  required 
to  fly  a shorter  distance),  capture  of  larger 
prey  items  (fewer  attempts  needed),  and  a 
greater  percentage  of  successful  foraging  at- 
tempts (less  wasted  energy  on  failed  foraging 
attempts).  An  increase  in  foraging  efficiency 
also  may  be  reflected  by  more  frequent  cap- 
tures per  unit  time,  even  if  success  rate  does 
not  improve.  Furthermore,  shrikes  may  select 
nest  locations  near  foraging  areas  to  decrease 
energy  expended  in  flight  while  tending  a nest. 


333 


334 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Shrike  foraging  efficiency  may  be  con- 
strained by  the  number  and  arrangement  of 
available  hunting  perches.  Prior  to  our  supple- 
mental perch  experiment,  we  had  found  a 
greater  number  of  trees  and  shrubs  at  sites  oc- 
cupied by  shrikes  on  SCI  than  at  sites  shrikes 
had  abandoned  within  the  past  10  years  (SL 
unpubl.  data).  If  hunting  perches  are  limited, 
then  it  seemed  reasonable  to  expect  that  the 
addition  of  supplemental  perches  within 
shrike  territories  would  allow  foraging  effi- 
ciency to  increase  by  providing  shrikes  a 
greater  choice  of  hunting  perches,  thereby  in- 
creasing their  opportunity  to  choose  the  best 
hunting  area.  Therefore,  we  designed  an  ex- 
periment to  determine  whether  the  addition  of 
supplemental  perches  to  shrike  territories 
would  increase  foraging  efficiency  and  the  ef- 
fective usable  area  of  a given  home  range.  We 
also  examined  whether  the  presence  of  sup- 
plemental perches  would  alter  shrike  breeding 
behavior  by  allowing  them  to  forage  nearer  to 
their  nests. 

METHODS 

Study  area. — San  Clemente  Island  (32°  50' 
N,  118°  30'  W),  the  southern-most  of  Califor- 
nia’s Channel  Islands,  is  located  about  100  km 
northwest  of  San  Diego,  California.  The  is- 
land is  28  km  long  (width  = 3-7  km,  area  = 
145  km2)  and  rises  abruptly  to  599  m in  ele- 
vation on  the  eastern  escarpment.  Numerous 
canyons  cut  through  marine  terraces  on  the 
southwestern  part  of  the  island.  Island  tem- 
peratures range  from  7—35°  C,  precipitation 
ranges  from  12-20  cm/year  (mainly  Novem- 
ber through  March),  and  fog  is  common,  es- 
pecially in  summer  months  (Jorgensen  and 
Ferguson  1984,  Scott  and  Morrison  1990). 

Native  vegetation  on  the  island  has  been 
substantially  altered  by  introduced  herbivores, 
including  sheep  (Ovis  aries),  goats,  and  pigs 
(Sus  scrofa ),  all  of  which  were  eradicated  by 
1993.  By  the  time  of  our  study,  the  dominant 
plant  community  comprised  native  and  non- 
native grasses  (including  Avena,  Bromus,  and 
Nassella  spp.)  interspersed  with  areas  of  re- 
cently recruited  coyote  brush  ( Baccharis  pi- 
lularis ),  which  covered  —33%  of  the  flatter 
upper  reaches  of  the  island  (U.S.  Department 
of  the  Navy  2001).  Shrubs  and  trees  were  pri- 
marily restricted  to  the  canyon  bottoms.  SCI 
is  operated  by  the  U.S.  Navy  as  a training 


base,  primarily  for  ship-to-shore  bombard- 
ment in  the  area  where  we  conducted  our 
study.  See  U.S.  Department  of  the  Navy 
(2001)  for  additional  information  on  the  is- 
land’s vegetation,  geography,  and  other  natu- 
ral resources. 

Site  selection  and  study  design. — In  1999, 
we  selected  four  (of  eight  total)  pairs  of  breed- 
ing shrikes  on  SCI  for  study.  None  of  the 
pairs’  home  ranges  overlapped,  and  the  dis- 
tance between  the  edge  of  each  pair’s  home 
range  and  its  closest  neighbor  ranged  from 
100-800  m.  Sample  size  was  constrained  by 
logistical  and  conservation  considerations, 
such  as  site  accessibility  and  concerns  about 
manipulating  the  breeding  sites  of  a highly  en- 
dangered population.  We  studied  shrike  be- 
havior and  recorded  their  responses  to  supple- 
mental perches  during  two  periods:  13  March 
through  4 June  (period  1)  and  5 June  through 
2 August  1999  (period  2).  On  13  March,  we 
installed  supplemental  perches  at  two  sites  (A 
and  D;  Fig.  1).  During  period  1,  we  observed 
at  least  75  foraging  attempts  at  the  sites  with 
supplemental  perches  and  also  at  two  sites  (B 
and  C;  Fig.  1)  without  supplemental  perches. 
On  5 June,  we  removed  the  perches  from  sites 
A and  D and  installed  them  at  sites  B and  C; 
during  period  2,  we  observed  another  75  + 
foraging  attempts  at  each  site.  This  paired 
sampling  design  controlled  for  seasonal  and 
individual  differences  in  behavior. 

The  shrike  breeding  season  typically  begins 
in  January  with  pair  formation  and  extends 
through  mid-August,  when  the  last  fledglings 
disperse  from  their  natal  territories.  Because 
we  were  concerned  that  different  breeding 
stages  might  elicit  differences  in  foraging  be- 
havior, we  recorded  the  shrikes’  breeding 
stage  throughout  the  study  and  mapped  the  lo- 
cations of  their  nests.  During  the  nestling  and 
fledgling  stages,  shrikes  may  alter  their  for- 
aging behavior  by  increasing  foraging  rates  to 
provide  for  their  young.  Therefore,  we  elimi- 
nated foraging  attempts  observed  during  these 
periods  to  avoid  biasing  our  results. 

At  sites  B and  C,  the  original  females  were 
replaced  by  captive-released  females  during 
the  breeding  season.  The  original  female  at 
site  B disappeared  between  1 1 and  17  April 
and  was  replaced  with  a released  female  on  1 
May.  We  collected  data  on  this  female  during 
both  study  periods.  At  site  C,  the  original  fe- 


Lynn  et  al.  • SUPPLEMENTAL  PERCHES  FOR  SHRIKES 


335 


FIG.  1.  Maps  of  minimum  convex  polygon  home-range  estimates,  encompassing  all  foraging  locations,  when 
supplemental  perches  were  present  (treatment)  and  not  present  (control)  within  San  Clemente  Loggerhead  Shrike 
territories,  San  Clemente  Island,  California,  1999. 


male  was  depredated  between  2 and  5 May 
and  replaced  with  a released  female  on  15 
May,  prior  to  the  installation  of  supplemental 
perches  at  that  site. 

At  all  sites,  we  installed  3 groups  of  5 sup- 
plemental perches,  arranged  linearly  where 
possible  (Fig.  1),  for  a total  of  15  perches  per 
site.  Within  a group  of  five,  we  spaced  sup- 


plemental perches  30  m apart,  which  was 
twice  the  average  attack  distance  for  a ground 
foraging  attempt  (SL  unpubl.  data),  and  >30 
m from  naturally  occurring,  elevated  (>2  m) 
perches.  We  placed  each  line  of  perches  at  a 
randomly  selected  distance  (1  to  200  m)  from 
the  shrike  activity  center  at  each  site,  and  we 
oriented  each  line  according  to  randomly  se- 


336 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


lected  compass  directions.  Supplemental 
perches  were  poles  of  aluminum  conduit  (3  m 
long,  1.3  cm  in  diameter)  slipped  over  a piece 
of  rebar  pounded  into  the  ground.  Attached  to 
each  pole  were  three  horizontal  cross  pieces 
(40  cm  long)  made  of  wooden  dowels  (0.3  cm 
in  diameter)  positioned  at  2.5,  1.5,  and  0.75 
m from  the  ground.  Barbed  wire  was  wound 
around  the  joint  of  the  cross  piece  and  upright 
conduit  to  serve  as  a site  for  shrikes  to  impale 
their  prey. 

Data  collection. — We  identified  all  shrikes 
by  unique  combinations  of  colored  leg  bands. 
Our  observation  points  were  >50  m away 
from  the  center  of  shrike  activity  to  avoid  dis- 
turbing the  shrikes;  at  sites  where  one  obser- 
vation point  was  not  sufficient  to  observe  the 
entire  area,  we  placed  additional  points  at  var- 
iable distances  from  the  activity  center.  We 
observed  each  shrike  pair  for  0.5— 1.0  hr  per 
visit.  In  addition  to  bird  identity  and  weather 
conditions,  for  each  foraging  attempt  we  re- 
corded perch  substrate,  perch  height,  type  of 
foraging  maneuver  (aerial  sally,  ground  forage 
[flight  to  the  ground  from  an  elevated  perch], 
or  vegetation  glean),  outcome,  foraging-at- 
tempt distance,  and  prey  captured  (mouse,  liz- 
ard, bird,  small  arthropod  [<10  mm,  i.e., 
smaller  than  the  length  of  a shrike  bill],  and 
large  arthropod  [>10  mm]).  Because  there 
were  significant  differences  between  male  and 
female  behaviors  (i.e.,  the  female  is  the  pri- 
mary incubator,  the  male  provisions  the  fe- 
male when  she  is  on  the  nest),  we  analyzed 
foraging  behavior  separately  by  sex. 

Statistical  analyses. — We  mapped  the  loca- 
tions of  perches  used  by  shrikes  during  for- 
aging attempts,  then  transferred  these  loca- 
tions to  ArcView,  v.  3.2a  (Environmental  Sys- 
tems Research  Institute,  Inc.  2000).  We  gen- 
erated minimum  convex  polygons  using 
ArcView  Animal  Movements  Extension,  v. 
2.0  beta  (Hooge  et  al.  1999)  for  locations 
mapped  when  supplemental  perches  were  pre- 
sent (treatment:  n — 73-85)  and  not  present 
(control:  n = 80-94).  We  used  paired  f-tests 
to  compare  the  sizes  of  minimum  convex 
polygons  between  treatments  and  controls.  To 
determine  whether  shrikes  shifted  their  for- 
aging areas  in  response  to  the  installation  or 
removal  of  supplemental  perches,  we  also 
mapped  the  locations  of  supplemental  perches 
used  by  shrikes  and  then  counted  the  number 


that  fell  within  the  polygons  generated  during 
treatment  and  control  periods.  We  used  Fish- 
er’s exact  test  of  independence  (Sokal  and 
Rohlf  1981)  to  compare  the  number  of  perch 
sites  used  during  control  and  treatment  peri- 
ods. 

To  determine  whether  supplemental  perches 
affected  the  selection  of  nest  sites,  at  each  site 
we  recorded  whether  each  nest  was  initiated 
during  treatment  or  control.  For  nests  initiated 
during  treatment,  we  measured  the  distance 
from  the  nest  to  all  supplemental  perches.  For 
nests  initiated  during  control,  we  measured  the 
distance  from  the  nest  to  where  the  supple- 
mental perches  were  installed  during  treat- 
ment. At  sites  where  shrikes  built  nests  during 
both  treatment  and  control,  we  compared  the 
mean  nest-to-supplemental  perch  distance 
during  treatment  to  the  mean  nest-to-supple- 
mental perch  distance  for  all  supplemental 
perch  sites  (i.e.,  perch  site  = location  where 
a supplemental  perch  would  be,  or  had  been, 
placed  during  treatment)  during  control.  We 
used  paired  Mests  to  ascertain  differences  in 
foraging-attempt  distances  between  treatment 
and  control.  Where  sample  sizes  were  large 
enough,  we  used  chi-square  tests  to  test  for 
treatment  versus  control  differences  in  forage- 
maneuver  type,  foraging  success,  and  size  of 
prey  item  captured;  otherwise  we  used  Fish- 
er’s exact  test.  Because  of  inherent  differences 
in  foraging-maneuver  type  (i.e.,  larger  prey 
items,  such  as  lizards  and  mice,  were  not  cap- 
tured during  aerial  sallies),  we  analyzed  size 
of  prey  and  foraging-attempt  distances  by  type 
of  foraging  maneuver.  Means  are  reported  ± 
SD.  We  considered  P < 0.05  to  be  statistically 
significant. 

RESULTS 

We  observed  a total  of  674  foraging  at- 
tempts, 338  of  which  occurred  during  the 
treatment  phase  (110  from  supplemental 
perches,  228  from  naturally  occurring  perch- 
es) and  336  during  the  control  phase  of  our 
study.  After  eliminating  foraging  attempts 
when  nestlings  or  fledglings  were  present,  we 
were  able  to  determine  whether  a foraging  at- 
tempt was  successful  for  447  attempts,  224 
during  treatment  (86  from  supplemental 
perches  and  138  from  naturally  occurring 
perches)  and  223  during  control. 

Pairs  at  sites  B and  C built  and  tended  one 


Lynn  et  al.  • SUPPLEMENTAL  PERCHES  FOR  SHRIKES 


337 


TABLE  1.  Distance  between  nests  and  supplemental  perches  installed  within  San  Clemente  Loggerhead 
Shrike  territories,  San  Clemente  Island,  California,  1999.  During  control  periods,  distances  were  measured  be- 
tween nests  and  the  pre-designated  locations  of  supplemental  perches,  which  were  present  only  during  treatment 
periods. 

Site 

Nest 

Period  when 
nest  initiated 

Distance  to  nearest 
supplemental  perch 

Mean  distance  (±  SD)  to 
supplemental  perches 

A 

A 

Pre-study 

41  m 

1 18  ± 52  m 

B 

Treatment 

31  m 

153  ± 63  m 

C 

Control 

121  m 

274  ± 97  m 

D 

Control 

132  m 

233  ± 68  m 

B 

A 

Control 

70  m 

149  ± 61  m 

C 

A 

Control 

72  m 

122  ± 35  m 

D 

A 

Pre-study 

80  m 

119  ± 29  m 

B 

Treatment 

73  m 

121  ± 27m 

C 

Control 

111  m 

145  ± 29  m 

D 

Control 

85  m 

126  ± 31  m 

nest  each.  Shrike  pairs  at  sites  A and  D,  how- 
ever, each  built  and  tended  four  consecutive 
nests,  none  of  which  were  successful.  One 
nest  at  each  of  these  two  sites  was  initiated 
during  treatment  (i.e.,  supplemental  perches 
were  present).  Both  of  the  nests  initiated  dur- 
ing treatment  were  closer  to  the  nearest  sup- 
plemental perch  site  than  any  other  nests  (Ta- 
ble 1).  The  mean  distance  from  each  of  these 
two  nests  to  all  supplemental  perch  sites,  how- 
ever, was  not  shorter  than  that  of  nests  initi- 
ated when  supplemental  perches  were  not  pre- 
sent (Table  1).  Shrike  home-range  size  did  not 
differ  between  treatment  and  control  (treat- 
ment: 8.5  ± 6.1  ha;  control:  7.7  ± 2.7  ha;  t3 
— 0.24,  P = 0.83).  However,  shrikes  shifted 
their  home  ranges  to  include  some  of  the  sup- 
plemental perches  when  they  were  present. 
Significantly  more  of  the  supplemental  perch 
sites  were  located  within  shrike  home  ranges 


Male  ground-forage  Male  aerial-forage  Female  ground-forage 
attempts  attempts  attempts 


FIG.  2.  Mean  ± SD  foraging-attempt  distances  of 
male  and  female  San  Clemente  Loggerhead  Shrikes  in 
territories  with  (treatment)  and  without  (control)  sup- 
plemental perches,  San  Clemente  Island,  California, 
1999. 


during  treatment  ( n = 40)  than  during  control 
(n  = 32;  P = 0.023,  df  = 3). 

The  addition  of  supplemental  perches  did 
not  affect  average  distance  of  foraging  at- 
tempts (Fig.  2).  For  male  shrikes,  attack  dis- 
tances for  ground-foraging  attempts  were  not 
affected  by  the  presence  of  supplemental 
perches  ( n w 300,  t3  — 1.06,  P = 0.37)  nor 
were  attack  distances  of  aerial  sallies  ( n = 
140,  t3  = 0.59,  P = 0.60;  Fig.  2).  Likewise, 
female  attack  distances  for  ground-foraging 
attempts  were  not  affected  by  the  presence  of 
supplemental  perches  (n  = 51,  t2  = 0.29,  P = 
0.79).  We  did  not  observe  a sufficient  number 
of  vegetation  gleans  for  analysis  of  attack  dis- 
tance. Also,  the  addition  of  supplemental 
perches  did  not  result  in  altered  proportions  of 
foraging  maneuver  types  used  by  males  ( n = 
471,  x2  — 0.48,  P = 0.79,  df  = 2)  or  females 
in  = 70,  x2  = 2.68,  P = 0.10,  df  = 1;  Fig. 

3) . 

Foraging  success  of  neither  males  ( n = 327, 
X2  = 1.53,  P = 0.22,  df  = 1)  nor  females  ( n 
= 52,  x2  — 0.79,  P = 0.38,  df  = 1)  improved 
when  supplemental  perches  were  present  (Fig. 

4) .  Shrikes  foraged  from  supplemental  perch- 
es 33%  of  the  time  when  they  were  present, 
and  we  found  no  difference  in  the  proportion 
of  successful  foraging  attempts  launched  from 
supplemental  and  naturally  occurring  perches 
(n  = 224,  x2=  1-43,  P = 0.23,  df  = 1).  Al- 
though shrikes  tended  to  capture  more  prey/ 
hr  when  using  supplemental  perches  (0.98  ± 
0.48  successful  foraging  attempts/hr)  than 
when  using  naturally  occurring  perches  (0.52 


338 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


FIG.  3.  Percentages  of  foraging-maneuver  types 
performed  by  San  Clemente  Loggerhead  Shrikes  in 
territories  with  (treatment)  and  without  (control)  sup- 
plemental perches,  San  Clemente  Island,  California, 
1999. 


± 0.15),  the  difference  was  not  significant  ( n 
= 159,  t3  = 1.84,  P = 0.16).  Shrikes  always 
perched  on  the  top-most  crossbar  before  for- 
aging, and  only  once  did  a shrike  use  a lower 
crossbar — briefly,  before  moving  up  to  the  top 
crossbar. 

During  ground-foraging  attempts,  neither 
males  (n  = 95,  x2  = 1-46,  P = 0.23,  df  = 1) 
nor  females  (n  = 14,  Fisher’s  exact  P = 0.46, 
df  = 1)  captured  larger  prey  (small/large: 
males  with  supplemental  perches  = 33/19, 
males  without  supplemental  perches  = 22/21, 
females  with  supplemental  perches  = 2/2,  fe- 
males without  supplemental  perches  = 7/3) 
when  supplemental  perches  were  present. 
During  aerial  sallies,  however,  males  captured 
more  small  arthropods  than  large  arthropods 
when  supplemental  perches  were  present  ( n — 
93,  Fisher’s  exact  P = 0.007,  df  = 1 ; small/ 
large:  with  supplemental  perches  = 43/3, 
without  supplemental  perches  = 34/13).  Veg- 
etation gleans  by  males  tended  to  yield  small- 
er prey  when  supplemental  perches  were  pres- 
ent ( n = 22,  Fisher’s  exact  P = 0.08,  df  = 1; 
small/large:  with  supplemental  perches  = 8/5, 
without  supplemental  perches  = 2/7). 

DISCUSSION 

Although  many  aspects  of  shrike  foraging 
efficiency  did  not  increase  when  we  installed 
supplemental  perches,  San  Clemente  Logger- 
head  Shrikes  responded  positively  to  the  pres- 
ence of  supplemental  perches  by  increasing 
their  use  of  the  areas  around  the  perches. 
Shrikes  readily  used  supplemental  perches, 
and  we  found  that  when  supplemental  perches 


□ Control 
■ T reatment 

P = 0.22 


Males  Females 


FIG.  4.  Percent  foraging  success  of  male  and  fe- 
male San  Clemente  Loggerhead  Shrikes  in  territories 
with  (treatment)  and  without  (control)  supplemental 
perches,  San  Clemente  Island,  California,  1999. 


were  added  to  a home  range,  shrikes  shifted 
their  foraging  habitat  to  include  the  area 
around  some,  but  not  all,  of  the  supplemental 
perches.  The  one  exception  to  this  pattern  was 
an  apparent  shift  toward  an  area  without  sup- 
plemental perches  that  was  burned  by  a late- 
season  fire  at  site  B. 

The  shift  in  areas  used  by  shrikes  when 
supplemental  perches  were  present  suggests 
that  some  areas  of  the  shrikes’  home  ranges 
contained  prey  resources  that  could  not  be 
used  due  to  a lack  of  appropriate  foraging 
perches.  Although  our  sample  size  was  insuf- 
ficient for  statistical  comparisons,  the  shrikes 
seemed  to  place  their  nests  closer  to  supple- 
mental perches  when  they  were  present  (Fig. 
1);  if  true,  shrikes  may  have  reduced  their  en- 
ergetic costs  by  taking  advantage  of  the  newly 
available  foraging  areas.  Tall  perches  may 
have  provided  other  benefits  to  shrikes,  in- 
cluding increased  capacity  for  predator  vigi- 
lance and  more  display  areas  for  territory  de- 
fense and  mate  attraction.  In  contrast,  Chavez- 
Ramirez  et  al.  (1994)  found  that  shrikes  in 
natural  grasslands  in  Texas  did  not  shift  their 
foraging  areas  as  densities  of  artificial  perches 
were  manipulated;  instead,  the  shrikes  in- 
creased their  use  of  herbaceous  perches,  and 
Chavez-Ramirez  et  al.  (1994)  concluded  that 
foraging  perches  were  not  a limiting  factor  in 
natural  grasslands. 

Habitat  enhancement  has  yielded  beneficial 
results  where  focal  species  lacked  certain  hab- 
itat components.  In  disturbed  landscapes  of 
Washington  state  (Rocklage  and  Ratti  2000), 
bird  species  diversity  increased  with  the  ad- 
dition of  irrigation  along  the  Snake  River  and, 
in  New  Zealand,  several  bird  species  in- 


Lynn  et  al.  • SUPPLEMENTAL  PERCHES  FOR  SHRIKES 


339 


creased  their  use  of  areas  cleared  of  willows 
along  braided  rivers  (Maloney  et  al.  1999). 
Probably  due,  in  part,  to  the  extremely  low 
number  of  shrikes  on  SCI,  we  did  not  see  a 
similar  increase  in  bird  density  with  the  ad- 
dition of  supplemental  perches.  Consequently, 
the  lack  of  intraspecific  competition  between 
San  Clemente  Loggerhead  Shrikes  allowed 
them  to  investigate  areas  that  were  previously 
unavailable  and  to  respond  opportunistically 
to  novel  structures.  We  did  not  find  a concur- 
rent increase  in  foraging  success  or  efficiency 
with  the  addition  of  supplemental  perches,  in- 
dicating that  the  areas  opened  up  for  foraging 
by  the  addition  of  perches  may  not  have  been 
superior  to  those  already  available.  This  idea 
was  supported  by  the  substantial  overlap  in 
areas  used  during  treatment  and  control  peri- 
ods (Fig.  1)  and  our  observation  that  shrikes 
did  not  use  all  of  the  supplemental  perches 
provided,  both  of  which  indicate  that  the  hab- 
itat quality  in  some  areas  was  poor  and  would 
not  be  enhanced  even  by  the  installation  of 
supplemental  perches. 

Shrikes  in  Florida  reduce  their  territory  size 
with  the  addition  of  foraging  perches,  and  new 
shrike  pairs  will  establish  territories  in  the  ar- 
eas vacated  (Yosef  and  Grubb  1994).  When  a 
limited  resource  (foraging  perches)  is  added, 
shrikes  are  able  to  decrease  the  energy  ex- 
pended on  moving  throughout  and  defending 
a large  territory  from  other  shrikes,  thereby 
potentially  improving  their  nutritional  status 
(Yosef  and  Grubb  1992).  With  the  decrease  in 
territory  size  defended,  and  the  density  in- 
crease in  pairs  of  shrikes,  the  addition  of  sup- 
plemental perches  potentially  increased  the 
carrying  capacity  of  shrike  habitat  in  Florida. 

Unlike  shrikes  in  Florida,  however,  home- 
range  size  of  San  Clemente  Loggerhead 
Shrikes  was  not  affected  by  the  presence  of 
additional  foraging  perches.  On  SCI,  the  low 
number  of  breeding  shrikes  (eight  pairs)  ne- 
gated the  advantage  of  decreasing  home-range 
size  to  reduce  energy  expenditure  on  territory 
defense.  Shrike  home-ranges  were  far  enough 
apart  (>100  m;  T.  Mader  unpubl.  data)  that 
territorial  defense  against  neighboring  shrike 
pairs  was  unlikely  to  limit  the  home-range 
size  of  the  resident  pair.  Furthermore,  because 
the  shrike  population  in  our  study  was  thor- 
oughly observed  and  color-marked,  we  are 
confident  that  no  additional  shrike  pairs  were 


breeding  nearby;  therefore,  little  competition 
for  breeding  resources  could  have  occurred. 

After  the  addition  of  supplemental  perches, 
San  Clemente  Loggerhead  Shrikes  incorporat- 
ed previously  unused  habitat  while  maintain- 
ing similarly  sized  home  ranges,  suggesting 
that  other  aspects  of  their  home  range  were 
still  important  to  their  survival.  Supplemental 
perches  provided  substrates  on  which  to  perch 
and  impale  captured  prey,  but  did  not  provide 
the  structure  and  foliage  of  trees — features  re- 
quired by  shrikes  for  nest  placement  and  for 
concealment  and  escape  from  predators.  Kim 
et  al.  (2003)  found  that  shrikes  were  more 
closely  associated  with  natural  woody  perches 
than  artificial  perches  and  attributed  this  as- 
sociation to  the  lack  of  escape  cover  at  arti- 
ficial perches.  In  Kansas,  the  number  of  po- 
tential nesting  trees  was  the  most  important 
predictive  variable  for  shrike  habitat  suitabil- 
ity (Lauver  et  al.  2002).  Trees  and  shrubs  on 
SCI  can  attain  heights  of  >10  m,  but  they  are 
limited  to  canyon  bottoms  and  other  areas  that 
were  protected  from  goat  herbivory.  Nonethe- 
less, shrikes  must  include  these  remnant  trees 
and  shrubs  in  their  breeding  home  ranges  for 
successful  reproduction  and  survival. 

In  contrast  to  Yosef  and  Grubb  (1994),  we 
did  not  find  evidence  that  the  availability  of 
suitable  foraging  perches  limits  shrikes  ener- 
getically, possibly  due  to  the  differences  in 
terrain  between  their  study  site  and  ours. 
Shrikes  on  SCI  typically  inhabit  steep,  rocky, 
topographically  complex  canyons,  although 
they  occasionally  forage  on  flat  mesas  be- 
tween canyons.  In  such  topographically  com- 
plex environments,  short  foraging  perches 
may  not  limit  the  area  available  that  shrikes 
can  search  for  prey  to  the  degree  that  they 
would  in  a flatter  environment.  Two  of  the 
shrike  territories  we  observed  were  in  typi- 
cally rocky  canyons,  and  two  were  in  shallow- 
er canyons  flanked  by  flat  mesas.  Our  results 
suggest  that  there  may  be  an  interaction  be- 
tween foraging-perch  availability  and  topog- 
raphy, although  our  sample  size  was  insuffi- 
cient to  demonstrate  this  conclusively. 

With  recent  increases  in  the  shrike  popula- 
tion resulting  from  intensive  population  man- 
agement— including  the  release  of  captive- 
bred  shrikes  into  the  wild — competition  may 
play  a greater  role  in  the  choice  of  defended 
foraging  areas.  To  accommodate  an  increasing 


340 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


population,  potential  shrike  habitat  should  be 
made  available  by  the  addition  of  hunting 
perches.  Long-term  improvement  of  shrike 
habitat  should  include  restoring  trees  and 
shrubs  to  SCI  to  increase  the  availability  of 
nesting  habitat.  Meanwhile,  the  lack  of  ele- 
vated hunting  perches  may  be  temporarily  al- 
leviated by  the  installation  of  artificial  perch- 
es. 

ACKNOWLEDGMENTS 

Thanks  to  D.  M.  Cooper  and  K.  M.  Wakelee,  who 
collected  much  of  our  data.  Thanks  to  T.  M.  Ostheimer, 
W.  A.  Ostheimer,  and  the  rest  of  the  PRBO  field  crew 
who  shared  their  observations  and  helped  follow  the 
shrikes.  Thanks  to  P.  Sharpe  for  reviewing  our  study 
design  and  to  E.  L.  Kershner  for  encouraging  the  de- 
velopment of  this  manuscript  and  for  insightful  com- 
ments and  suggestions.  Thanks  also  to  T.  J.  Cade,  F. 
Chavez-Ramirez,  and  an  anonymous  reviewer  for  their 
critiques  of  this  manuscript.  This  study  was  part  of  the 
recovery  effort  for  the  San  Clemente  Loggerhead 
Shrike  and  was  funded  by  the  Commander  in  Chief, 
Pacific  Fleet,  Pearl  Harbor,  Hawaii. 

LITERATURE  CITED 

Bent,  A.  C.  1950.  Loggerhead  Shrike.  Pages  131-148 
in  Life  histories  of  North  American  wagtails, 
shrikes,  vireos,  and  their  allies.  U.S.  National  Mu- 
seum Bulletin,  no.  197. 

Cade,  T.  J.  and  C.  P.  Woods.  1997.  Changes  in  dis- 
tribution and  abundance  of  the  Loggerhead 
Shrike.  Conservation  Biology  11:21—31. 
Chavez-Ramirez,  F,  D.  E.  Gawlik,  F.  G.  Prieto,  and 
R.  D.  Slack.  1994.  Effects  of  habitat  structure  on 
patch  use  by  Loggerhead  Shrikes  wintering  in  a 
natural  grassland.  Condor  96:228-231. 

Coblentz,  B.  E.  1980.  Effects  of  feral  goats  on  the 
Santa  Catalina  Island  ecosystem.  Pages  167-170 
in  The  California  Islands:  proceedings  of  a mul- 
tidisciplinary symposium  (D.  M.  Power,  Ed.). 
Santa  Barbara  Museum  of  Natural  History,  Santa 
Barbara,  California. 

Environmental  Systems  Research  Institute,  Inc. 
2000.  Arc  View  3.2a.  Environmental  Systems  Re- 
search Institute,  Redlands,  California. 

Esely,  J.  D.,  Jr.,  and  E.  K.  Bollinger.  2001.  Habitat 
selection  and  reproductive  success  of  Loggerhead 
Shrikes  in  northwest  Missouri:  a hierarchical  ap- 
proach. Wilson  Bulletin  113:290-296. 

Grinnell,  J.  1897.  Report  on  the  birds  recorded  during 
a visit  to  the  Islands  of  Santa  Barbara,  San  Nicolas 
and  San  Clemente  in  the  spring  of  1897.  Pasadena 
Academy  of  Sciences  1:1-26. 

Hooge,  P.  N.,  W.  Eichenlaub,  and  E.  Solomon.  1999. 
The  animal  movement  program.  U.S.  Geological 


Survey,  Alaska  Biological  Science  Center,  An- 
chorage, Alaska. 

Jorgensen,  P.  D.  and  H.  L.  Ferguson.  1984.  The  birds 
of  San  Clemente  Island.  Western  Birds  15:1 1 1 — 
130. 

Kay,  B.  J.,  L.  E.  Twigg,  T.  J.  Korn,  and  H.  I.  Nicol. 
1994.  The  use  of  artificial  perches  to  increase  pre- 
dation on  house  mice  ( Mus  domesticus)  by  rap- 
tors. Wildlife  Research  21:95-106. 

Kim,  D.  H.,  F.  Chavez-Ramirez,  and  R.  D.  Slack. 
2003.  Effects  of  artificial  perches  and  interspecific 
interactions  on  patch  use  by  wintering  raptors.  Ca- 
nadian Journal  of  Zoology  81:2038-2047. 

Lauver,  C.  L.,  W.  H.  Busby,  and  J.  L.  Whistler. 
2002.  Testing  a GIS  model  of  habitat  suitability 
for  a declining  grassland  bird.  Environmental 
Management  30:88-97. 

Linton,  C.  B.  1908.  Notes  from  San  Clemente  Island. 
Condor  10:82-86. 

Maloney,  R.  E,  R.  J.  Keedwell,  N.  J.  Wells,  A.  L. 
Rebergen,  and  R.  J.  Nilsson.  1999.  Effect  of  wil- 
low removal  on  habitat  use  by  five  birds  of  braid- 
ed rivers,  Mackenzie  Basin,  New  Zealand.  New 
Zealand  Journal  of  Ecology  23:53-60. 

Mills,  G.  S.  1979.  Foraging  patterns  of  kestrels  and 
shrikes  and  their  relation  to  an  optimal  foraging 
model.  Ph.D.  dissertation,  University  of  Arizona, 
Tucson. 

Rocklage,  A.  M.  and  J.  T.  Ratti.  2000.  Avian  use  of 
evolved  riparian  habitat  on  the  lower  Snake  River, 
Washington.  Northwest  Science  74:286-293. 

Schoener,  T.  W.  1971.  Theory  of  feeding  strategies. 
Annual  Review  of  Ecology  and  Systematics  2: 
369-404. 

Scott,  T.  A.  and  M.  L.  Morrison.  1990.  Natural  his- 
tory and  management  of  the  San  Clemente  Log- 
gerhead Shrike.  Western  Foundation  of  Vertebrate 
Zoology  4:23-57. 

Sokal,  R.  R.  and  F.  J.  Rohlf.  1981.  Biometry.  W.  H. 
Freeman  and  Company,  New  York. 

U.S.  Department  of  the  Navy,  Southwest  Division. 
2001.  San  Clemente  Island  integrated  natural  re- 
sources management  plan.  Prepared  by  Tierra 
Data  Systems,  Escondido,  California. 

Wolff,  J.  O.,  T.  Fox,  R.  R.  Skillen,  and  G.  Wang. 
1999.  The  effects  of  supplemental  perch  sites  on 
avian  predation  and  demography  of  vole  popula- 
tions. Canadian  Journal  of  Zoology  77:535-541. 

Yosef,  R.  1994.  Evaluation  of  the  global  decline  in  the 
true  shrikes  (Family  Laniidae).  Auk  1 1 1 :228-233. 

Yosef,  R.  and  T.  C.  Grubb,  Jr.  1992.  Territory  size 
influences  nutritional  condition  in  nonbreeding 
Loggerhead  Shrikes  ( Lanius  ludovicianus ):  a ptil- 
ochronology  approach.  Conservation  Biology  6: 
447_449. 

Yosef,  R.  and  T.  C.  Grubb,  Jr.  1994.  Resource  de- 
pendence and  territory  size  in  Loggerhead  Shrikes 
(Lanius  ludovicianus).  Auk  1 1 1 :465— 469. 


The  Wilson  Journal  of  Ornithology  1 18(3):34 1—352,  2006 


DO  AMERICAN  ROBINS  ACQUIRE  SONGS  BY  BOTH  IMITATING 

AND  INVENTING? 

STEVEN  L.  JOHNSON1 


ABSTRACT. — Although  the  majority  of  oscine  species  acquire  a song  repertoire  by  imitating  songs  they  have 
been  exposed  to,  some  species  also  improvise  and  invent  songs.  To  test  the  hypothesis  that  American  Robins 
{Turdus  migratorius ) both  imitate  and  invent  the  elements  of  their  whistle  songs,  I analyzed  the  song  repertoires 
of  wild  robins  at  three  locations  in  western  Massachusetts  and  the  song  development  of  five  tutor-trained  nestling 
robins.  Robins  appear  to  invent  or  improvise  most  of  the  elements  in  their  repertoires  (75-82%),  but  as  fledglings 
and  juveniles  they  acquire  the  remaining  elements  by  imitating  the  songs  of  neighboring  birds.  Received  29 
April  2005,  accepted  1 February  2006. 


Although  it  is  generally  agreed  that  bird- 
song serves  two  basic  functions,  mate  attrac- 
tion and  territory  maintenance  (Catchpole  and 
Slater  1995),  there  are  striking  differences  in 
how  various  songbirds  acquire  the  songs 
needed  for  these  functions.  In  many  species, 
young  males  imitate  only  conspecific  songs 
heard  during  a sensitive  period  of  song  ac- 
quisition (Marler  1981,  Catchpole  and  Slater 
1995).  In  contrast,  several  species  mimic  het- 
erospecific songs  (e.g.,  Northern  Mocking- 
bird, Mimus  polyglottos\  Howard  1974, 
Owen-Ashley  et  al.  2002).  Others  not  only 
mimic,  but  also  create  new  versions  of  song 
through  progressive  modification  of  previous- 
ly memorized  song,  known  as  improvisation, 
and/or  through  invention  of  entirely  new 
songs  unlike  anything  heard  by  the  young  bird 
(Marler  and  Peters  1982)  (e.g..  Gray  Catbird, 
Dumetella  carolinensis,  Kroodsma  et  al. 
1997).  There  are  also  species  that  rely  almost 
entirely  on  improvisation  or  invention  to  de- 
velop songs  (e.g..  Sedge  Wren,  Cistothorus 
platensis,  Kroodsma  et  al.  1999a).  While  im- 
itation and  mimicry  are  widespread  among  all 
taxa  with  vocal  learning  (e.g.,  dolphins,  Tyack 
1986;  hummingbirds,  Baptista  and  Schuch- 
mann  1990;  songbirds.  Nelson  et  al.  1995; 
parrots,  Hile  et  al.  2000),  improvisation  or  in- 
vention has  been  documented  in  only  a few 
songbird  species  (e.g..  Nightingale,  Luscinia 
megarhynchos,  Hultsch  and  Kopp  1989;  In- 
digo Bunting,  Passerina  cyanea , Payne  1996; 
Sedge  Wren,  Kroodsma  et  al.  1999a,  Hughes 


1 Graduate  Program  in  Organismic  and  Evolutionary 
Biology,  Dept,  of  Biology,  Univ.  of  Massachusetts, 
Amherst,  MA  01003,  USA;  e-mail: 
sjohnson@bio.umass.edu 


et  al.  2002)  and  possibly  the  signature  whis- 
tles of  dolphins  (Sayigh  1990). 

It  is  not  understood  why  some  species  im- 
provise or  invent  (Kroodsma  1996),  nor  is  it 
known  how  extensive  these  tendencies  are 
among  songbirds  or  how  many  times  they 
have  evolved.  A better  understanding  of  the 
selective  forces  for  improvising  and  inventing 
will  emerge  only  after  additional  species  are 
studied  and  only  after  life  history  traits  are 
correlated  to  particular  styles  of  song  devel- 
opment. A challenge  to  such  studies  is  that 
distinguishing  between  songs  generated  by 
improvisation,  invention,  or  inaccurate  imita- 
tion is  difficult  and  often  rather  subjective.  To 
distinguish  improvisation  from  invention,  the 
researcher  must  be  able  to  document  song  el- 
ements changing  over  time,  from  something 
closely  resembling  tutor  song  to  songs  that 
may  not  resemble  the  tutor  song  at  all.  If, 
however,  this  period  of  improvisation  is  oc- 
curring during  the  winter  months  when  a bird 
may  be  only  mentally  rehearsing  song,  it 
would  be  impossible  to  distinguish  between 
these  two  types  of  song  learning. 

It  has  been  suspected  that  American  Robins 
( Turdus  migratorius ) improvise  or  invent 
when  acquiring  song.  An  early  study  of  robin 
song  found  no  shared  song  elements  between 
any  of  the  wild  robins  studied,  even  among 
neighbors  (Konishi  1965).  Konishi  proposed 
two  possible  reasons  for  this  lack  of  shared 
elements:  (1)  young  robins  improvise  or  in- 
vent the  elements  of  their  repertoires  during 
the  song  acquisition  phase,  or  (2)  robins  learn 
through  imitation,  but  then  disperse  to  breed- 
ing grounds  where  their  song  elements  are 
unique  (Konishi  1965).  Later  studies  revealed 


341 


342 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


FIG.  1 . A representative  segment  of  American  Robin  song,  recorded  in  western  Massachusetts,  2002,  show- 
ing the  various  structural  units  and  their  associated  terms.  Notes  range  from  25  to  250  msec  in  length  and  have 
a frequency  range  of  300  to  1,500  Hz.  Elements  range  from  150  to  350  msec  in  length,  and  can  have  a frequency 
range  of  1,000  to  7,000  Hz  or  wider.  The  time  intervals  between  elements  (250  to  2,000  msec)  are  always  longer 
than  the  intervals  between  notes  within  an  element  (10  to  125  msec).  Whistle  elements  have  a narrow  frequency 
range  (mean  frequency  range  = 1.78  ± 0.03  kHz,  n = 46;  Dziadosz  1977),  with  individual  notes  ranging  from 
a low  frequency  of  1.5  kHz  to  a high  of  4 kHz  (Dziadosz  1977,  Tsipoura  1985;  SLJ  pers.  obs.).  Hisselly  elements 
have  a wider  frequency  range  (mean  frequency  range  = 4.74  ± 0.24,  n — 46  kHz,  Dziadosz  1977)  and  more 
rapid  frequency  modulation  (Konishi  1965).  Some  hisselly  elements  also  show  evidence  of  both  syrinxes  being 
used  simultaneously,  as  found  in  other  thrush  species. 


that  robins  shared  one  to  five  elements  with 
neighboring  robins  (Dziadosz  1977,  Thomas 
1979,  Tsipoura  1985,  Sousa  1999),  whereas 
most  elements  were  unique  (Tsipoura  1985). 
The  fact  that  robins  share  a few  elements  with 
close  neighbors  but  not  with  males  from  more 
distant  locations  (Dziadosz  1977,  Sousa  1999) 
suggests  that  the  shared  elements  are  imitated, 
but  that  the  unique  elements  are  either  impro- 
vised, invented,  or  learned  elsewhere.  Because 
of  the  difficulties  in  distinguishing  between 
improvisation  and  invention,  I refer  to  the 
song  learning  processes  of  robins  in  terms  of 
imitation  and  invention,  but  with  the  under- 
standing that  robins  may  actually  be  impro- 
vising some  song  elements.  Here  I provide  ev- 
idence that  robins  both  imitate  and  invent/im- 
provise song  elements,  based  on  research  with 
both  wild  populations  of  robins  and  hand- 
reared  nestlings. 

METHODS 

Description  of  robin  song. — The  song  of 
the  American  Robin  is  composed  of  sequenc- 
es of  “song  elements”  that  are  made  up  of 
one  or  more  “notes”  shown  as  continuous 
markings  on  a spectrogram  (Fig.  1).  Male  rob- 


ins sing  two  song  element  types  (Konishi 
1965,  Dziadosz  1977,  Hsu  1991).  The  more 
common  is  the  familiar  whistle-like  song  usu- 
ally described  as  some  variation  of  cheerily, 
cheer  up,  cheer  up,  cheerily,  cheer  up  (Sal- 
labanks  and  James  1999).  These  elements 
generally  sound  like  clear  whistles,  but  can 
blend  into  buzzes  or  trills.  Male  robins  typi- 
cally have  between  6 and  25  whistle  elements 
in  their  repertoires  (Sallabanks  and  James 
1999;  SLJ  unpubl.  data).  The  second  type  of 
element,  described  as  the  hisselly,  or  whisper, 
song  (W.  M.  Tyler,  as  quoted  in  Bent  1949, 
and  Young  1955,  respectively),  is  generally 
sung  very  softly  and  has  a much  more  com- 
plex structure.  Robins  tend  to  combine  both 
whistle  and  hisselly  elements  to  form  groups 
typically  consisting  of  3-8  elements  (Fig.  2). 
Although  robins  have  a larger  repertoire  of 
hisselly  than  whistle  elements,  they  typically 
sing  whistle  elements  5 to  10  times  more  fre- 
quently than  hisselly  elements  (Konishi  1965; 
SLJ  unpubl.  data).  Therefore,  I chose  to  look 
for  evidence  of  imitation  and  invention  in  the 
whistle  elements  of  both  wild  and  hand-reared 
robins. 

Recording  and  analyzing  songs  of  wild  rob- 


Johnson  • ROBINS  IMITATE  AND  INVENT  SONGS 


343 


N 

X 


10 

9 

8 

7 ■ 
6 


>,  5 
o 

c 4 
0 
D 

cr  3 
0 

it  2 
1 
0 


Hisselly 

* 


Whistle  i 

rS 

* 


Hisselly 

rh 


Hisselly 

rS 


Whistle 
Whistle ii 

rS'T 


•|h 


Whistle 

rS 


If 


group 


Whistle 

r*- 

A. 


Time  (sec) 

FIG.  2.  Spectrogram  showing  the  typical  grouping  of  song  elements  by  an  American  Robin  in  western 
Massachusetts.  Robins  combine  both  whistle  and  hisselly  elements  to  form  groups  typically  consisting  of  3-8 
elements. 


ins. — I recorded  the  pre-dawn  song  of  42  male 
robins  throughout  the  2002  breeding  season  at 
three  locations  in  Hampshire  County,  western 
Massachusetts:  16  birds  at  the  Quabbin  Cem- 
etery (42°  16'  48"  N,  72°  18'  32"  W),  16  birds 
at  Mt.  Pollux  Conservation  Area  (42°  19'  39" 
N,  72°  30'  06"  W),  and  1 1 birds  at  Wildwood 
Cemetery  (42°  23'  23"  N,  72°  30'  44"  W).  The 
three  sites  were  between  6 and  21  km  apart 
and  consisted  of  open,  mowed  grassy  areas 
with  trees,  shrubs,  and  wooded  edges.  From 
18  April  through  4 August  2002,  I recorded 
twice  per  week  at  each  of  the  three  sites,  be- 
ginning each  day  with  the  first  robin  song 
heard,  generally  1-2  hr  before  sunrise,  and 
ending  at  the  first  lull  in  singing  after  sunrise. 
Recording  typically  began  at  approximately 
04:30  EST  and  ended  before  07:00.  Record- 
ings were  made  with  a Marantz  PMD430  ste- 
reo cassette  recorder  and  two  Sennheiser 
ME62  microphones  mounted  on  a Dan  Gibson 
or  a Telinga  parabola.  I attempted  to  record 
all  the  robins  singing  at  each  site  each  day  and 
recorded  two  birds  at  a time  whenever  possi- 
ble. I attempted  to  focus  on  any  birds  for 
which  I had  fewer  recordings  (i.e.,  less  vocal 
individuals),  and  generally  limited  my  record- 
ings of  the  more  vocal  birds  to  20  to  30  min 
each  day. 

I cataloged  the  song  repertoires  of  individ- 
ual birds  by  using  field  recordings  made  be- 
tween 18  April  and  16  May.  During  this  pe- 


riod, I recorded  1 to  29  bouts  per  bird  (mean 
= 8.5),  with  total  recording  time  per  bird 
ranging  from  3 to  218  min  (mean  = 46  min). 
Because  the  robins  were  not  banded  and  I 
conducted  most  recording  when  it  was  dark,  I 
relied  on  the  precise  recording  locations  and 
the  recordings  themselves  to  determine  indi- 
vidual repertoires.  I began  by  noting  the  lo- 
cation of  each  bird  as  I recorded  it,  and  then 
I determined  the  repertoire  of  song  elements 
for  each  individual  recording.  I digitized  the 
recordings  (sample  rate  = 23,952.1  Hz)  and 
then  printed  continuous  spectrograms  through 
Signal  sound  analysis  software  (Beeman 
2003)  with  the  settings  as  follows:  transform 
length  = 256  points,  frequency  resolution  = 
93.6  Hz,  time  resolution  = 10.7  msec,  and 
number  of  transformations  = 2000.  From  the 
spectrogram  of  each  recording,  I determined 
the  song  element  repertoire.  The  repertoires 
were  very  distinct,  each  being  a unique  com- 
bination of  song  elements  primarily  composed 
of  elements  found  in  no  other  repertoire.  An- 
other distinct  feature  of  each  repertoire  was 
the  order  in  which  the  elements  were  sung. 
During  each  recording  of  a specific  repertoire, 
certain  element  combinations  were  sung  much 
more  than  would  be  expected  by  chance;  these 
combinations  were  very  distinct  and  consis- 
tent over  time.  I also  found  that  each  reper- 
toire of  song  elements  was  sung  only  in  a 
small  portion  of  the  recording  site.  I recorded 


344 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


each  repertoire  repeatedly  within  a specific 
area,  and  these  areas  corresponded  to  approx- 
imate territories  of  robins  observed  after  sun- 
rise. 

To  verify  that  I had  sufficient  samples  of 
each  individual  to  allow  me  to  determine  com- 
plete repertoires,  I randomly  selected  200  sec 
of  recording  from  each  bird  for  which  I had 
ample  recordings,  (and  180  sec  from  the  one 
bird  for  which  I had  only  3 min  of  recording), 
and  next  plotted  the  number  of  different  ele- 
ments sung  over  time.  In  each  case,  element 
diversity  reached  an  asymptote  after  50  to  100 
sec,  suggesting  that  the  complete  repertoire 
was  revealed.  My  results  were  similar  to  those 
of  Konishi  (1965),  who  found  that  American 
Robin  repertoires  were  usually  exhausted  ev- 
ery 100  elements.  During  the  robin’s  pre-dawn 
chorus,  an  individual  will  typically  sing  100 
elements  in  under  100  sec.  The  number  of 
song  elements  revealed  within  each  of  the 
200-sec  samples  was  the  same  as  the  number 
of  elements  found  for  that  individual  through- 
out the  total  recordings  made  during  the  first 
half  of  the  breeding  season,  and,  in  most  cas- 
es, throughout  the  entire  breeding  season. 
Therefore,  I feel  confident  that  I had  deter- 
mined the  complete  repertoire  of  each  bird 
sampled. 

Next,  I printed  representative  spectrograms 
(11  X 14  cm)  of  all  song  elements  in  each 
bird’s  repertoire  from  the  best-quality  record- 
ings. Only  a few  of  the  elements  showed  any 
variability,  and  these  were  represented  by 
multiple  spectrograms.  To  assess  repertoire 
overlap  among  males,  five  naive  observers 
were  provided  with  a total  of  315  spectro- 
grams representing  the  song  elements  from  all 
the  recorded  repertoires.  Observers  laid  out  all 
spectrograms  and  sorted  the  images  by  gen- 
eral similarities  before  searching  for  matching 
pairs  of  song  elements,  which  generally  took 
8 to  10  hr.  Identified  pairs  were  then  scored — 
rating  their  similarity  on  a six-level  scale  (0 
to  5) — according  to  written  instructions  spec- 
ifying the  criteria  for  each  level.  A simplified 
version  of  the  criteria  follows:  0 = no  simi- 
larity; 1 = elements  have  same  general  char- 
acter, but  <20%  overlap;  2 = elements  have 
some  similarity,  20-49%  overlap;  3 = ele- 
ments are  similar,  50-79%  overlap;  4 = ele- 
ments are  very  similar,  80—90%  overlap;  5 = 


elements  essentially  the  same,  91-100%  over- 
lap. 

Because  of  the  large  number  of  potential 
comparisons,  it  was  rare  for  all  observers  to 
identify  a specific  match;  instead,  typically 
two  to  four  observers  noted  a given  match.  To 
ensure  that  the  identified  matches  did  repre- 
sent very  similar  song  elements,  I and  one  of 
the  original  observers  scored  each  match  iden- 
tified by  one  or  more  naive  observers,  and  re- 
jected any  matches  that  did  not  receive  a score 
of  3 or  higher  from  both  of  us. 

To  determine  whether  robins  change  their 
song  elements  or  repertoires  within  the  breed- 
ing season,  I also  evaluated  repertoires  in  a 
second  set  of  recordings  made  from  18  June 
through  4 August  2002.  I compared  the  ele- 
ments in  the  repertoires  for  each  individual 
recorded  during  these  later  periods  to  the  rep- 
ertoires from  the  beginning  of  the  2002  breed- 
ing season. 

Analyzing  repertoire  development  in  hand- 
reared  robins. — In  July  2002,  I collected  14 
nestling  robins  (4  to  14  days  old)  from  six 
nests  in  Hampshire,  Franklin,  and  Berkshire 
counties,  Massachusetts.  The  nestlings  were 
hand-reared  in  an  animal  care  facility  at  the 
University  of  Massachusetts,  Amherst,  where 
they  were  fed  a diet  adapted  from  Lanyon 
(1979).  Nest  mates  were  initially  raised  to- 
gether in  the  same  cages.  Soon  after  the  young 
robins  fledged,  I placed  each  bird  in  its  own 
cage  and  divided  the  birds  into  two  groups  of 
seven,  separating  siblings  as  much  as  possible 
and  attempting  to  create  similar  sex  ratios  in 
the  two  groups.  The  apparent  sex  of  each  bird 
was  based  on  the  intensity  of  plumage  color 
on  the  head  and  breast.  Male  robins  generally 
have  darker  plumage  in  both  of  these  regions. 
There  were  four  apparent  males  in  Group  1, 
and  three  males  in  Group  2.  Because  female 
American  Robins  also  sing  occasionally 
(Wauer  1999),  I monitored  all  birds.  Each 
group  was  housed  in  a separate  isolation 
chamber  (Acoustic  Systems,  Austin,  Texas), 
and  experienced  daily  periods  of  illumination 
mimicking  the  natural  photoperiod. 

Each  group  of  robins  was  exposed  to  four 
tutor  tapes,  each  containing  the  songs  of  a dif- 
ferent wild  robin.  I created  each  tape  from  ap- 
proximately 10  min  of  high-quality  recording 
from  one  of  four  robins  recorded  in  Amherst, 
Massachusetts.  Each  recording  was  repeated 


Johnson  • ROBINS  IMITATE  AND  INVENT  SONGS 


345 


four  to  five  times  to  fill  one  45-min  side  of  a 
cassette  tape.  The  tapes  were  broadcast  over 
two  periods.  The  first  tutor  period  began  in 
August  2002,  soon  after  the  youngest  birds 
fledged,  at  which  time  they  ranged  in  age  from 
14  to  40  days;  each  group  was  exposed  to  two 
of  the  four  tutor  tapes  during  this  period.  On 
alternating  days,  tapes  1 and  2 were  played  in 
Chamber  1 , and  tapes  3 and  4 were  played  in 
Chamber  2.  Tapes  were  played  for  the  first  30 
min  of  each  daylight  period  and  for  15  min  at 
the  end  of  the  day.  Each  robin  heard  tutor 
song  for  75  days  during  this  first  period. 

The  second  tutor  period  began  in  early  Feb- 
ruary 2003,  at  which  time  I switched  the  tapes 
between  the  two  chambers,  exposing  the 
young  birds  to  new  song  elements.  The  goal 
of  exchanging  the  tapes  was  to  evaluate 
whether  the  robins  imitated  sounds  heard  in 
their  first  spring  as  sub-adults.  The  young 
birds  began  singing  on  day  21  of  this  tutor 
period.  I continued  to  play  the  tutor  tapes  for 
5 more  days  and  then  began  recording  the 
young  birds. 

Using  a preamplifier  and  two  microphones, 
I recorded  the  young  birds  with  a Nakamichi 
DR-3  cassette  deck.  To  reduce  the  chances  of 
recording  birds  other  than  the  focal  subject,  I 
placed  5-cm  acoustic  foam  around  each  mi- 
crophone and  cage,  and,  when  recording  qui- 
eter birds,  I removed  louder  birds  from  the 
chamber.  The  young  birds  were  recorded  for 
two  30-min  periods  each  day:  the  first  30  min 
of  daylight  and  30  min  after  feeding,  when  the 
birds  often  increased  their  rate  of  vocalization. 
I recorded  the  birds  for  62  days  from  late  Feb- 
ruary to  early  May. 

Five  of  the  birds  identified  as  males  pro- 
duced song  elements  similar  to  those  of  wild 
robins;  the  remaining  birds  made  only  call 
notes.  Four  of  the  singing  birds  were  in  Group 
1 , and  one  was  in  Group  2.  Two  of  the  singing 
males  in  Group  1 were  nest  mates,  while  a 
third  bird  had  a nest  mate  in  Group  2.  The 
song  elements  in  each  bird’s  repertoire  re- 
mained stable  throughout  the  2.5-month  re- 
cording period,  and  so  appeared  to  represent 
crystallized  song. 

I digitized  the  recordings  of  the  hand-reared 
birds  and  the  tutor  tapes,  sampling  at  a rate  of 
20,000  Hz.  I selected  a representative  example 
of  each  song  element  from  each  robin,  and 
printed  spectrograms  using  the  same  methods 


described  above  for  the  field  recordings.  Five 
naive  observers  compared  331  representative 
spectrograms  from  the  hand-reared  and  tutor 
repertoires.  The  same  conditions  and  criteria 
for  scoring  similarity  were  followed  as  de- 
scribed above. 

To  determine  whether  the  young  robins  had 
imitated  adult  song  heard  near  their  nest  sites 
prior  to  capture,  I compared  each  young  bird’s 
repertoire  to  that  of  adult  robins  (n  — 3 to  6) 
from  each  nest  site,  as  assessed  from  record- 
ings made  on  the  morning  of  capture  or  the 
day  after.  Representative  spectrograms  were 
printed  and  scored  for  similarity  by  two  naive 
observers,  as  described  above.  Means  are  pre- 
sented ± SD. 

RESULTS 

Element  similarity,  repertoire  delivery,  and 
stability  in  wild  robins. — Males  from  the  same 
sites  shared  more  song  elements  than  those 
from  different  sites  (Mann-Whitney  test:  P < 
0.001,  n = 42),  suggesting  that  robins  imitate 
some  of  the  elements  of  local  robins.  The  na- 
ive observers  identified  59  element  pairs  out 
of  a possible  49,455  pairs,  for  which  a major- 
ity of  observers  gave  a similarity  score  of  3 
or  higher.  Fifty-six  of  these  identified  pairs 
represented  birds  from  the  same  recording 
site;  their  average  similarity  score  was  3.7. 
The  remaining  three  pairs  represented  ele- 
ments recorded  at  different  locations;  no  ob- 
server, however,  gave  a score  higher  than  3 
for  these  pairs,  and  their  average  similarity 
score  was  2.3.  All  matches  found  between 
multiple  representatives  of  a single  element 
type  from  within-bird  repertoires  were  scored 
4 or  higher  by  the  observers.  Thirty-six  of  the 
42  birds  shared  elements  with  other  birds 
within  their  site.  The  percentage  of  elements 
in  a bird’s  repertoire  that  were  similar  to  ele- 
ments in  other  repertoires  at  the  same  site 
ranged  from  0 to  50%  (mean  = 25  ± 15% 
SD).  In  contrast,  only  five  birds  had  elements 
that  were  judged  as  similar  to  elements  of 
birds  from  different  locations  (Fig.  3).  In  each 
bird’s  repertoire,  the  percentage  of  elements 
that  were  similar  to  elements  in  the  repertoires 
of  birds  from  different  sites  ranged  from  0 to 
16.6%. 

Most  elements  within  each  bird’s  repertoire 
were  judged  to  be  unique  to  that  individual 
(mean  = 75  ± 15%  SD),  indicating  that  the 


346 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


FIG.  3.  Comparison  of  the  percent  of  each  American  Robin  repertoire  shared  within  and  between  three  sites 
in  western  Massachusetts,  2002.  Each  bar  represents  a single  robin’s  repertoire.  American  Robins  share  far  more 
elements  with  neighboring  robins  than  with  robins  from  different  sites.  The  percent  of  shared  elements  in  the 
repertoires  of  42  robins  is  shown  for  both  within  and  between  sites.  Note  that  37  of  42  birds  share  0%  of  their 
repertoire  with  birds  from  other  sites. 


robins  either  invented  most  of  their  song  ele- 
ments, learned  them  elsewhere,  or  learned 
them  from  a bird  no  longer  present.  In  later 
recordings,  these  unique  elements  made  it 
possible  to  identify  each  bird  by  its  songs 
alone.  The  repertoires  recorded  during  both 
the  early  and  late  periods  retained  the  majority 
(mean  = 98  ± 14%;  n = 15  birds)  of  their 
elements  throughout  the  entire  season.  How- 
ever, the  repertoires  of  six  well-sampled  birds 
(>440  sec  of  recording  each  period)  did  ap- 
pear to  change.  One  to  two  elements  were 
added  to  two  repertoires,  and  one  to  four  el- 
ements were  dropped  from  four  repertoires. 
Two  of  these  fluctuations  may  have  been  ar- 
tifacts of  unequal  recording  time  between  the 
two  periods  (i.e.,  the  increase  or  decrease  in 
repertoire  size  paralleled  the  increase  or  de- 
crease in  sample  size  between  the  two  time 
periods),  but  the  remaining  four  repertoire 
changes  trend  in  the  opposite  direction  from 
changes  in  the  sample  sizes  between  the  two 
periods.  For  example,  four  of  the  elements  in 
bird  W3’s  early  repertoire  were  missing  in  the 
later  repertoire,  despite  an  increase  in  record- 
ing time.  Conversely,  a new  element  was 
found  in  the  late  repertoire  of  Q3,  despite  a 
97%  reduction  in  recording  time. 

Some  robins  clearly  modified  individual  el- 


ements over  the  course  of  the  breeding  season. 
Birds  P6  and  Q5  each  sang  one  element  that 
changed  over  the  course  of  the  breeding  sea- 
son (Fig.  4).  In  both  cases,  the  new  form  com- 
pletely replaced  the  old  form.  What  was  par- 
ticularly striking  about  the  change  in  Q5’s 
case  was  that  the  later  version  was  a much 
closer  match  to  elements  in  three  other  rep- 
ertoires from  the  same  location  (Fig.  5). 

Song  learning  in  hand-reared  robins. — The 
tape-tutoring  experiment  provided  evidence  of 
both  invention  and  imitation  during  song 
learning.  The  percentage  of  shared  elements 
varied  greatly  among  the  five  hand-reared  rob- 
ins that  produced  song.  Two  nest  mates  shared 
between  55.5  and  65%  of  their  repertoires 
with  each  other,  two  other  birds  in  this  group, 
and  the  tutor  tapes,  whereas  there  were  fewer 
shared  elements  in  repertoires  of  the  remain- 
ing three  birds  (range  = 0-30%,  mean  = 14 
± 15%  SD).  There  was  almost  no  evidence  of 
imitation  of  songs  heard  at  the  nest;  one  ele- 
ment of  a single  hand-reared  bird  was  consid- 
ered similar  (average  score  3)  to  an  element 
recorded  at  that  bird’s  nest  site.  These  may 
have  matched  by  chance,  since  both  elements 
were  simple  descending  whistles. 

The  remaining  elements  produced  by  the 
five  birds  did  not  match  elements  from  the 


Johnson  • ROBINS  IMITATE  AND  INVENT  SONGS 


347 


N 


X 

> 

o 

c 

CD 

=3 

CT 

0 


Time  (sec) 


FIG.  4.  Modifications  of  song  elements  over  time  from  two  wild  American  Robins  (P6,  element  N;  Q5, 
element  B).  Subjects  were  recorded  in  April  and  July  2002  in  western  Massachusetts. 


nest  sites,  the  tutor  tapes,  or  other  hand-reared 
birds,  suggesting  that  the  unique  elements 
were  either  improvised  or  invented  (Marler 
and  Peters  1982,  Nowicki  et  al.  2002).  I com- 
pared examples  of  these  elements  at  different 
times  throughout  the  62-day  recording  period 
and  found  no  change  over  time,  suggesting 
that  the  unique  elements  were  invented,  rather 
than  improvised;  however,  I cannot  eliminate 


the  possibility  that  the  young  birds  improvised 
changes  during  the  winter  silent  period  or  be- 
fore I began  recording.  I also  compared  the 
elements  produced  by  the  hand-reared  birds  to 
spectrograms  of  Konishi’s  (1965)  isolated  and 
deafened  robins.  I found  that  the  elements 
produced  by  my  hand-reared  birds  showed  lit- 
tle or  no  within-element  variability  and  con- 
sisted of  whistle  notes  similar  to  those  of  wild 


Q5B 

Q8G 

Q12L 

Q16C 

?\j\ 



a 

0 0.5  1.0  1.5  2.0 

Time  (sec) 


FIG.  5.  Song  elements  of  four  American  Robins  recorded  at  the  Quabbin  Cemetery  in  Hampshire  County, 
western  Massachusetts,  2002.  The  late  (July)  version  of  bird  Q5’s  element  B is  a closer  match  to  elements  in 
three  local  birds’  repertoires  than  the  early  (April)  version  of  bird  Q5’s  element  B in  Figure  4. 


348 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


TABLE  1.  The  number  of  song  elements  that  four 
hand-reared  birds  (A2,  FI,  Dl,  and  D2  in  columns) 
within  one  isolation  chamber  shared  among  them- 
selves and  two  tutor  tapes  (T1 A and  TIB).  The  highest 
incidence  of  sharing  was  between  hand-reared  siblings 
Dl  and  D2.  FI  did  not  share  any  elements  with  two 
siblings  raised  in  a separate  chamber.  All  birds  were 
reared  and/or  recorded  in  western  Massachusetts, 
2002. 


Bird  ID 

A2 

FI 

Dl 

D2 

T1A 

1 

1 

0 

0 

TIB 

0 

1 

0 

0 

A2 

— 

0 

1 

0 

FI 

0 

— 

3 

2 

Dl 

1 

3 

— 

15 

D2 

0 

2 

15 

— 

robins,  whereas  Konishi’s  birds  produced 
songs  with  a high  degree  of  within-element 
variability;  elements  consisted  of  wavering 
whistle  notes.  This  suggests  that  the  song  el- 
ements produced  by  the  hand-reared  birds 
were  fully  crystallized,  invented/improvised 
songs,  rather  than  the  basic  acoustic  features 
of  song  that  can  be  produced  by  isolated  birds. 

Although  most  of  the  elements  were  in- 
vented/improvised, imitation  was  also  evident 
in  four  of  the  young  birds’  repertoires.  The 
young  birds  tended  to  share  more  elements 
with  other  hand-reared  birds  than  with  the  tu- 
tor tapes  (Table  1).  The  naive  observers  iden- 
tified 24  pairs  of  elements,  the  average  simi- 
larity scores  of  which  were  >3,  indicating  a 
high  degree  of  similarity.  Fifteen  of  the  24 
identified  pairs  were  between  two  siblings 
housed  in  the  same  chamber  (see  Fig.  6 for 
examples).  Two  of  the  elements  shared  by 
these  siblings  were  also  sung  by  non-siblings 
housed  within  the  same  chamber.  Six  pairs 
were  between  non-siblings  within  the  same 
chamber,  and  three  pairs  were  between  tutors 
and  young  birds  (see  Fig.  7 for  example).  The 
imitated  tutor  elements  were  from  tapes 
played  only  during  the  first  tutoring  period, 
whereas  the  elements  shared  between  birds 
could  not  have  been  heard  until  the  birds  were 
old  enough  to  sing.  No  elements  were  shared 
between  the  birds  in  Group  1 and  the  single 
singing  bird  in  Group  2,  even  though  this  bird 
had  two  male  siblings  in  Group  1. 

The  percentage  of  shared  elements  in  each 
bird’s  repertoire  varied  greatly.  Bird  A2 


shared  30%  of  its  repertoire.  Bird  Dl  65%, 
Bird  D2  55.5%,  Bird  FI  13%,  and  Bird  F2 
0%  (mean  = 32.8  ± 27.5%  SD).  The  degree 
of  sharing  in  A2,  FI,  and  F2  falls  within  the 
range  of  sharing  I found  for  wild  robins;  how- 
ever, that  of  the  siblings  D 1 and  D2  was  much 
greater  due  to  the  percentage  of  elements  they 
shared  with  each  other  (63%  and  42%,  re- 
spectively). 

DISCUSSION 

The  field  recording  and  tape-tutoring  com- 
ponents of  this  study  indicate  that  American 
Robins  can  and  do  imitate  song  elements. 
Among  repertoires  of  wild  robins,  closely 
matching  song  elements  were  found  within 
sites,  but  only  weak  similarities  were  found 
between  sites,  indicating  that  the  matching  el- 
ements were  imitated.  Additional  evidence  of 
imitation  was  found  in  the  case  of  one  bird  at 
the  Quabbin  site  that  changed  one  element  to 
more  closely  match  an  element  shared  by 
three  other  birds  from  that  site,  indicating  that 
robins  can  change  their  repertoires  to  match 
other  birds.  Because  the  ages  of  the  recorded 
robins  were  not  known,  it  has  yet  to  be  deter- 
mined whether  this  ability  is  restricted  to  the 
first  breeding  season. 

A similar  pattern  was  found  in  the  reper- 
toires of  hand-reared  birds,  which  together 
produced  three  close  matches  to  elements 
from  tutor  tapes.  In  addition,  birds  kept  within 
a single  chamber  produced  21  closely  match- 
ing elements,  but  there  were  no  matching  el- 
ements between  birds  raised  in  separate  cham- 
bers. The  fact  that  the  21  matching  elements 
between  birds  could  not  have  been  learned  un- 
til the  birds  began  singing  also  supports  the 
idea  that  adult  robins — at  least  in  their  first 
breeding  season — can  change,  or  add  to,  their 
repertoires.  Closely  related  Blackbirds  ( Tur - 
dus  merula)  also  appear  to  continue  learning 
songs  as  adults  (Rasmussen  and  Dabelsteen 
2002).  A possible  limitation  on  the  interpre- 
tation of  these  results  is  that  tutor  tapes,  rather 
than  live  tutors,  were  used,  and  the  stimulus 
of  live  tutors,  as  experienced  in  nature,  may 
elicit  a higher  degree  of  imitation. 

Robins  may  have  a tendency  to  learn  song 
elements  that  are  heard  more  often,  either  be- 
cause they  are  sung  by  multiple  birds,  or  are 
sung  by  a highly  vocal  bird.  My  data  offer 
some  support  for  this  tendency.  Two  of  the 


Johnson  • ROBINS  IMITATE  AND  INVENT  SONGS 


349 


Time  (sec) 

FIG.  6.  Four  examples  of  song  element  sharing  between  three  hand-reared  American  Robins  raised  in  one 
chamber  in  western  Massachusetts,  2002.  Birds  D1  and  D2  are  brothers  and  shared  more  elements  than  any 
other  hand-reared  birds.  The  lower  two  elements  were  shared  only  by  D1  and  D2,  not  by  FI. 


song  elements  sung  by  the  hand-reared  robins 
were  shared  by  three  individuals,  and  many  of 
the  elements  shared  by  wild  robins  were 
shared  by  three  or  more  individuals.  It  also 
appears  that  one  wild  robin  altered  one  ele- 
ment in  his  repertoire  to  more  closely  match 
that  of  three  other  robins  within  his  particular 
recording  area. 

Robins  also  appear  to  invent  or  improvise 
song  elements.  The  majority  of  elements  pro- 
duced by  the  tape-tutored  birds  were  unique 


for  each  individual,  indicating  that  the  ele- 
ments were  invented/improvised  by  the  tu- 
tored birds.  The  majority  of  elements  in  the 
wild  robin  repertoires  were  also  unique  to 
each  individual,  which  suggests  that  invention 
or  improvisation  also  could  be  involved  in 
song  acquisition  in  the  wild.  However,  I can- 
not rule  out  the  possibility  that  at  least  some 
of  these  elements  may  have  been  learned  else- 
where or  from  birds  no  longer  present  at  the 
local  site. 


350 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


N 

X 

o 

c 

0 

D 

CT 

0 

i_ 

LL 


Tutor  1 A Bird  A2 


4 


1 

0 

0 0.5  1.0 

Time  (sec) 


FIG.  7.  Example  of  song  element  matching  between  tutor  tape  1A  and  hand-reared  American  Robin  A2, 
western  Massachusetts,  2002. 


My  results  are  not  completely  consistent 
with  either  of  Konishi’s  (1965)  hypotheses  on 
robin  song  development.  Konishi  found  no 
evidence  of  element  matching,  and  he  ex- 
plained this  by  suggesting  that  either  robins 
improvise/invent  the  elements  of  their  reper- 
toires during  song  acquisition,  or  they  learn 
through  imitation  and  then  disperse  to  breed- 
ing grounds  where  their  song  elements  are 
unique  (Konishi  1965).  My  results  suggest 
that  robins  do  improvise/invent  songs,  but 
also  imitate  songs  of  nearby  robins,  and  that 
these  imitations  occur  during  both  early  song 
acquisition  and  after  robins  settle  on  breeding 
territories,  allowing  adult  birds  to  share  song 
elements  with  local  males. 

Song  sharing  plays  an  important  role  in  the 
communication  of  several  species.  For  exam- 
ple, neighboring  males  in  many  species  song- 
match  during  territory  defense  as  a warning 
of  potential  escalation  (Krebs  et  al.  1981,  Falls 
et  al.  1982,  Beecher  et  al.  2000a).  A benefit 
of  this  system  is  illustrated  in  Song  Sparrows 
by  the  positive  correlation  between  how  long 
a male  holds  a territory  and  his  ability  to  share 
songs  with  his  neighbors  (Beecher  et  al. 
2000b).  Robins  also  may  benefit  from  sharing 
elements  in  their  repertoire;  although  they 
may  not  song-match,  most  robins  sing  the 
shared  elements  in  their  repertoire  more  than 
would  be  expected  by  chance  (SLJ  unpubl. 
data).  It  is  also  worth  noting  that  only  three 


robins  recorded  during  the  first  third  of  the 
breeding  season  did  not  share  elements  with 
other  birds  at  their  sites,  and  that  none  of  these 
birds  could  be  found  in  the  last  third  of  the 
season. 

The  results  of  my  tape-tutoring  experiment 
indicated  that  social  interaction  with  live  birds 
provided  stronger  stimulation  for  imitation 
than  tutor  tapes — as  found  in  many  studies 
(e.g.,  Beecher  1996),  suggesting  that  the  ben- 
efit of  sharing  elements  is  tied  to  social  inter- 
actions. A particularly  interesting  result  of  this 
experiment  is  the  high  percentage  of  element 
sharing  between  the  two  siblings  with  visual 
and  acoustical  access  to  each  other.  This  con- 
trasts with  the  lower  percentage  of  sharing 
with  other,  equally  accessible  birds  in  the 
same  chamber,  and  with  the  complete  lack  of 
sharing  between  the  siblings  raised  in  differ- 
ent chambers.  It  appears  unlikely  that  this 
high  degree  of  sharing  is  a result  of  songs 
learned  and  imitated  from  parents  or  neigh- 
bors during  the  nestling  period.  One  possible 
interpretation  is  that  there  is  a predisposition 
to  learn  from  one’s  relatives  (Nelson  and  Mar- 
ler  2005).  Further  research  into  the  social  in- 
teractions between  adult  and  fledgling  robins, 
particularly  between  closely  related  birds, 
may  provide  additional  clues  to  the  impor- 
tance of  shared  elements  in  American  Robins. 

Why  American  Robins  both  imitate  and  in- 
vent during  song  development  remains  a mys- 


Johnson  • ROBINS  IMITATE  AND  INVENT  SONGS 


351 


tery.  A key  to  unraveling  this  mystery  is  the 
fact  that  song  development  evolves  in  re- 
sponse to  selection  pressures  brought  about  by 
other  life-history  traits  (Kroodsma  1983).  For 
example,  some  highly  migratory  or  nomadic 
species  tend  to  improvise  or  invent  a higher 
percentage  of  their  songs  than  closely  related 
species  and  subspecies  that  are  non-migratory 
and/or  exhibit  greater  philopatry  (Kroodsma 
et  al.  1999a,  b;  Nelson  et  al.  2001;  Handley 
and  Nelson  2005).  We  can  address  the  ques- 
tion of  why  a species  invents  and/or  imitates 
by  looking  for  correlations  between  song  de- 
velopment and  life-history  traits  (e.g.,  migra- 
tory status,  philopatry)  among  closely  related 
groups  (e.g..  Read  and  Weary  1992,  Nelson  et 
al.  1995).  The  American  Robin,  with  seven 
subspecies,  including  one  that  is  non-migra- 
tory,  promises  to  be  an  excellent  subject  for 
such  a comparative  study.  With  65  congeners 
(Phillips  1991),  the  robin  could  also  be  part 
of  a much  broader  study  that  incorporates  a 
wide  range  of  traits  in  song  development  and 
life  history. 

ACKNOWLEDGMENTS 

I thank  D.  E.  Kroodsma  and  B.  E.  Byers  for  their 
invaluable  advice  and  Jeff  Podos  for  providing  the 
acoustic  isolation  chambers  and  room  for  raising 
young  robins.  I also  wish  to  thank  L.  Johnson,  S.  Hub- 
er, K.  Belinsky,  J.  Southall,  C.  Kennedy,  M.  Miller, 
and  J.  Claude  Razafimahaimodison  for  their  long  hours 
searching  for  element  matches.  This  project  was  fund- 
ed in  part  by  a GAANN  Fellowship  Grant.  The  Mas- 
sachusetts Department  of  Fish  and  Wildlife  and  the 
U.S.  Fish  and  Wildlife  Service  provided  scientific  col- 
lecting permits.  I also  wish  to  thank  three  anonymous 
reviewers  for  their  comments  and  suggestions. 

LITERATURE  CITED 

Baptista,  L.  F.  and  K.  L.  Schuchmann.  1990.  Song 
learning  in  the  Anna  Hummingbird  ( Calypte 
anna).  Ethology  84:15-26. 

Beecher,  M.  D.  1996.  Birdsong  learning  in  the  labo- 
ratory and  field.  Pages  61-78  in  Ecology  and  evo- 
lution of  acoustic  communication  in  birds  (D.  E. 
Kroodsma  and  E.  H.  Miller,  Eds.).  Cornell  Uni- 
versity Press,  Ithaca,  New  York. 

Beecher,  M.  D.,  S.  E.  Campbell,  J.  M.  Burt,  C.  E. 
Hill,  and  J.  C.  Nordby.  2000a.  Song-type  match- 
ing between  neighbouring  Song  Sparrows.  Animal 
Behaviour  59:29-37. 

Beecher,  M.  D.,  S.  E.  Campbell,  and  J.  C.  Nordby. 
2000b.  Territory  tenure  in  Song  Sparrows  is  re- 
lated to  song  sharing  with  neighbours,  but  not  to 
repertoire  size.  Animal  Behaviour  59:29-37. 


Beeman,  K.  2003.  Signal  for  Windows,  ver.  4.02.03e. 
Engineering  Design,  Berkeley,  California. 

Bent,  A.  C.  1949.  Life  histories  of  North  American 
thrushes,  kinglets,  and  their  allies.  U.S.  National 
Museum  Bulletin,  no.  196.  Smithsonian  Institu- 
tion, Washington,  D.C. 

Catchpole,  C.  K.  and  P.  J.  B.  Slater.  1995.  Bird 
song:  biological  themes  and  variations.  Cam- 
bridge University  Press,  Cambridge,  United  King- 
dom. 

Dziadosz,  V.  M.  1977.  The  vocalizations  of  the  Amer- 
ican Robin.  Ph.D.  dissertation,  Ohio  State  Uni- 
versity, Columbus. 

Falls,  J.  B.,  J.  R.  Krebs,  and  P.  K.  McGregor.  1982. 
Song  matching  in  the  Great  Tit  ( Parus  major):  the 
effect  of  similarity  and  familiarity.  Animal  Behav- 
iour 30:997-1009. 

Handley,  H.  G.  and  D.  A.  Nelson.  2005.  Ecological 
and  phylogenetic  effects  on  song  sharing  in  song- 
birds. Ethology  111:221-238. 

Hile,  A.  G.,  T.  K.  Plummer,  and  G.  F.  Striedter. 
2000.  Male  vocal  imitation  produces  call  conver- 
gence during  pair  bonding  in  Budgerigars,  Mel- 
opsittacus  undulatus.  Animal  Behaviour  59:1209— 
1218. 

Howard,  R.  D.  1974.  The  influence  of  sexual  selection 
and  interspecific  competition  on  Mockingbird 
song  ( Mimus  polyglottos).  Evolution  28:428-483. 

Hsu,  Y.  1991.  The  function  of  aggressive  interactions 
and  singing  behavior  in  the  American  Robin  ( Tur - 
dus  migratorius).  M.Sc.  thesis.  State  University  of 
New  York,  Syracuse. 

Hughes,  M.,  H.  Hultsch,  and  D.  Todt.  2002.  Imita- 
tion and  invention  in  song  learning  in  nightingales 
{Luscinia  megarhynchos  B.,  Turdidae).  Ethology 
108:97-113. 

Hultsch,  H.  and  M.  L.  Kopp.  1989.  Early  auditory 
learning  and  song  improvisation  in  nightingales, 
Luscinia  megarhynchos.  Animal  Behaviour  37: 
510-512. 

Konishi,  M.  1965.  Effects  of  deafening  on  song  de- 
velopment in  American  Robins  and  Black-headed 
Grosbeaks.  Zeitschrift  fur  Tierpsychologie  22: 
584-599. 

Krebs,  J.  R.,  R.  Ashcroft,  and  K.  Vanorsdol.  1981. 
Song  matching  in  the  Great  Tit  ( Parus  major). 
Animal  Behaviour  29:918-923. 

Kroodsma,  D.  E.  1983.  The  ecology  of  avian  vocal 
learning.  BioScience  33:165-171. 

Kroodsma,  D.  E.  1996.  Ecology  of  passerine  song  de- 
velopment. Pages  3-19  in  Ecology  and  evolution 
of  acoustic  communication  in  birds  (D.  E.  Kroods- 
ma and  E.  H.  Miller,  Eds.).  Cornell  University 
Press,  Ithaca,  New  York. 

Kroodsma,  D.  E.,  P.  W.  Houlihan,  P.  A.  Fallon,  and 
J.  A.  Wells.  1997.  Song  development  by  Grey 
Catbirds.  Animal  Behaviour  54:457-464. 

Kroodsma,  D.  E.,  W.-C.  Liu,  E.  Goodwin,  and  P.  A. 
Bedell.  1999a.  The  ecology  of  song  improvisa- 
tion as  illustrated  by  North  American  Sedge 
Wrens.  Auk  116:373-386. 


352 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Kroodsma,  D.  E.,  J.  Sanchez,  D.  W.  Stemple,  E. 
Goodwin,  M.  L.  da  Silva,  and  J.  M.  E.  Viel- 
liard.  1999b.  Sedentary  life  style  of  Neotropical 
Sedge  Wrens  promotes  song  imitation.  Animal 
Behaviour  57:855-863. 

Lanyon,  W.  E.  1979.  Development  of  song  in  the 
Wood  Thrush  ( Hylocichla  murtelina),  with  notes 
on  a technique  for  hand-rearing  passerines  from 
the  egg.  American  Museum  Novitates  2666:1-27. 

Marler,  R 1981.  Sparrows  learn  adult  song  and  more 
from  memory.  Science  213:780-782. 

Marler,  P.  and  S.  Peters.  1982.  Subsong  and  plastic 
song:  their  role  in  the  vocal  learning  process.  Pag- 
es 25-50  in  Acoustic  communication  in  birds  (D. 
E.  Kroodsma  and  E.  H.  Miller,  Eds.).  Academic 
Press,  New  York. 

Nelson,  D.  A.,  H.  Khanna,  and  P.  Marler.  2001. 
Learning  by  instruction  or  selection:  implications 
for  patterns  of  geographic  variation  in  bird  song. 
Behaviour  138:1137-1160. 

Nelson,  D.  A.  and  P.  Marler.  2005.  Do  bird  nest- 
mates  learn  the  same  songs?  Animal  Behaviour 
69:1007-1010. 

Nelson,  D.  A.,  P.  Marler,  and  A.  Palleroni.  1995. 
A comparative  approach  to  vocal  learning:  intra- 
specific variation  in  the  learning-process.  Animal 
Behaviour  50:83-97. 

Nowicki,  S.,  W.  A.  Searcy,  and  S.  Peters.  2002. 
Quality  of  song  learning  affects  female  response 
to  male  bird  song.  Proceedings  of  the  Royal  So- 
ciety of  London,  Series  B 269:1949-1954. 

Owen-Ashley,  N.  T.,  S.  J.  Schoech,  and  R.  L. 
Mumme.  2002.  Context-specific  response  of  Flor- 
ida Scrub-Jay  pairs  to  Northern  Mockingbird  vo- 
cal mimicry.  Condor  104:858-865. 

Payne,  R.  B.  1996.  Song  traditions  in  Indigo  Buntings: 
origin,  improvisation,  dispersal,  and  extinction  in 
cultural  evolution.  Pages  198-220  in  Ecology  and 
evolution  of  acoustic  communication  in  birds  (D. 
E.  Kroodsma  and  E.  H.  Miller,  Eds.).  Cornell  Uni- 
versity Press,  Ithaca,  New  York. 


Phillips,  A.  R.  1991.  The  known  birds  of  North  and 
Middle  America:  distributions  and  variation,  mi- 
grations, changes,  hybrids,  etc.,  part  II.  Bomby- 
cillidae;  Sylviidae  to  Sturnidae;  Vireonidae.  Allan 
R.  Phillips,  Denver,  Colorado. 

Rasmussen,  R.  and  T.  Dabelsteen.  2002.  Song  rep- 
ertoires and  repertoire  sharing  in  a local  group  of 
blackbirds.  Bioacoustics  13:63-76. 

Read,  A.  F.  and  D.  M.  Weary.  1992.  The  evolution 
of  bird  song:  comparative  analyses.  Philosophical 
Transactions  of  the  Royal  Society  of  London,  Se- 
ries B 338:165-187. 

Sallabanks,  R.  and  F.  C.  James.  1999.  American 
Robin  ( Turdus  migratorius).  The  Birds  of  North 
America,  no.  462. 

Sayigh,  L.  S.,  Peter  L.  Tyack,  Randall  S.  Wells, 
and  Michael  D.  Scott.  1990.  Signature  whistles 
of  free-ranging  bottlenose  dolphins  Tursiops  trun- 
catus : stability  and  mother-offspring  comparisons. 
Behavioral  Ecology  and  Sociobiology  26:247- 
260. 

Sousa,  C.  M.  1999.  How  male  singing  behaviors  affect 
extra-pair  copulations  in  a population  of  American 
Robins.  Senior  Honors  thesis.  University  of  Mas- 
sachusetts, Amherst. 

Thomas,  D.  K.  1979.  An  analysis  of  the  morning  song 
of  the  American  Robin  ( Turdus  migratorius ) 
throughout  the  breeding  season.  M.Sc.  thesis, 
Bloomsburg  State  College,  Bloomsburg,  Pennsyl- 
vania. 

Tsipoura,  N.  K.  1985.  Individual  variation  in  the  song 
of  the  American  Robin.  M.Sc.  thesis,  Washington 
State  University,  Pullman. 

Tyack,  P.  1986.  Whistle  repertoires  of  2 bottle-nosed 
dolphins,  Tursiops  truncatus:  mimicry  of  signa- 
ture whistles?  Behavioral  Ecology  and  Sociobi- 
ology 18:251-257. 

Wauer,  R.  H.  1999.  The  American  Robin.  Corrie  Her- 
ring Hooks  Series,  no.  39.  Austin,  Texas. 

Young,  H.  1955.  Breeding  behavior  and  nesting  of  the 
eastern  robin.  American  Midland  Naturalist  53: 
329-352. 


The  Wilson  Journal  of  Ornithology  1 18(3):353— 363,  2006 


EFFECTS  OF  MOWING  AND  BURNING  ON  SHRUBLAND  AND 
GRASSLAND  BIRDS  ON  NANTUCKET  ISLAND, 
MASSACHUSETTS 

BENJAMIN  ZUCKERBERG1 24  AND  PETER  D.  VICKERY13 


ABSTRACT. — Throughout  the  United  States,  declines  in  breeding  populations  of  grassland  and  shrubland 
birds  have  prompted  conservation  agencies  and  organizations  to  manage  and  restore  early-successional  habitats. 
These  habitats  support  a variety  of  birds,  some  of  which  have  been  classified  as  generalists;  thus,  often  these 
birds  are  thought  to  be  less  affected  by  habitat  manipulation.  More  information,  however,  is  needed  on  the 
response  of  early-successional  generalists  to  habitat  management,  because  conservation  agencies  are  increasing 
their  focus  on  the  regional  preservation  and  management  of  common  species.  On  Nantucket  Island,  Massachu- 
setts, the  goal  of  the  Partnership  for  Harrier  Habitat  Preservation  (PHHP)  has  been  to  restore  more  than  373  ha 
of  grassland  for  the  island’s  population  of  Northern  Harriers  ( Circus  cyaneus).  This  management  program  has 
entailed  methods  such  as  prescribed  burning  and  mowing  (e.g.,  brushcutting)  to  restore  and  maintain  grassland 
habitat.  Over  a 3-year  period,  we  found  that  songbird  response  to  burning  and  mowing  varied  among  species, 
depending  on  subtle  habitat  preferences  and  the  intensity  and  type  of  management.  In  shrublands,  Eastern  Towhee 
( Pipilo  erythrophthalmus)  and  Common  Yellowthroat  ( Geothlypis  trichas ) abundance  declined  in  mowed  areas 
but  were  unaffected  by  prescribed  burning.  In  grasslands.  Savannah  Sparrow  ( Passerculus  sandwichensis ) abun- 
dance showed  no  response  to  either  burning  or  mowing,  whereas  Song  Sparrows  ( Melospiza  melodia ) preferred 
unmanaged  grasslands.  In  shrublands,  mowing  was  the  most  effective  method  for  restoring  grassland  habitat, 
whereas  prescribed  burning  had  little  effect  on  abundances  of  shrubland  birds  and  vegetation  structure.  In 
grasslands,  both  mowing  and  burning  were  successful  in  restricting  shrubland  encroachment  and  maintaining 
grassland  habitat.  Received  27  June  2005,  accepted  1 March  2006. 


Between  1966  and  2004,  there  have  been 
significant  population  declines  in  10  of  14 
(71%)  grassland  and  16  of  36  (44%)  shrub- 
land bird  species  within  the  eastern  Breeding 
Bird  Survey  region  (Sauer  et  al.  2005) — a re- 
sult of  habitat  loss  and  fragmentation  (Vickery 
1992,  Askins  2002,  Confer  and  Pascoe  2003, 
Dettmers  2003,  Vickery  et  al.  2005).  Because 
of  these  population  declines,  prescribed  burn- 
ing and  mowing  have  become  increasingly 
important  conservation  tools  in  managing 
grasslands  and  shrublands  throughout  the 
northeastern  United  States  (Vickery  et  al. 
2005). 

Efforts  to  restore  and  maintain  early-suc- 
cessional areas  traditionally  focused  on  pro- 
viding habitat  for  rare  and  threatened  grass- 
land specialists.  Consequently,  researchers  of- 
ten emphasize  the  effects  of  habitat  distur- 


1 Dept,  of  Natural  Resources  Conservation,  Hold- 
sworth  Natural  Resources  Center,  Univ.  of  Massachu- 
setts, Amherst,  MA  01003,  USA. 

2 Current  address:  State  Univ.  of  New  York,  College 
of  Environmental  Science  and  Forestry,  1 Forestry  Dr., 
Syracuse,  NY  13210,  USA. 

3 Current  address:  Center  for  Ecological  Research, 
P.O.  Box  127,  Richmond,  ME  04357,  USA. 

4 Corresponding  author;  e-mail:  bzuckerb@syr.edu 


bance  on  single  species  that  tend  to  be  habitat 
specialists  (i.e.,  species  with  rigid  habitat  re- 
quirements) rather  than  habitat  generalists 
(i.e.,  species  with  broad  habitat  requirements; 
Bayne  and  Hobson  2001,  Fort  and  Otter 
2004).  As  regional  programs,  such  as  Partners 
in  Flight  (Rich  et  al.  2004)  and  the  National 
Gap  Analysis  Program  (Scott  et  al.  1993), 
continue  to  advocate  a conservation  approach 
of  “keeping  common  species  common,”  there 
is  a greater  need  to  study  the  effects  of  habitat 
disturbance  and  management  on  generalist 
species.  Although  studies  have  addressed  the 
effects  of  rangeland  management  on  early- 
successional  songbirds  in  the  western  United 
States  (e.g.,  Wiens  and  Rotenberry  1985, 
Wiens  et  al.  1986)  and  the  effects  of  manage- 
ment on  grassland  birds  in  northeastern  and 
midwestern  sectors  of  the  country  (Bollinger 
et  al.  1990,  Herkert  et  al.  1999,  Johnson  et  al. 
2004),  no  studies  have  focused  on  the  effects 
of  large-scale  grassland  restoration  on  both 
grassland  and  shrubland  generalists  in  the 
northeastern  United  States. 

Massachusetts’  coastal  sandplain  grass- 
lands, heathlands,  and  shrublands  are  impor- 
tant regional  conservation  priorities  because 
they  support  unique  regional  biodiversity 


353 


354 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


(Barbour  et  al.  1999).  It  is  estimated  that  more 
than  90%  of  coastal  heathlands  and  grasslands 
in  the  northeastern  United  States  have  been 
lost  since  the  middle  of  the  19th  century  due 
to  development,  cultivation,  and  shrubland  en- 
croachment (Barbour  et  al.  1999).  The  largest 
remaining  contiguous  areas  of  sandplain 
grasslands  and  coastal  heathlands  in  the 
Northeast  are  found  on  Nantucket  Island 
(hereafter  Nantucket;  Tiffney  and  Eveleigh 
1985,  Dunwiddie  1989).  Currently,  Nantuck- 
et’s grasslands  and  heathlands  are  being  lost 
to  increasing  residential  development  and 
shrubland  encroachment  (Tiffney  and  Evel- 
eigh 1985,  Dunwiddie  and  Caljouw  1990, 
Barbour  et  al.  1999),  the  latter  representing  an 
important  cause  of  both  habitat  loss  and  deg- 
radation for  grassland  birds. 

Many  of  Nantucket’s  shrubland  and  grass- 
land areas  have  been  targeted  for  restoration 
and  management.  In  1996,  the  Partnership  for 
Harrier  Habitat  Preservation  (PHHP)  was 
formed  to  develop  a large-scale  vegetation 
management  program  aimed  at  restoring 
>373  ha  of  grassland  to  create  and  sustain 
habitat  for  Northern  Harriers  ( Circus  cy- 
aneus),  an  obligate  grassland  species  that  re- 
quires relatively  open  areas  for  most  of  its 
breeding  cycle  (Christiansen  and  Reinert 
1990,  Dechant  et  al.  2003).  This  program  has 
entailed  two  basic  methods  of  restoration  and 
management:  prescribed  burning  and  mechan- 
ical restoration  (i.e.,  brush  cutting  and  repeat- 
ed mowing;  Combs-Beattie  and  Steinauer 
2001).  Although  the  goals  of  the  PHHP  em- 
phasize the  creation  of  habitat  for  Northern 
Harriers,  Nantucket’s  shrublands  and  grass- 
lands support  several  regionally  declining 
generalist  species  whose  habitat  preferences 
are  relatively  broad,  including  Eastern  Tow- 
hees  ( Pipilo  erythrophthalmus ; Greenlaw 
1996),  Savannah  Sparrows  ( Passerculus  sand- 
wichensis;  Wheelwright  and  Rising  1993), 
Common  Yellowthroats  ( Geothlypis  trichas\ 
Guzy  and  Ritchison  1999),  and  Song  Spar- 
rows ( Melospiza  melodia\  Arcese  et  al.  2002). 

Our  goal  was  to  document  the  effects  of 
prescribed  burning  and  mowing  on  Nantuck- 
et’s assemblage  of  shrubland  and  grassland 
songbirds.  In  so  doing,  our  objectives  were  to 
(1)  document  changes  in  vegetation  structure 
in  response  to  management,  (2)  identify  hab- 
itat associations  of  shrubland  and  grassland 


songbirds,  and  (3)  analyze  the  response  of 
shrubland  and  grassland  generalists  to  habitat 
alteration.  Habitat  restoration  can  be  a pow- 
erful conservation  tool,  but  considering  the  re- 
gional goals  and  objectives  of  many  conser- 
vation programs  aimed  at  preserving  common 
species,  we  believe  that  it  is  important  to 
study  the  effects  of  habitat  management  on 
habitat  generalists,  as  well  as  specialists. 

METHODS 

Study  areas. — Nantucket  (41°  28.3' N,  70° 
1'  W)  is  about  48  km  south  of  Cape  Cod  and 
measures  1 1 X 24  km  (Litchfield  1994).  The 
island  contains  naturally  occurring  and  re- 
gionally rare  sandplain  grasslands,  scrub  oak 
shrublands,  and  sandplain  heathlands  (Swain 
and  Kearsley  2001).  The  sandplain  grasslands 
are  dominated  by  graminoids,  primarily  little 
bluestem  {Schizachyrium  scoparium),  Penn- 
sylvania sedge  ( Carex  pensylvanica),  and 
poverty  oatgrass  ( Danthonia  spicata ).  Scrub 
oak  shrublands  are  dominated  by  bear  oak 
{Quercus  ilicifolia)  and  have  an  understory  of 
black  huckleberry  ( Gaylussacia  baccata ), 
bearberry  ( Arctostaphylos  uva-ursi),  and  low- 
bush  blueberry  ( Vaccinium  angustifolium\ 
Dunwiddie  and  Sorrie  1996).  Heathlands  sup- 
port many  of  the  same  plant  species  as  those 
found  in  grasslands  and  scrub  oak  shrublands, 
but  are  dominated  by  low-growing  black 
huckleberry,  bearberry,  and  lowbush  blueber- 
ry (Swain  and  Kearsley  2001).  Despite  shar- 
ing many  of  the  same  characteristic  plant  spe- 
cies as  shrublands,  heathlands  found  along  the 
coastline  are  noticeably  shorter  and  often  in- 
termix and  overlap  with  grassland  communi- 
ties; consequently,  we  defined  grassland/ 
heathland  areas  as  grassland  for  subsequent 
analyses  (Dunwiddie  and  Sorrie  1996). 

From  1998  to  2001,  the  PHHP  targeted 
>373  ha  of  shrubland  and  grassland  for  res- 
toration and  maintenance  (Table  1).  Manage- 
ment plans  have  included  prescribed  burning 
on  142  ha  of  scrub  oak  shrubland  and  >26  ha 
of  grassland/heathland,  and  repeated  mowing 
and  brush  cutting  on  205  ha  of  shrubland  (Ta- 
ble 1).  The  frequency  of  management  differed 
among  study  sites:  shrubland  areas  were 
burned  no  more  than  once,  and  mowing  fre- 
quency ranged  from  0 (control  areas)  to  1—3 
cuts  annually.  In  addition  to  these  activities, 
the  Nantucket  Land  Bank  Commission  began 


Zuckerberg  and  Vickery  • EARLY-SUCCESSIONAL  BIRDS  ON  NANTUCKET  ISLAND 


355 


TABLE  1 . Management  areas  and  restoration  histories  of  grassland  and  shrubland  study  sites  on  Nantucket 
Island,  Massachusetts,  1999-2001. 

Site  name 

Area  (ha) 

No.  bird  survey  plots 

Restoration  history 

Years  sampled 

Shrublands 

D 

19.4 

6 

Control/burn  (2000) 

1999-2001 

El 

19.3 

8 

Control 

1999-2001 

SHRUB 

14.2 

5 

Control 

1999-2001 

BC 

68.0 

12 

Mow  (1998-2001) 

1999-2001 

A 

10.5 

4 

Mow  (1998,  1999) 

1999 

LB  1 

19.8 

5 

Mow  (1999-2001) 

2000-2001 

LB  2 

19.0 

5 

Mow  (1999-2001) 

2000-2001 

A2 

9.7 

3 

Mow  (2000) 

2000-2001 

TRI 

6.9 

3 

Mow  (2000,  2001) 

2000-2001 

LB4 

21.0 

8 

Mow  (1999-2001) 

2001 

ABURN 

10.9 

4 

Burn  (2000) 

2001 

E2 

16.2 

4 

Burn  (1994) 

1999-2001 

E3 

0.8 

1 

Burn  (1998) 

1999-2001 

F 

4.9 

3 

Burn  (1996) 

1999-2001 

Grasslands 

LRAM 

4.5 

3 

Control/burn  (2001) 

1999-2001 

HPLAIN 

19.0 

6 

Control 

1999-2001 

LB3 

12.1 

5 

Control 

2000-2001 

RAM 

30.8 

6 

Mow  (1999,  2000)/burn  (2001) 

1999-2001 

GOLF 

6.1 

4 

Mow  (1998-2001) 

1999-2001 

AIR 

7.7 

4 

Mow  (1998-2001) 

1999-2001 

similar  brush-cutting  efforts  in  three  separate 
areas  comprising  >74  ha.  Study  sites  consist- 
ed of  areas  that  were  either  controls  (grass- 
lands, shrublands,  or  heathlands  that  had  not 
been  managed  for  at  least  10  years)  or  areas 
that  had  received  or  are  receiving  manage- 
ment through  mowing  or  prescribed  burning 
since  1988.  Given  the  duration  of  the  man- 
agement plan,  the  number  of  areas  being  man- 
aged and  surveyed  changed  each  year  (Table 
1).  Management  areas  were  typically  discrete 
subsets  of  larger,  more  contiguous  habitats 
that  were  receiving  a particular  treatment.  No 
two  adjacent  study  areas  shared  the  same 
treatment  history,  and  study  areas  were  spa- 
tially separated  by  other  habitat  types  or  bar- 
riers (e.g.,  wetlands,  open  water,  or  roads).  To 
avoid  disruption  due  to  treatment  activities, 
we  collected  data  only  in  those  areas  that  were 
not  being  actively  managed  during  the  sum- 
mer months  of  this  study.  Due  to  unexpected 
summer  management  activities  on  some  study 
sites,  we  did  not  sample  every  site  in  each 
year;  thus,  the  number  of  observations  dif- 
fered among  study  sites  and  sample  data  were 
unbalanced  (Table  1). 

Bird  censuses. — In  the  breeding  seasons  of 


1999-2001,  we  determined  avian  abundance 
of  shrubland  and  grassland  songbirds  by  con- 
ducting 10-min  avian  surveys  in  fixed-radius, 
50-m  circular  plots  along  pre-established  par- 
allel transects,  the  length  and  number  of 
which  varied,  depending  on  the  size  and  con- 
figuration of  each  site  (Table  1;  Bibby  et  al. 
2000).  Survey  plots  were  >100  m from  any 
habitat  edges  and  >200  m from  other  plots 
(Hutto  et  al.  1986,  Bibby  et  al.  2000).  From 
22  May  to  10  August  during  the  breeding  sea- 
sons of  1999-2001,  we  visited  14  shrubland 
and  6 grassland  sites  three  times  (Vickery  et 
al.  1994).  We  conducted  surveys  between  06: 
00  and  10:00  EDT  and  began  surveys  2 min 
after  arriving  at  the  site,  but  we  did  not  survey 
birds  during  inclement  weather,  such  as  rain 
or  high  wind  (>15  km/hr;  Vickery  et  al. 
1994).  Because  our  focus  was  limited  to  avian 
and  vegetation  changes  only  within  manage- 
ment areas,  our  protocol  purposely  did  not  ac- 
count for  changes  along  or  near  habitat  edges. 
For  a given  breeding  season,  we  considered 
the  maximum  number  of  singing  males  de- 
tected during  our  three  visits  as  a measure  of 
avian  abundance,  and  combined  these  data  to 


356 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


derive  a mean  for  all  survey  plots  within  a 
particular  management  area. 

Vegetation  surveys. — At  each  survey  plot, 
we  sampled  the  vegetation  at  0.5-m  intervals 
along  four  50-m  transects  that  radiated  from 
the  center  of  each  survey  plot  in  the  four  car- 
dinal directions  (Brower  and  Zar  1977).  This 
resulted  in  400  vegetation  sampling  points  per 
survey  plot.  At  each  sampling  point,  we  re- 
corded the  dominant  vegetation  type  and 
height.  We  classified  vegetation  cover  into 
four  type  categories  (sparse  vegetation,  litter, 
grass/forb.  and  shrub)  and  seven  height  cate- 
gories (0,  >0-0.1,  >0. 1-0.5,  >0. 5-1.0, 
>1. 0-2.0,  >2.0-5. 0,  and  >5.0  m.  Vegetation 
data  were  converted  to  relative  frequencies 
and,  for  a given  parameter  in  a given  survey 
plot,  we  averaged  all  values  from  the  four 
transects.  This  method  allowed  us  to  establish 
a basic  portrait  of  vegetation  height  and  type 
for  each  point  count  and  study  site. 

Statistical  analyses. — Our  null  hypothesis 
was  that  that  bird  densities  within  control 
shrublands  and  grasslands  would  be  the  same 
as  those  in  managed  shrublands  and  grass- 
lands, respectively.  We  used  univariate  meth- 
ods to  determine  species-specific  responses  to 
restoration  techniques  and  vegetation  charac- 
teristics. We  were  unable  to  randomize  our 
treatments  because  management  of  this  large, 
multi-agency  restoration  program  was  con- 
strained by  multiple  factors  beyond  our  con- 
trol. This  is  not  uncommon  in  “natural  exper- 
iments” and  we  employed  matching  in  lieu  of 
a controlled  experimental  design;  that  is,  we 
compared  managed  units  with  units  that  were 
not  managed  (i.e.,  control),  but  were  similar 
to  the  treated  units  in  terms  of  proximity  and 
environmental  conditions  (Johnson  2002). 

We  used  a proportional  odds  logistic  re- 
gression model  with  forward  selection  to 
identify  significant  vegetation  predictors  of 
avian  occurrence  (Hosmer  and  Lemeshow 
1989;  PROC  LOGISTIC;  SAS  Institute,  Inc. 
1990).  Heavily  skewed  data  on  vegetation  and 
uncommon  bird  species  that  did  not  satisfy 
normality  requirements  were  converted  to  de- 
tection/non-detection (i.e.,  presence/absence) 
data  for  further  analysis.  For  these  data,  we 
used  chi-square  analysis  to  determine  which 
vegetation  variables  influenced  the  detection/ 
non-detection  (i.e.,  presense/absence)  of  se- 
lected bird  species  (Kleinbaum  et  al.  1998); 


only  vegetation  variables  that  were  significant 
(a  < 0.05)  in  this  analysis  were  used  in  the 
logistic  regression  models  (Hosmer  and  Le- 
meshow 1989). 

We  used  repeated-measures  analysis  of  var- 
iance (ANOVA)  to  determine  bird  species- 
specific  responses  to  management  (Sokal  and 
Rohlf  1995).  Due  to  the  unbalanced  nature  of 
the  study  design,  we  used  SAS  (PROC 
MIXED;  SAS  Institute,  Inc.  1990),  which  al- 
lows for  interval-independent  variables  and 
uses  the  maximum  likelihood  method  to  esti- 
mate parameters  (Kleinbaum  et  al.  1998). 
Study  sites  that  received  the  prescribed  burn- 
ing treatment  were  categorized  by  two  post- 
bum classifications:  1 year  post-bum  and  2-7 
years  post-bum.  One-way  ANOVAs  were 
used  to  determine  differences  in  vegetation 
variables  within  grasslands  and  shrublands 
treated  with  different  methods  and,  because  all 
pairwise  comparisons  were  of  interest,  we 
used  the  Tukey-Kramer  method  for  all  multi- 
ple-comparison tests  (Kleinbaum  et  al.  1998). 
We  conducted  ANOVAs  separately  on  grass- 
land/heathland  and  shrubland  areas  for  both 
bird  abundance  and  vegetation  data.  The  den- 
sities of  three  species — Eastern  Towhee,  Sa- 
vannah Sparrow,  and  Song  Sparrow — were 
adequate  to  meet  the  requirements  for  repeat- 
ed measures  ANOVA.  We  set  ( a priori ) a sig- 
nificance level  of  P — 0.05  and  a “marginal” 
significance  level  of  0.10  > P > 0.05.  We 
conducted  power  analyses  on  ANOVA  results 
at  a significance  level  of  P — 0.05.  Means  are 
presented  ± SE. 

RESULTS 

Changes  in  vegetation  structure. — Mowing 
and  burning  had  different  effects  on  vegeta- 
tion structure  and  composition  (Table  2). 
Mowing  in  shrublands  produced  the  most  no- 
table difference.  Mowed  shrublands  had  a 
greater  percent  cover  of  litter  (37.7%  ± 17.5) 
than  burned  (2.3%  ± 2.1)  or  control  areas 
(1.9%  ± 1.8;  F2A2  = 15.22,  P < 0.001).  Me- 
dium-height shrubs  (1. 0-2.0  m)  were  common 
in  control  (44.4%  ± 12.1)  and  burned  shrub- 
lands (47.3%  ± 14.5)  but  significantly  less  in 
mowed  shrublands  (11.1%  ± 8.3;  F2l2  = 
17.82,  P < 0.001).  We  documented  similar 
findings  for  tall  shrubs  (2.0-5. 0 m;  F2l2  = 
9.17,  P = 0.004).  Although  not  significant  at 
the  0.05  alpha  level,  medium-height  grasses 


Zuckerberg  and  Vickery  • EARLY-SUCCESSIONAL  BIRDS  ON  NANTUCKET  ISLAND  357 


TABLE  2.  Percent  cover  (SE)  for  vegetation  variables,  and  results  of  one-way  analysis  of  variance  (ANO- 
VA),  testing  treatment  effects  in  shrubland  and  grassland  habitats  on  Nantucket  Island,  Massachusetts,  1999- 
2001.  Several  vegetation  variables  changed  in  response  to  mowing  and  prescribed  burning  in  shrubland  and 
grassland  study  sites.  In  shrubland  sites,  mowed  areas  had  greater  proportions  of  litter  and  short  shrubs  and 
lower  proportions  of  medium  and  tall  shrubs.  In  grassland  sites,  unmanaged  grasslands  had  higher  proportions 
of  medium  shrubs.  Significant  values  (P  < 0.05)  are  in  bold. 

Variable  entered 

Control 

Bum 

Mow 

p 

Shrublands 

Sparse  vegetation 

0.04  (0.04) 

0.08  (0.04) 

0.03  (0.04) 

0.091 

Litter  (0-0.1  m) 

0.02  (0.02) 

0.02  (0.02) 

0.38  (0.17) 

<0.001 

Short  grass  (0—0.1  m) 

0.01  (0.02) 

0.00  (0.01) 

0.07  (0.07) 

0.10 

Medium-height  grass  (0. 1-0.5  m) 

0.16  (0.03) 

0.11  (0.11) 

0.28  (0.13) 

0.079 

Short  shrub  (0-0.1  m) 

0.50  (0.19) 

0.34  (0.32) 

0.24  (0.22) 

0.36 

Short  shrub  (0. 1-0.5  m) 

0.46  (0.06) 

0.50  (0.10) 

0.72  (0.14) 

0.006 

Medium-height  shrub  (0.5-1. 0 m) 

0.39  (0.11) 

0.33  (0.10) 

0.37  (0.14) 

0.82 

Medium-height  shrub  (1. 0-2.0  m) 

0.44  (0.12) 

0.47  (0.15) 

0.11  (0.08) 

<0.001 

Tall  shrub  (2.0-5. 0 m) 

0.44  (0.09) 

0.46  (0.17) 

0.15  (0.13) 

0.004 

Tall  shrub  (>5.0  m) 

0.04  (0.04) 

0.07  (0.11) 

0.06  (0.06) 

0.88 

Grasslands 

Short  grass  (0-0.1  m) 

0.13  (0.12) 

0.30  (0.00) 

0.53  (0.17) 

0.046 

Medium-height  grass  (0. 1-0.5  m) 

0.66  (0.11) 

0.75  (0.01) 

0.65  (0.07) 

0.43 

Short  shrub  (0-0.1  m) 

0.26  (0.02) 

0.37  (0.10) 

0.32  (0.25) 

0.73 

Short  shrub  (0. 1-0.5  m) 

0.67  (0.10) 

0.55  (0.02) 

0.39  (0.19) 

0.13 

Medium-height  shrub  (0.5-1. 0 m) 

0.38  (0.07) 

0.14  (0.10) 

0.13  (0.04) 

0.025 

Medium-height  shrub  (1. 0-2.0  m) 

0.08  (0.00) 

0.00  (0.00) 

0.03  (0.01) 

0.67 

Tall  shrub  (2.0-5. 0 m) 

0.01  (0.03) 

0.00  (0.00) 

0.04  (0.07) 

0.67 

(0. 1-0.5  m),  which  were  uncommon  in  control 
(15.6%  ± 3.3)  and  burned  (11.3%  ± 11.2) 
shrublands,  were  slightly  more  common  in 
mowed  areas  (27.7%  ± 13.1;  F2l2  = 3.14,  P 
= 0.080). 

In  grasslands,  burning  and  mowing  pro- 
duced notable  differences  in  vegetation  (Table 
2).  Compared  with  grasslands  that  had  been 
burned  or  mowed,  control  grasslands  were 
characterized  by  a relatively  greater  percent 
cover  of  short-shrub  vegetation.  Medium- 
height  shrubs  (0.5-1. 0 m)  were  more  abun- 
dant in  control  grasslands  (37.6%  ± 6.7),  and 
less  abundant  in  burned  (13.7%  ± 10.1)  or 
mowed  grasslands  (12.7%  ± 4.4;  F2A  = 8.37, 
P = 0.025).  Mowed  grasslands  had  higher 
proportions  of  short  grass  (0.0-0. 1 m;  52.6% 
± 17.0)  compared  with  burned  (30.0  ± 0.0%) 
and  control  grasslands  (13.0%  ± 12.0;  F24  = 
6.08,  P = 0.046). 

Avian  response  to  vegetation. — Shrubland 
and  grassland  bird  communities  on  Nantucket 
were  relatively  depauperate,  a common  char- 
acteristic of  faunal  communities  on  islands 
(Brown  and  Lomolino  1998).  Important  veg- 
etation predictors  of  Eastern  Towhee,  Com- 


mon Yellowthroat,  Song  Sparrow,  and  Savan- 
nah Sparrow  presence  varied  by  species  (Ta- 
ble 3).  Towhees  were  positively  associated 
with  litter  (0-0.1  m)  and  medium  (1. 0-2.0  m) 
and  tall  (2.0-5. 0 m)  shrubs,  but  they  were 
negatively  associated  with  medium-height 
grass  (0. 1-0.5  m;  Table  3).  Unlike  towhees, 
Common  Yellowthroats  were  negatively  as- 
sociated with  litter  (0-0.1  m)  but  positively 
associated  with  medium  shrubs  (1. 0-2.0  m). 
Song  Sparrows  were  positively  associated 
with  medium-height  grass  (0. 1-0.5  m)  and 
medium  shrubs  (0.5- 1.0  m),  but  they  were 
negatively  associated  with  litter  (0-0.1  m). 
Savannah  Sparrows  were  positively  associated 
with  medium  grass  (0. 1-0.5  m)  but  negatively 
associated  with  litter  (0-0. 1 m)  and  tall  shrubs 
(2.0-5. 0 m;  Table  3). 

Avian  response  to  management  within 
shrublands. — Within  shrubland  areas,  we  re- 
corded Eastern  Towhees,  Common  Yellow- 
throats,  Song  Sparrows,  Gray  Catbirds  (9m- 
metella  carolinensis ),  Eastern  Kingbirds 
( Tyrannus  tyrannus).  Blue  Jays  ( Cyanocitta 
cristata ),  American  Crows  ( Corvus  brachy- 
rhynchos ),  and  Prairie  Warblers  ( Dendroica 


358 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  3.  Proportional  odds  logistic  regression  using  percent  cover  of  vegetation  predictors  to  model  the 
probability  of  bird  species  presence  in  shrubland  and  grassland  habitat  on  Nantucket  Island,  Massachusetts, 
1999-2001.  Significant  values  ( P < 0.05)  are  in  bold. 


Variable  entered 

Estimate 

Standard  error 

p 

Eastern  Towhee 

Bare  ground 

-0.57 

0.40 

0.15 

Litter  (0-0.1  m) 

1.35 

0.41 

0.001 

Short  grass  (0-0.1  m) 

0.26 

0.49 

0.60 

Medium-height  grass  (0. 1-0.5  m) 

-0.85 

0.43 

0.05 

Tall  grass  (0.5-1. 0 m) 

-1.55 

1.07 

0.15 

Medium-height  shrub  (0. 5-1.0  m) 

-0.10 

0.69 

0.88 

Medium-height  shrub  (1. 0-2.0  m) 

1.20 

0.50 

<0.001 

Tall  shrub  (2.0-5. 0 m) 

1.67 

0.39 

<0.001 

Tall  shrub  (>5.0  m) 

0.31 

0.78 

0.69 

Common  Yellowthroat 

Litter  (0-0.1  m) 

-0.88 

0.38 

0.02 

Short  grass  (0-0.1  m) 

-0.34 

0.61 

0.57 

Medium-height  grass  (0. 1-0.5  m) 

-0.26 

0.42 

0.54 

Medium-height  shrub  (1. 0-2.0  m) 

1.18 

0.62 

0.05 

Tall  shrub  (2.0-5.0  m) 

0.64 

0.48 

0.18 

Song  Sparrow 

Litter  (0-0. 1 m) 

-1.09 

0.37 

0.004 

Medium-height  grass  (0. 1-0.5  m) 

1.97 

0.50 

<0.001 

Medium-height  shrub  (0.5- 1.0  m) 

1.63 

0.54 

0.003 

Tall  shrub  (>5.0  m) 

-1.03 

0.83 

0.22 

Savannah  Sparrow 

Litter  (0-0. 1 m) 

-2.85 

0.74 

<0.001 

Short  grass  (0-0.1  m) 

0.14 

0.45 

0.80 

Medium-height  grass  (0. 1-0.5  m) 

2.13 

0.89 

0.02 

Short  shrub  (0-0.1  m) 

-0.26 

0.61 

0.68 

Medium-height  shrub  (0.5-1. 0 m) 

-0.32 

0.46 

0.49 

Medium-height  shrub  ( 1 .0-2.0  m) 

-0.53 

0.48 

0.26 

Tall  shrub  (2.0-5.0  m) 

-2.78 

0.75 

<0.001 

discolor).  Eastern  Towhees  showed  a clear 
response  to  management  practices  in  shrub- 
lands  (Table  4).  In  two  out  of  the  three  breed- 
ing seasons.  Eastern  Towhee  abundance  was 
greater  in  control  or  burned  shrublands  com- 
pared with  shrublands  that  had  been  mowed. 
Overall,  towhee  abundance  was  greatest  in 
areas  that  had  been  burned  (1.42/ha  ± 0.49), 
and  there  was  no  difference  in  densities  be- 
tween controls  (1.12/ha  ± 0.37)  and  mowed 
areas  (0.66/ha  ± 0.50;  Fig.  1);  however,  our 
power  to  detect  this  difference  was  low  ((3  = 
0.09).  The  abundance  of  towhees  differed 
significantly  among  years  (Table  4),  decreas- 
ing in  every  season  from  an  average  of  1.48 
± 0.86  in  1999  to  0.86  ± 0.75  in  2000  to 
0.71  ± 0.64  in  2001. 

Towhee  abundance  decreased  as  the  fre- 
quency of  mowing  increased  between  sites 


(Table  4).  After  a single  mowing  event,  tow- 
hee abundance  dropped  from  an  average  of 
1.13/ha  ± 0.17  to  0.85/ha  ± 0.17.  After  a sec- 
ond mowing,  abundance  further  declined  to 
0.53/ha  ± 0.18,  although  this  decrease  was 
not  significant;  again,  however,  our  power  to 
detect  significant  differences  was  limited  ((3  = 
0.3). 

We  found  no  significant  differences  in  to- 
whee abundance  in  relation  to  time  since  the 
most  recent  bum  (Table  4),  but  power  was  low 
((3  = 0.21).  Although  towhee  abundance  de- 
clined slightly  in  the  first  year  after  a bum, 
this  decline  was  not  significant,  and  abun- 
dance in  sites  that  had  been  burned  2-7  years 
earlier  was  not  significantly  different  than  the 
abundance  in  control  areas. 

Among  the  less  common  shrubland  birds. 
Common  Yellowthroats  preferred  control  and 


Zuckerberg  and  Vickery  • EARLY-SUCCESSIONAL  BIRDS  ON  NANTUCKET  ISLAND  359 


TABLE  4.  Repeated  measures  analysis  of  variance  (ANOVA)  testing  treatment  effects  on  Eastern  Towhees 
in  shrubland  habitats  on  Nantucket  Island,  Massachusetts,  1999-2001.  Densities  of  Eastern  Towhees  were  most 
affected  by  mowing  and  the  frequency  of  mowing  within  shrubland  sites;  prescribed  burning  had  little  effect  on 
Eastern  Towhee  abundance.  Significant  values  ( P < 0.05)  are  in  bold. 

Variable  entered3 

df 

Estimate 

Standard 

error 

F or  t 

p 

Treatment  comparisons 

2,  12 

4.25 

0.040 

Control  versus  bum 

12 

0.30 

0.31 

0.94 

0.63 

Bum  versus  mow 

12 

-0.76 

0.29 

2.84 

0.037 

Control  versus  mow 

12 

-0.47 

0.28 

1.64 

0.27 

Mowing  frequency 

2,  4 

5.25 

0.035 

Control  versus  1 mowing/season 

8 

0.28 

0.24 

1.22 

0.47 

Control  versus  2 mowings/season 

8 

-0.78 

0.24 

3.22 

0.030 

1 mowing  versus  2 mowings/season 

8 

-0.50 

0.24 

2.04 

0.17 

Years  post-bumb 

2,  2 

0.78 

0.51 

Year 

2,  1 

14.56 

<0.001 

3 Within-treatment  comparisons  were  tested  using  the  Tukey-Kramer  comparison  (i.e.,  mowing  frequency  and  years  post-bum). 

b Within-treatment  comparisons  were  not  included  for  prescribed  burning  because  the  overall  model  was  not  significant,  and  the  yearly  differences  were 
not  significant. 


burned  shrublands  and  avoided  shrublands 
that  had  been  mowed  (x2  = 14.43,  df  = 2,  P 
< 0.001;  Fig.  1).  As  with  Eastern  Towhees, 
the  frequency  of  mowing  within  a season  had 
a significant  effect  on  Common  Yellowthroat 
presence  (x2  = 17.47,  df  = 2,  P < 0.001), 
which  was  greater  than  expected  in  shrublands 
that  had  not  been  mowed,  but  lower  than  ex- 
pected after  one  mowing;  no  Common  Yel- 
lowthroats  were  recorded  in  shrublands  that 
were  mowed  two  or  more  times  within  a sea- 
son. 

Song  Sparrow  abundance  did  not  differ 
among  shrublands  that  had  been  mowed, 
burned,  or  left  unmanaged  (x2  = 1.97,  df  = 
2,  P = 0.37;  |3  = 0.20;  Fig.  1).  In  addition. 
Song  Sparrow  presence  did  not  change  sig- 
nificantly with  respect  to  the  frequency  of 
mowing  (x2  = 1.66,  df  = 2,  P = 0.44).  Nei- 
ther Common  Yellowthroat  (x2  — 3.41,  df  = 
2,  P = 0.18)  nor  Song  Sparrow  (x2  = 0.25, 
df  = 2,  P = 0.88)  presence  differed  with  re- 
spect to  years  since  burning. 

Avian  response  to  grassland  manage- 
ment.— Within  grassland  areas,  we  recorded 
Savannah  Sparrows,  Song  Sparrows,  and 
American  Goldfinches  ( Carduelis  tristis).  Sa- 
vannah Sparrow  abundance  did  not  differ 
among  grasslands  that  had  been  burned, 
mowed,  or  left  unmanaged  (F24  = 0.04,  P = 
0.96;  (3  = 0.06;  Fig.  2).  Song  Sparrow  abun- 
dance was  greatest  in  unmanaged  grasslands 
(0.60/ha  ± 0.09),  but  was  similar  in  burned 


(0.11/ha  ± 0.08)  or  mowed  (0.11/ha  ± 0.09; 
F2>4  = 8.35,  P = 0.025)  grasslands  (Fig.  2). 

DISCUSSION 

Management  in  shrublands. — Our  findings 
suggest  that  the  effects  of  grassland  restora- 
tion on  generalist  species  will  vary  with  man- 
agement type  and  the  subtle  habitat  preferenc- 
es of  the  affected  species.  Not  surprisingly, 
mowing  produced  the  most  noticeable  chang- 
es in  vegetation  by  reducing  tall  shrub  cover. 
Mowed  areas  were  dominated  by  litter  and 
short  shrubs  and  contained  greater  grass  cover. 
Shrublands  that  were  left  unmanaged  or 
burned  once  were  not  noticeably  different  and 
were  characterized  by  tall  shrubs.  Due  to  lo- 
gistical difficulties,  such  as  the  availability  of 
adequate  bum  days  and  trained  personnel,  sin- 
gle burns  are  common  in  prescribed  burning 
programs  (Combs-Beattie  and  Steinauer 
2001);  thus,  the  results  we  observed  in  shrub- 
lands burned  once  could  be  expected  in  other 
prescribed  fire  programs. 

Although  several  generalist  species  inhab- 
ited the  same  habitat  type,  a different  suite  of 
vegetation  variables  affected  the  presence  of 
each  species.  Eastern  Towhees  were  positively 
associated  with  litter  and  medium  and  tall 
shrubs  (1. 0-5.0  m),  and  they  were  negatively 
associated  with  medium-height  grass.  Com- 
mon Yellowthroats  preferred  habitats  charac- 
terized by  no  litter  cover  and  medium-height 
shrubs  (1.0— 2.0  m).  Song  Sparrows  preferred 


Relative  abundance 


360 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


CD 


Control  Burn  Mow 


Control  Burn  Mow 


FIG.  1.  In  shrubland  study  sites,  bird  species  re- 
sponded differently  to  both  burning  and  mowing  man- 
agement. The  abundance  (±1  SE)  of  Eastern  Towhees 
(A)  and  Common  Yellowthroats  (B)  was  most  affected 
by  mowing  management,  but  was  similar  in  burned  and 
unmanaged  shrublands.  Song  Sparrows  (C)  showed  lit- 
tle response  to  management  activities.  Data  collected  on 
Nantucket  Island,  Massachusetts,  1999-2001. 


FIG.  2.  In  grassland  study  sites,  Savannah  Spar- 
row (A)  densities  (±1  SE)  were  unaffected  by  man- 
agement type,  whereas  Song  Sparrow  (B)  densities  (± 
1 SE)  were  lower  in  both  mowed  or  burned  grasslands. 
Data  collected  on  Nantucket  Island,  Massachusetts, 
1999-2001. 


areas  that  had  grass  and  short  shrub  vegeta- 
tion. 

Despite  being  generalists,  several  bird  spe- 
cies appeared  to  respond  differently  to  burn- 
ing and  mowing  treatments  in  shrublands,  as 
has  been  found  in  other  studies  (e.g.,  Wiens 
and  Rotenberry  1985,  Wiens  et  al.  1986). 
Eastern  Towhee  and  Common  Yellowthroat 
densities  were  greater  in  shrublands  that  had 
been  burned  or  left  unmanaged,  whereas  Song 
Sparrow  densities  showed  no  response  to  ei- 
ther restoration  technique  (Fig.  1).  The  effects 
of  mowing  frequency  were  more  immediate 
for  Common  Yellowthroats;  they  disappeared 
after  the  initial  mowing  event. 

Grassland  management. — In  grassland  hab- 


Zuckerberg  and  Vickery  • EARLY-SUCCESSIONAL  BIRDS  ON  NANTUCKET  ISLAND  361 


itats,  prescribed  burning  and  mowing  pro- 
duced similar  results.  The  purpose  of  burning 
and  mowing  in  grasslands  was  to  maintain 
grassland.  Consequently,  management  in 
grassland  had  less  impact  on  vegetation  struc- 
ture than  similar  restoration  techniques  used 
in  dense  shrublands.  Dunwiddie  and  Caljouw 
(1990)  found  that  burning  and  mowing  of 
Nantucket  grasslands  were  equally  effective  in 
suppressing  shrubs  and  enhancing  grasses.  In 
this  study,  unmanaged  grasslands  had  greater 
cover  of  short  shrubs  compared  with  burned 
and  mowed  grasslands,  and  low-growing 
shrubs  often  dominated  grasslands  that  were 
left  unmanaged  for  >6  years  (Dunwiddie  and 
Caljouw  1990).  Mowing  resulted  in  grass- 
lands with  the  greatest  percentages  of  short- 
to  medium-height  grass  cover.  These  findings 
suggest  that,  for  a limited  number  of  years, 
grasslands  left  unmanaged  will  continue  to 
provide  habitat  for  some  species  of  grassland- 
dependent  songbirds,  but  that  eventually  these 
grasslands  will  be  succeeded  by  shrublands 
(Dunwiddie  and  Caljouw  1990). 

Similar  to  shrubland  generalists,  the  re- 
sponse of  grassland  generalists  to  manage- 
ment practices  varied  among  bird  species  (Fig. 
2).  Savannah  Sparrow  abundance  was  similar 
in  grasslands  that  had  been  mowed,  burned, 
or  left  unmanaged.  Song  Sparrows,  which 
were  present  in  both  grassland  and  shrubland 
habitats,  occurred  at  significantly  greater  den- 
sities in  unmanaged  grasslands.  Both  Savan- 
nah and  Song  sparrows  were  negatively  as- 
sociated with  litter  and  positively  associated 
with  medium  to  tall  grass  cover.  Song  Spar- 
rows also  were  associated  positively  with 
short  shrubs,  whereas  Savannah  Sparrows 
were  negatively  associated  with  tall  shrubs. 
Song  Sparrows  required  short  to  medium 
shrubs,  and  any  grassland  management  that 
substantially  reduced  shrub  cover  also  reduced 
Song  Sparrow  abundance  significantly. 

Some  researchers  have  suggested  that  site 
fidelity  may  preclude  birds  from  responding 
immediately  to  management  practices  (Wiens 
and  Rotenberry  1985,  Wiens  et  al.  1986,  but 
see  Vickery  et  al.  1999).  Our  findings  suggest 
that  species-specific  habitat  requirements  and 
the  magnitude  of  the  management,  especially 
mowing,  appeared  to  outweigh  any  affects  of 
site  tenacity  for  Common  Yellowthroats  and 
Eastern  Towhees.  The  Eastern  Towhee’s  pref- 


erence for  foraging  habitat  (i.e.,  litter;  Green- 
law 1996)  may  make  towhees  less  susceptible 
to  burning  and  mowing  than  Common  Yel- 
lowthroats. In  the  case  of  Song  Sparrows, 
their  lack  of  dependence  on  tall  shrubs  and 
their  preference  for  grass  cover  may  explain 
why  their  densities  were  not  affected  by  either 
restoration  technique. 

The  lack  of  avian  response  to  management 
may  have  been  a product  of  the  spatial  and 
temporal  scales  at  which  this  study  was  con- 
ducted. Many  avian  species  respond  to  habitat 
alteration  at  both  landscape  and  patch  scales 
(Herkert  et  al.  1994,  Donovan  and  Flather 
2002,  McGarigal  and  Cushman  2002).  The  fo- 
cus of  our  research,  however,  was  patch-scale 
disturbances  and  responses,  and  not  land- 
scape-scale changes.  In  addition,  many  grass- 
land birds  are  area-sensitive  and  require  rela- 
tively large  grassland  habitats  (>25  ha;  Win- 
ter and  Faaborg  1999,  Mitchell  et  al.  2000, 
Johnson  and  Igl  2001).  Because  the  average 
size  of  the  grassland  habitats  included  in  this 
study  was  13.4  ha  (Table  1),  many  of  the 
grassland  areas  may  not  have  been  large 
enough  to  support  a diverse  community  of 
grassland  birds,  regardless  of  management  in- 
tensity and/or  duration.  In  the  future,  restora- 
tion activities  within  the  shrubland  study  areas 
may  produce  relatively  large  grassland  habi- 
tats, but  our  study  was  focused  on  the  initial 
years  of  management  as  opposed  to  the  long- 
term effects  of  restoration. 

Management  implications.  — Conservation 
agencies  must  address  several  issues  regarding 
the  restoration  or  management  of  early-suc- 
cessional  areas,  including  the  response  of  gen- 
eralist species  and  the  type  and  spatial  scale 
of  the  management.  Despite  sharing  similar 
habitat  requirements,  individual  bird  species 
will  respond  differently  to  management  due  to 
subtle  preferences  in  vegetation  structure  and 
composition.  In  the  case  of  habitat  restoration 
on  Nantucket,  much  of  the  management  had 
the  unforeseen  effect  of  making  common  spe- 
cies less  common.  Considering  these  species- 
specific  responses  to  mowing  and  burning 
(even  among  habitat  generalists),  managers 
must  proceed  cautiously  and  consider  the  re- 
gional declines  of  the  affected  bird  species. 
This  is  especially  true  of  grassland  restoration 
aimed  at  shrubland  areas,  as  managers  are 
faced  with  the  dilemma  of  managing  one  re- 


362 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


gionally  rare  community  at  the  expense  of  an- 
other. In  this  scenario,  a dynamic  and  diverse 
set  of  strategies  must  be  integrated  into  man- 
agement such  that  sites  are  rotated,  allowing 
some  to  succeed  to  later  stages  before  they  are 
disturbed,  to  provide  habitat  for  both  shrub- 
land  and  grassland  songbird  communities. 

ACKNOWLEDGMENTS 

Logistical  and  financial  support  for  this  research  was 
provided  by  the  University  of  Massachusetts  at  Am- 
herst, Massachusetts  Audubon  Society,  Nantucket 
Conservation  Foundation,  Maria  Mitchell  Association, 
Partnership  for  Harrier  Habitat  Preservation,  A.  V. 
Stout  Fund,  Sweet  Water  Trust,  Quebec  Labrador 
Foundation,  Conservation  and  Research  Foundation, 
and  National  Fish  and  Wildlife  Foundation.  The  Nan- 
tucket Conservation  Foundation,  Nantucket  Land 
Bank  Commission,  and  Nantucket  Airport  provided 
access  to  their  properties.  A.  L.  Jones,  C.  R.  Griffin, 
D.  E.  Kroodsma,  R.  A.  Askins,  E.  M.  Steinauer,  K.  P. 
Combs-Beattie,  R.  Newman,  J.  F.  Lentowski,  E.  F.  An- 
drews, W.  T.  Maple,  B.  C.  McComb,  T.  Hosmer,  and 
K.  McGarigal  all  provided  helpful  comments  related 
to  this  study.  F.  M.  Zuckerberg  provided  critical  logis- 
tical support  and  is  warmly  acknowledged.  We  would 
like  to  thank  D.  H.  Johnson  and  two  anonymous  re- 
viewers for  their  thorough  reviews  of  this  manuscript. 

LITERATURE  CITED 

Arcese,  R,  M.  K.  Sogge,  A.  B.  Marr,  and  M.  A. 
Patten.  2002.  Song  Sparrow  ( Melospiza  melo- 
dia).  The  Birds  of  North  America,  no.  704. 
Askins,  R.  A.  2002.  Restoring  North  America’s  birds: 
lessons  from  landscape  ecology,  2nd  ed.  Yale  Uni- 
versity Press,  New  London,  Connecticut. 
Barbour,  H.,  T.  Simmons,  P.  Swain,  and  H.  Woolsey. 
1999.  Our  irreplaceable  heritage:  protecting  bio- 
diversity in  Massachusetts.  Natural  Heritage  and 
Endangered  Species  Program,  Massachusetts  Di- 
vision of  Fish  and  Wildlife  and  the  Massachusetts 
Chapter  of  The  Nature  Conservancy,  Boston, 
Massachusetts. 

Bayne,  E.  M.  and  K.  A.  Hobson.  2001.  Effects  of 
habitat  fragmentation  on  pairing  success  of  Ov- 
enbirds:  importance  of  male  age  and  floater  be- 
havior. Auk  1 18:380-388. 

Bibby,  C.  J.,  N.  D.  Burgess,  D.  A.  Hill,  and  S.  H. 
Mustoe.  2000.  Bird  census  techniques,  2nd  ed. 
Academic  Press,  New  York. 

Bollinger,  E.  K.,  P.  B.  Bollinger,  and  T.  B.  Gavin. 
1990.  Effects  of  hay-cropping  on  eastern  popula- 
tions of  Bobolink.  Wildlife  Society  Bulletin  18: 
142-150. 

Brower,  J.  E.  and  J.  H.  Zar.  1977.  Field  and  labo- 
ratory methods  for  general  ecology,  2nd  ed.  Wm. 
C.  Brown  Publishers,  Dubuque,  Iowa. 

Brown,  J.  H.  and  M.  V.  Lomolino.  1998.  Biogeog- 


raphy, 2nd  ed.  Sinauer,  Sunderland,  Massachu- 
setts. 

Christiansen,  D.  A.,  Jr.,  and  S.  E.  Reinert.  1990. 
Habitat  use  of  the  Northern  Harrier  in  a coastal 
Massachusetts  shrubland  with  notes  on  population 
trends  in  southeastern  New  England.  Journal  of 
Raptor  Research  24:84-90. 

Combs-Beattie,  K.  and  E.  M.  Steinauer.  2001.  An- 
nual report  of  the  Partnership  for  Harrier  Habitat 
Preservation  (PHHP)  Nantucket,  Massachusetts. 
Unpublished  Report,  Nantucket  Golf  Club,  Nan- 
tucket, Massachusetts. 

Confer,  J.  L.  and  S.  M.  Pascoe.  2003.  Avian  com- 
munities on  utility  rights-of-ways  and  other  man- 
aged shrublands  in  the  northeastern  United  States. 
Forest  Ecology  and  Management  185:193-205. 

Dechant,  J.  A.,  M.  L.  Sondreal,  D.  H.  Johnson,  L. 
D.  Igl,  C.  M.  Goldade,  M.  P.  Nenneman,  and  B. 
R.  Euliss.  2003.  Effects  of  management  practices 
on  grassland  birds:  Northern  Harrier.  Northern 
Prairie  Wildlife  Research  Center,  Jamestown, 
North  Dakota,  www.npwrc.usgs.gov/resource/ 
literatr/grasbird/noha/noha.htm  (ver.  12AUG2004; 
accessed  2 June  2005). 

Dettmers,  R.  2003.  Status  and  conservation  of  shrub- 
land  birds  in  the  northeastern  US.  Forest  Ecology 
and  Management  185:81-93. 

Donovan,  T.  M.  and  C.  H.  Flather.  2002.  Relation- 
ships among  North  American  songbird  trends, 
habitat  fragmentation,  and  landscape  occupancy. 
Ecological  Applications  12:364-374. 

Dunwiddie,  P.  W.  1989.  Forest  and  heath:  the  shaping 
of  the  vegetation  on  Nantucket  Island.  Journal  of 
Forest  History  33:126-133. 

Dunwiddie,  P.  W.  and  C.  Cauouw.  1990.  Prescribed 
burning  and  mowing  of  coastal  heathlands  and 
grasslands  in  Massachusetts.  Ecosystem  Manage- 
ment: rare  species  and  significant  habitats.  New 
York  State  Museum  Bulletin  471:271-275. 

Dunwiddie,  P.  W.  and  B.  A.  Sorrie.  1996.  A flora  of 
the  vascular  and  non-vascular  plants  of  Nantucket, 
Tuckernuck,  and  Muskeget  islands.  Rhodora  98: 
94-98. 

Fort,  K.  T.  and  K.  A.  Otter.  2004.  Effects  of  habitat 
disturbance  on  reproduction  in  Black-capped 
Chickadees  ( Poecile  atricapillus ) in  northern  Brit- 
ish Columbia.  Auk  121:1070-1080. 

Greenlaw,  J.  S.  1996.  Eastern  Towhee  ( Pipilo  ery- 
throphthalmus).  The  Birds  of  North  America,  no. 
262. 

Guzy,  M.  J.  and  G.  Ritchison.  1999.  Common  Yel- 
lowthroat  ( Geothlypis  trichas).  The  Birds  of  North 
America,  no.  448. 

Herkert,  J.  R.  1994.  The  effects  of  habitat  fragmen- 
tation on  midwestern  grassland  bird  communities. 
Ecological  Applications  4:461-471. 

Herkert,  J.  R.,  S.  A.  Simpson,  R.  L.  Westemeier,  T. 
L.  Esker,  and  J.  W.  Walk.  1999.  Response  of 
Northern  Harriers  and  Short-eared  Owls  to  grass- 
land management  in  Illinois.  Journal  of  Wildlife 
Management  63:517-523. 


Zuckerberg  and  Vickery  • EARLY-SUCCESSIONAL  BIRDS  ON  NANTUCKET  ISLAND  363 


Hosmer,  D.  W.  and  S.  Lemeshow.  1989.  Applied  lo- 
gistic regression.  John  Wiley  and  Sons,  New  York. 

Hutto,  R.  L.,  S.  M.  Pletschet,  and  P.  Hendricks. 
1986.  A fixed-radius  point  count  method  for  non- 
breeding and  breeding  season  use.  Auk  103:593- 
602. 

Johnson,  D.  H.  2002.  The  importance  of  replication  in 
wildlife  research.  Journal  of  Wildlife  Management 
66:919-932. 

Johnson,  D.  H.  and  L.  D.  Igl.  2001.  Area  require- 
ments of  grassland  birds:  a regional  perspective. 
Auk  118:24-34. 

Johnson,  D.  H.,  L.  D.  Igl,  and  J.  A.  Dechant  Shaffer 
(Series  Coord.).  2004.  Effects  of  management 
practices  on  grassland  birds.  Northern  Prairie 
Wildlife  Research  Center,  Jamestown,  North  Da- 
kota. www.npwrc.usgs.gov/resource/literatr/ 
grasbird/grasbird.htm  (ver.  12AUG2004;  accessed 
5 June  2005). 

Kleinbaum,  D.  G.,  L.  L.  Kupper,  K.  E.  Muller,  and 
A.  Nizam.  1998.  Applied  regression  analysis  and 
other  multivariable  methods.  Duxbury  Press,  New 
York. 

Litchfield,  M.  J.  1994.  Cape  Cod  and  the  islands: 
Nantucket.  A birder’s  guide  to  eastern  Massachu- 
setts. American  Birding  Association,  Colorado 
Springs,  Colorado. 

McGarigal,  K.  and  S.  A.  Cushman.  2002.  Compar- 
ative evaluation  of  experimental  approaches  to  the 
study  of  habitat  fragmentation  effects.  Ecological 
Applications  12:335-345. 

Mitchell,  L.  R.,  C.  R.  Smith,  and  R.  A.  Malecki. 
2000.  Ecology  of  grassland  breeding  birds  in  the 
northeastern  United  States:  a literature  review 
with  recommendations  for  management.  U.S. 
Geological  Survey,  Biological  Resources  Divi- 
sion, New  York  Cooperative  Fish  and  Widlife  Re- 
search Unit,  Dept,  of  Natural  Resources,  Cornell 
University,  Ithaca,  New  York. 

Rich,  T.  D.,  C.  J.  Beardmore,  H.  Berlanga,  P.  J. 
Blancher,  M.  S.  W.  Bradstreet,  G.  S.  Butcher, 

D.  W.  Demarest,  E.  H.  Dunn,  W.  C.  Hunter,  E. 

E.  Inigo-Elias,  et  al.  2004.  Partners  in  Flight 
North  American  landbird  conservation  plan.  Cor- 
nell Lab  of  Ornithology,  Ithaca,  New  York. 

SAS  Institute,  Inc.  1990.  SAS/STAT  user’s  guide. 
SAS  Institute,  Inc.,  Cary,  North  Carolina. 

Sauer,  J.  R.,  J.  E.  Hines,  and  J.  Fallon.  2005.  The 
North  American  Breeding  Bird  Survey,  results 
and  analysis  1966-2004,  ver.  2005.2.  U.S.  Geo- 
logical Survey  Patuxent  Wildlife  Research  Center, 


Laurel,  Maryland,  www.mbr-pwrc.usgs.gov/bbs/ 
html  (accessed  5 June  2005). 

Scott,  J.  M.,  F.  Davis,  B.  Csuti,  R.  Noss,  B.  Butter- 
field, C.  Groves,  H.  Anderson,  S.  Caicco,  F. 
D’Erchia,  T.  C.  Edwards,  Jr.,  et  al.  1993.  Gap 
analysis:  a geographic  approach  to  protection  of 
biological  diversity.  Wildlife  Monographs,  no. 
123. 

Sokal,  R.  R.  and  F.  J.  Rohlf.  1995.  Biometry.  W.  H. 
Freeman  and  Company,  New  York. 

Swain,  P.  C.  and  J.  B.  Kearsley.  2001.  Classification 
of  the  natural  communities  of  Massachusetts.  Nat- 
ural Heritage  and  Endangered  Species  Program, 
Massachusetts  Division  of  Fisheries  and  Wildlife, 
Westborough,  Massachusetts. 

Tiffney,  W.  N.,  Jr.,  and  D.  E.  Eveleigh.  1985.  Nan- 
tucket’s endangered  maritime  heaths.  Pages  1093- 
1109  in  Coastal  zone  ’85,  vol.  1 (O.  T.  Magoon, 
H.  Converse,  D.  Miner,  D.  Clark,  and  L.  T.  Tobin, 
Eds.).  American  Society  of  Civil  Engineers,  New 
York. 

Vickery,  P.  D.  1992.  A regional  analysis  of  endan- 
gered, threatened,  and  special-concern  birds  in  the 
northeastern  United  States.  Transactions  of  the 
Northeast  Section  of  The  Wildlife  Society  48:1- 
10. 

Vickery,  P.  D.,  M.  L.  Hunter,  Jr.,  and  S.  M.  Melvin. 
1994.  Effects  of  habitat  area  on  the  distribution  of 
grassland  birds  in  Maine.  Conservation  Biology 
8:1087-1097. 

Vickery,  P.  D.,  M.  L.  Hunter,  Jr.,  and  J.  V.  Wells. 
1999.  Effects  of  fire  and  herbicide  treatment  on 
habitat  selection  in  grassland  birds  in  southern 
Maine.  Studies  in  Avian  Biology  19:149-159. 

Vickery,  P.  D.,  B.  Zuckerberg,  A.  L.  Jones,  W.  G. 
Shriver,  and  A.  P.  Weik.  2005.  Influence  of  fire 
and  other  anthropogenic  practices  on  grassland 
and  shrubland  birds  in  New  England.  Studies  in 
Avian  Biology  30:139-146. 

Wheelwright,  N.  T.  and  J.  D.  Rising.  1993.  Savannah 
Sparrow  ( Passerculus  sandwichensis).  The  Birds 
of  North  America,  no.  45. 

Wiens,  J.  A.  and  J.  T.  Rotenberry.  1985.  Response 
of  breeding  passerine  birds  to  rangeland  alteration 
in  a North  American  shrubsteppe  locality.  Journal 
of  Applied  Ecology  22:655-668. 

Wiens,  J.  A..  J.  T.  Rotenberry,  and  B.  Van  Horne. 
1986.  A lesson  in  the  limitations  of  field  experi- 
ments: shrubsteppe  birds  and  habitat  alteration. 
Ecology  67:365-376. 

Winter,  M.  and  J.  Faaborg.  1999.  Patterns  of  area 
sensitivity  in  grassland-nesting  birds.  Conserva- 
tion Biology  13:1424-1436. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):364-373,  2006 


SPATIAL  BEHAVIOR  OF  EUROPEAN  ROBINS  DURING 
MIGRATORY  STOPOVERS:  A TELEMETRY  STUDY 

NIKITA  CHERNETSOV1 3 AND  ANDREY  MUKHIN1  2 


ABSTRACT. — We  studied  the  movement  patterns  of  European  Robins  ( Erithacus  rubecula)  at  stopovers 
during  spring  and  fall  migration  on  the  southeastern  Baltic  Coast,  Russia.  On  the  1st,  and  sometimes  the  2nd, 
day  after  arrival  at  a stopover  site,  robin  movements  were  less  aggregated  than  those  made  on  subsequent  days. 
Search/settling  time  varied  between  several  hours  and  2 days.  During  this  period,  migrants  either  occupied  a 
defined  stopover  area  or  left  the  site.  Stopover  duration  was  1 to  12  days  in  spring  (mean  = 2.4  days  ± 0.31 
SE)  and  1 to  14  days  in  fall  (mean  = 3.4  days  ± 0.50).  The  home-range  size  of  European  Robins  on  the 
southeastern  Baltic  Coast  did  not  differ  between  seasons  (spring:  4,320  m1 2 3  ± 545,  n = 15;  fall:  3,562  m2  ± 
598,  n = 15)  and  was  similar  to  that  at  a central  European  site  in  fall  (4,264  m2  ± 241,  n = 14).  These  home 
ranges  were  not  defended  territories.  We  found  no  relationship  between  the  robins’  spatial  behavior  and  their 
fat  stores  on  arrival,  although  in  spring  more  lean  than  fat  robins  stopped  for  >2  days.  The  pattern  of  movements 
at  the  stopover  was  variable,  both  in  birds  that  arrived  lean  and  those  that  arrived  with  much  more  fat.  Stopover 
duration  estimates  based  on  radio-tagging  are  superior  to  those  based  on  capture-mark-recapture.  Received  27 
December  2004,  accepted  23  January  2006. 


Passerines  spend  at  least  90%  of  their  time 
during  migration  at  migratory  stopover  sites. 
Stopover  variables  (e.g.,  rates  of  fat  deposi- 
tion, predation  risk,  habitat  suitability)  strong- 
ly influence  migration  strategies  and  tactics 
(Lindstrom  2003).  Another  important  aspect 
of  migrant  stopover  ecology  is  spatial  behav- 
ior— territoriality  versus  broader  movements, 
size  of  temporary  home  ranges,  and  sharing 
of  home  ranges  versus  defending  them  from 
conspecifics  (Chernetsov  2003,  Chernetsov 
and  Bolshakov  in  press).  Some  migrants  oc- 
cupy temporary  territories  at  stopovers  (Rap- 
pole  and  Warner  1976;  Kodric-Brown  and 
Brown  1978;  Bibby  and  Green  1980,  1981; 
Carpenter  et  al.  1983,  1993a,  1993b),  whereas 
others  move  broadly  across  a given  stopover 
area.  Intraspecific  variation  in  spatial  behavior 
has  also  been  reported;  some  individuals  oc- 
cupy relatively  small  home  ranges,  whereas 
others  move  over  much  broader  areas  (Aborn 
and  Moore  1997,  Delingat  and  Dierschke 
2000).  Until  recently,  capture-recapture  anal- 
ysis has  been  the  main  method  for  studying 
the  pattern  of  movements  made  by  passerines 
at  stopovers  (Titov  1999a,  1999b;  Chernetsov 
and  Titov  2001;  Chernetsov  2002),  and  these 


1 Biological  Station  Rybachy,  Rybachy  238535.  Ka- 
liningrad Region,  Russia. 

2 Max  Planck  Research  Inst,  for  Ornithology,  Von- 
Der-Tann-Str.  7,  D-82346  Andechs,  Germany. 

3 Corresponding  author;  e-mail: 
nchernetsov@bioryb.koenig.ru 


analyses  suggest  that — during  fall  (south- 
bound) migration — European  Robins  (Eritha- 
cus rubecula)  occupy  defined  stopover  areas 
(DSA).  Robins  spend  up  to  2 days  occupying 
a DSA  (Titov  1999a)  and,  after  a maximum 
of  2 days,  either  resume  migration  or  settle  in 
a defined  home  range. 

An  important  weakness  of  capture-recap- 
ture analysis  is  that  the  capture  probability  of 
passerine  migrants  at  stopovers  is  usually  low 
(Chernetsov  and  Titov  2000)  and  most  likely 
differs  between  groups  of  birds  (e.g.,  fat  ver- 
sus lean  birds,  those  refueling  versus  those 
losing  weight,  and  new  arrivals  versus  those 
occupying  a DSA).  Radio-tracking  has  been 
used  more  recently  (Aborn  and  Moore  1997, 
Lajda  2001),  which  makes  it  possible  to  as- 
certain the  location  of  a bird  without  having 
to  capture  it  or  otherwise  influence  its  behav- 
ior. 

We  investigated  movement  patterns  of  ra- 
dio-tagged European  Robins  during  spring 
(northbound)  and  fall  migration  stopovers  on 
the  southeastern  Baltic  Coast,  Russia.  Our  ob- 
jectives were  (1)  to  test  the  hypothesis  that 
individuals  remain  within  defined  areas  at 
stopover  sites;  (2)  to  estimate  home-range 
area  and  settling  time;  and  (3)  to  assess  the 
impact  of  initial  fat  stores  on  robins’  spatial 
behavior.  Understanding  patterns  of  spatial 
use  by  migrants  within  habitats,  including 
habitats  being  lost  or  fragmented,  is  crucial 
for  understanding  the  importance  of  relatively 


364 


Chernetsov  and  Mukhin  • STOPOVER  BEHAVIOR  OF  EUROPEAN  ROBINS 


365 


TABLE  1.  Number  and  condition  of  European  Robins  radio-tagged  and  followed  during  spring  (northbound) 
and  fall  (southbound)  migration  stopover,  2002-2003,  on  the  Courish  Spit,  southeastern  Baltic  Coast,  Russia. 


Season 

No.  tagged 
at  stopover 

No.  followed 
from  the  1st  day 

No.  followed  from 
the  1st  to  the  last  day 

No.  fat  birds3 

No.  lean  birds3 

Spring 

2002 

21 

12 

10 

13 

4 

2003 

30 

30 

29 

16 

14 

Total  spring 

51 

42 

39 

29 

18 

Fall 

2002 

29 

25 

24 

10 

19 

2003 

36 

36 

35 

17 

19 

Total  fall 

65 

61 

59 

27 

38 

aBody  mass  of  “lean”  birds  exceeded  their  calculated  lean  body  mass  by  <1.2  g;  body  mass  of  “fat”  birds  exceeded  their  calculated  lean  body  mass 
by  >1.5  g. 


large  versus  small  habitat  patches.  Habitat  use 
and  spatial  behavior  of  migratory  landbirds 
have  not  been  studied  adequately,  in  spite  of 
their  importance  as  conservation  issues  (Petit 
2000). 

METHODS 

Study  site. — We  conducted  our  study  during 
spring  and  fall,  2002-2003,  at  Biological  Sta- 
tion Rybachy  on  Cape  Rossitten  on  the  Cour- 
ish Spit,  Russia  (southeastern  Baltic  coast, 
55°  09'  N,  20°  51'  E).  Our  study  periods  were 
1 April  to  4 May  2002,  13  April  to  7 May 
2003,  2 September  to  29  October  2002,  and  6 
September  to  8 November  2003.  The  overall 
area  of  the  study  site  is  6 ha.  Vegetation  at 
the  study  site  is  a mosaic  of  willow  (Salix 
spp.)  scrub  and  common  reed  ( Phragmites 
communis ),  and  some  trees,  including  rowan 
trees  ( Sorbus  aucuparia),  white  willows  ( Salix 
alba),  and  bird  cherry  ( Prunus  racemosa).  We 
mist-netted  European  Robins — the  most  com- 
monly occurring  migratory  species  captured  at 
this  site  (Bolshakov  et  al.  2002) — and  banded 
them  with  aluminum  leg-bands  (Moscow 
Ringing  Center  bands). 

Radio-tagged  birds. — We  fitted  117  Euro- 
pean Robins  with  radio  transmitters  (Table  1). 
To  obtain  unbiased  estimates  of  stopover  du- 
ration, we  made  every  effort  to  tag  birds  just 
after  their  arrival.  The  rate  of  daily  captures 
of  small  passerines,  including  European  Rob- 
ins, at  our  study  site  are  highly  variable  (due 
to  occurrence  of  migration  waves),  as  it  is  at 
many  other  coastal  sites  (Dolnik  1975,  Titov 
and  Chernetsov  1999,  Chernetsov  and  Titov 
2000).  Results  of  seniority  analysis  (i.e.,  cap- 


ture-mark-recapture models  applied  back- 
wards in  time;  Pradel  1996)  indicate  that  the 
vast  majority  of  European  Robins  initially 
captured  on  days  when  many  new  birds  are 
banded  (following  a day  of  few  captures)  have 
just  arrived  (Titov  and  Chernetsov  1999, 
Chernetsov  and  Titov  2000). 

In  2003,  all  birds  were  radio-tagged  on  the 
1st  day  of  a migration  wave  ( n = 66).  In  2002, 
most  European  Robins  were  radio-tagged  on 
the  1st  day  of  a migration  wave  ( n = 37), 
while  others  were  radio-tagged  upon  recapture 
on  the  2nd  or  3rd  day  after  their  initial  band- 
ing ( n = 13).  We  assume  that  our  estimates 
of  stopover  duration  of  tagged  birds  are  un- 
biased. 

All  birds  radio-tagged  in  fall  were  in  their 
hatching  year;  in  spring,  all  birds  were  in  their 
2nd  calendar  year.  Bolshakov  et  al.  (2003) 
used  linear  regression  of  body  mass  on  wing 
length  to  calculate  lean  body  mass  of  Euro- 
pean Robins  that  had  no  visible  subcutaneous 
fat  (fat  score  0,  after  Kaiser  1993);  they  made 
separate  calculations  for  September,  October 
(fall)  and  April  (spring).  Based  on  those  cal- 
culations, all  radio-tagged  robins  in  our  study 
were  categorized  as  either  “fat”  or  “lean” 
(Table  1);  lean  birds  exceeded  their  calculated 
lean  body  mass  by  <1.2  g (<0.5  g in  63.4% 
of  birds),  and  fat  birds  exceeded  their  calcu- 
lated lean  body  mass  by  >1.5  g (>2.0  g in 
93.8%  of  birds).  If  a bird  was  radio-tagged 
when  recaptured  rather  than  when  it  was  first 
captured  (which  occurred  in  spring  2002),  its 
fat  score  at  the  time  of  radio-tagging  was  used 
to  assign  it  to  the  fat  or  lean  group.  The  mass 
and  wing  length  of  birds  at  capture  were  re- 


366 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


corded  to  the  nearest  0.1  g and  0.5  mm,  re- 
spectively. 

Telemetry  protocol. — We  radio-tagged  Eu- 
ropean Robins  with  LB-2  transmitters  (Holo- 
hil  Systems,  Carp,  Ontario,  Canada).  The 
measured  life  span  of  the  transmitters  was  at 
least  10  days  during  spring  passage  and  21 
days  during  fall  migration.  Transmitters  were 
fitted  as  backpacks  with  a Rappole  harness 
(Rappole  and  Tipton  1991).  The  weight  of  a 
transmitter  with  harness  was  0.61  g,  and  the 
body  mass  of  radio-tagged  European  Robins 
varied  between  14.8  and  19.2  g;  thus,  the  mass 
of  transmitters  represented  3.2-4. 1%  of  a 
bird’s  body  mass  (<5%  is  believed  to  be  the 
upper  limit  permissible;  Caccamise  and  Hedin 
1985,  Naef-Daenzer  1993). 

We  used  receivers  with  Yagi  antennae  from 
Wildlife  Materials  (Carbondale,  Illinois)  and 
Advanced  Telemetry  Systems  (Isanti,  Minne- 
sota). The  location  of  birds  was  estimated  by 
biangulation  and  triangulation.  For  each  indi- 
vidual, one  location  per  hr  was  taken  between 
the  onset  of  daytime  activity  (dawn)  and  even- 
ing civil  twilight.  The  number  of  observations 
per  individual  per  day  varied  between  1 1 and 
17,  depending  on  the  duration  of  the  daylight 
period.  Locations  were  plotted  on  a digitized 
map  of  the  study  area.  From  sunset  to  dawn, 
all  birds  were  surveyed  continuously  from  a 
stationary  watch  point  15  m above  ground 
level;  therefore,  migratory  departure  time  was 
usually  detected  to  the  nearest  1-3  min  and 
the  exact  night  of  departure  was  known.  Mi- 
gratory departures  invariably  occurred  during 
the  nighttime.  Generally,  birds  were  absolute- 
ly stationary  during  the  night  (no  signal 
change  caused  by  movements);  thus,  an  abrupt 
signal  change  indicated  take-off.  The  signal 
could  usually  be  received  from  the  flying  bird 
for  some  time  (1-20  min),  but  it  later  disap- 
peared. As  the  range  of  transmitter  detectabil- 
ity did  not  exceed  1.5  km,  signal  reception 
from  a flying  bird  for  more  than  3-4  min 
clearly  indicated  that  a bird  was  flying  in  cir- 
cles before  choosing  a direction.  This  behav- 
ior was  very  distinctive,  and  the  probability 
that  some  other  nocturnal  activity  was  mistak- 
en for  a migratory  departure  was  small.  If  a 
bird  left  the  study  area  and  occupied  a home 
range  elsewhere,  the  data  for  that  bird  were 
included  only  in  qualitative  estimates  of 
whether  or  not  the  bird  occupied  a DSA.  If  a 


bird  spent  the  night  far  enough  from  the  sta- 
tionary watch  point  to  preclude  signal  recep- 
tion at  the  stationary  site,  we  attempted  to  lo- 
cate it  every  1-2  hr  until  dawn.  A bird  was 
assumed  to  have  departed  if  the  signal  could 
not  be  detected  during  that  night. 

Data  analyses. — We  tested  the  locations  for 
statistical  independence  by  using  the  Schoener 
index  (Swihart  and  Slade  1985).  The  data 
were  not  formally  independent  (i.e.,  consecu- 
tive locations  were  aggregated  with  a greater- 
than-chance  probability);  nevertheless,  we  as- 
sumed that  our  data  could  be  used  for  the 
analysis  of  spatial  distribution.  We  based  our 
assumption  on  the  empirical  rule  suggested  by 
White  and  Garrott  (1990),  which  states  that  if 
enough  time  has  elapsed  between  two  consec- 
utive observations  for  an  animal  to  move  from 
one  end  of  its  home  range  to  another,  the  ob- 
servations in  question  may  be  considered  sta- 
tistically independent.  In  our  study,  at  least  45 
min  elapsed  between  observations,  during 
which  each  individual  would  have  had  ample 
time  to  move  to  any  point  in  its  stopover  area. 

When  locating  birds,  every  effort  was  made 
to  approach  them  as  closely  as  possible  to 
minimize  location  error.  We  believe  that  in 
most  cases  we  located  their  positions  to  the 
nearest  5 m and,  following  Lajda  (2001),  as- 
sumed a standard  deviation  of  10  m.  Home- 
range  area  was  estimated  on  the  basis  of  all 
locations  available  as  95%  kernel  by  Animal 
Movement  Extension  in  ArcView  (Hooge  and 
Eichenlaub  2000).  The  estimated  home-range 
area  increases  with  an  increasing  number  of 
locations  until  that  number  reaches  40—50 
(Lajda  2001);  therefore,  we  did  not  estimate 
the  home-range  area  of  birds  with  <38  loca- 
tions. Due  to  this  limitation,  we  only  estimat- 
ed home-range  area  for  the  entire  stopover  pe- 
riod and  for  the  birds  that  stopped  for  >4  days 
(n  = 30).  To  estimate  the  aggregation  of  lo- 
cations from  birds  that  were  followed  during 
shorter  periods  of  time,  we  used  the  linearity 
index  as  applied  in  Animal  Movement  Exten- 
sion of  ArcView  (Hooge  and  Eichenlaub 
2000);  this  is  the  linear  distance  moved  (i.e., 
the  distance  between  the  initial  and  final  lo- 
cations) divided  by  cumulative  distance  be- 
tween all  successive  locations.  The  maximum 
value  of  the  linearity  index  is  1 (i.e.,  if  a bird 
is  moving  along  a straight  line).  This  index 
may  be  calculated  for  a given  time  interval 


Chernetsov  and  Mukhin  • STOPOVER  BEHAVIOR  OF  EUROPEAN  ROBINS 


367 


FIG.  1.  Frequency  distribution  of  stopover  dura- 
tions of  European  Robins  assessed  by  radio  tracking 
in  spring  (northbound)  and  fall  (southbound),  2002- 
2003,  on  the  Courish  Spit,  southeastern  Baltic  Coast, 
Russia.  Only  birds  radio-tagged  on  the  1st  day  after 
arrival  and  known  to  depart  by  nocturnal  flight  are 
included.  Spring:  2.4  days  ± 0.31,  median  = 2,  n = 
40;  fall:  3.4  days  ± 0.50,  median  = 2,  n = 59. 


(e.g.,  the  total  observation  period  or  a single 
day)  and  is  a measure  of  area-restricted  move- 
ment. The  linearity  index  is  reciprocal  to  the 
meander  ratio  (Williamson  and  Gray  1975) 


0.5 


0.4 


. 29 


Spring 


and  was  preferred  to  it  due  to  the  statistical 
properties  of  the  linearity  index.  We  used  the 
arbitrarily  selected  threshold  of  0.10  as  an  in- 
dication that  a bird  occupied  a DSA;  we  as- 
sumed that  birds  showing  linearity  index  val- 
ues below  this  threshold  remained  in  a DSA. 
For  comparison,  Aborn  and  Moore  (1997) 
found  that  the  meander  ratio  for  Summer  Tan- 
agers  ( Piranga  rubra ) “settled”  at  stopovers 
on  the  Gulf  of  Mexico  coast  averaged  4.8, 
which  corresponds  to  a linearity  index  of  0.21 . 
Thus,  our  threshold  was  rather  conservative. 

We  used  r-tests  to  compare  pairs  of  means 
when  the  assumption  of  population  normality 
was  not  violated,  and  we  used  nonparametric 
Mann- Whitney  U- tests  when  normality  was 
clearly  violated  (e.g.,  distribution  of  stopover 
duration  values.  Fig.  1).  We  also  used  Spear- 
man’s rank  correlation  when  the  normality  as- 
sumption was  violated.  We  used  ANOVA  to 
compare  multiple  samples,  and  we  used  Tu- 
key’s  honestly  significant  difference  tests  for 
post-hoc  analyses.  All  tests  were  two-tailed; 
the  null  hypothesis  was  rejected  if  P < 0.05; 
means  are  presented  ± SE.  Data  analyses 
were  performed  using  SPSS  version  11.0 
(SPSS,  Inc.  1999). 


0.3-  20 

■ m 6 64 

II  lliiiiilli 


Days  since  arrival 


FIG.  2.  Daily  linearity  index  values  of  European 
Robins  during  spring  (northbound)  and  fall  (south- 
bound) migration  stopovers,  2002-2003,  on  the  Cour- 
ish Spit,  southeastern  Baltic  Coast,  Russia.  Sample  siz- 
es are  shown  above  the  histogram  bars.  Days  with 
mean  linearity  index  values  significantly  different  from 
the  remaining  days  (one-way  ANOVA  with  post-hoc 
tests)  are  shown  by  open  bars. 


RESULTS 

Spring  Migration 

Stopover  duration  and  establishing  a 
DSA. — The  stopover  duration  of  European 
Robins  during  spring  migration  varied  from  1 
to  12  days  (Fig.  1).  Twelve  of  40  birds  radio- 
tagged  on  the  1st  day  after  arrival  (30%) 
stopped  for  >2  days.  The  mean  stopover 
length  was  2.4  days  ± 0.31. 

We  plotted  the  movements  of  33  birds  from 
the  1st  until  the  last  day  of  stopover.  We  ob- 
tained at  least  6,  and  up  to  92,  locations  over 
1-6  days  from  these  birds.  The  linearity  index 
for  these  birds  varied  from  0.008  (very  aggre- 
gated locations)  to  0.65  (nearly  straight-line 
movement)  and  was  negatively  correlated 
with  both  number  of  locations  (Spearman’s 
rank  correlation:  rs  = -0.69,  P < 0.001)  and 
stopover  duration  in  days  (rs  = —0.58,  P < 
0.001).  The  longer  a bird  remained  at  stop- 
over, the  more  aggregated  its  locations  were. 

We  also  calculated  the  linearity  index  for 
each  stopover  day  (Fig.  2).  The  pattern  was 
rather  obvious:  during  the  1st  day  of  stopover. 


368 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


B 


0 


FIG.  3.  Examples  of  the  distributions  of  locations 
of  two  different  birds  during  spring  (northbound)  and 
fall  (southbound)  migration  stopovers,  2002-2003,  on 
the  Courish  Spit,  southeastern  Baltic  Coast,  Russia. 
Each  dot  represents  a single  location.  (A)  All  locations 
are  in  the  defined  stopover  area  (DSA).  (B)  Some  lo- 
cations are  associated  with  the  search/settling  period; 
others  are  in  the  DSA. 


robins  moved  broadly,  and  from  the  2nd  day 
on  they  began  to  remain  in  a more  restricted 
area  (one-way  ANOVA:  F1097  = 6.85,  P < 
0.001).  The  linearity  index  for  day  1 differed 
from  that  of  all  other  days  (Tukey’s  honestly 
significant  difference  test;  all  P < 0.008).  For 
movements  during  the  first  day,  the  linearity 
index  did  not  differ  between  birds  continuing 
with  migration  on  the  1st  night  and  those  that 
remained  for  more  than  1 day  (t  = 1.21,  P = 
0.20,  nx  = 14,  n2  = 15).  This  means  that  on 
the  1st  day  of  stopover,  the  birds  behaved  the 
same  as  they  did  on  subsequent  days:  their 
movement  patterns  were  not  indicative  of 
their  subsequent  decisions  to  remain  or  depart. 

The  movements  of  European  Robins  that 
remained  for  several  days  showed  varying 
patterns.  In  some  cases,  all  locations  were  ag- 
gregated (Fig.  3A).  In  others,  first  locations, 
presumably  from  the  search/settling  period, 
were  more  dispersed  (Fig.  3B).  We  were  able 
to  estimate  home-range  area  for  15  European 
Robins  (where  n > 38  telemetry  locations;  Ta- 
ble 2).  DSA  size  was  negatively  correlated 
with  the  number  of  locations  (r  = —0.54,  P 
= 0.036).  Birds  that  stopped  over  for  a long 
time  (and  thus  yielded  many  location  points) 
tended  to  remain  within  a more  clearly  defined 
area. 

Behavior  of  fat  and  lean  birds. — Of  the  51 
European  Robins  included  in  the  analysis  of 
spatial  behavior,  1 8 were  lean  at  radio-tagging 
(fat  stores  <0.5  g),  29  were  fat  (fat  stores  >2 
g),  and  4 had  intermediate  fat  stores.  The 
transmitter  was  removed  from  one  lean  bird, 
so  its  stopover  duration  was  unknown.  Of  the 
remaining  17  lean  birds,  10  (59%)  stopped  for 
>2  days,  and  mean  stopover  length  was  3.8 
days  ± 0.75.  The  linearity  index  values  of  all 
these  10  birds  were  <0.10,  and  we  assumed 
that  they  occupied  a DSA.  Of  seven  lean  birds 
that  stopped  for  1-2  days,  two  remained  with- 


TABLE  2.  Home-range  size  (m2)  of  European  Robins  during  spring  (northbound)  and  fall  (southbound) 
migration  stopovers  on  the  Courish  Spit  (Rybachy),  southeastern  Baltic  Coast,  Russia  (this  study)  and  during 
fall  migration  in  southwestern  Germany  (Mettnau;  Lajda  2001).  There  was  no  significant  difference  between 
Rybachy  and  Mettnau  in  fall  (t  = 0.95,  P = 0.35)  nor  between  seasons  in  Rybachy  ( t = 0.94,  P = 0.38). 


Range  (m2) 

Mean  (m2) 

Median  (m2) 

SE 

» 

Source 

Spring,  Rybachy 

1,932-9,215 

4,320 

4,091 

545 

15 

This  study 

Fall,  Rybachy 

1,060-10,083 

3,562 

2,801 

598 

15 

This  study 

Fall,  Mettnau 

1,900-7,600 

4,264 

4,400 

421 

14 

Lajda  (2001) 

Chernetsov  and  Mukhin  • STOPOVER  BEHAVIOR  OF  EUROPEAN  ROBINS 


369 


in  a small  defined  area,  three  roamed  broadly, 
and  two  yielded  too  few  locations  to  assign 
their  spatial  behavior  as  either  DSA  owners  or 
roamers. 

Of  the  29  initially  fat  birds,  seven  (24%) 
remained  for  >2  days;  the  mean  stopover  du- 
ration was  2.6  days  ± 0.53.  All  seven  birds 
that  stopped  over  for  >2  days  occupied  a 
DSA.  Of  21  birds  that  departed  after  1-2  days, 
1 1 moved  broadly  (linearity  index  >0.25). 
The  difference  in  stopover  duration  between 
fat  and  lean  birds  was  not  significant  (Mann- 
Whitney  U- test:  z = 155,  P = 0.12);  however, 
the  proportion  of  birds  that  stopped  for  >2 
days  was  greater  among  lean  birds  (Yates-cor- 
rected x2  — 4.15,  P = 0.041). 

Home-range  area  in  birds  that  arrived  fat 
(4,101  m2  ± 493,  n = 5)  and  those  that  arrived 
lean  (4,683  m2  ± 976,  n — 8)  did  not  differ 
(/-test,  t = 0.44,  P = 0.67);  however,  we  could 
only  estimate  home-range  area  in  individuals 
that  stopped  over  for  >4  days.  The  linearity 
index  did  not  differ  between  birds  that  arrived 
lean  and  those  that  arrived  fat  on  either  the 
1st  day  of  stopover  (fat:  0.34  ± 0.039,  n = 
16;  lean:  0.32  ± 0.059,  n = 11;  median  test: 
X2  = 0.30,  P = 0.58)  or  on  the  2nd  day  (fat: 
0.18  ± 0.037,  n = 11;  lean:  0.15  ± 0.040,  n 
= 6;  median  test:  x2  = 0.03,  P = 0.86).  Ap- 
parently, both  lean  and  fat  birds  can  show  var- 
ious spatial  patterns  in  the  first  days  after  ar- 
rival. We  did  not  compare  linearity  indices  of 
initially  lean  and  initially  fat  birds  in  the  sub- 
sequent (>2)  days  after  arrival,  because  the 
chance  was  too  high  that  the  nutritional  status 
of  the  birds  had  already  changed. 

Fall  Migration 

Stopover  duration  and  establishing  a 
DSA. — Fall  stopover  duration  varied  between 
1 and  14  days  (Fig.  1).  Twenty-three  European 
Robins  of  the  59  tracked  since  the  1st  day  of 
stopover  remained  over  for  >2  days.  The 
mean  stopover  length  was  3.4  days  ± 0.50 
(Fig.  1),  which  did  not  differ  significantly 
from  the  duration  of  spring  stopovers  (2.4 
days  ± 0.31;  Mann- Whitney  fZ-test:  z = 0.03, 
P = 0.97). 

Of  the  birds  that  stopped  for  >2  days  (n  = 
23),  all  but  one  occupied  a DSA.  One  bird  that 
stopped  for  3 days  in  fall  2003  covered  a lin- 
ear distance  of  ~4  km,  moving  during  day- 
time before  it  departed.  Home-range  size  was 


estimated  for  15  individuals  for  which  at  least 
39  locations  were  obtained  per  bird  (Table  2). 
The  number  of  locations  was  not  significantly 
correlated  with  home-range  size  (r  = —0.43, 
P = 0.1 1).  The  area  of  DSAs  occupied  during 
fall  migration  did  not  differ  significantly  from 
the  area  of  DSAs  occupied  in  spring  (Table 
2). 

In  fall,  European  Robins  spent  from  several 
hr  to  1 .5  days  moving  around  before  settling. 
In  one  case,  a European  Robin  that  settled  in 
a DSA  on  the  1st  day  changed  its  DSA  on  the 
morning  of  the  4th  day.  This  individual  de- 
parted by  nocturnal  flight  after  a 5 -day  stop- 
over. 

We  tracked  42  birds  from  the  1st  until  the 
last  day  of  stopover.  We  obtained  4-172  lo- 
cations over  1—14  days  from  these  birds.  The 
linearity  index  of  their  movements  varied 
from  0.003  to  0.93  and  was  negatively  related 
to  both  number  of  locations  (Spearman’s  rank 
correlation:  rs  = —0.55,  P < 0.001)  and  stop- 
over duration  in  days  (rs  m —0.56,  P < 
0.001).  Individuals  that  stopped  over  for  lon- 
ger periods  showed  more  area-restricted 
movement. 

In  fall,  the  linearity  index  differed  between 
the  days  of  stopover  (one-way  ANOVA:  F9  149 
= 6.69,  P < 0.001).  The  days  with  linearity 
index  values  different  from  the  others  were 
days  1 and  2 (both  different  from,  e.g.,  day  4, 
Tukey’s  honestly  significant  difference  test:  P 
< 0.001  in  both  cases).  Beginning  with  the 
3rd  day  of  stopover,  there  was  no  significant 
between-day  variation  in  the  linearity  index 
(post-hoc  tests;  all  P > 0.05).  The  linearity 
index  did  not  differ  between  the  1st  and  the 
2nd  day  of  stopover  (Tukey’s  HSD  test:  P = 
0.56).  On  the  1st  day,  the  linearity  index  did 
not  differ  between  birds  continuing  migration 
on  the  next  night  and  those  that  remained  for 
more  than  1 day  ( t = 0.97,  P = 0.34,  nx  = 
28,  n2  — 27). 

Behavior  of  fat  and  lean  birds. — Of  65  Eu- 
ropean Robins  radio  tracked  in  fall,  38  were 
lean  when  radio-tagged  and  27  were  fat  (Table 
1).  Of  the  38  lean  birds,  19  (50%)  stopped 
over  for  >2  days.  Mean  stopover  duration  was 
4.1  days  ± 0.67  (median  = 2 days,  n = 36); 
for  two  birds,  stopover  duration  was  not 
known  exactly,  but  was  >2  days.  Of  the  19 
lean  robins  that  stopped  over  for  >2  days,  18 
occupied  a DSA  (linearity  index  <0.10).  The 


370 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


only  bird  with  a higher  linearity  index  (0.22), 
stopped  for  3 days.  Of  the  19  lean  birds  that 
spent  1-2  days  at  the  stopover,  the  movements 
of  10  were  not  very  area-restricted  (linearity 
index  >0.25).  Of  the  27  initially  fat  robins,  9 
(33%)  stopped  for  >2  days,  and  mean  stop- 
over duration  was  3.2  days  ± 0.69  (median  = 

1 day).  The  difference  in  stopover  duration 
between  fat  and  lean  birds  was  not  significant 
(Mann-Whitney  U- test:  z = 0.74,  P = 0.43). 
The  difference  in  the  proportion  of  fat  and 
lean  birds  that  stopped  over  for  >2  days  also 
was  not  significant  (Yates-corrected  x2  — 1.17, 
P = 0.28). 

As  in  spring,  there  was  no  difference  in  the 
size  of  DSAs  between  initially  fat  (2,970  m2 
± 518,  n = 6)  and  initially  lean  (3,957  m2  ± 
939,  n = 9)  birds  ( t = 0.80,  P = 0.44).  Stop- 
over area  could  be  estimated  only  for  robins 
that  made  longer  stopovers  (>4  days),  during 
which  their  nutritional  status  might  have 
changed.  All  birds  that  carried  large  fat  stores 
at  arrival  and  stopped  over  for  >2  days  ( n — 
9)  occupied  a DSA.  The  linearity  index  was 
<0.10  in  all  cases  in  which  it  was  possible  to 
calculate  ( n — 6).  Fat  robins  that  stayed  for 
1-2  days  (n  = 19)  moved  across  a large  area 
(linearity  index  >0.25  in  10/14  cases).  Five 
birds  were  tracked  for  too  short  a time  to  es- 
timate their  spatial  status. 

DISCUSSION 

Even  though  the  maximum  stopover  dura- 
tion assessed  by  radio  tracking  was  12  days 
in  spring  and  14  days  in  fall,  the  medians  were 

2 days  and  1 day,  respectively.  In  spring  and 
in  fall,  70%  and  61%,  respectively,  of  Euro- 
pean Robins  resumed  migration  after  1 or  2 
days  of  stopover.  Even  though  there  was  a 
weak  tendency  among  lean  birds  to  make 
longer  stopovers,  it  was  not  statistically  sig- 
nificant. Optimal  migration  theory  predicts 
that  in  time-minimizing  migrants,  stopover 
duration  should  depend  on  migrant  fuel  status 
and  fat-deposition  rate  (Alerstam  and  Lind- 
strom  1990).  Wind  direction  and  strength  are 
also  of  paramount  importance  (Liechti  and 
Bruderer  1998).  Our  data,  like  that  of  some 
other  studies  (e.g.,  Rguibi-Idrissi  et  al.  2003), 
indicate  that  relationships  between  individual 
stopover  parameters  (e.g.,  stopover  duration 
and  fat  status)  are  often  not  as  straightforward 


as  predicted  by  the  necessarily  simplified 
models. 

Our  telemetry  study  of  European  Robins  at 
a migratory  stopover  showed  that  all  birds  that 
stopped  over  for  >2  days  occupied  a DSA. 
Previously,  this  pattern  has  been  predicted  on 
the  basis  of  capture-recapture  analysis  (Szulc- 
Olech  1965,  Titov  1999b);  however,  analysis 
based  on  recaptures  is  an  indirect  method  that 
is  strongly  dependent  on  the  recapture  prob- 
abilities of  the  birds.  Our  telemetry  data, 
which  are  independent  of  recapture  probabil- 
ity, confirmed  the  hypothesis  that  European 
Robins  first  move  around  broadly,  and,  after 
1-2  days,  either  settle  in  a DSA  or  resume 
migration.  During  the  first  2 days  after  arrival, 
roughly  one-half  of  the  birds  remained  within 
a restricted  area  and  one-half  moved  broadly 
(high  linearity  index).  The  latter  pattern  was 
especially  typical  of  the  1st  day  after  arrival. 
The  maximum  linear  range  of  European  Rob- 
in movements  was  ~4  km.  We  suggest  that 
these  movements  were  associated  with  the 
search/settling  period  when  fat-deposition 
rates  may  have  been  low  or  even  negative  (Ti- 
tov 1999a,  Chernetsov  et  al.  2004b).  Normal- 
ly, positive  fat-deposition  rates  are  not 
achieved  until  the  birds  settle  and  occupy  a 
DSA  (Titov  1999a). 

Direct  visual  observations  of  radio-tagged 
European  Robins  suggested  that  their  DSAs 
were  not  defended  territories,  either  in  spring 
or  in  fall.  We  frequently  observed  “intruders” 
in  the  core  parts  of  occupied  home  ranges, 
quite  near  the  owner  and  causing  no  aggres- 
sion. In  the  vast  majority  of  cases,  Lajda 
(2001)  observed  no  aggressive  responses  to  a 
mounted  European  Robin  presented  to  DSA 
owners  during  migration.  In  our  study,  home 
ranges  of  neighbors  often  overlapped,  a pat- 
tern also  reported  by  Lajda  (2001).  Territorial 
behavior  in  birds  is  known  to  be  context-de- 
pendent (Davies  and  Houston  1983)  and 
might  or  might  not  occur,  depending  on  food 
distribution  and  availability,  density  of  com- 
petitors, or  exposure  to  predators.  Although 
we  did  not  observe  territorial  behavior  in  Eu- 
ropean Robins  during  migratory  stopovers,  we 
cannot  rule  out  that,  in  some  situations  (e.g., 
low  density  of  conspecifics),  they  might  be 
territorial  at  stopovers.  The  DSA  size  used  by 
European  Robins  during  fall  migration  stop- 
overs at  Cape  Rossitten  did  not  differ  between 


Chernetsov  and  Mukhin  • STOPOVER  BEHAVIOR  OF  EUROPEAN  ROBINS 


371 


seasons  (Table  2).  The  size  of  home  ranges 
occupied  during  fall  stopovers  on  the  Courish 
Spit  did  not  differ  from  the  values  reported 
from  the  Mettnau  peninsula  in  southwestern 
Germany  (Lajda  2001).  It  is  worth  noting, 
however,  that  fall  stopovers  at  Rybachy  (3.4 
days  ± 0.50)  were  significantly  shorter  than 
those  reported  in  southwestern  Germany  (6.7 
days  ± 1.04,  Mann- Whitney  U- test:  z = 2.79, 
P = 0.003;  Lajda  2001). 

In  our  study,  European  Robins  spent  up  to 
2 days  settling.  Two  days  seems  to  be  the 
maximum  length  of  search/settling  time,  after 
which  a robin  must  either  establish  a DSA,  or 
leave  the  area.  Our  estimate  of  search/settling 
time,  an  important  stopover  parameter  for  op- 
timal migration  models  (Weber  and  Houston 
1997a,  1997b;  Houston  1998;  Chernetsov  et 
al.  2004b),  ranges  from  several  hours  up  to  2 
days.  In  some  cases,  birds  that  seemed  to  have 
occupied  a DSA  for  several  days  would  then 
move  up  to  1 km  and  occupy  a new  DSA. 
Even  though  settling  within  2 days  is  a gen- 
eral rule  for  migrating  European  Robins,  there 
may  be  exceptions. 

We  did  not  find  a relationship  between  spa- 
tial behavior  of  European  Robins  and  their  fat 
stores  on  arrival.  The  only  difference  was  that, 
in  spring,  more  lean  birds  than  fat  birds 
stopped  for  >2  days.  Because  fat  status  of  mi- 
grants is  known  to  affect  their  foraging  be- 
havior (Loria  and  Moore  1990),  which  is 
closely  related  to  spatial  behavior,  we  had  ex- 
pected a difference  in  average  stopover  dura- 
tion. The  pattern  of  movements  at  the  stop- 
over could  have  been  quite  varied  in  either 
group.  It  is  most  likely  that  during  stopover 
the  fat  stores  of  the  birds  changed:  most  in- 
dividuals probably  refueled,  but  some  may 
have  lost  mass,  especially  during  the  initial 
phase  of  stopover,  as  observed  by  Rappole 
and  Warner  (1976),  Moore  and  Kerlinger 
(1987),  Moore  and  Yong  (1991),  and  Yong 
and  Moore  (1997).  European  Robins  that 
stopped  over  for  longer  periods  probably 
gained  mass,  but  the  low  number  of  recaptures 
after  >3-4  days  of  stopover  precluded  us 
from  estimating  fat-deposition  rates. 

The  proportion  of  birds  stopping  over  for 
>2  days  (30%  in  spring  and  39%  in  fall)  was 
much  greater  than  that  estimated  by  capture- 
mark-recapture  models  (8.4%  for  birds  first 
captured  during  a wave  of  arrivals;  Chernet- 


sov and  Titov  2000).  The  reason  for  this  dis- 
agreement is  probably  not  a delayed  departure 
due  to  the  effect  of  radio-tags  (our  study),  but 
the  fact  that  birds  that  leave  the  immediate 
vicinity  of  the  release  site — but  remain  within 
500-1,000  m — are  assumed  in  capture-mark- 
recapture  estimates  to  have  departed.  We  sug- 
gest that  capture-mark-recapture  estimates, 
and  not  the  estimates  based  on  telemetry  data, 
are  biased. 

Occupation  of  DSAs,  which  we  found  in 
the  European  Robin — or  occupation  of  terri- 
tories, as  reported  by  a number  of  authors  for 
several  other  passerine  species  (Rappole  and 
Warner  1976;  Kodric-Brown  and  Brown 
1978;  Bibby  and  Green  1980,  1981) — is  just 
one  possible  tactic  employed  by  migrants  at 
stopovers.  Other  nocturnal  passerine  migrants, 
for  example,  Blackcap  ( Sylvia  atricapilla ; 
Chernetsov  2002),  Sedge  Warbler  {Acroce- 
phalus  schoenobaenus;  Bibby  and  Green 
1981,  Chernetsov  and  Titov  2001),  and  Eur- 
asian Reed  Warbler  (A.  scirpaceus\  Chernet- 
sov and  Titov  2001),  occupy  larger  areas  than 
do  robins.  In  some  species,  authors  have  ob- 
served birds  making  broad  movements,  and  in 
others  they  have  observed  birds  occupying 
DSAs  or  even  defending  territories — e.g.,  the 
Pied  Flycatcher  ( Ficedula  hypoleuca ; Bibby 
and  Green  1980,  Chernetsov  et  al.  2004a)  and 
the  Eurasian  Reed  Warbler  (Bibby  and  Green 
1981,  Chernetsov  and  Titov  2001).  Interspe- 
cific comparisons  suggest  that  spatial  pattern 
and  territorial  behavior  of  stopover  migrants 
are  probably  related  to  the  pattern  of  food  dis- 
tribution (Chernetsov  and  Bolshakov  in  press) 
and  possibly  to  the  density  of  conspecific  and 
heterospecific  competitors.  European  Robins 
forage  mainly  on  terrestrial  invertebrates, 
which  are  relatively  evenly  distributed  across 
space  and  time  (Titov  2000,  Chernetsov  and 
Titov  2003),  and  may  occupy  a DSA,  at  least 
when  they  make  a longer  stopover.  Species 
whose  prey  are  more  unpredictable  (e.g.,  Eur- 
asian Reed  and  Sedge  warblers,  Chernetsov 
and  Titov  2001;  Pied  Flycatchers,  Chernetsov 
et  al.  2004a),  move  more  broadly. 

ACKNOWLEDGMENTS 

The  authors  are  grateful  to  D.  Leoke  for  his  help  in 
the  field  and  to  W.  Fiedler  for  logistical  assistance. 
Constructive  criticism  by  three  anonymous  reviewers 
helped  to  improve  an  earlier  draft.  This  study  was  sup- 


372 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


ported  by  the  Russian  Foundation  for  Basic  Research 

(grant  no.  02-04-48608). 

LITERATURE  CITED 

Aborn,  D.  A.  and  F.  R.  Moore.  1997.  Pattern  of 
movement  by  Summer  Tanagers  ( Piranga  rubra ) 
during  migratory  stopover:  a telemetry  study.  Be- 
havior 134:1077-1100. 

Alerstam,  T.  and  A Lindstrom.  1990.  Optimal  bird 
migration:  the  relative  importance  of  time,  energy, 
and  safety.  Pages  331-351  in  Bird  migration: 
physiology  and  ecophysiology  (E.  Gwinner,  Ed.). 
Springer,  Berlin,  Germany. 

Bibby,  C.  J.  and  R.  E.  Green.  1980.  Foraging  behavior 
of  migrant  Pied  Flycatchers,  Ficedula  hypoleuca, 
on  temporary  territories.  Journal  of  Animal  Ecol- 
ogy 49:507-521. 

Bibby,  C.  J.  and  R.  E.  Green.  1981.  Autumn  migration 
strategies  of  Reed  and  Sedge  warblers.  Ornis 
Scandinavica  12:1-12. 

Bolshakov,  C.,  V.  Bulyuk,  and  N.  Chernetsov. 
2003.  Spring  nocturnal  migration  of  Reed  War- 
blers Acrocephalus  scirpaceus:  departure,  landing 
and  body  condition.  Ibis  145:106-112. 

Bolshakov,  C.  V.,  A.  P.  Shapoval,  and  N.  P.  Zele- 
nova.  2002.  Results  of  bird  trapping  and  ringing 
by  the  Biological  Station  “Rybachy”  on  the  Cour- 
ish  Spit  in  2000.  Avian  Ecology  and  Behaviour  8: 
109-166. 

Caccamise,  D.  F.  and  R.  F.  Hedin.  1985.  An  aerody- 
namic basis  for  selecting  transmitter  loads  in 
birds.  Wilson  Bulletin  97:306-318. 

Carpenter,  F.  L.,  M.  A.  Hixon,  R.  W.  Russel,  D.  C. 
Paton,  and  E.  J.  Temeles.  1993a.  Interference 
asymmetries  among  age-classes  of  Rufous  Hum- 
mingbirds during  migratory  stopover.  Behavioral 
Ecology  and  Sociobiology  22:297-304. 

Carpenter,  F.  L.,  M.  A.  Hixon,  E.  J.  Temeles,  R.  W. 
Russel,  and  D.  C.  Paton.  1993b.  Exploitative 
compensation  by  subordinate  age-classes  of  mi- 
grant Rufous  Hummingbirds.  Behavioral  Ecology 
and  Sociobiology  22:305-312. 

Carpenter,  F.  L.,  D.  C.  Paton,  and  M.  A.  Hixon. 
1983.  Weight  gain  and  adjustment  of  feeding  ter- 
ritory size  in  migrant  hummingbirds.  Proceedings 
of  the  National  Academy  of  Sciences  USA  80: 
7259-7263. 

Chernetsov,  N.  2002.  Spatial  behavior  of  first-year 
Blackcaps  ( Sylvia  atricapilla)  during  the  pre-mi- 
gratory  period  and  during  autumn  migratory  stop- 
overs. Journal  fur  Omithologie  143:424-429. 

Chernetsov,  N.  2003.  Stopover  ecology  and  behavior 
of  migrating  passerines:  problems  and  perspec- 
tives of  research.  Ornithologia  (Moscow)  30:136- 
146.  [In  Russian] 

Chernetsov,  N.  and  C.  V.  Bolshakov.  In  press.  Spa- 
tial behavior  of  some  nocturnal  passerine  migrants 
during  stopovers.  Acta  Zoologica  Sinica. 

Chernetsov,  N.,  A.  Mukhin,  and  P.  Ktitorov.  2004a. 
Contrasting  spatial  behaviour  of  two  long-distance 


passerine  migrants  at  spring  stopovers.  Avian 
Ecology  and  Behaviour  12:53-61. 

Chernetsov,  N.  S.,  E.  A.  Skutina,  V.  N.  Bulyuk,  and 
A.  L.  Tsvey.  2004b.  Optimal  stopover  decisions 
of  migrating  birds  under  variable  stopover  quality: 
model  predictions  and  the  field  data.  Zhurnal  Ob- 
schey  Biologii  65:211-217. 

Chernetsov,  N.  and  N.  Titov.  2000.  Design  of  a trap- 
ping station  for  studying  migratory  stopovers  by 
capture-mark-recapture  analysis.  Avian  Ecology 
and  Behaviour  5:27-33. 

Chernetsov,  N.  and  N.  Titov.  2001.  Movement  pat- 
terns of  European  Reed  Warblers  Acrocephalus 
scirpaceus  and  Sedge  Warblers  A.  schoenobaenus 
before  and  during  autumn  migration.  Ardea  89: 
509-515. 

Chernetsov,  N.  S.  and  N.  V.  Titov.  2003.  Foraging 
and  spring  migratory  strategy  of  the  Robin  Eri- 
thacus  rubecula  (Aves,  Turdidae)  in  the  south- 
eastern Baltic  Sea  region.  Zoologicheskii  Zhurnal 
82:1525-1529.  [In  Russian] 

Davies,  N.  B.  and  A.  I.  Houston.  1983.  Time  allo- 
cation between  territories  and  flocks  and  owner- 
satellite  conflict  in  foraging  Pied  Wagtails.  Journal 
of  Animal  Ecology  52:621-624. 

Delingat,  J.  and  V.  Dierschke.  2000.  Habitat  utili- 
zation by  Northern  Wheatears  ( Oenanthe  oenan- 
the)  stopping  over  on  an  offshore  island  during 
migration.  Vogelwarte  40:271-278. 

Dolnik,  V.  R.  1975.  Migratory  disposition  in  birds. 
Nauka,  Moscow,  Russia.  [In  Russian] 

Hooge,  P.  N.  and  B.  Eichenlaub.  2000.  Animal  move- 
ment extension  to  Arcview,  ver.  2.0.  Alaska  Sci- 
ence Center,  U.S.  Geological  Survey,  Anchorage, 
Alaska. 

Houston,  A.  I.  1998.  Models  of  optimal  avian  migra- 
tion: state,  time  and  predation.  Journal  of  Avian 
Biology  29:395-404. 

Kaiser,  A.  1993.  A new  multi-category  classification 
of  subcutaneous  fat  deposits  in  song  birds.  Journal 
of  Field  Ornithology  64:246-255. 

Kodric-Brown,  A.  and  J.  H.  Brown.  1978.  Influence 
of  economics,  interspecific  competition,  and  sex- 
ual dimorphism  on  territoriality  in  migrant  Rufous 
Hummingbirds.  Ecology  49:285-296. 

Lajda,  M.  2001.  Telemetrische  Untersuchung  zum 
Rastverhalten  des  Rotkehlchens  ( Erithacus  rube- 
cula) in  Siidwestdeutschland  wahrend  des  Herb- 
stzuges.  Diploma  thesis.  University  of  Zurich,  Zu- 
rich, Switzerland.  [In  German] 

Liechti,  F.  and  B.  Bruderer.  1998.  The  relevance  of 
wind  for  optimal  migration  theory.  Journal  of  Avi- 
an Biology  29:561-568. 

Lindstrom,  A.  2003.  Fuel  deposition  rates  in  migrating 
birds:  causes,  constraints  and  consequences.  Pages 
307-320  in  Avian  migration  (P.  Berthold,  E. 
Gwinner,  and  E.  Sonnenschein,  Eds.).  Springer, 
Berlin,  Germany. 

Loria,  D.  E.  and  F.  R.  Moore.  1990.  Energy  demands 
of  migration  in  Red-eyed  Vireos,  Vireo  olivaceus. 
Behavioral  Ecology  1:24-35. 


Chernetsov  and  Mukhin  • STOPOVER  BEHAVIOR  OF  EUROPEAN  ROBINS 


373 


Moore,  F.  R.  and  P.  Kerlinger.  1987.  Stopover  and 
fat  deposition  by  North  American  wood-warblers 
(Parulinae)  following  spring  migration  over  the 
Gulf  of  Mexico.  Oecologia  74:47-54. 

Moore,  F.  R.  and  W.  Yong.  1991.  Evidence  of  food- 
based  competition  among  passerine  migrants  dur- 
ing stopover.  Behavioral  Ecology  and  Sociobiol- 
ogy 28:85-90. 

Naef-Daenzer,  B.  1993.  A new  transmitter  for  small 
animals  and  enhanced  methods  of  home  range 
analysis.  Journal  of  Wildlife  Management  57: 
680-689. 

Petit,  D.  R.  2000.  Habitat  use  of  landbirds  along  Ne- 
arctic-Neotropical  migration  routes:  implications 
for  conservation  of  stopover  habitats.  Studies  in 
Avian  Biology  20:15-33. 

Pradel,  R.  1996.  Utilization  of  capture-mark-recapture 
for  the  study  of  recruitment  and  population  growth 
rate.  Biometrics  52:703-709. 

Rappole,  J.  H.  and  A.  R.  Tipton.  1991.  New  harness 
design  for  attachment  of  radio-transmitters  to 
small  passerines.  Journal  of  Field  Ornithology  62: 
335-337. 

Rappole,  J.  H.  and  D.  W.  Warner.  1976.  Relation- 
ships between  behavior,  physiology  and  weather 
in  avian  transients  at  a migration  stopover  site. 
Oecologia  26:193-212. 

Rguibi-Idrissi,  H.,  R.  Julliard,  and  F.  Bairlein.  2003. 
Variation  in  the  stopover  duration  of  Reed  War- 
blers Acrocephalus  scirpaceus  in  Morocco:  effects 
of  season,  age  and  site.  Ibis  145:650-656. 

SPSS,  Inc.  1999.  SPSS  Base  10.0  user’s  guide.  SPSS, 
Inc.,  Chicago,  Illinois. 

Swihart,  R.  E.  and  N.  A.  Slade.  1985.  Testing  for 


independence  of  observations  in  animal  move- 
ments. Ecology  66:1 176-1 184. 

Szulc-Olech,  B.  1965.  The  resting  period  of  migrant 
robins  on  autumn  passage.  Bird  Study  12:1-7. 

Titov,  N.  1999a.  Home  ranges  in  two  passerine  noc- 
turnal migrants  at  a stopover  site  in  autumn.  Avian 
Ecology  and  Behaviour  3:69-78. 

Titov,  N.  1999b.  Individual  home  ranges  of  Robins 
Erithacus  rubecula  at  stopovers  during  autumn 
migration.  Vogelwelt  120:237-242. 

Titov,  N.  2000.  Interaction  between  foraging  strategy 
and  autumn  migratory  strategy  in  the  Robin  Eri- 
thacus rubecula.  Avian  Ecology  and  Behaviour  5: 
35-44. 

Titov,  N.  V.  and  N.  S.  Chernetsov.  1999.  Stochastic 
models  as  a new  method  for  estimating  length  of 
migratory  stopovers  in  birds.  Uspekhi  Sovremen- 
noi  Biologii  119:396-403.  [In  Russian] 

Weber,  T.  P.  and  A.  I.  Houston.  1997a.  Flight  costs, 
flight  range  and  the  stopover  ecology  of  migrating 
birds.  Journal  of  Animal  Ecology  66:297-306. 

Weber,  T.  P.  and  A.  I.  Houston.  1997b.  A general 
model  for  time-minimising  avian  migration.  Jour- 
nal of  Theoretical  Biology  185:447-458. 

White,  G.  C.  and  R.  A.  Garrott.  1990.  Analysis  of 
wildlife  tracking  data.  Academic  Press,  London, 
United  Kingdom. 

Williamson,  P.  and  L.  Gray.  1975.  Foraging  behavior 
of  the  Starling  ( Sturnus  vulgaris ) in  Maryland. 
Condor  77:84-89. 

Yong,  W.  and  F.  R.  Moore.  1997.  Spring  stopover  of 
intercontinental  migratory  thrushes  along  the 
northern  coast  of  the  Gulf  of  Mexico.  Auk  114: 
263-278. 


The  Wilson  Journal  of  Ornithology  1 18(3):374-379,  2006 


AGE-RELATED  TIMING  AND  PATTERNS  OF  PREBASIC  BODY 
MOLT  IN  WOOD  WARBLERS  (PARULIDAE) 

CHRISTINE  A.  DEBRUYNE,1  JANICE  M.  HUGHES,13  AND 
DAVID  J.  T.  HUSSELL2 3 


ABSTRACT. — We  compared  timing  and  patterns  of  prebasic  body  molt  between  hatch-year  (HY)  and  after- 
hatch-year (AHY)  American  Redstarts  ( Setophaga  ruticilla)  and  Yellow  Warblers  ( Dendroica  petechia)  in  On- 
tario, Canada.  In  each  body  region  of  both  species,  there  was  no  age-related  difference  in  the  proportion  of 
individuals  undergoing  molt.  Furthermore,  there  was  no  difference  between  HY  and  AHY  American  Redstarts 
in  the  overall  timing  of  body  molt;  molt  started  in  early  July  and  lasted  until  early  September.  In  contrast,  HY 
Yellow  Warblers  started  body  molt  in  late  June  to  early  July,  while  adults  began  body  molt  in  mid-July.  Both 
American  Redstarts  and  Yellow  Warblers  displayed  age-class  differences  in  the  intensity  and  timing  of  molt 
among  specific  body  regions.  External  factors  (e.g.,  food  availability  and  geographical  distribution),  and  internal 
factors  (e.g.,  physiological  status)  may  contribute  to  variations  in  body  molt  timing  observed  in  these  two  species. 
Received  2 December  2004,  accepted  13  March  2006. 


Molt  plays  an  important  role  in  the  life  cy- 
cle of  birds  because  feathers  have  multiple 
functions,  such  as  display  during  courtship 
(e.g.,  Beehler  1983),  thermoregulation 
(Schieltz  and  Murphy  1997),  and  protection 
from  dermal  parasites  (Post  and  Enders  1970). 
Most  importantly,  birds  must  replace  their 
feathers  before  progressive  wear  impedes 
flight  (Ginn  and  Melville  1983).  However, 
molting  consumes  large  amounts  of  energy 
and  protein  reserves  to  produce  new  feathers 
and  to  compensate  for  the  effects  of  reduced 
insulation  and  decreased  flight  efficiency 
(Dolnik  and  Gavrilov  1979;  Murphy  and  King 
1991,  1992).  To  minimize  energetic  constraints 
and  avoid  undue  overlap  with  other  energeti- 
cally demanding  activities,  such  as  reproduc- 
tion and  migration,  many  birds  molt  during 
times  when  food  is  abundant  (Payne  1972). 

Typically,  adult  (after-hatch-year  or  AHY) 
wood  warblers  attain  basic  plumage  by  un- 
dergoing a complete  prebasic  molt — which  re- 
places nearly  all  feathers — while  still  on  the 
breeding  grounds  prior  to  migration.  Hatch- 
year  (HY)  wood  warblers  with  juvenal  plum- 
age body  feathers — which  are  weaker  and 
looser  in  texture — attain  their  winter  plumage 
through  a first  prebasic  molt,  replacing  only 


1 Dept,  of  Biology,  Lakehead  Univ.,  955  Oliver  Rd., 
Thunder  Bay,  ON  P7B  5E1,  Canada. 

2 Wildlife  Research  and  Development  Section,  On- 
tario Ministry  of  Natural  Resources,  300  Water  St., 
Peterborough,  ON  K9J  8M5,  Canada. 

3 Corresponding  author;  e-mail: 
janice.hughes@lakeheadu.ca 


body  contour  feathers  and  most  of  the  wing 
coverts  (Pyle  1997). 

After  breeding,  most  warblers  prepare  for 
the  flight  to  their  wintering  grounds  by  in- 
creasing their  nutritional  intake  and  molting 
prior  to  migration.  We  compared  the  body 
molt  patterns  and  timing  of  HY  versus  AHY 
Yellow  Warblers  {Dendroica  petechia ) and 
American  Redstarts  {Setophaga  ruticilla)  to 
determine  whether  any  age-related  differences 
in  chronology  and  rate  of  molt  could  be  attri- 
buted to  constraints  inherent  to  the  breeding 
cycle.  HY  warblers  do  not  molt  as  extensively 
as  AHYs;  hence,  their  preparations  for  migra- 
tion, including  molt,  may  be  limited  by  the 
timing  of  fledging.  Thus,  we  would  expect 
AHYs — constrained  by  both  nesting  respon- 
sibilities and  the  timing  of  migration — to  be- 
gin molting  later  than  HYs  but,  once  initiated, 
to  undergo  a more  rapid  body  molt. 

METHODS 

Study  areas. — Yellow  Warbler  and  Ameri- 
can Redstart  molt  data  were  obtained  at  Innis 
Point  Bird  Observatory  (IPBO)  and  Thunder 
Cape  Bird  Observatory  (TCBO),  respectively. 
IPBO  is  located  approximately  12  km  west  of 
Ottawa,  Ontario  (45°  22' N,  75°53'W)  near 
Shirley’s  Bay  on  Department  of  National  De- 
fense property  along  the  southwestern  bank  of 
the  Ottawa  River.  The  surrounding  habitat  in- 
cludes deciduous  forest  and  regenerating  farm 
fields  dotted  with  small  trees  and  shrubs. 
TCBO  is  situated  at  the  tip  of  the  Sibley  Pen- 
insula, on  the  northwest  shore  of  Lake  Supe- 


374 


Debruyne  et  al.  • BODY  MOLT  IN  WOOD  WARBLERS 


375 


rior,  approximately  80  km  from  Thunder  Bay, 
Ontario  (48°  18' N,  88°  56' W).  The  area  is 
predominantly  forested,  consisting  mostly  of 
coniferous  trees  and  shrubs. 

Field  procedures. — From  6 July  to  10  Sep- 
tember 1998-2002,  we  captured  113  Ameri- 
can Redstarts  (85  HYs  and  28  AHYs)  and  68 
Yellow  Warblers  (43  HYs  and  25  AHYs)  us- 
ing mist  nests  (30-mm  mesh  size)  and  Heli- 
goland traps  according  to  TCBO  and  IPBO 
standard  protocols.  Ninety-four  American 
Redstarts  (71  HYs  and  23  AHYs)  and  50  Yel- 
low Warblers  (27  HYs  and  23  AHYs)  were 
actively  molting  when  captured.  We  obtained 
body  molt  data  for  five  body  regions  (head, 
back,  belly,  uppertail  coverts,  and  undertail 
coverts).  To  satisfy  sample  size  and  distribu- 
tion requirements  of  log-linear  models  (Sokal 
and  Rohlf  1995,  Yuri  and  Rohwer  1997),  each 
body  region  was  scored  on  an  ordinal  scale  of 
0 to  5 based  on  the  estimated  proportion  of 
actively  molting  feathers  (molt  score  of  0 = 
no  molt;  1 = 0-20%  complete;  2 = 21-40%; 
3 = 41-60%;  4 = 61-80%;  and  5 = 81- 
100%).  A total  body  molt  score  for  each  in- 
dividual was  determined  by  summing  the  in- 
dividual molt  scores  for  all  five  body  regions; 
thus,  total  body  molt  scores  ranged  from  0 to 
25.  To  obtain  a representative  sample,  body 
molt  was  scored  on  all  birds  captured,  whether 
they  were  molting  or  not. 

AHY  warblers  were  differentiated  from  HY 
warblers  on  the  basis  of  plumage  and  bill  col- 
or, and  extent  of  skull  pneumatization.  HY 
Yellow  Warblers  are  typically  duller  in  col- 
oration than  AHYs  in  definitive  basic  plum- 
age, and  they  have  tapered  outer  primary  co- 
verts with  narrow  or  indistinct  buffy  edging 
(Pyle  1997).  Also,  AHYs  have  dark  lower 
mandibles  (Mundy  and  McCracken  1997).  Fe- 
male AHY  American  Redstarts  were  distin- 
guished from  HYs  of  both  sexes  by  their  trun- 
cate, dusky  brown  outer  primary  coverts  (not 
tipped  with  buff)  and  the  large  yellow  patch 
on  their  rectrices.  In  addition,  the  AHY’s  outer 
rectrices  of  both  species  have  truncated  inner 
webs.  AHYs  were  also  identified  by  their  fully 
pneumatized  skulls;  skulls  of  HYs  were  in- 
completely pneumatized  (Pyle  1997). 

Statistical  analyses. — We  categorized  cap- 
ture dates  for  American  Redstarts  into  three 
consecutive,  17-day  blocks  (22  July  to  7 Au- 
gust: n = 21  HYs  and  12  AHYs;  8 to  24  Au- 


gust: n = 45  HYs  and  9 AHYs;  25  August  to 
10  September:  n = 5 HYs  and  2 AHYs).  To 
satisfy  sample  size  and  distribution  require- 
ments of  log-linear  models  (Sokal  and  Rohlf 
1995,  Yuri  and  Rohwer  1997),  molt  scores  of 
0 to  1 were  combined.  Capture  dates  for  Yel- 
low Warblers  were  divided  into  three  consec- 
utive, 16-day  blocks  to  provide  a feasible  dis- 
tribution of  captures  (6  to  21  July:  n = 13 
HYs  and  3 AHYs;  22  July  to  6 August:  n = 
8 HYs  and  9 AHYs;  7 to  22  August:  n = 6 
HYs  and  11  AHYs).  Due  to  an  unequal  dis- 
tribution of  molt  scores  among  body  regions, 
molt  scores  were  grouped  into  only  three  clas- 
ses (0-3,  4,  and  5). 

To  determine  peak  molt  interval  and  the 
progression  and  rate  of  molt,  we  first  ran  log- 
linear  models  with  a G-test  using  Williams’ 
correction  (Sokal  and  Rohlf  1995)  to  deter- 
mine whether  overall  body  molt  scores  (i.e., 
five  body  regions;  for  HY  and  AHY  warblers) 
were  independent  of  capture  date  (i.e.,  three 
consecutive  17-day  blocks  for  American  Red- 
starts, 16-day  blocks  for  Yellow  Warblers; 
Yuri  and  Rohwer  1997).  We  then  used  one- 
way analyses  of  covariance  (ANCOVA;  Sokal 
and  Rohlf  1995) — using  total  body  molt  score 
as  the  dependent  variable,  age  as  the  indepen- 
dent variable,  and  date  captured  as  the  covar- 
iate— to  test  for  age  class  differences  in  the 
timing  of  body  molt  (all  body  regions  com- 
bined). We  used  SPSS  (Release  10.07a)  for 
Macintosh  (SPSS,  Inc.  2000),  and  set  statis- 
tical significance  at  P < 0.05. 

RESULTS 

Yellow  Warblers. — Body  molt  of  all  regions 
occurred  from  early  July  to  mid- August.  With- 
in this  period,  molt  progressed  uniformly  with 
no  peak  interval,  which  would  have  been  ex- 
pressed as  a greater  proportion  of  individuals 
undergoing  molt.  For  example,  whether  age 
classes  were  pooled  or  analyzed  separately, 
there  was  no  difference  in  the  proportion  of 
molting  individuals  with  respect  to  date  (G- 
tests:  P > 0.99  in  all  cases). 

The  timing  of  body  molt  depended  on  age; 
HYs  began  body  molt  earlier  than  AHY  in- 
dividuals (F1;49  = 11.23,  P = 0.002,  n = 50; 
Fig.  1).  Molt  scores  across  body  regions  dif- 
fered between  age  classes  (Gadj  = 16.49,  df  = 
8,  P < 0.05);  the  greatest  differences  were 
observed  in  the  crown  (HY  mean  molt  score 


376 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


g>  20  ■ 
o 
8 


20  30  40 

Capture  date  (1  = 1 July) 


oo  ▲ 
o o 

Qfc 


20  30  40  50  60 

Capture  date  (1  = 1 July) 


FIG.  1.  Relationship  between  total  body  molt 
score  and  capture  date  (6  July  to  22  August)  for  hatch- 
year  (HY;  triangles)  and  after-hatch-year  (AHY;  cir- 
cles) Yellow  Warblers.  HY  birds  typically  began  molt 
earlier  than  AHY  individuals  ( Fl49  = 11.23,  P = 
0.002). 


FIG.  2.  Relationship  between  total  body  molt 
score  and  capture  date  (22  July  to  10  September)  for 
hatch-year  (HY;  triangles)  and  after-hatch-year  (AHY; 
circles)  American  Redstarts.  The  timing  of  molt  did 
not  differ  between  HY  and  AHY  birds  (F193  = 1.34, 
P = 0.25). 


= 4.4;  AHY  mean  molt  score  = 3.5)  and  back 
(HY  mean  molt  score  = 4.0;  AHY  mean  molt 
score  = 3.4)  regions.  The  progression  of  body 
molt  for  HY  individuals  was  crown,  back,  un- 
dertail coverts,  uppertail  coverts,  and  belly; 
for  AHY  birds,  the  sequence  was  undertail  co- 
verts, uppertail  coverts,  crown,  back,  and  bel- 
ly. In  both  age  classes,  undertail  and  uppertail 
covert  molt  occurred  almost  simultaneously. 

With  respect  to  timing,  molt  scores  differed 
between  age  classes  (Gadj  — 17.74,  df  = 4,  P 
< 0.005).  The  greatest  difference  occurred 
from  6 to  21  July,  during  which  the  estimated 
mean  molt  score  (i.e.,  mean  value  of  the  molt 
score  for  all  five  body  regions)  was  3.9  for 
HYs  and  3.0  for  AHYs,  indicating  that  molt 
begins  earlier  among  HYs  than  among  AHYs 
during  that  date  block.  From  22  July  to  6 Au- 
gust, HY  body  molt  decreased  slightly  (mean 
molt  score  3.8),  but  in  AHYs  it  increased 
(mean  AHY  molt  scores  in  time  blocks  1,  2, 
and  3 were  3.0,  3.7,  and  3.8,  respectively). 

From  6 to  21  July,  the  percentage  of  indi- 
viduals that  had  not  started  molting  (molt 
score  0)  was  33%  for  HYs  and  35%  for 
AHYs.  By  22  July,  however,  all  individuals 
had  initiated  molt.  By  date  block,  the  per- 
centage of  individuals  that  had  completed 
their  molt  (molt  score  25)  was  0%  for  both 
age  classes  (6  to  21  July),  47%  for  HYs  and 
0%  for  AHYs  (22  July  to  6 August),  and  25% 
for  HYs  and  0%  for  AHYs  (7  to  22  August). 
All  AHYs  were  in  active  body  molt  from  22 
July  to  22  August;  however,  AHYs  captured 


from  7 to  22  August  had  total  molt  scores  of 
23  or  24,  indicating  that  their  molt  was  almost 
completed  by  then. 

American  Redstarts. — Body  molt  in  all  re- 
gions occurred  from  mid-July  to  early  Sep- 
tember. Whether  age  classes  were  pooled  or 
analyzed  separately,  there  was  no  age-class 
difference  in  the  proportion  of  individuals  un- 
dergoing body  molt  (G-tests:  P > 0.50  in  all 
cases). 

Analyses  of  the  effect  of  age  class — with 
total  body  molt  score  as  the  dependent  vari- 
able and  capture  date  as  a covariate — indicat- 
ed no  difference  in  timing  of  molt  within  any 
date  block  (F193  = 1.34,  P — 0.25,  n — 94; 
Fig.  2).  Although  body  molts  in  HYs  and 
AHYs  were  concurrent,  molt  scores  across 
body  regions  differed  between  age  classes 
(Gadj  = 79.17,  df  = 16,  P < 0.001):  HY  molt 
was  more  advanced  than  that  of  AHYs  in  all 
three  date  blocks.  The  greatest  difference  in 
molt  scores  between  age  classes  was  in  the 
undertail  covert  region  (mean  HY  molt  score 
= 3.6;  mean  AHY  molt  score  = 3.1).  The 
progression  of  body  molt  for  HYs  was  back, 
undertail  coverts,  uppertail  coverts,  belly,  and 
crown;  for  AHY  birds  it  was  back,  uppertail 
coverts,  undertail  coverts,  belly,  and  crown.  In 
both  age  classes,  undertail  coverts  and  belly 
molts  occurred  almost  simultaneously. 

With  respect  to  timing,  American  Redstarts 
displayed  age-related  differences  in  molt 
scores  (Gadj  = 42.14,  df  = 8,  P < 0.001). 
From  25  August  to  10  September,  there  was 


Debruyne  et  al  • BODY  MOLT  IN  WOOD  WARBLERS 


377 


a large  age-related  difference  in  molt  scores; 
the  estimated  mean  molt  score  (i.e.,  mean  val- 
ue of  the  molt  score  for  all  five  body  regions) 
was  3.2  for  HYs  and  4.6  for  AHYs,  indicating 
that  AHYs  initiate  molt  earlier  than  HYs  dur- 
ing that  date  block.  In  addition,  body  molt  of 
HYs  was  most  intense  in  early  August  (mean 
molt  score  3.6);  however,  in  AHYs  it  in- 
creased linearly  with  time  (mean  molt  scores 
in  time  blocks  1,  2,  and  3 were  2.5,  3.2,  and 
4.6,  respectively). 

Within  the  three  date  blocks,  the  percentage 
of  individuals  that  had  not  initiated  molt  (molt 
score  0)  was  0%  for  HYs  and  20%  for  AHYs 
(22  July  to  7 August),  2%  for  HYs  and  0% 
for  AHYs  (8  to  24  August),  and  0%  in  both 
age  classes  (25  August  to  10  September).  The 
percentage  of  individuals  that  had  completed 
molt  (molt  score  25)  was  4%  for  HYs  and  0% 
for  AHYs  (22  July  to  7 August),  16%  for  HYs 
and  0%  for  AHYs  (8  to  24  August),  and  38% 
for  HYs  and  50%  for  AHYs  (25  August  to  10 
September).  All  AHYs  were  actively  molting 
from  8 to  24  August. 

DISCUSSION 

Ginn  and  Melville  (1983)  emphasized  the 
need  to  examine  body  molt  because  body 
feathers  account  for  more  than  half  of  a bird’s 
feather  mass.  Consequently,  their  replacement 
may  lead  to  greater  overall  energetic  require- 
ments than  the  molt  of  flight  feathers.  Molt 
must  be  timed  to  minimize  energetic  losses 
while  progressing  adequately  enough  to  pre- 
pare for  fall  migration;  thus,  a bird’s  annual 
cycle  must  be  structured  to  optimize  repro- 
ductive, migratory,  and  molt  requirements. 
Factors  such  as  arrival  on  the  breeding 
grounds  will  set  the  timeline  that  AHY  war- 
blers require  to  fulfill  all  the  tasks  associated 
with  breeding.  On  the  other  hand,  the  molt 
timeline  for  HY  birds  is  probably  established 
by  hatch  dates,  with  factors  such  as  nutritional 
provisioning  by  adults  determining  the  opti- 
mal physiological  conditions  for  molt.  Fur- 
thermore, to  maximize  flight  efficiency,  both 
age  classes  must  complete  adequate  feather 
replacement  prior  to  departure  for  the  winter- 
ing grounds. 

Molt  in  relation  to  breeding. — One  may  as- 
sume that  HY  warblers  would  be  more  likely 
to  initiate  molt  earlier  than  AHY  birds  be- 
cause they  do  not  expend  time  or  energy  pro- 


ducing offspring.  In  addition,  the  first  prebasic 
molt  of  wood  warblers  does  not  include  most 
of  the  flight  feathers  (Pyle  1997);  hence,  phys- 
iological demands  of  feather  replacement  in 
HY  birds  should  be  considerably  less  than  that 
of  AHY  individuals.  As  predicted,  our  study 
demonstrates  that  HY  Yellow  Warblers  do  ini- 
tiate molt  earlier  than  AHYs.  Body  molt  be- 
gan in  late  June  to  early  July  for  HYs  and  mid 
to  late  July  in  AHYs,  with  greatest  age-related 
differences  in  molt  scores  occurring  in  the  6 
to  21  July  date  block.  In  Ontario,  records  of 
active  Yellow  Warbler  nests  peak  during  the 
first  2 weeks  of  June  (Peck  and  James  1987), 
suggesting  that  HY  birds  may  begin  prebasic 
body  molt  while  still  in  the  nest.  Lowther  et 
al.  (1999)  also  indicated  that  prebasic  molt  in 
Yellow  Warblers  often  begins  before  fledging. 
Peak  fledging  of  Yellow  Warblers  in  Ontario 
occurs  in  late  June  (Peck  and  James  1987). 
Our  early  captures  demonstrated  that  body 
molt  in  most  HY  Yellow  Warblers  was  well 
underway  during  the  first  week  of  July;  67% 
of  individuals  were  in  active  molt  and  had  a 
mean  molt  score  of  3.9. 

Differences  in  body  molt  schedules  in  Yel- 
low Warblers  relative  to  ongoing  energetic  ex- 
penditures other  than  molt  also  may  explain 
differences  observed  in  molt  intensity  over 
time.  For  example,  in  early  to  mid-July,  HY 
birds  had  considerably  higher  molt  scores  than 
AHYs  for  all  body  regions  combined.  Molt  in 
AHY  Yellow  Warblers  overlapped  with  breed- 
ing; consequently,  they  may  be  compensating 
with  a less  intensive  body  molt  early  in  the 
molting  period.  Nolan  (1978)  suggested  that 
Prairie  Warblers  ( Dendroica  discolor ) with 
dependent  young  underwent  slower  molt  than 
birds  that  were  not  tending  to  offspring.  Our 
results  showed  increased  molt  intensity  in 
AHYs  in  mid-to  late  July  when  young  are  less 
dependent  on  their  parents  (Lowther  et  al. 
1999).  At  James  Bay  in  northern  Ontario, 
Rimmer  (1988)  concluded  that  molt  among 
Yellow  Warblers  typically  overlaps  fledgling 
care  because  the  young  are  relatively  indepen- 
dent at  that  time,  thereby  reducing  parental 
demands. 

We  found  that  body  molt  for  both  age  clas- 
ses of  American  Redstarts  occurred  concur- 
rently in  all  body  regions  from  mid- July  to 
early  September.  Other  warbler  species,  in- 
cluding Hermit  ( Dendroica  occidentalis ) and 


378 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Townsend’s  ( D . toxvnsendi ) warblers,  also  dis- 
play a lack  of  age-related  differences  in  the 
timing  of  body  molt  (Jackson  et  al.  1992). 
Similar  to  that  of  Yellow  Warblers,  body  molt 
in  AHY  American  Redstarts  overlapped  with 
breeding.  In  Ontario,  records  imply  that  peak 
fledging  of  American  Redstarts  occurs  during 
mid-to  late  June  (Peck  and  James  1987); 
therefore,  young  would  continue  to  be  depen- 
dent on  parents  through  July  (Sherry  and 
Holmes  1997).  The  parallel  timing  of  body 
molt  between  HY  and  AHY  American  Red- 
starts could  explain  the  similarities  in  their 
molt  intensity  during  the  first  month  of  the 
molt  period,  in  which  case  the  adult  birds 
must  have  sufficient  energetic  reserves  to 
complete  their  parental  duties  when  initiating 
molt.  However,  the  considerable  age-related 
difference  in  molt  intensity  from  25  August  to 
10  September  might  reflect  the  termination  of 
breeding  duties,  allowing  for  more  energy  to 
be  allocated  to  the  molting  process. 

Molt  in  relation  to  migration. — Most  mi- 
gratory birds  complete  a substantial  portion  of 
their  prebasic  molt  before  leaving  the  breed- 
ing grounds;  some  warbler  species  delay  their 
departures  for  several  days  until  their  feathers 
are  adequately  grown  (Rimmer  1988).  It  has 
been  shown  that  body  molt  and  primary  feath- 
er molt  can  occur  simultaneously  (Sherry  and 
Holmes  1997,  Lowther  et  al.  1999).  Further- 
more, the  rate  and  chronology  of  primary  molt 
in  many  species  of  wood  warblers  are  typi- 
cally correlated  with  time  of  southbound  mi- 
gration, such  that  earlier  departure  from  the 
breeding  grounds  is  associated  with  a shorter 
and  more  rapid  molt  (Debruyne  2003). 

Yellow  Warblers  are  among  the  earliest  of 
wood  warblers  to  begin  their  southbound  mi- 
gration, with  most  departing  from  their  breed- 
ing grounds  in  eastern  Canada  by  early  Au- 
gust (Lowther  et  al.  1999).  Peak  migration  of 
HY  individuals  may  occur  1-2  weeks  earlier 
(Rimmer  1988).  This  is  consistent  with  our 
study,  which  demonstrates  that  body  molt  oc- 
curs earlier  in  Yellow  Warblers  than  it  does  in 
American  Redstarts,  which  begin  migration  in 
late  August  to  early  September  (Sherry  and 
Holmes  1997),  and  that  molt  in  HY  Yellow 
Warblers  occurs  earlier  than  it  does  in  AHY 
individuals.  HY  birds  would  be  able  to  molt 
and  migrate  earlier  than  AHYs  because  they 
do  not  have  the  energetic  demands  of  raising 


young  and  their  first  prebasic  molt  does  not 
include  most  of  the  flight  feathers.  Further- 
more, both  HY  and  AHY  Yellow  Warblers 
may  begin  migrating  while  undergoing  the  fi- 
nal stages  of  body  molt.  Rimmer  (1988)  noted 
that  AHY  Yellow  Warblers  in  northern  Ontar- 
io begin  migrating  during  the  final  stages  of 
molt  (i.e.,  final  stages  of  growth  of  the  last 
two  primaries);  he  suggested  that  the  energetic 
costs  associated  with  this  stage  of  molt  were 
not  significant  enough  to  preclude  simulta- 
neous migration.  Rimmer  also  found  that  Yel- 
low Warblers  lost  body  weight  during  the  later 
stages  of  molt  because  individuals  departed 
without  the  typical  premigration  accumulation 
of  fat.  He  concluded  that  migration  timing 
may  be  regulated  by  flight  efficiency  rather 
than  physiological  readiness.  This  relief  from 
the  constraint  of  premigratory  preparedness 
would  favor  an  early  departure  from  the 
breeding  grounds,  particularly  if  suitable  food 
resources  are  exhausted. 

The  timing  of  body  molt  for  HY  and  AHY 
American  Redstarts  at  TCBO  is  consistent 
with  the  timing  of  southbound  migration  in 
late  August;  many  individuals  had  completed 
body  molt  by  this  time.  Additionally,  both  age 
classes  arrive  synchronously  at  the  banding 
stations  of  LPBO  and  the  Allegheny  Front  Mi- 
gration Observatory  in  West  Virginia  (Hall 
1981,  Woodrey  and  Chandler  1997).  This  sup- 
ports the  lack  of  age-related  differences  in  the 
timing  of  prebasic  molt  among  American 
Redstarts.  Jackson  et  al.  (1992)  observed  that 
male  Hermit  and  Townsend’s  warblers  com- 
plete most  of  their  prebasic  molt  on  their 
breeding  grounds  prior  to  migration,  and  sug- 
gested that  their  breeding  areas — moist  mon- 
tane and  lowland  habitats,  respectively — still 
provided  sufficient  food  resources  after  breed- 
ing to  allow  birds  to  molt  before  departure. 
American  Redstarts  also  prefer  moist,  produc- 
tive habitats  that  offer  abundant  food  resourc- 
es in  late  summer  (Sherry  and  Holmes  1997), 
perhaps  explaining  similarities  in  the  timing 
of  molt  among  these  species.  In  addition, 
American  Redstarts  demonstrate  substantial 
flexibility  in  both  dietary  choices  and  foraging 
strategies,  which  would  allow  both  HY  and 
AHY  individuals  to  linger  on  the  breeding 
grounds  during  molt. 

This  study  provides  a foundation  for  future 
research  on  body  molt  in  two  wood  warbler 


Debruyne  et  al.  • BODY  MOLT  IN  WOOD  WARBLERS 


379 


species  found  throughout  eastern  North  Amer- 
ica. External  factors,  including  food  availabil- 
ity, and  internal  factors,  such  as  physiological 
readiness  to  molt  and  migrate,  may  provide 
some  explanation  for  the  timing  of  body  molt. 
Continued  examination  of  the  many  biological 
and  environmental  aspects  affecting  molt  and 
migration  will  contribute  to  a better  under- 
standing of  body  molt  patterns  in  wood  war- 
blers. 

ACKNOWLEDGMENTS 

We  thank  the  staff  and  volunteers  at  Thunder  Cape 
and  Innis  Point  bird  observatories  for  collecting  molt 
data  over  many  years;  this  study  would  not  have  been 
possible  without  their  long  hours  of  hard  work.  The 
senior  author  is  particularly  grateful  to  J.  Allair,  J. 
Woodcock,  M.  Woodcock,  K.  Burrell,  P.  Biedermann, 
C.  Friis,  and  A.  Blake  for  their  assistance  with  her  field 
work  at  Thunder  Cape  in  2002.  The  Thunder  Cape 
Bird  Observatory  is  a joint  project  of  the  Thunder  Bay 
Field  Naturalists,  Bird  Studies  Canada,  and  the  Ontario 
Ministry  of  Natural  Resources  (OMNR),  with  major 
funding  from  OMNR’s  Wildlife  Assessment  Program. 
We  would  also  like  to  thank  C.  C.  Rimmer  and  two 
anonymous  reviewers  for  their  helpful  suggestions. 
This  research  was  supported,  in  part,  by  a Natural  Sci- 
ences and  Engineering  Research  Council  grant  to 
JMH.  This  paper  is  a contribution  of  the  Ontario  Min- 
istry of  Natural  Resources,  Wildlife  Research  and  De- 
velopment Section. 

LITERATURE  CITED 

Beehler,  B.  1983.  Lek  behavior  of  the  Lesser  Bird  of 
Paradise.  Auk  100:992-995. 

Debruyne,  C.  A.  2003.  Pattern  and  chronology  of  pre- 
basic  moult  in  wood-warblers  (Parulidae).  M.Sc. 
thesis,  Lakehead  University,  Thunder  Bay,  Ontar- 
io. 

Dolnik,  V.  R.  and  V.  M.  Gavrilov.  1979.  Bioener- 
getics of  molt  in  the  Chaffinch  ( Fringilla  coelebs). 
Auk  96:253-264. 

Ginn,  H.  B.  and  D.  S.  Melville.  1983.  Moult  in  birds. 
British  Trust  for  Ornithology  Guide,  no.  19. 
Maund  & Irvine,  Tring,  United  Kingdom. 

Hall,  G.  A.  1981.  Fall  migration  patterns  of  wood 
warblers  in  the  southern  Appalachians.  Journal  of 
Field  Ornithology  52:43-49. 

Jackson,  W.  M.,  C.  S.  Wood,  and  S.  Rohwer.  1992. 
Age-specific  plumage  characters  and  annual  molt 


schedules  of  Hermit  Warblers  and  Townsend’s 
Warblers.  Condor  90:490-501. 

Lowther,  P.  E.,  C.  Celada,  N.  K.  Klein,  C.  C.  Rim- 
mer, and  D.  A.  Spector.  1999.  Yellow  Warbler 
(Dendroica  petechia ).  The  Birds  of  North  Amer- 
ica, no.  454. 

Mundy,  R.  P.  and  J.  D.  McCracken.  1997.  Bill  color 
as  an  age  character  in  Yellow  Warblers.  North 
American  Bird  Bander  22:116-118. 

Murphy,  M.  E.  and  J.  R.  King.  1991.  Nutritional  as- 
pects of  avian  molt.  Acta  Congressus  Internation- 
alis  Ornithologici  20:2186-2194. 

Murphy,  M.  E.  and  J.  R.  King.  1992.  Energy  and  nu- 
trient use  during  molt  by  White-crowned  Spar- 
rows Zonotrichia  leucophrys  gambelii.  Ornis 
Scandanavica  23:304-313. 

Nolan,  V.,  Jr.  1978.  The  ecology  and  behavior  of  the 
Prairie  Warbler  Dendroica  discolor.  Ornithologi- 
cal Monographs,  no.  26. 

Payne,  R.  B.  1972.  Mechanisms  and  control  of  molt. 
Pages  103-155  in  Avian  biology,  vol.  2 (D.  D. 
Farner  and  J.  R.  King,  Eds.).  Academic  Press, 
New  York. 

Peck,  G.  and  R.  James.  1987.  Breeding  birds  of  On- 
tario: nidiology  and  distribution,  vol.  2.  Passer- 
ines. Royal  Ontario  Museum  of  Life  Sciences 
Miscellaneous  Publications,  Toronto,  Canada. 

Post,  W.  and  F.  Enders.  1970.  The  occurrence  of  Mal- 
lophaga  on  two  bird  species  occupying  the  same 
habitat.  Ibis  112:539-540. 

Pyle,  P.  1997.  Identification  guide  to  North  American 
birds,  part  1 . Slate  Creek  Press,  Bolinas,  Califor- 
nia. 

Rimmer,  C.  C.  1988.  Timing  of  the  definitive  pre-basic 
molt  in  Yellow  Warblers  at  James  Bay,  Ontario. 
Condor  90:141-156. 

Schieltz,  P.  C.  AND  M.  E.  Murphy.  1997.  The  contri- 
bution of  insulation  changes  to  the  energy  cost  of 
avian  molt.  Canadian  Journal  of  Zoology  75:396- 
400. 

Sherry,  T.  W.  and  R.  T.  Holmes.  1997.  American 
Redstart  ( Setophaga  ruticilla ).  The  Birds  of  North 
America,  no.  277. 

Sokal,  R.  R.  and  F.  J.  Rohlf.  1995.  Biometry.  W.  H. 
Freeman  and  Company,  New  York. 

SPSS,  Inc.  2000.  SPSS  for  Macintosh,  ver.  10.07a. 
SPSS,  Inc.,  Chicago,  Illinois. 

Woodrey,  M.  S.  and  C.  R.  Chandler.  1997.  Age- 
related  timing  of  migration:  geographic  and  inter- 
specific patterns.  Wilson  Bulletin  109:52-67. 

Yuri,  T.  and  S.  Rohwer.  1997.  Molt  and  migration  in 
the  Northern  Rough-winged  Swallow.  Auk  114: 
249-262. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):380 — 390,  2006 


FORAGING  ECOLOGY  OF  BALD  EAGLES  AT  AN 
URBAN  LANDFILL 

KYLE  H.  ELLIOTT,1  JASON  DUFFE,2 3  SANDI  L.  LEE,1  PIERRE  MINEAU,2  AND 

JOHN  E.  ELLIOTT1 3 


ABSTRACT. — We  observed  Bald  Eagles  ( Haliaeetus  leucocephalus)  foraging  at  the  landfill  in  Vancouver, 
British  Columbia,  Canada,  1994-1996  and  2001-2002,  to  determine  (1)  diet  and  time  budgets  of  eagles  visiting 
the  landfill;  (2)  whether  food  taken  from  the  landfill  provided  a significant  energy  source  for  local  eagle  popu- 
lations; and  (3)  the  effects  of  eagle  density  and  weather  on  eagle  behavior.  Eagles  fed  primarily  on  human  refuse 
(95%,  n = 628),  but  food  items  taken  from  the  landfill  accounted  for  only  10  ± 3%  of  their  daily  energy  needs. 
Subadults  foraged  at  the  landfill  more  often  than  adults,  and  most  “refuse  specialists”  appeared  to  be  subadults. 
Eagle  time  budgets  consisted  of  mostly  resting  (91%),  the  remainder  largely  spent  drinking  (2.6%),  scavenging 
(2.3%),  and  pirating  (1.8%).  Resting  increased  with  wind  speed,  and  foraging  efficiency  declined  with  precipi- 
tation, consistent  with  the  hypothesis  that  the  landfill  is  primarily  a location  for  resting  during  inclement  weather. 
Foraging  efficiency  decreased  when  number  of  eagles  and  piracies  increased,  and  percent  of  eagles  foraging 
decreased  with  increased  numbers  of  eagles.  The  home  ranges  of  only  2 of  1 1 radio-tagged  eagles,  both  subadults, 
consisted  largely  (>20%)  of  the  landfill;  home-range  size  and  percent  of  the  home  range  that  included  the 
landfill  were  negatively  correlated,  suggesting  that  most  eagles  visited  the  landfill  occasionally  while  a few  spent 
most  of  their  time  there.  We  concluded  that  (1)  the  Vancouver  landfill  was  not  a major  energy  source  for  eagles, 
in  part  because  their  foraging  is  inefficient  due  to  the  large  number  of  potential  pirates;  (2)  most  eagles  apparently 
used  the  landfill  primarily  as  a site  for  resting  during  inclement  weather  (the  landfill  is  protected  from  the  wind, 
is  slightly  warmer  than  surrounding  areas  due  to  decomposing  refuse  and  the  surrounding  conifer  trees,  and  is 
relatively  free  of  human  activity);  and  (3)  a small  population  of  largely  subadult  refuse  specialists  appeared  to 
gain  much  or  all  of  their  energy  from  the  landfill.  Received  14  December  2004,  accepted  2 March  2006. 


Landfills  can  provide  a constant  and  abun- 
dant food  source  for  birds,  potentially  increas- 
ing reproductive  success  at  nearby  nesting  col- 
onies (Pons  and  Migot  1995,  Tortosa  et  al. 
2003)  and  allowing  some  regions  to  support 
otherwise  unsustainable  populations  (Sibly  and 
McCleery  1983).  During  the  breeding  season, 
landfills  are  particularly  important  for  several 
species,  including  American  Crow  ( Corvus 
brachyrhynchos , Stouffer  and  Caccamise 
1991),  Alpine  [currently  Yellow-billed] 
Chough  ( Pyrrhocorax  graculus,  Delestrade 
1994),  White  Stork  ( Ciconia  ciconia,  Tortosa 
et  al.  2003),  Black  Kite  ( Milvus  migrans , Blan- 
co 1997)  and  Common  Raven  ( Corvus  corax , 
Restani  et  al.  2001).  Foraging  at  landfills,  how- 
ever, can  lower  avian  survivorship  and  repro- 
duction (Pierotti  and  Annett  1991,  Smith  and 
Carlile  1993,  Annett  and  Pierotti  1999)  due  to 


1 Canadian  Wildlife  Service,  Pacific  Wildlife  Re- 
search Centre,  5421  Robertson  Rd.,  Delta,  BC  V4K 
3N2,  Canada. 

2 Canadian  Wildlife  Service,  National  Wildlife  Re- 
search Centre,  Carleton  Univ.,  Ottawa,  ON  KIM  2A6, 
Canada. 

3 Corresponding  author;  e-mail: 
john.elliott@ec.gc.ca 


poor  food  quality  (Smith  and  Carlile  1993,  An- 
nett and  Pierotti  1999),  increased  transmission 
of  disease  (Durrant  and  Beatson  1981,  Mon- 
aghan et  al.  1985,  Ortiz  and  Smith  1994),  in- 
gestion of  synthetics  (Inigo  Elias  1987),  and 
contamination  by  toxins  (Millsap  et  al.  2005). 
During  the  nonbreeding  season,  some  popula- 
tions of  Bald  Eagles  ( Haliaeetus  leucocephal- 
us) are  highly  mobile  foragers,  traveling  thou- 
sands of  km  to  congregate  where  food  is  abun- 
dant (Knight  and  Knight  1983,  Knight  and 
Skagen  1988,  Restani  et  al.  2000).  Because 
food  availability  during  late  winter  is  critical  to 
eagle  survivorship  (Sherrod  et  al.  1976,  Stal- 
master  and  Gessaman  1984),  the  additional 
food  available  at  landfills  might  contribute  to 
increases  in  local  eagle  populations  (Hancock 
2003).  Sherrod  et  al.  (1976)  and  Jackson 
(1981)  attributed  a population  increase  of  ea- 
gles to  increased  food  supply  at  a landfill. 

Understanding  the  population  effects  of 
landfills  in  British  Columbia  is  important  for 
several  reasons.  Moul  and  Gebauer  (2002), 
Sullivan  et  al.  (2002),  and  Vennesland  (2004) 
suggested  that  landfills  increased  eagle  car- 
rying capacities,  which,  in  turn,  impacted  wa- 
terbird  populations.  Increased  eagle  numbers 


380 


Elliott  et  al.  • EAGLE  FORAGING  ECOLOGY  AT  VANCOUVER  LANDFILL 


381 


in  the  Pacific  Northwest  (Dunwiddie  and 
Kuntz  2001,  Watson  et  al.  2002),  purportedly 
due  to  anthropogenic  food  sources,  has  led 
some  First  Nation  groups  of  British  Columbia 
to  request  permission  to  harvest  eagles.  The 
Vancouver  landfill  manager  is  considering  a 
number  of  bird-harassment  techniques,  includ- 
ing covering  the  active  area  with  netting,  to 
reduce  bird  numbers  and  the  potential  for  air- 
craft-bird collisions  at  a nearby  airport  (P. 
Henderson  pers.  comm.).  The  potential  con- 
sequence of  such  practices  on  eagle  popula- 
tions is  unknown. 

On  the  other  hand,  eagles  have  died  from 
pentobarbital  poisoning  after  eating  eutha- 
nized animals  that  were  improperly  wrapped 
at  landfills  on  Vancouver  Island,  Canada 
(three  poisoned;  Wilson  et  al.  1997),  and  at 
numerous  locations  in  the  United  States  (50 
cases  nationwide;  Millsap  et  al.  2005).  Mill- 
sap  et  al.  (2005)  reported  reduced  survival  of 
“suburban”  eagles  compared  with  “rural”  ea- 
gles, with  11%  ( n = 18)  of  mortality  occur- 
ring at  landfills.  While  no  eagle  mortality  has 
been  reported  at  the  Vancouver  landfill  (Elliott 
et  al.  1996,  1997),  dozens  of  Glaucous-winged 
Gulls  ( Larus  glaucescens ) died  in  1999  fol- 
lowing ingestion  of  chocolate  at  this  landfill. 

Despite  the  abundance  of  literature  con- 
cerning eagle  foraging  ecology  and  the  large 
number  of  eagles  that  frequent  landfills 
throughout  North  America  (Stalmaster  1987, 
Gerrard  and  Bortolotti  1988,  Buehler  2000), 
there  are  few  published  reports  on  the  rele- 
vance of  landfills  to  eagle  foraging  and  pop- 
ulation ecology.  We  initiated  a study  to  deter- 
mine (1)  diet  and  time  budgets  of  eagles  vis- 
iting the  Vancouver  landfill;  (2)  whether  food 
from  the  landfill  provided  a large  energy 
source  for  local  eagle  populations;  and  (3)  ef- 
fects of  eagle  density,  age,  and  weather  on  ea- 
gle behavior.  Because  eagles  in  the  Pacific 
Northwest  are  primarily  avivores  in  late  win- 
ter (Watson  et  al.  1991,  Hunt  et  al.  1992,  Pe- 
terson et  al.  2001),  we  suspected  that  eagles 
at  the  Vancouver  landfill  fed  primarily  on  the 
gulls  (>30,000)  that  regularly  visit  the  site  in 
mid- winter  (Ward  1973).  We  expected  that  in- 
traspecific pirating  also  would  play  an  impor- 
tant role  at  the  landfill,  as  it  does  along  salmon 
streams  (Stalmaster  and  Gessaman  1984,  Han- 
sen 1986,  Knight  and  Skagen  1988). 


METHODS 

Study  area. — The  Vancouver  landfill  (49° 
15'  N,  123°  10'  W),  located  near  Vancouver, 
British  Columbia,  Canada,  is  a 10-ha  disposal 
site  for  urban  and  commercial  waste.  Sur- 
rounding the  landfill  are  agricultural  lands 
where  eagles  often  hunt  or  scavenge  ducks 
foraging  on  winter  cover  crops.  Boundary 
Bay — where  eagles  often  hunt  and  scavenge 
wintering  waterfowl  numbering  in  the  hun- 
dreds of  thousands — is  5 km  south  of  the 
landfill.  During  1994-1998,  there  were  five 
major  eagle  roosts  within  a 5-km  radius  of  the 
landfill  (Peterson  et  al.  2001),  including  one 
at  Deas  Island  (49°  18'  N,  123°  10'  W)  and 
South  Arm  (49°  18'  N,  123°  108'  W). 

The  landfill  included  an  active  refuse-de- 
position  area  (~1  ha),  where  most  eagle  for- 
aging occurred.  Many  additional  eagles 
perched  in  the  trees  and  on  fence  posts  sur- 
rounding the  landfill.  The  location  of  the  ac- 
tive area  changed  yearly.  Although  eagles  at 
the  landfill  were  continually  surrounded  by 
loud  machines,  the  machines  did  not  deter  the 
birds,  as  they  regularly  perched  on  active  ma- 
chinery or  grabbed  food  as  it  was  being 
dumped,  compacted,  or  moved.  By  contrast, 
eagles  in  surrounding  areas  were  often  ha- 
rassed by  dogs,  photographers,  eagle-watch- 
ers, and  automobiles,  and  there  have  been  a 
number  of  recent  instances  where  eagles  have 
been  shot  in  Greater  Vancouver.  For  example, 
during  1998-2001,  three  large  roost  sites — in- 
cluding Deas  Island  and  South  Arm — are  be- 
lieved to  have  been  abandoned  (the  birds 
moving  elsewhere)  due  to  nearby  housing  de- 
velopments. 

Observations. — To  determine  diets,  time 
budgets,  and  foraging  behaviors,  we  visited 
the  Vancouver  landfill  at  least  once  per  week 
from  11  January  to  18  April  1994  (total  ob- 
servation = 132  hr),  25  January  to  1 March 
1995  (48  hr),  13  February  to  28  March  1996 
(68  hr),  and  10  November  2001  to  28  April 
2002  (224  hr).  Observations  took  place  be- 
tween 06:00  and  20:00  PST  in  4-hr,  randomly 
chosen  blocks.  All  observations  were  made  by 
at  least  two  observers  inside  a vehicle  ap- 
proximately 50  m from  the  active  area.  Due 
to  topography  of  the  active  area,  we  were  un- 
able to  make  observations  from  elsewhere. 
Eagles  were  habituated  to  vehicles  and  heavy 


382 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


machinery,  which  were  always  present  and  of- 
ten <50  m from  eagles,  so  it  seemed  unlikely 
that  we  influenced  eagle  behavior.  Because 
virtually  all  foraging  occurred  within  the  ac- 
tive area  (>99%),  and  because  we  could  mon- 
itor most  of  the  entire  landfill  from  our  van- 
tage point  atop  the  landfill,  we  concluded  that 
our  observations  included  all  foraging  events. 

Once  each  hour,  we  drove  around  the  rim 
of  the  landfill,  counted  adult  and  subadult  ea- 
gles, and  classified  eagle  behaviors  as  resting, 
bathing,  preening,  pirating,  eating,  scaveng- 
ing, drinking,  or  hunting.  We  classified  all  ea- 
gles <5  years  old  as  subadults  according  to 
the  methods  outlined  in  McCollough  (1989). 
We  classified  eagle  behavior  as  follows:  pi- 
rating (chasing  or  harassing  another  bird  car- 
rying or  eating  food),  scavenging  (picking 
through  the  garbage  in  the  landfill  active 
area),  and  foraging  (carrying  food,  pirating, 
scavenging,  or  hunting).  We  classified  the 
number  of  food  items  obtained  per  eagle  for- 
aging attempt  as  “foraging  efficiency.”  Dur- 
ing 1994—1998,  we  also  visited  two  roost  sites 
(Deas  Island  and  South  Arm)  beginning  an 
hour  prior  to  sunset  twice  a week  and  record- 
ed direction  of  arrival  to  determine  whether 
the  eagles  at  the  landfill  were  using  these  roost 
sites. 

We  recorded  wind  speed,  precipitation, 
temperature,  and  percent  cloud  cover  at  the 
active  site  at  the  beginning  and  end  of  each 
observation  period.  For  analysis,  beginning 
and  ending  values  were  averaged.  Detection 
probabilities  for  adult  versus  subadult  eagles 
can  vary,  especially  when  the  birds  are 
perched  (Anthony  et  al.  1999).  However,  the 
proportion  of  subadults  seen  flying  and  for- 
aging at  the  landfill  was  similar  to  the  pro- 
portion seen  roosting  in  the  surrounding  trees 
(KHE  unpubl.  data);  thus,  we  concluded  that 
we  counted  all  eagles  present  (Hancock  1964, 
Anthony  et  al.  1999).  We  recorded  the  direc- 
tion of  arrival  or  departure  of  all  incoming  or 
outgoing  eagles. 

Energy  consumption. — Following  the  pro- 
tocol set  out  by  Dykstra  et  al.  (1998),  Warnke 
et  al.  (2002),  and  Gill  and  Elliott  (2003),  we 
identified  any  item  an  eagle  attempted  to  eat 
during  the  observation  period  and  estimated 
its  size  relative  to  the  eagle’s  talons  or  man- 
dibles. At  the  beginning  of  each  field  season, 
we  spent  10  hr  practicing  food-item  identifi- 


cation. Based  on  104  items  retrieved  later,  we 
obtained  accuracies  of  >95%  for  classifying 
type  and  size  and  80%  for  estimating  food 
mass  based  on  size  estimates.  We  assumed, 
therefore,  that  our  mass  estimates  were  accu- 
rate to  within  20%.  We  estimated  the  mass 
and  caloric  value  of  each  food  item  based  on 
its  size  by  using  a sample  of  food  items  col- 
lected at  the  landfill  or  from  a local  grocery 
store.  We  classified  each  food  item  as  red  meat 
waste  (mammalian  origin,  including  bones 
and  suet),  chicken,  gull,  rat,  garbage,  or  fish. 
To  estimate  post-assimilation  energetic  effi- 
ciencies, we  used  the  mass-specific  energetic 
and  percent  edible  values  provided  in  Stal- 
master  and  Gessaman  (1982)  for  captive  ea- 
gles feeding  on  mammalian  meat  (black-tailed 
jackrabbit,  Lepus  califomicus),  birds  (Mal- 
lard, Anas  platyrhynchos ),  and  fish  (chum 
salmon,  Oncorhynchus  keta).  We  necessarily 
assumed  that  bone  and  suet  had  mass-specific 
post-assimilation  energetic  values  identical  to 
jackrabbit.  Thus,  we  (1)  estimated  size  and 
categorized  food  items;  (2)  used  regressions 
on  a sample  of  items  we  collected  and 
weighed  to  develop  an  item-specific  relation- 
ship between  size  and  mass;  (3)  used  the  re- 
gression between  size  and  mass  on  a subsam- 
ple of  measured  items  to  estimate  the  mass  of 
each  food  item  observed;  (4)  used  mass-spe- 
cific caloric  values  from  the  literature  to  es- 
timate actual  caloric  values  of  each  food  item 
observed;  and  (5)  estimated  digestive  efficien- 
cy from  Stalmaster  and  Gessaman’s  (1982) 
post-assimilation  energetic  efficiencies  to  de- 
termine actual  energy  absorbed. 

Since  the  main  factors  influencing  energy 
intake  and  number  of  eagles  present  were  time 
of  day  and  date,  respectively  (see  Results), 
and  because  both  of  these  relationships  were 
clearly  nonlinear,  we  used  Akaike’s  Informa- 
tion Criterion  (AICc)  to  determine  what  high- 
er-order polynomial  best  described  the  rela- 
tionships between  energy  intake  versus  time 
of  day,  and  number  of  eagles  present  versus 
date  (Burnham  and  Anderson  1998:66-67).  In 
both  cases,  quadratic  polynomials  provided 
the  best  fit  (energy  intake:  AAICc  =8.5;  num- 
ber of  birds:  AAICc  = 26.1,  compared  to  the 
null  model).  Thus,  we  used  the  relationship 
between  energy  intake  and  time  of  day  ob- 
served during  our  random  observation  periods 


Elliott  et  al.  • EAGLE  FORAGING  ECOLOGY  AT  VANCOUVER  LANDFILL 


383 


to  estimate  the  total  number  of  food  items  tak- 
en for  each  day: 

+ $Ti  + lT2o 

where  a,  (3  and  y are  the  coefficients  for  the 
quadratic  regression  of  number  of  prey  items 
eaten  per  hour  against  number  of  hours  after 
sunrise  (Tt).  The  summation  was  taken  over 
all  hours  between  0.5  hr  before  sunrise  and 
0.5  hr  after  sunset.  Energy  intake  per  day  is 
the  product  of  average  energetic  value  of  food 
items,  n,  and  the  number  of  food  items  per 
day,  assuming  energy  content  of  food  items 
does  not  change  with  time  of  day  or  date: 

2 «(<*  + PT,  + Y^2/)- 

Finally,  energy  intake  per  day  is  divided  by 
the  predicted  number  of  eagles  to  determine 
the  energy  intake  per  eagle  per  day: 

y n(a  + (37)  + y T2.) 

1 j a + bDj  + cD2j  ’ 

where  a,  b and  c are  the  coefficients  for  the 
quadratic  regression  of  the  number  of  eagles 
present  against  date  (Dy).  The  summation  was 
taken  over  all  dates  between  1 February  and 
31  March.  An  alternative  formula,  which  av- 
eraged energy  intake  for  each  observation  pe- 
riod over  the  entire  season,  provided  almost 
identical  results  (KHE  unpubl.  data). 

To  estimate  the  population  increase  result- 
ing from  energy  obtained  at  the  landfill,  we 
used  Stalmaster’s  (1983)  model,  which  con- 
verts salmon  carcass  availability  into  “Eagle 
Use  Days.”  We  modified  the  “consumable 
salmon  biomass”  section  of  the  model  to  rep- 
resent the  average  energy  intake  of  eagles  at 
the  landfill  (207  ± 62  kJ/day;  see  Results).  We 
set  the  flight  time  to  0.084  hr/day  (0.7%  of  a 
12-hr  day;  see  Results)  and  human  distur- 
bance to  0 hr  (human  disturbance  at  the  land- 
fill was  minimal);  otherwise,  we  used  default 
values  reported  in  Stalmaster  (1983).  The  20% 
error  estimate  associated  with  food  energy  es- 
timates and  the  error  estimate  (SD)  associated 
with  the  quadratic  regression  coefficients  were 
propagated  through  the  formula  following 
Stalmaster  (1983).  This  uncertainty  was  then 
increased  by  19%  to  account  for  error  within 
the  model  itself  (Stalmaster  1983). 

Radio  telemetry. — In  the  agricultural  fields 


surrounding  the  landfill,  we  radio-tagged  nine 
eagles  (four  adults,  five  subadults)  during  22- 
31  January  1997  and  three  subadult  eagles  on 
18  January  1998.  We  used  172  mHz  backpack 
transmitters  weighing  90  g (Advanced  Telem- 
etry Systems,  Isanti,  Minnesota).  Half-inch 
Teflon  Ribbon  (Bally  Ribbon  Mills,  Bally. 
Pennsylvania)  was  used  to  attach  transmitters 
in  the  backpack  “X”  configuration,  as  de- 
scribed by  Buehler  et  al.  (1995).  Birds  were 
caught  using  floating  fish  snares  or  padded 
leg-hold  traps.  Birds  were  tracked  for  0-17 
days  over  the  next  3 months.  Only  verified 
(triangulated)  locations  were  included  in  the 
analysis.  To  reduce  bias,  we  only  included  the 
1 1 individuals  for  which  we  had  >15  samples. 
The  fixed  kernel  density  estimator  (set  at 
95%),  using  least-squares  cross  validation, 
was  calculated  using  the  ArcView  3.2  Animal 
Movement  Analysis  extension  (Hooge  2005) 
for  individual  birds.  Fixed  kernel  calculates 
utilization  distributions  using  a probabilistic 
model  and  infers  the  relative  amount  of  time 
the  animal  spends  in  any  one  place.  We  cal- 
culated home-range  size  and  the  percent  of  the 
home  range  consisting  of  the  landfill. 

Statistical  analysis. — For  each  behavior 
(resting,  bathing,  preening,  pirating,  eating, 
scavenging,  drinking,  and  hunting),  we  con- 
structed a linear  model  in  which  hours  after 
sunrise,  date,  weather  (cloud  cover,  precipita- 
tion. wind,  and  temperature),  and  number  of 
eagles  present  were  the  independent  variables. 
We  also  constructed  linear  models — with 
number  of  eagles,  percent  of  eagles  foraging 
or  pirating,  and  foraging  efficiency  as  depen- 
dent variables — and  weather  (cloud  cover, 
precipitation,  wind,  and  temperature),  date, 
hours  after  sunrise,  number  of  eagles,  number 
of  pirating  events,  and  percent  of  eagles  for- 
aging as  independent  variables.  We  inserted 
quadratic  terms  into  the  models  to  account  for 
the  dependence  of  eagle  numbers  on  date  and 
foraging  on  time  of  day,  as  described  above. 
For  each  model  we  used  a positive  stepwise 
method  to  remove  all  nonsignificant  factors 
(at  P < 0.05).  We  report  the  R2  values  for  the 
model  that  included  only  significant  factors. 
We  used  contingency  tables  with  Yates’  cor- 
rection for  continuity  to  compare  behaviors  of 
subadults  and  adults  (Zar  1999).  We  used 
Rayleigh’s  Test  to  determine  whether  the  di- 
rections of  birds  coming  in  to  roosts  coincided 


384 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  1.  Foods  consumed  by  Bald  Eagles  at  the  Vancouver  landfill,  British  Columbia,  Canada,  during 
1993-1996  and  2001-2002.  Eagles  consumed  primarily  red  meat  waste  (mammalian  origin)  and  bones. 


Food  item 

No.  consumed 

Percent  of  total  diet 

Wet  mass  (g)a 

Energetic  value  (kJ)a’b 

Red  meat  waste 

194 

30.7 

320  (35) 

1,160  (130) 

Bones 

142 

22.4 

450  (35) 

1,625  (125) 

Garbagec 

42 

6.6 

210  (50) 

0 

Fat/suet 

26 

4.1 

340  (70) 

1,230  (250) 

Glaucous-winged  Gulld 

14 

2.2 

980  (90) 

5,505  (500) 

Fish 

3 

0.4 

310  (80) 

920  (240) 

Rat 

2 

0.3 

245  (80) 

890  (290) 

Chicken 

1 

0.2 

480 

2,700 

Unknown 

204 

32.3 

a Mean  value  (SE). 

b Based  on  the  mean  estimated  mass,  using  the  percent  edibility  from  Stalmaster  and  Gessaman  (1982)  and  mass-specific  caloric  information  provided 
by  the  appropriate  food  labels  from  nearby  grocery  stores  or  the  literature. 
c Includes  inedible  items,  largely  paper. 
d Includes  10  scavenged  and  4 killed  gulls. 


with  directions  from  the  landfill  (Batschelet 
1981).  We  performed  all  tests  in  STATISTI- 
CA  (StatSoft,  Inc.  2004).  We  tested  for  nor- 
mality (Kolmogrov-Smirnov)  and  homogene- 
ity of  variance  (Levine’s  test)  before  using 
parametric  statistics,  and  we  used  arcsine 
transformations  prior  to  doing  statistical  tests 
on  percentages.  Our  P- values  include  Bonfer- 
roni  adjustments  for  multiple  comparisons,  as 
calculated  by  STATISTICA.  If  analysis  of  co- 
variance  provided  no  significant  variation  be- 
tween years,  data  from  separate  years  were 
pooled.  Results  were  considered  significant  if 
P < 0.05.  Results  are  presented  as  means  ± 
SE. 

RESULTS 

Diet  and  energy  intake. — Household  food 
refuse,  particularly  red  meat  waste  and  bones, 
made  up  95%  of  known  food  items  of  Bald 
Eagles  foraging  at  the  landfill  (Table  1).  Al- 
though some  meat  was  identifiable  (e.g.,  sau- 
sage or  hamburger),  most  was  unidentifiable 
and  clearly  putrid  or  decomposing.  Eagles 
also  consumed  garbage,  including  paper  tow- 
els and  plastic  bags.  Glaucous-winged  Gulls 
(10  scavenged,  4 captured  live)  composed 
only  2.2%  of  the  diet.  Average  energy  intake 
per  eagle  was  207  ± 62  kJ/day,  which  was  10 
± 3%  of  the  required  daily  energy  intake.  The 
number  of  “Eagle  Use  Days”  (1,300  ± 400) 
at  the  landfill  during  the  winter  was  equivalent 
to  17  ± 5 eagles  over  the  peak  period  of  use 
from  February— March. 

Time  budgets  and  behavior. — Eagles  at  the 
landfill  spent  most  (91.0%)  of  their  time  rest- 


ing. Resting  occurred  primarily  later  in  the 
day  and  when  more  eagles  were  present.  Rest- 
ing was  linearly  related  ( R 2 = 0.21)  to  number 
of  hours  after  sunrise  (r185  — —4.4,  P < 0.001) 
and  wind  (f186  = 4.0,  P = 0.004).  Percent  time 
bathing  (0.06%),  drinking  (2.6%),  eating 
(1.2%),  flying  (0.7%),  hunting  (0.3%),  pirat- 
ing (1.8%),  preening  (0.6%)  and  scavenging 
(2.3%)  were  not  explained  by  environmental 
variables. 

Peak  numbers  at  both  the  landfill  and  near- 
by roosts  occurred  in  late  winter  (Fig.  1 ),  after 
eagle  numbers  had  peaked  at  local  salmon 
spawning  streams  (Dunwiddie  and  Kuntz 
2001).  The  highest  count  was  453  on  26  Feb- 
ruary 2001  (Fig.  1).  The  percentage  of  adults 
present  at  both  the  landfill  and  nearby  roosts 
declined  with  date  at  similar  rates  (Fig.  1). 
The  percentage  of  eagles  foraging  declined  as 
the  number  of  eagles  present  increased  and 
when  precipitation  fell  (Table  2),  and  was 
greatest  during  the  first  3 hr  after  sunrise  (Fig. 
2).  Foraging  efficiency  increased  as  wind 
speed  increased,  and  it  declined  with  date, 
number  of  eagles  pirating,  number  of  eagles 
present,  and  when  precipitation  fell  (Table  2). 
Overall,  60%  of  food  items  obtained  were  lat- 
er pirated;  84%  of  theft  attempts  were  directed 
against  other  eagles;  and  16%  were  directed 
against  gulls.  The  percentage  of  eagles  pirat- 
ing increased  as  the  percentage  of  eagles  for- 
aging increased,  and  decreased  with  the  num- 
ber of  eagles  present  (Table  2).  The  likelihood 
of  a food  item  being  pirated  increased  with 
size  of  the  food  item  ( R 2 = 0.45,  P < 0.001). 


Elliott  et  al.  • EAGLE  FORAGING  ECOLOGY  AT  VANCOUVER  LANDFILL 


385 


Nov  Dec  Jan  Feb  Mar  Apr 


225 

200 

175 

150 

125 

100 

75 

50 

25 

0 


O 

00 


9L 

Q_ 

m 

03 

CQ 


FIG.  1.  Bald  Eagle  numbers  (solid  lines)  and  the  percentage  of  adult  (as  opposed  to  subadult)  eagles  (hatched 
lines)  present  at  the  Vancouver,  British  Columbia,  Canada,  landfill  (diamonds)  and  at  two  nearby  roost  sites 
(squares)  during  the  weeks  after  1 November.  Eagle  numbers  are  weekly  averages  of  daily  peak  numbers,  and 
percentages  of  adult  eagles  are  weekly  averages.  Values  were  averaged  over  1993-1996  and  2001-2002  (landfill) 
and  1993-1996  (roosts)  winters.  Roosts  were  inactive  in  2001-2002. 


Subadults  spent  more  time  pirating,  scaveng- 
ing, flying,  and  bathing,  whereas  adults  spent 
more  time  hunting  and  resting  (Table  3);  how- 
ever, foraging  efficiency  and  pirating  success 
were  similar  between  adults  and  subadults 
(Table  3). 

Eagles  arriving  to  roost  at  the  South  Arm 
and  Deas  Island  sites  came  from  significantly 
different  directions  than  that  of  the  landfill  (Z 


= 14.5,  P < 0.001).  Eagles  arrived  at  the 
landfill  primarily  from  adjacent  agricultural 
fields  and  not  from  the  South  Arm  and  Deas 
Island  roosts  (Z  = 18.6,  P < 0.001). 

Radio  telemetry. — Six  of  the  1 1 radio- 
tagged  eagles  had  home  ranges  that  included 
the  landfill  (Table  4,  Fig.  3).  There  was  no 
relationship  between  number  of  points  used 
for  analysis  and  home-range  size.  The  two  in- 


TABLE  2.  Number  of  eagles  present  at  the  Vancouver  landfill,  British  Columbia,  Canada,  1993-2002.  Eagle 
numbers  increased  with  increasing  wind,  precipitation,  and  cloud  cover.  The  percentage  of  eagles  foraging 
decreased  with  precipitation  and  number  of  eagles  present.  The  percentage  of  eagles  pirating  decreased  with 
number  of  eagles  but  increased  with  number  of  eagles  foraging.  Foraging  efficiency  increased  with  wind  and 
decreased  with  precipitation,  date,  number  of  eagles  present,  and  number  of  eagles  pirating. 


Effect 

No.  eagles 

Eagles  foraging  (%) 

Eagles  pirating  (%) 

Foraging  efficiency" 

487  P 

7*186 

p 

487 

p 

486 

p 

Wind 

4.2  <0.001 

NSb 

NS 

2.6 

0.012 

Precipitation 

4.1  <0.001 

-2.1 

0.010 

NS 

-2.4 

0.019 

Cloud  cover 

7.5  <0.001 

NS 

NS 

NS 

Date 

c 

NS 

NS 

-3.1 

0.002 

Temperature 

NS 

NS 

NS 

NS 

Hour  after  sunrise 

NS 

d 

NS 

NS 

No.  eagles  present 

— 

-2.7 

0.007 

-2.7 

0.008 

-2.4 

0.02 

Eagles  foraging  (%) 

NS 

— 

3.0 

<0.001 

NS 

No.  eagles  pirating 

NS 

NS 

— 

-9.9 

<0.001 

Re 

0.46 

0.48 

0.08 

0.53 

a Number  of  food  items  taken  per  foraging  attempt. 
b Not  significant  (P  > 0.05). 

c The  linear  model  for  number  of  eagles  fitted  to  a quadratic  term  to  account  for  the  effect  of  date. 

d The  linear  model  for  percentage  of  eagles  foraging  fitted  to  a quadratic  term  to  account  for  the  effect  of  hours  after  sunrise. 
e Refers  to  the  total  linear  model  once  nonsignificant  factors  have  been  removed  (positive  stepwise). 


386 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


Hours  after  sunrise 

LIG.  2.  Percent  of  eagles  foraging  in  relation  to  hours  after  sunrise  at  the  Vancouver,  British  Columbia, 
Canada,  landfill  during  the  winters  of  1993-1996  and  2001-2002.  Peak  foraging  occurred  in  early  and  late  hours 
of  the  day.  Based  on  these  data,  the  quadratic  regression  for  percent  of  eagles  foraging  = 0.18(time  of  day)2 
-2.8(time  of  day)  ± 16;  R2  = 0.67.  Error  bars  represent  SE;  sample  sizes  appear  above,  below,  or  to  the  right 
of  data  points. 


dividuals  whose  home  ranges  largely  consist- 
ed of  the  Vancouver  landfill  (e.g.,  >10%  of 
their  home  range  was  the  Vancouver  landfill) 
had  the  smallest  home  ranges,  and  home- 
range  size  was  negatively  correlated  with  the 
percentage  of  the  home  range  that  encom- 
passed the  landfill  (t5  = —3.05,  P = 0.04,  r2 
= 0.70). 


TABLE  3.  Percent  time  adult  and  subadult  Bald 
Eagles  spent  engaged  in  various  behaviors  at  the  Van- 
couver landfill,  British  Columbia,  Canada,  1993-2002. 
Adults  spent  more  time  resting  and  hunting  than  sub- 
adults, whereas  subadults  spent  more  time  scavenging, 
pirating,  flying,  and  bathing.  Foraging  efficiency,  pi- 
rating success,  and  percent  time  spent  drinking  and 
preening  were  equivalent  between  the  two  groups. 


Behavior 

Adult 

Subadult 

V2 

p 

Resting 

93.1 

88.2 

3.7 

0.048 

Drinking 

2.4 

2.7 

NSa 

Scavenging 

1.0 

5.4 

22.4 

0.001 

Pirating 

0.5 

4.9 

33.7 

0.001 

Preening 

0.6 

0.6 

NS 

Flying 

0.2 

1.5 

9.2 

0.001 

Hunting 

0.8 

0.1 

5.5 

0.016 

Bathing 

0.02 

0.1 

6.2 

0.014 

Foraging  efficiencyb 

0.31 

0.33 

NS 

Pirating  success0 

0.48 

0.49 

NS 

a Not  significant  (P  > 0.05). 

b Number  of  food  items  taken  per  foraging  attempt. 
c Percentage  of  pirating  attempts  that  were  successful. 


DISCUSSION 

Contrary  to  initial  expectations,  the  Van- 
couver landfill  accounted  for  only  10  ± 3% 
of  the  energy  intake  of  the  eagles  that  frequent 
the  landfill.  Furthermore,  the  actual  intake  was 
likely  <10%  because  we  assumed  liberal  val- 
ues for  major  food  items,  such  as  bone  and 
rancid  foods,  and  the  eagles  wasted  consid- 
erable amounts  of  food  that  we  could  not 
quantify.  Eagle  behavior  was  similar  to  that  of 
Herring  Gulls  ( Larus  argentatus),  which  use 
landfills  primarily  for  social  interaction  and 
loafing,  especially  when  higher-quality  food  is 
available  elsewhere  (Belant  et  al.  1993).  Near- 
by waterfowl  concentrations  probably  repre- 
sented a higher-quality  food  base  (Peterson  et 


TABLE  4.  Home-range  sizes  of  eagles  radio- 
tagged  near  the  Vancouver  landfill,  British  Columbia, 
Canada,  decreased  during  winter  1997  and  1998  as  the 
landfill  portion  of  their  home  range  increased. 


Bird 

Year 

Age 

Area  in 
landfill  (%) 

Home  range 
(km2) 

373 

1997 

Subadult 

1.7 

20.4 

241 

1997 

Subadult 

1.5 

27.8 

190 

1997 

Adult 

0.9 

37.3 

210 

1998 

Subadult 

3.4 

14.2 

072 

1998 

Subadult 

20.4 

2.5 

062 

1998 

Subadult 

50.6 

1.5 

Elliott  et  al.  • EAGLE  FORAGING  ECOLOGY  AT  VANCOUVER  LANDFILL 


387 


FIG.  3.  Home  ranges  of  two  (A,  B)  “refuse  specialist”  Bald  Eagles  (>20%  of  their  home  ranges  comprised 
the  Vancouver  landfill)  radiotagged  near  the  landfill  in  Vancouver,  British  Columbia,  Canada,  during  the  winters 
of  1997  and  1998.  Forward  slashes  (III)  represent  eagle  home  ranges;  crosshatching  represents  the  Vancouver 
landfill. 


al.  2001),  and  most  eagles  may  have  foraged 
on  waterfowl.  Consistent  with  this  hypothesis, 
resting  and  overall  numbers  of  eagles  peaked 
during  periods  of  inclement  weather  because 
the  landfill  is  protected  from  the  wind,  is 
slightly  warmer  due  to  decomposing  refuse 
and  surrounding  conifer  trees,  and  is  relatively 
free  of  human  disturbance — all  of  which  are 
known  to  reduce  the  energetic  costs  associated 
with  resting  (Stalmaster  and  Newman  1979, 
Keister  et  al.  1985).  The  possibility  of  feeding 
at  the  landfill  was  likely  an  added  bonus. 

It  is  improbable  that  the  landfill  contributed 
significantly  to  an  increased  eagle  carrying  ca- 
pacity in  the  region,  as  the  observed  energy 
intake  only  accounted  for  an  additional  17  ± 
5 eagles  during  peak  eagle  use.  This  is  a very 
small  number  compared  to  the  500-1,000  ea- 
gles that  use  the  surrounding  area  in  late  win- 
ter, and  it  does  not  account  for  the  30-fold 
population  increase  that  has  occurred  over  the 
last  30  years.  Percent  of  eagles  foraging  de- 
clined with  a decrease  in  the  number  of  eagles 


present,  suggesting  that  the  number  of  forag- 
ers stayed  relatively  constant  and  the  remain- 
der only  visited  to  rest.  Thus,  some  eagles  (the 
refuse  specialists)  may  have  foraged  primarily 
at  the  landfill  and  obtained  much  of  their  en- 
ergy needs  there.  Furthermore,  the  standard 
deviation  for  average  energy  intake  (264  kJ / 
day)  was  greater  than  the  average  intake  rate 
(207  kJ/day)  itself,  indicating  wide  variation 
among  individuals. 

Consistent  with  the  existence  of  refuse  spe- 
cialists, 2 of  11  (18%)  radio-tagged  eagles  had 
a fixed  kernel  home  range  that  mostly  (>20%) 
included  the  landfill,  whereas  another  4 visited 
the  landfill  only  occasionally  (Table  4).  Visual 
inspection  of  the  home  ranges  of  the  two  re- 
fuse specialists  suggests  that  they  rarely  left 
the  landfill;  most  of  the  points  outside  the 
landfill  appeared  to  be  in  adjacent  conifer 
trees,  which  are  used  for  resting  (Fig.  3).  The 
refuse  specialist  estimate  (18%)  is  quite  close 
to  our  estimate  for  the  proportion  of  the  local 
population  that  was  supported  by  energy  in- 


388 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


take  from  the  landfill  (10  ± 2%).  It  appears 
that  younger  eagles  were  the  refuse  special- 
ists, because  they  spent  more  of  their  time  for- 
aging and  older  eagles  spent  more  time  resting 
at  the  landfill — possibly  because  younger  ea- 
gles are  less  efficient  hunters  than  the  adults 
(Stalmaster  and  Gessaman  1984,  Brown  1993, 
Bennetts  and  McClelland  1997).  A similar 
study  at  a nearby  salmon  stream  in  late  winter 
showed  a strong  relationship  between  pirating 
success  and  age  (Stalmaster  and  Gessaman 
1984),  and,  at  the  Vancouver  landfill,  sub- 
adults pirated  more  than  adults;  this  may  re- 
flect a change  in  dominance  structure  associ- 
ated with  the  predictability  of  anthropogenic 
food  sources  (e.g.,  Restani  et  al.  2001).  More- 
over, home  ranges  of  refuse  specialists  in  a 
wide  variety  of  taxa  are  much  smaller  than  the 
average  home  range  size,  and  reduced  home 
range  size  is  often  associated  with  a change 
in  social  structure  due  to  increased  density  at 
landfills  (e.g.,  Blanchard  and  Knight  1991, 
Delestrade  1994,  Gilchrist  and  Otali  2002). 

Pirating  was  common  at  the  Vancouver 
landfill,  which  may  partially  explain  why  few 
eagles  forage  there.  Foraging  efficiency  and 
the  percent  of  birds  foraging  declined  as  the 
number  of  birds  present  and  pirating  in- 
creased. Although  piracy  is  also  common  at 
waterfowl  carcasses  (Peterson  et  al.  2001)  and 
salmon  streams  (Stalmaster  and  Gessman 
1984),  it  may  be  that  the  higher  quality  of 
those  food  types  makes  pirating  them  more 
worthwhile  energetically.  Eagles  at  the  landfill 
pirated  primarily  conspecifics;  thus,  although 
both  gulls  and  eagles  competed  for  the  same 
resource  (human  refuse),  there  appeared  to  be 
few  interactions  between  them. 

At  both  the  landfill  and  nearby  roosts,  the 
timing  of  peak  eagle  numbers  and  the  per- 
centage of  adults  present  were  similar,  sup- 
porting our  assumption  (based  on  radiotelem- 
etry) that  individuals  regularly  moved  be- 
tween these  sites  (this  study,  Servheen  and 
English  1979,  Hunt  et  al.  1992).  The  percent- 
age of  subadults  increased  over  the  winter  at 
both  locations,  not  only  because  subadults 
learned  about  food  concentrations  from  adults 
(Knight  and  Knight  1983,  Bennetts  and 
McClelland  1997,  Restani  et  al.  2000).  but 
also  because  many  breeders  returned  to  their 
territories  in  late  fall  (Stalmaster  and  Kaiser 
1997). 


Eagles  spent  most  of  their  time  resting 
(91%).  At  the  landfill,  they  rested  more  than 
they  did  at  the  Columbia  River  estuary  (54%; 
Watson  et  al.  1991),  and  they  spent  less  time 
flying  (0.7%  versus  6%).  Overall,  time  spent 
flying  was  similar  to  that  reported  on  the 
Nooksack  River  (1.0%;  Stalmaster  and  Ges- 
saman 1984).  In  previous  studies,  eagles 
(Sherrod  et  al.  1976)  and  gulls  (Sibly  and 
McCleery  1983,  Coulson  et  al.  1987)  at  sev- 
eral landfills  foraged  whenever  the  landfills 
were  active,  with  peak  foraging  occurring 
when  the  landfill  machinery  activities 
stopped.  In  contrast,  eagles  at  the  Vancouver 
landfill — where  food  was  available  almost 
continuously  because  refuse  dumping  started 
every  day  before  sunrise  (06:30)  and  did  not 
end  until  after  sunset  (18:30) — foraged  pri- 
marily during  early  morning  and  late  after- 
noon (Fig.  2).  This  reflects  the  typical  diurnal 
feeding  patterns  of  eagles  (Watson  et  al.  1991; 
Elliott  et  al.  2003,  2005),  as  well  as  the  short 
day  length  during  Vancouver  winters. 

ACKNOWLEDGMENTS 

We  thank  R Henderson  and  the  Greater  Vancouver 
Regional  District  for  permission  to  enter  the  landfill. 
M.  Porter  provided  interesting  anecdotal  information. 
L.  L.  Jordison  helped  with  some  of  the  surveys.  A. 
Fabro  provided  us  with  copies  of  some  of  the  more 
obscure  literature.  J.  M.  Touchton  and  the  Smithsonian 
Tropical  Research  Institute  provided  computer  support. 
R.  W.  Butler,  W.  G.  Hunt,  A.  R.  Harmata,  J.  W.  Wat- 
son, and  an  anonymous  reviewer  provided  valuable 
comments  on  an  earlier  draft  of  this  manuscript. 

LITERATURE  CITED 

Annett,  C.  A.  and  R.  Pierotti.  1999.  Long-term  re- 
productive output  in  Western  Gulls:  consequences 
of  alternate  tactics  in  diet  choice.  Ecology  80: 
288-297. 

Anthony,  R.  G.,  M.  G.  Garrett,  and  F.  B.  Isaacs. 
1999.  Double-survey  estimates  of  Bald  Eagle  pop- 
ulations in  Oregon.  Journal  of  Wildlife  Manage- 
ment 63:794-802. 

Batschelet,  E.  1981.  Circular  statistics  in  biology. 

Academic  Press,  New  York. 

Belant,  J.  L.,  T.  W.  Seamans,  S.  W.  Gabrey,  and  S. 
K.  Ickes.  1993.  Importance  of  landfills  to  nesting 
Herring  Gulls.  Condor  95:817-830. 

Bennetts,  R.  E.  and  B.  R.  McClelland.  1997.  Influ- 
ence of  age  and  prey  availability  on  Bald  Eagle 
foraging  behavior  at  Glacier  National  Park,  Mon- 
tana. Wilson  Bulletin  109:393-409. 

Blanchard,  B.  M.  and  R.  R.  Knight.  1991.  Move- 
ments of  Yellowstone  grizzly  bears.  Biological 
Conservation  58:41-68. 


Elliott  et  al.  • EAGLE  FORAGING  ECOLOGY  AT  VANCOUVER  LANDFILL 


389 


Blanco,  G.  1997.  Role  of  refuse  as  food  for  migrant, 
floater  and  breeding  Black  Kites  ( Milvus  mig- 
rans).  Journal  of  Raptor  Research  31:71-76. 

Brown,  B.  T.  1993.  Wintering  foraging  ecology  of 
Bald  Eagles  in  Arizona.  Condor  95:132-138. 

Buehler,  D.  A.  2000.  Bald  Eagle  ( Halieaeetus  leu- 
cocephalus ).  The  Birds  of  North  America,  no. 
564. 

Buehler,  D.  A.,  J.  D.  Fraser,  M.  R.  Fuller,  L.  S. 
McAllister,  and  J.  K.  D.  Seegar.  1995.  Captive 
and  field-tested  radio  transmitter  attachments  for 
Bald  Eagles.  Journal  of  Field  Ornithology  66: 
173-180. 

Burnham,  K.  P.  and  D.  R.  Anderson.  1998.  Model 
selection  and  multimodel  inference.  Springer- Ver- 
lag,  New  York. 

Coulson,  J.  C.,  J.  Butterfield,  N.  Duncan,  and  C. 
Thomas.  1987.  Use  of  refuse  tips  by  adult  British 
Flerring  Gulls  Larus  argentatus  during  the  week. 
Journal  of  Applied  Ecology  24:789-800. 

Delestrade,  A.  1994.  Factors  affecting  flock  size  in 
the  Alpine  Chough  Pyrrhocorax  graculus.  Ibis 
136:91-96. 

Dunwiddie,  P.  W.  and  R.  C.  Kuntz,  II.  2001.  Long- 
term trends  of  Bald  Eagles  in  winter  on  the  Skagit 
River,  Washington.  Journal  of  Wildlife  Manage- 
ment 65:290-299. 

Durrant,  D.  S.  and  S.  H.  Beatson.  1981.  Salmonellae 
isolated  from  domestic  meat  waste.  Journal  of  Hy- 
giene 86:259-264. 

Dykstra,  C.  J.  R.,  M.  W.  Meyer,  D.  K.  Warnke,  W. 
H.  Karasov,  D.  E.  Andersen,  W.  W.  Bowerman, 
IV,  and  J.  P.  Giesy.  1998.  Low  reproductive  rates 
of  Lake  Superior  Bald  Eagles:  low  food  delivery 
rates  or  environmental  contamination?  Journal  of 
Great  Lakes  Research  24:22-34. 

Elliott,  J.  E.,  K.  M.  Langelier,  P.  Mineau,  and  L. 
K.  Wilson.  1996.  Poisoning  of  Bald  Eagles  and 
Red-tailed  Hawks  by  carbofuran  and  fensulfothion 
in  the  Fraser  Delta  of  British  Columbia,  Canada. 
Journal  of  Wildlife  Diseases  32:486-491. 

Elliott,  J.  E.,  L.  K.  Wilson,  K.  M.  Langeler,  P.  Mi- 
neau, and  P.  H.  Sinclair.  1997.  Secondary  poi- 
soning of  birds  of  prey  by  the  organophosphorus 
insecticide  phorate.  Ecotoxicology  6:219-231. 

Elliott,  K.  H.,  C.  E.  Gill,  and  J.  E.  Elliott.  2005. 
Influence  of  tide  and  weather  on  Bald  Eagle  pro- 
visioning rates  in  coastal  British  Columbia.  Jour- 
nal of  Raptor  Research  39:99-108. 

Elliott,  K.  H.,  C.  L.  Struik,  and  J.  E.  Elliott.  2003. 
Bald  Eagles,  Haliaeetus  leucocephalus,  feeding 
on  spawning  plainfin  midshipman,  Porichthys  no- 
tatus,  at  Crescent  Beach,  British  Columbia.  Ca- 
nadian Field-Naturalist  117:601-604. 

Gerrard,  J.  M.  and  G.  R.  Bortolotti.  1988.  The 
Bald  Eagle:  haunts  and  habits  of  a wilderness 
monarch.  Smithsonian  Institution  Press,  Washing- 
ton, D.C. 

Gilchrist,  J.  S.  and  E.  Otali.  2002.  The  effects  of 
refuse-feeding  on  home-range  use,  group  size,  and 


intergroup  encounters  in  the  banded  mongoose. 
Canadian  Journal  of  Zoology  80:1795-1802. 

Gill,  C.  E.  and  J.  E.  Elliott.  2003.  Influence  of  food 
supply  and  chlorinated  hydrocarbon  contaminants 
on  breeding  success  of  Bald  Eagles.  Ecotoxicol- 
ogy 12:95-111. 

Hancock,  D.  1964.  Bald  Eagles  wintering  in  the 
southern  Gulf  Islands,  British  Columbia.  Wilson 
Bulletin  76:117-120. 

Hancock,  D.  2003.  The  Bald  Eagle  of  Alaska,  BC  and 
Washington.  Hancock  House,  Surrey,  British  Co- 
lumbia, Canada. 

Hansen,  A.  J.  1986.  Fighting  behavior  in  Bald  Eagles: 
a test  of  game  theory.  Ecology  67:787-797. 

Hooge,  P.  N.  and  B.  Eichenlaub.  2005.  Animal  Move- 
ment Extension  to  ArcView,  ver.  3.2.  U.S.  Geo- 
logical Survey  Alaska  Science  Center,  Juneau 
Alaska. 

Hunt,  W.  G.,  B.  S.  Johnson,  and  R.  J.  Jackman.  1992. 
Carrying  capacity  for  Bald  Eagles  wintering  along 
a northwestern  river.  Journal  of  Raptor  Research 
26:49-60. 

Inigo  Elias,  E.  E.  1987.  Feeding  habits  and  ingestion 
of  synthetic  products  in  a Black  Vulture  popula- 
tion from  Chiapas,  Mexico.  Acta  Zoologica  Mex- 
icana  Nueva  Serie  22:1-16. 

Jackson,  F.  L.  1981.  King  of  the  heap.  National  Wild- 
life 19:36-39. 

Keister,  G.  P,  R.  G.  Anthony,  and  H.  R.  Holbo. 
1985.  A model  of  energy  consumption  in  Bald 
Eagles  Haliaeetus  leucocephalus : an  evaluation  of 
night  communal  roosting.  Wilson  Bulletin  97: 
148-160. 

Knight,  R.  L.  and  S.  K.  Skagen.  1988.  Agonistic 
asymmetries  and  the  foraging  ecology  of  Bald  Ea- 
gles. Ecology  69:1188-1194. 

Knight,  S.  K.  and  R.  L.  Knight.  1983.  Aspects  of 
food  finding  by  wintering  Bald  Eagles.  Auk  100: 
477-484. 

McCollough,  M.  A.  1989.  Molting  sequence  and  ag- 
ing of  Bald  Eagles.  Wilson  Bulletin  101:1-10. 

Mills ap,  B.,  T.  Breen,  E.  McConnell,  T.  Steffer,  L. 
Phillips,  N.  Douglass,  and  S.  Taylor.  2005. 
Comparative  fecundity  and  survival  of  Bald  Ea- 
gles fledged  from  suburban  and  rural  natal  areas 
in  Florida.  Journal  of  Wildlife  Management  68: 
1018-1031. 

Monaghan,  P,  C.  B.  Shedden,  K.  Ensor,  C.  R.  Frick- 
er,  and  R.  W.  A.  Girdwood.  1985.  Salmonella 
carriage  by  Herring  Gulls  in  the  Clyde  area  of 
Scotland  in  relation  to  their  feeding  ecology.  Jour- 
nal of  Applied  Ecology  22:669-680. 

Moul,  I.  E.  and  M.  B.  Gebauer.  2002.  Status  of  the 
Double-crested  Cormorant  in  British  Columbia. 
Wildlife  Working  Report,  no.  105.  Ministry  of 
Water,  Land  and  Air  Protection,  Surrey,  British 
Columbia,  Canada. 

Ortiz,  N.  E.  and  G.  R.  Smith.  1994.  Landfill  sites, 
botulism  and  gulls.  Epidemiology  and  Infection 
112:385-391. 

Peterson,  C.  A.,  S.  L.  Lee,  and  J.  E.  Elliott.  2001. 


390 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  3,  September  2006 


Scavenging  of  waterfowl  carcasses  by  birds  in  ag- 
ricultural fields  of  British  Columbia.  Journal  of 
Field  Ornithology  72:150-159. 

Pierotti,  R.  and  C.  A.  Annett.  1991.  Diet  choice  in 
the  Herring  Gull:  constraints  imposed  by  repro- 
ductive and  ecological  factors.  Ecology  72:319- 
328. 

Pons,  J.  M.  and  P.  Migot.  1995.  Life-history  strategy 
of  the  Herring  Gull:  changes  in  survival  and  fe- 
cundity in  a population  subjected  to  various  feed- 
ing conditions.  Journal  of  Animal  Ecology  64: 
592-599. 

Restani,  M.,  A.  R.  Harmata,  and  E.  M.  Madden. 
2000.  Numerical  and  functional  responses  of  mi- 
grant Bald  Eagles  exploiting  a seasonally  concen- 
trated food  source.  Condor  102:561-568. 

Restani,  M.,  J.  M.  Marzluff,  and  R.  E.  Yates.  2001. 
Effects  of  anthropogenic  food  sources  on  move- 
ments, survivorship,  and  sociality  of  Common  Ra- 
vens in  the  Arctic.  Condor  103:399-404. 

Servheen,  C.  and  W.  English.  1979.  Movements  of 
rehabilitated  Bald  Eagles  and  proposed  seasonal 
movement  patterns  of  Bald  Eagles  in  the  Pacific 
Northwest.  Raptor  Research  13:79-88. 

Sherrod,  S.  K.,  C.  M.  White,  and  F.  S.  L.  William- 
son. 1976.  Biology  of  the  Bald  Eagle  on  Amchit- 
ka  Island,  Alaska,  USA.  Living  Bird  15:143-182. 

Sibly,  R.  M.  and  R.  H.  McCleery.  1983.  Increase  in 
weight  of  Herring  Gulls  while  feeding.  Journal  of 
Animal  Ecology  52:35-50. 

Smith,  G.  C.  and  N.  Carlile.  1993.  Food  and  feeding 
ecology  of  breeding  Silver  Gulls  ( Larus  novae- 
hollandiae ) in  urban  Australia.  Colonial  Water- 
birds  16:9-17. 

Stalmaster,  M.  V.  1983.  An  energetics  simulation 
model  for  managing  wintering  Bald  Eagles.  Jour- 
nal of  Wildlife  Management  47:349-359. 

Stalmaster,  M.  V.  1987.  The  Bald  Eagle.  Universe 
Books,  New  York. 

Stalmaster,  M.  V.  and  J.  A.  Gessaman.  1982.  Food 
consumption  and  energy  requirements  of  captive 
Bald  Eagles.  Journal  of  Wildlife  Management  46: 
646-654. 

Stalmaster,  M.  V.  and  J.  A.  Gessaman.  1984.  Eco- 
logical energetics  and  foraging  behavior  of  over- 
wintering Bald  Eagles.  Ecological  Monographs 
54:407-428. 

Stalmaster,  M.  V.  and  J.  L.  Kaiser.  1997.  Winter 
ecology  of  Bald  Eagles  in  the  Nisqually  River 


drainage,  Washington.  Northwest  Science  71:2 14 — 
223. 

Stalmaster,  M.  V.  and  J.  R.  Newman.  1979.  Perch- 
site  preferences  of  wintering  Bald  Eagles  in  north- 
west Washington.  Journal  of  Wildlife  Manage- 
ment 43:221-224. 

StatSoft,  Inc.  2004.  STATISTICA  7.0  user’s  guide. 
StatSoft,  Inc.,  Tulsa,  Oklahoma. 

Stouffer,  P.  C.  and  D.  F.  Caccamise.  1991.  Roosting 
and  diurnal  movements  of  radio-tagged  American 
Crows.  Wilson  Bulletin  103:387-400. 

Sullivan,  T.  M.,  S.  L.  Hazlitt,  and  M.  J.  F.  Lemon. 
2002.  Population  trends  of  nesting  Glaucous- 
winged Gulls,  Larus  glaucescens,  in  the  southern 
Strait  of  Georgia,  British  Columbia.  Canadian 
Field-Naturalist  116:603-606. 

Tortosa,  F.  S.,  L.  Perez,  and  L.  Hillstrom.  2003. 
Effect  of  food  abundance  on  laying  date  and 
clutch  size  in  the  White  Stork  Ciconia  ciconia. 
Bird  Study  50:112-115. 

Vennesland,  R.  G.  2004.  Great  Blue  Heron.  Accounts 
and  measures  for  managing  identified  wildlife,  ac- 
counts V.  Ministry  of  Water,  Land  and  Air  Pro- 
tection, Surrey,  British  Columbia,  Canada. 

Vennesland,  R.  G.  and  R.  W.  Butler.  2004.  Factors 
influencing  Great  Blue  Heron  nesting  productivity 
on  the  Pacific  coast  of  Canada  from  1998  to  1999. 
Waterbirds  27:289-296. 

Ward,  J.  G.  1973.  Reproductive  success,  food  supply 
and  evolution  of  clutch  size  in  the  Glaucous- 
winged Gull.  Ph.D.  dissertation,  University  of 
British  Columbia,  Vancouver. 

Warnke,  D.  K.,  D.  E.  Andersen,  C.  J.  R.  Dykstra, 
M.  W.  Meyer,  and  W.  H.  Karasov.  2002.  Pro- 
visioning rates  and  time  budgets  of  adult  and  nest- 
ling Bald  Eagles  at  inland  Wisconsin  nests.  Jour- 
nal of  Raptor  Research  36:121-127. 

Watson,  J.  W.,  M.  G.  Garrett,  and  R.  G.  Anthony. 
1991.  Foraging  ecology  of  Bald  Eagles  in  the  Co- 
lumbia River  estuary,  USA.  Journal  of  Wildlife 
Management  55:492-499. 

Watson,  J.  W.,  D.  Stinson,  K.  R.  McAllister,  and 
T.  E.  Owens.  2002.  Population  status  of  Bald  Ea- 
gles breeding  in  Washington  at  the  end  of  the  20th 
century.  Journal  of  Raptor  Research  36:161-169. 

Wilson,  L.  K..  J.  E.  Elliott,  and  M.  McAdie.  1997. 
Barbiturate  poisoning  of  Bald  Eagles.  Wildlife 
Health  Centre  Newsletter  (Canadian  Cooperative 
Wildlife  Health  Centre)  5(1  ):8. 

Zar,  J.  H.  1999.  Biostatistical  analysis.  Prentice-Hall, 
New  York. 


The  Wilson  Journal  of  Ornithology  1 18(3):39 1-398,  2006 


TERRITORY  SELECTION  BY  UPLAND  RED-WINGED 
BLACKBIRDS  IN  EXPERIMENTAL  RESTORATION  PLOTS 

MARIA  A.  FUREY1 3 AND  DIRK  E.  BURHANS24 


ABSTRACT. — We  examined  territory  selection  of  Red-winged  Blackbirds  ( Agelaius  phoeniceus)  in  experi- 
mental treatments  with  varied  groundcovers  and  densities  of  planted  and  naturally  occurring  oaks  ( Quercus  spp.) 
used  by  blackbirds  for  perching.  We  also  compared  vegetation  parameters  between  blackbird  territories  and 
unused  (i.e.,  unoccupied  by  Red-winged  Blackbirds)  areas.  Although  perch  densities  were  greater  in  blackbird 
territories  in  unplanted  controls  and  oak-planted  treatments  without  redtop  grass  ( Agrostis  gigantea)  than  they 
were  in  unused  areas,  the  low  densities  of  perches  in  territories  planted  with  redtop  grass  indicate  that  perch 
density  is  not  limiting  above  some  lower  threshold.  Territories,  particularly  in  treatments  with  no  redtop,  tended 
to  have  greater  mean  grass  cover  and  taller  grass  heights  than  unused  areas.  Our  results  are  consistent  with  other 
studies  in  finding  that  Red-winged  Blackbirds  prefer  areas  having  tall  vegetation  and  dense  grass.  Received  14 
July  2005,  accepted  21  February  2006. 


A large  body  of  observational  studies  has 
documented  relationships  between  avian 
abundance,  or  territory  use,  and  vegetation  pa- 
rameters. Examples  include  studies  comparing 
differences  among  songbird  territories  with  re- 
spect to  vegetation  height  or  litter  depth 
(Wiens  1969)  and  grass  or  shrub  cover  (Ro- 
tenberry  and  Wiens  1980),  and  those  that  re- 
late avian  abundance  to  vegetation  density 
(Orians  and  Wittenberger  1991)  or  grass 
(Scott  et  al.  2002).  However,  important  rela- 
tionships between  vegetation  and  habitat  use 
can  be  obscured  if  the  variation  among  study 
sites  (or  plots)  is  minimal  (Orians  and  Witten- 
berger 1991,  Pribil  and  Pieman  1997).  One 
way  to  elucidate  habitat  variation  and  distin- 
guish factors  important  in  habitat  selection  is 
by  comparing  sites  that  differ  explicitly  in 
terms  of  vegetation  management.  For  exam- 
ple, Shochat  et  al.  (2005),  Wood  et  al.  (2004), 
and  Murkin  et  al.  (1997)  evaluated  avian  re- 
sponses among  plots  that  varied  with  respect 
to  management  regime,  and  were  able  to  make 
clear  inferences  that  may  have  been  obscured 
had  they  studied  only  unmanaged  habitats. 

Even  where  variation  among  plots  is  made 
explicit,  however,  the  influences  of  vegetative 


1 Dept,  of  Forestry,  203  Natural  Resources  Bldg., 
Univ.  of  Missouri,  Columbia,  MO  65211-7270,  USA. 

2 Dept,  of  Fisheries  and  Wildlife  Sciences,  Natural 
Resources  Bldg.,  Univ.  of  Missouri,  Columbia,  MO 
65211-7270,  USA. 

3 Current  address:  P.O.  Box  7021,  Columbia,  MO 
65205-7021,  USA. 

4 Corresponding  author;  e-mail: 
burhansd@missouri.edu 


factors  on  avian  settlement  patterns  may  be 
masked  if  measurements  are  made  at  inappro- 
priate scales  (Orians  and  Wittenberger  1991, 
Pribil  and  Pieman  1997).  For  example,  Orians 
and  Wittenberger  (1991)  found  that  Yellow- 
headed Blackbirds  ( Xanthocephalus  xantho- 
cephalus)  settle  according  to  food  supplies  at 
the  scale  of  an  entire  marsh,  a relationship  that 
was  not  apparent  at  the  territory  scale.  Simi- 
larly, Burhans  (1997)  found  that  some  factors 
explaining  brood  parasitism  at  the  nest-site 
scale  were  relevant  only  when  considered  at 
the  larger  scale  of  habitat. 

We  investigated  the  role  of  vegetation  struc- 
ture in  the  selection  of  breeding  territories  by 
Red-winged  Blackbirds  (. Agelaius  phoeniceus ) 
in  two  experimentally  manipulated  restoration 
sites  of  floodplain  oak  ( Quercus  spp.)  near  the 
Missouri  River.  Numerous  researchers  have 
investigated  habitat  selection  by  Red-winged 
Blackbirds  (Albers  1978,  Joyner  1978,  Pribill 
and  Pieman  1997,  Turner  and  McCarthy 
1998),  and  some  have  examined  responses  of 
Red-winged  Blackbirds  and  other  species 
within  plots  characterized  by  differing  man- 
agement regimes  (Herkert  1994,  McCoy  et  al. 
2001,  LaPointe  et  al.  2003);  however,  our 
study  is  the  only  one  we  know  of  in  which 
more  than  one  factor  varied  (i.e.,  perch  avail- 
ability and  grass  cover)  among  adjoining 
treatment  plots  within  the  same  sites.  These 
plots  varied  with  respect  to  densities  of  plant- 
ed trees,  which  blackbirds  used  as  perches, 
and  the  presence  or  absence  of  a planted  cover 
crop.  Typically,  managed  plots  in  other  song- 
bird studies  have  been  geographically  sepa- 


391 


392 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


rated  (Herkert  1994,  Swengel  1996,  McCoy  et 
al.  2001);  however,  our  plots  shared  common 
boundaries  to  allow  comparisons  of  habitat  se- 
lection without  the  confounding  effects  of  be- 
tween-site  variation. 

We  specifically  wished  to  determine  (1) 
how  the  availability  of  perches  and  vegetation 
determines  Red- winged  Blackbird  territory 
use  and  density  at  the  treatment  scale,  and  (2) 
how  within-treatment  vegetation  composition 
and  structure  in  territories  would  compare 
with  unused  (i.e.,  unoccupied  by  Red-winged 
Blackbirds)  areas.  We  were  particularly  inter- 
ested in  determining  the  importance  of  grass 
cover  and  density,  because  a dense,  short-stat- 
ure cover  crop  of  grass  (redtop,  Agrostis  gi- 
gantea)  planted  at  our  sites  had  suppressed  in- 
vading vegetation  but  was  unsuitable  for  nest- 
ing, whereas  the  common  invasive — Johnson- 
grass  ( Sorghum  halepense ),  which  was  also 
present — potentially  provided  a tall  nesting 
substrate  and  cover.  Because  blackbirds  in  up- 
land settings  prefer  dense,  tall  cover  (Albers 
1978,  Bollinger  1995),  we  predicted  that  den- 
sity of  blackbird  territories  would  be  greater 
in  treatments  not  planted  with  redtop.  Within 
treatments,  we  predicted  that  blackbird  terri- 
tories would  be  characterized  by  denser,  taller 
cover  than  unused  areas.  Based  on  previous 
studies  establishing  the  importance  of  perches 
(Joyner  1978,  Payne  et  al.  1998),  we  predicted 
that  densities  of  Red- winged  Blackbird  terri- 
tories would  be  greater  in  treatments  planted 
with  oaks,  and  that  territories  would  have 
perches  located  at  greater  heights  and  at  great- 
er densities  than  unused  areas. 

METHODS 

Study  site. — Our  research  was  conducted  in 
central  Missouri  at  two  sites  located  within  the 
Missouri  River  Floodplain.  Plowboy  Bend 
Conservation  Area  (38°  48' 5"  N,  92°  24'  17" 
W),  a landscape  dominated  by  row-crop  ag- 
riculture, is  located  west  of  the  Missouri  Riv- 
er’s main  channel  within  a levee-protected 
floodplain.  Smoky  Waters  Conservation  Area 
(38°  35'  9"  N,  91°  58'  3"  W)  is  located  72  km 
southeast  of  Plowboy  Bend,  between  the  Mis- 
souri River’s  main  channel  and  the  Osage  Riv- 
er. Smoky  Waters’  floodplain  has  not  been 
protected  since  a levee  was  breached  there  in 
the  1993  and  1995  floods;  thus,  it  is  subject 
to  occasional  flooding. 


Both  study  sites  encompassed  three  16.2- 
ha,  adjacent  experimental  treatments  (hereaf- 
ter, “blocks”)  that  differed  with  respect  to 
vegetation  treatments.  The  blocks — formerly 
row-cropped — were  established  in  1999  for  an 
ongoing  research  project  to  evaluate  the  res- 
toration of  hard  mast  (oak  acorn;  Dey  et  al. 
2003).  Oaks  were  planted  at  a density  of  1 19 
trees/ha  (Dey  et  al.  2003).  During  our  study, 
half  of  the  planted  oaks  were  >1.5  m high  and 
were  often  used  as  perches  by  Red-winged 
Blackbirds  (MAF  pers.  obs.).  Each  site  had 
three  treatment  blocks  with  varying  densities 
of  planted  and  natural  perches.  (1)  “Redtop” 
blocks,  seeded  with  a uniform  cover  of  redtop 
grass,  were  planted  with  saplings  of  swamp 
white  ( Quercus  bicolor ) and  pin  ( Q . palustris) 
oaks  distributed  in  planting  units  that  varied 
in  terms  of  planting  methods  but  had  a uni- 
form ground  cover  of  redtop  grass  (for  details, 
see  Dey  et  al.  2003).  The  redtop  grass  pro- 
duced a low,  dense  ground  cover  that  largely 
suppressed  invasion  by  other  herbaceous  and 
woody  vegetation  that  otherwise  may  have 
been  used  as  perches  or  nest  sites  by  Red- 
winged Blackbirds;  thus,  redtop  blocks  con- 
tained some  planted  oak  perches  but  few  or 
no  natural  perches.  (2)  “No  redtop”  blocks 
contained  the  same  configuration  of  oak  plant- 
ings described  above  for  redtop  blocks,  but 
they  were  not  seeded  with  a ground  cover; 
therefore,  over  time  they  contained  taller, 
denser  shrubs,  trees,  and  herbaceous  vegeta- 
tion and  more  “natural”  unplanted  perches 
than  redtop  blocks.  (3)  “Control”  blocks  con- 
tained only  natural  perches,  such  as  invading 
forbs  and  shrubs,  and  no  oak  plantings  or  any 
of  the  vegetation  treatments  listed  above. 

Delineation  of  breeding  territories. — We 
identified  breeding  territories  from  March  to 
May  in  2001  and  2002  by  monitoring  male 
Red-winged  Blackbirds  exhibiting  mating  be- 
haviors, such  as  the  “song  spread”  (Yasukawa 
and  Searcy  1995)  and  territory  defense.  To  de- 
lineate territories,  we  conducted  consecutive 
flushing  (Wiens  1969),  a technique  in  which 
males  are  approached  and  followed  until  they 
alight  on  the  perches  that  define  their  territory. 
Territories  were  delineated  by  identifying  and 
flagging  at  least  four  perches  used  consecu- 
tively by  each  male  (mean  number  of  perches 
flagged/territory  = 7.12  ± 1.97  SD). 

Vegetation  measurements. — Once  a breed- 


Furey  and  Burhans  • RED-WINGED  BLACKBIRD  TERRITORY  SELECTION 


393 


ing  territory  was  completely  flagged,  we  re- 
corded the  location,  species,  and  height  (m) 
for  each  perch.  We  established  two  1-m-wide 
belt  transects  in  each  territory  to  estimate  den- 
sity of  potential  perches  (no.  stems  >1.5  m 
tall/m2)  and  determined  average  maximum 
stem  height  (m).  To  establish  the  first  transect, 
the  center  of  the  territory  was  visually  located 
and  staked;  then  a random  azimuth  was  de- 
termined to  establish  the  direction  of  the  tran- 
sect across  the  territory.  The  second  transect 
location  was  established  perpendicular  to  the 
first.  Using  a 1-m  stick  held  horizontally  at 
1.5  m above  ground,  we  walked  the  territory 
end-to-end  along  each  transect,  recording  the 
number  of  stem  contacts  and  the  maximum 
vegetation  height  (m)  at  1-m  intervals.  Two 
vertical  density-board  measurements  were 
taken  at  random  locations  along  each  transect, 
resulting  in  four  individual  measurements  of 
vertical  vegetation  structure  for  each  breeding 
territory.  The  proportion  of  vertical  vegetation 
was  estimated  using  a 9-increment  density 
board  (2.25  m tall  X 0.25  m wide).  At  each 
0.25-m  increment,  we  estimated  the  proportion 
of  living  and  dead  vegetation  from  a distance 
of  15  m.  We  estimated  the  proportion  in  each 
increment  for  woody,  forb  (herbaceous),  and 
grass  vegetation  and  combined  them  to  gener- 
ate an  estimate  of  mean  total  proportion. 

We  randomly  located  unused  plots  (unoc- 
cupied by  Red-winged  Blackbirds)  by  using  a 
100-m  interval  grid  of  UTM  (Universal  Trans- 
mercator)  coordinates  placed  over  the  resto- 
ration sites  where  there  were  no  active  terri- 
tories. Sampling  of  vegetation  structure  was 
identical  to  that  conducted  within  blackbird 
territories,  with  the  exception  that  belt-transect 
length  within  a given  site  was  based  on  the 
average  belt-transect  length  of  all  breeding 
territories  found  at  the  site. 

Statistical  analyses. — For  each  year,  we  cal- 
culated territory  density  for  each  block  type 
(redtop,  no  redtop,  control)  by  summing  the 
numbers  of  territories  found  in  each  block 
type  and  dividing  by  16.2  ha.  If  a territory 
straddled  more  than  one  block  type,  we  placed 
it  in  the  block  type  in  which  the  majority  of 
its  area  occurred. 

We  averaged  vegetation  variables  for  the 
four  samples  taken  within  each  blackbird  ter- 
ritory. For  vertical  vegetation  measurements, 
the  mean  was  calculated  from  all  of  the  0.25- 


m increments  for  each  vegetation  type  of  in- 
terest (woody,  forb,  grass,  and  total  vertical 
vegetation).  Of  the  vertical  vegetation  mea- 
surements, we  included  only  mean  total  ver- 
tical cover,  mean  vertical  grass  cover,  and 
mean  grass  height,  which  was  defined  as  the 
last-recorded  increment  having  grass  cover  on 
the  vertical  density  board.  We  reasoned  that 
mean  total  vertical  cover  was  important  if 
blackbirds  were  assessing  territories  based  on 
cover  without  regard  to  vegetation  type.  We 
examined  grass  cover  and  height  because  of 
the  apparent  differences  in  grass  cover  be- 
tween redtop  blocks  and  the  other  block  types. 

We  also  used  the  vertical  vegetation  mea- 
surements to  create  a variable  called  “thresh- 
old nest-cover  height,”  defined  as  the  lowest 
height  at  which  mean  total  vertical  cover 
(based  on  the  density  board  samples)  was 
>60%.  The  latter  value  was  based  upon  a 
2001  sample  of  vegetation  measured  (using 
the  same  vertical  density  board  methodology 
described  above)  at  99  Red- winged  Blackbird 
nests.  At  the  99  nests,  we  determined  that  the 
mean  total  vertical  cover  at  nest  height 
(viewed  15  m from  the  board)  was  60%; 
therefore,  we  assumed  that  blackbirds  select 
nest  sites  with  at  least  60%  total  vertical  cover. 
Typically,  total  vertical  cover  approached 
100%  near  the  ground,  but  decreased  with  dis- 
tance above  ground;  thus,  a high  value  of 
threshold  nest-cover  height  (i.e.,  >60%)  usu- 
ally indicated  denser  cover  below  the  thresh- 
old height,  but  less  cover  above.  High  values 
of  threshold  nest-cover  height  do  not  indicate 
that  vertical  cover  was  denser;  rather,  they  in- 
dicate that  the  vertical  height  at  which  cover 
equaled  or  exceeded  60%  was  greater. 

We  used  a general  linear  model  (PROC 
MIXED;  SAS  Institute,  Inc.  2003)  to  test  for 
differences  in  territory  density  among  block 
types.  We  nested  block  within  site  as  a ran- 
dom effect  to  account  for  differences  in  site, 
and  included  “year”  in  the  model  to  account 
for  additional  variation.  We  used  the  likeli- 
hood ratio  test  to  test  the  overall  model 
against  a null  model  that  included  only  the 
intercept.  If  the  overall  model  was  significant, 
we  used  the  LSMEANS  statement  to  examine 
whether  territory  densities  varied  among  the 
three  block  types  (control,  no  redtop,  redtop); 
we  considered  differences  at  P < 0.05  to  be 
significant. 


394 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


We  analyzed  vegetation  differences  among 
block  types,  by  site,  using  PROC  MIXED 
models  as  above,  again  using  likelihood  ratio 
tests  to  compare  models  against  a null  model. 
Because  there  were  a large  number  of  vege- 
tation variables,  for  which  one  or  several  tests 
could  be  significant  by  chance,  we  used  the 
sequential  Bonferroni  method  to  interpret 
overall  model  significance  (Rice  1989).  Al- 
though the  sequential  Bonferroni  test  has  been 
criticized  as  overly  conservative  in  circum- 
stances where  numerous  individual  tests  show 
moderately  significant  results  (Moran  2003), 
in  this  circumstance  we  feel  that  it  was  a suit- 
able compromise  between  having  no  control 
for  type  I error  and  the  simple  Bonferroni  test, 
which  is  even  more  conservative  (Rice  1989). 
If  the  overall  model  was  significant,  we  used 
the  LSMEANS  statement  to  determine  wheth- 
er territory  area  and  vegetation  variables  var- 
ied among  the  three  block  types  (control,  no 
redtop,  redtop),  by  site;  within  each  model,  we 
considered  differences  at  adjusted  P < 0.05  to 
be  significant. 

We  also  compared  parameters  of  vegetation 
structure  between  areas  occupied  (“territo- 
ries”) and  unoccupied  (“unused”  plots)  by 
Red- winged  Blackbirds  to  describe  local  veg- 
etation differences  affecting  blackbird  habitat 
selection  within  blocks.  Because  flooding 
events  in  2001  prevented  us  from  sampling 
unused  plots  at  both  sites,  only  2002  field  data 
were  used  for  this  analysis,  and  we  removed 
territory  and  unused  samples  entirely  if  any 
data  values  were  missing.  We  used  a general 
linear  model  (PROC  MIXED;  SAS  Institute, 
Inc.  2003)  with  an  LSMEANS  statement  to 
calculate  means  and  standard  errors  for  each 
variable  of  interest.  We  determined  that  there 
were  differences  among  territories  and  unused 
plots  if  likelihood  ratio  tests  indicated  overall 
model  significance,  based  on  sequential  Bon- 
ferroni adjustments  for  the  six  vegetation  var- 
iables analyzed.  If  the  overall  model  was  sig- 
nificant, we  evaluated  multiple  comparisons 
among  different  combinations  of  block,  terri- 
tory, and  unused  plots  (15  comparisons  per 
model)  with  sequential  Bonferroni  tests  to 
control  for  type  I error. 

RESULTS 

We  analyzed  81  Red-winged  Blackbird 
breeding  territories  across  both  sites  and 


years.  Mean  breeding  territory  area  in  2001 
was  1,667  ± 195  m2  (n  = 19),  1,897  ± 221 
m2  ( n = 17),  and  2,310  ± 464  m2  ( n = 10) 
in  redtop,  no  redtop,  and  control  blocks,  re- 
spectively, and  in  2002  it  was  1,648  ± 173 
m2  (n  = 14),  1,808  ± 269  m2  ( n = 17),  and 
771  ±83  m2  ( n = 4).  We  found  no  differences 
in  territory  area  by  block  type  (likelihood  ratio 
test:  x2  = 2.3,  df  = 3,  P = 0.51).  In  2001, 
mean  territory  density  across  both  sites  was 
0.71  ± 0.74,  0.67  ± 0.26,  and  0.31  ± 0.26 
territories/ha  in  redtop,  no  redtop,  and  control 
blocks,  respectively.  In  2002,  mean  territory 
density  across  both  sites  was  0.46  ± 0.66, 
0.56  ± 0.17,  and  0.12  ± 0.17  territories/ha; 
there  were  no  blackbird  territories  in  redtop  or 
control  blocks  at  Plowboy  Bend  during  this 
year.  Territory  density  did  not  differ  among 
blocks  or  years  (likelihood  ratio  test:  x2  — 5.8, 
df  = 3,  P = 0.12). 

We  did  not  find  differences  among  the  three 
block  types  for  mean  perch  density,  mean  total 
vertical  cover,  or  mean  vertical  grass  cover 
(Fig.  1A,  C,  E).  The  model  for  mean  perch 
height  differed  significantly  from  the  null 
model  (x2  = 39.0,  df  = 3,  adj.  P < 0.001), 
but  the  differences  were  among  years  (2001: 
2.16  ± 0.03  m;  2002:  1.82  ± 0.04  m;  t = 
6.73,  df  = 74,  P < 0.001);  there  were  no  dif- 
ferences in  perch  height  among  blocks  (Fig. 
IB).  Similarly,  models  for  mean  threshold 
nest-cover  height  and  grass  height  differed 
from  null  models,  but  again  differences  were 
among  years  rather  than  blocks  (threshold 
nest-cover  height  model:  overall  x2  = 17.0,  df 
= 3,  adj.  P < 0.01;  mean  grass  height  model: 
overall  x2  — 28.6,  df  = 2,  adj.  P < 0.008;  Fig. 
ID,  F).  Mean  grass  height  across  all  territory 
blocks  was  greater  in  2001  (2001:  0.53  ± 0.02 
m;  2002:  0.36  ± 0.02  m;  t = 5.53,  df  = 74, 
P < 0.001),  whereas  mean  threshold  nest-cov- 
er  height  was  shorter  in  2001  (2001:  0.40  ± 
0.06  m;  2002:  0.63  ± 0.06  m;  t = -4.30,  df 
= 74,  P < 0.001). 

We  used  samples  from  35  Red-winged 
Blackbird  breeding  territories  and  35  unused 
plots  (2002  data  only)  to  compare  vegetation 
in  breeding  territories  with  that  in  unused 
plots  ( n = 10,  13,  and  12  unused  plots  sam- 
pled from  both  sites  combined  in  redtop,  no 
redtop,  and  control  blocks,  respectively). 
Models  testing  for  differences  between  terri- 
tories and  unused  plots  did  not  differ  from 


Furey  and  Burhans  • RED-WINGED  BLACKBIRD  TERRITORY  SELECTION 


395 


FIG.  1.  Vegetation  cover  (expressed  as  a proportion),  height,  and  perch  density  comparisons  (±SE)  among 
treatment  blocks  at  Plowboy  Bend  and  Smoky  Waters  Conservation  Areas,  Missouri,  2001-2002. 


null  models  with  respect  to  perch  height  (Fig. 
2B),  mean  total  vertical  cover  (Fig.  2C),  or 
threshold  nest-cover  height  (Fig.  2D).  Overall, 
mean  perch  density  varied  among  combina- 
tions of  block  and  territory  or  unused  plots  (x2 
= 28.5,  df  = 4,  adj.  P < 0.008;  Fig.  2A). 
Territories  in  control  blocks  had  greater  perch 
densities  than  in  all  other  block  types,  al- 
though there  were  only  four  control  territories 
in  the  analysis  (all  adj.  P < 0.005;  Fig.  2A). 
Perch  densities  did  not  differ  between  other 
combinations  of  block  and  territory  or  unused 
plots,  except  that  perch  densities  were  greater 
in  no  redtop  territories  than  they  were  in  red- 
top  territories  and  redtop  unused  plots  (no  red- 
top  territories  versus  redtop  territories:  t = 
3.42,  df  = 61,  adj.  P < 0.005;  no  redtop  ter- 
ritories versus  redtop  unused  plots:  t = 3.01, 
df  = 61,  adj.  P < 0.006). 


Overall,  mean  vertical  grass  cover  varied 
among  combinations  of  block  and  territory  or 
unused  plots  (x2  = 21.5,  df  = 5,  adj.  P < 
0.01).  Grass  cover  was  greater  in  no  redtop 
territories  compared  with  no  redtop  unused 
plots,  control  unused  plots,  and  redtop  terri- 
tories and  unused  plots  (all  adj.  P ^ 0.004; 
Fig.  2E).  Grass  height  varied  overall  among 
combinations  of  block  and  territory  or  unused 
plots  (x2  = 15.4,  df  = 5,  adj.  P < 0.01).  Grass 
height  was  greater  in  no  redtop  territories  than 
in  redtop,  no  redtop,  and  control  unused  plots 
(all  adj.  P < 0.004;  Fig.  2F). 

DISCUSSION 

We  found  no  significant  differences  in  ter- 
ritory density  or  area  among  treatment  blocks, 
nor  did  we  find  differences  among  vegetation 
variables  by  territory  treatment  block  when 


396 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Control  No  redtop  Redtop 


2.5 


Control  No  redtop  Redtop  Control  No  redtop  Redtop 

FIG.  2.  Vegetation  cover  (expressed  as  a proportion),  height,  and  perch  density  comparisons  (±SE)  of  Red- 
winged Blackbird  territories  (used)  and  unused  plots  at  Plowboy  Bend  and  Smoky  Waters  Conservation  Areas, 
Missouri,  2002. 


2001  and  2002  data  were  combined.  We  did 
find  differences,  however,  between  territories 
and  unused  plots;  generally,  blackbird  terri- 
tories were  characterized  by  denser  or  taller 
grass  cover  than  unused  plots,  and  territories 
in  control  and  no  redtop  blocks  tended  to  con- 
tain more  perches  than  unused  plots. 

In  the  analysis  of  territories  versus  unused 
plots,  the  greater  perch  density  in  territory 
blocks  with  no  cover  crop  (no  redtop  and  con- 
trol blocks)  compared  with  those  that  had  a 
cover  crop  (redtop)  may  be  a reflection  of  red- 
top’s  ability  to  suppress  invasion  by  trees  and 
shrubs.  However,  perch  density  did  not  differ 
among  territory  blocks  or  years  when  data 
from  both  years  were  combined  (Fig.  1A), 
whereas  the  territory/unused  analysis,  which 
included  only  2002  data,  revealed  extreme  dif- 
ferences in  perch  density  among  territory 
blocks  (Fig.  2A).  In  the  case  of  control  terri- 


tories, perch  density  could  have  been  an  arti- 
fact of  small  sample  size,  as  there  were  only 
4 territories  in  control  blocks  in  2002  com- 
pared to  10  in  2001.  However,  upon  visual 
inspection,  we  detected  similar  between-year 
differences  in  mean  perch  density  in  redtop 
blocks  (Fig.  1A  versus  2A),  and  in  this  case 
sample  sizes  were  19  and  14  in  redtop  terri- 
tories in  2001  and  2002,  respectively.  Such 
inter-annual  inconsistencies  in  bird-vegetation 
relationships  are  common  and  often  prevent 
researchers  from  reaching  direct  conclusions 
in  studies  of  avian-habitat  associations  (Riffell 
et  al.  2001),  including  studies  of  Red-winged 
Blackbirds  (Erckmann  et  al.  1990)  and  other 
blackbirds  (Orians  and  Wittenberger  1991). 
Red- winged  Blackbirds  may  require  only  a 
few  perches  for  territory  defense.  We  noted 
that  blackbirds  typically  reused  the  same 
perches,  sometimes  frequenting  only  four  or 


Furey  and  Burhans  • RED- WINGED  BLACKBIRD  TERRITORY  SELECTION 


397 


five  perches  repeatedly  (MAP  pers.  obs.).  It 
may  be  that  perch  availability  limits  blackbird 
territory  settlement  only  at  some  lower  thresh- 
old, in  which  case  even  territories  with  very 
low  perch  densities  at  our  sites  (e.g.,  redtop; 
Fig.  2A)  may  have  met  this  requirement. 
Perches  have  been  shown  not  to  limit  habitat 
use  by  some  songbirds  (Vickery  and  Hunter 
1995),  but  at  least  one  study  suggests  that  they 
are  necessary  for  Red- winged  Blackbirds; 
Joyner  (1978)  found  that  even  in  areas  with  a 
preferred  grass  cover  type,  blackbirds  did  not 
establish  territories  if  fence  posts — used  as 
perches — were  totally  lacking. 

In  addition  to  variation  in  perch  density,  we 
also  found  differences  in  grass  cover  and 
height  between  territories  and  unused  plots 
within  and  among  treatment  blocks.  Variable 
grass  cover,  at  least  within  no  redtop  blocks, 
suggests  that  blackbirds  may  have  settled  in  a 
non-uniform  fashion  with  regard  to  grass 
patches.  Although  our  data  did  not  permit  us 
to  relate  territories  to  grass  patchiness  spatial- 
ly, overall  we  did  not  notice  obvious  patterns 
in  territory  settlement;  there  were  two  possible 
exceptions;  (1)  the  only  two  blackbird  terri- 
tories in  the  Plowboy  Bend  redtop  block  were 
very  close  to  blackbird  territories  on  the  ad- 
joining no  redtop  block,  from  which  forbs, 
shrubs,  and  Johnsongrass  had  spread  into  the 
redtop  block  (MAF  pers.  obs.);  and  (2)  black- 
birds tended  to  avoid  settlement  along  one 
edge  of  the  Smoky  Water  control  block  (MAF 
pers.  obs.).  In  the  second  case,  we  are  not  sure 
why  blackbirds  avoided  the  block  edge,  but 
we  believe  that  settlement  in  redtop  at  Plow- 
boy  Bend  may  have  been  influenced  both  by 
the  rampant  growth  of  Johnsongrass  and  by 
redtop’s  ability  to  suppress  Johnsongrass  and 
other  vegetation.  Redtop  cover  was  particu- 
larly uniform  at  Plowboy  Bend,  where  black- 
bird use  of  the  redtop  block  was  minimal, 
whereas  the  redtop  block  at  Smoky  Waters  un- 
derwent extensive  invasion  of  shrubs  and 
forbs  (MAF  and  DEB  pers.  obs.).  Johnson- 
grass, a dense,  stout-stemmed  grass  that  grows 
to  1.8  m high,  was  also  used  as  a nesting  sub- 
strate, whereas  redtop  was  not.  Of  more  than 
250  Red- winged  Blackbird  nests  found  from 
2001-2003,  none  were  anchored  in  redtop 
grass,  whereas  Johnsongrass  was  among  the 
five  most  commonly  used  nest  substrates 
(DEB  unpubl.  data). 


The  pattern  of  denser  and  taller  grass  cover 
in  territories,  especially  in  no  redtop  blocks, 
generally  agrees  with  other  findings  in  studies 
of  Red-winged  Blackbirds.  Bollinger  (1995) 
believed  that  blackbirds  occupied  his  upland 
sites  due  to  the  availability  of  suitable  nest 
cover  and  vegetation  with  stems  strong 
enough  to  support  their  nests;  results  of  other 
studies  also  indicate  that,  where  stout  plants 
are  available,  blackbirds  choose  them  as  nest 
sites  or  for  territorial  activity  (Albers  1978, 
Joyner  1978,  Turner  and  McCarthy  1998,  Ko- 
bal  et  al.  1999).  Bollinger  (1995)  found  a pos- 
itive relationship  between  presence  of  grass 
and  blackbirds,  and  Camp  and  Best  (1994) 
found  a positive  relationship  between  grass 
cover  and  nest  densities.  Other  studies  have 
shown  that  Red-winged  Blackbirds  favor 
dense  vegetation  (LaPointe  et  al.  2003);  Al- 
bers (1978)  found  that  blackbird  territories 
had  significantly  taller,  denser  vegetation  than 
unused  areas,  and  Bollinger  (1995)  found  that 
Red- winged  Blackbirds  were  most  abundant 
in  fields  with  dense  cover.  However,  in  a sur- 
vey of  Illinois  grassland  species,  Herkert 
( 1 994)  found  no  correlates  of  vegetation  struc- 
ture and  occupancy  by  Red- winged  Black- 
birds, which  were  present  on  93%  of  his  tran- 
sects, and  Scott  et  al.  (2002)  found  that  black- 
birds were  negatively  associated  with  grass 
cover  on  reclaimed  surface  mines  in  Indiana. 

Although  our  2002  data  revealed  differenc- 
es in  perch  density  when  comparing  territories 
with  unused  plots,  our  results  suggest  that 
perch  density  does  not  influence  Red-winged 
Blackbird  territory  selection  as  long  as  perch 
density  is  above  some  lower  limit.  However, 
particularly  in  no  redtop  blocks,  blackbirds 
tended  to  choose  territories  that  had  denser, 
taller  grass  cover  than  that  observed  in  unused 
plots.  This  finding  is  in  agreement  with  other 
studies,  which  have  shown  that  Red-winged 
Blackbirds  appear  to  favor  dense  vegetation 
(Albers  1978,  Kobal  et  al.  1999),  including 
tall  or  dense  grass  cover  (Camp  and  Best 
1994,  Bollinger  1995). 

ACKNOWLEDGMENTS 

This  work  was  funded  through  the  University  of 
Missouri’s  Center  for  Agroforestry  under  cooperative 
agreements  58-6227-1-004  with  the  Agricultural  Re- 
search Service  and  C R 826704-01-2  with  the  U.S. 
Environmental  Protection  Agency  (EPA).  The  results 


398 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


presented  are  the  sole  responsibility  of  the  Principal 
Investigators  and/or  University  of  Missouri  and  may 
not  represent  the  policies  or  positions  of  the  EPA.  Any 
opinions,  findings,  conclusions,  or  recommendations 
expressed  in  this  publication  are  those  of  the  author(s) 
and  do  not  necessarily  reflect  the  view  of  the  U.S. 
Department  of  Agriculture.  Equipment  and  computer 
resources  support  were  provided  courtesy  of  Univer- 
sity of  Missouri-Columbia’s  Quantitative  Silviculture 
Laboratory  and  the  U.S.  Department  of  Agriculture 
Forest  Service  North  Central  Research  Station.  We 
thank  F.  R.  Thompson  for  statistical  advice  and  M.  A. 
Gold  and  D.  C.  Dey  for  support  and  professional  con- 
siderations. Comments  by  G.  H.  Orians,  K.  Yasukawa, 
B.  B.  Steele,  P.  D.  Vickery,  and  an  anonymous  review- 
er greatly  improved  earlier  versions  of  the  manuscript. 
We  gratefully  thank  field  assistants  H.  Morris,  H. 
Heitz,  A.  Schultz,  and  T.  Altnether  for  data  collection. 

LITERATURE  CITED 

Albers,  P.  H.  1978.  Habitat  selection  by  breeding  Red- 
winged Blackbirds.  Wilson  Bulletin  90:619-634. 
Bollinger,  E.  K.  1995.  Successional  changes  and  hab- 
itat selection  in  hayfield  bird  communities.  Auk 
112:720-730. 

Burhans,  D.  E.  1997.  Habitat  and  microhabitat  fea- 
tures associated  with  cowbird  parasitism  in  two 
forest  edge  cowbird  hosts.  Condor  99:866-872. 
Camp,  M.  and  L.  B.  Best.  1994.  Nest  density  and 
nesting  success  of  birds  in  roadsides  adjacent  to 
rowcrop  fields.  American  Midland  Naturalist  131: 
347-358. 

Dey,  D.  C.,  J.  M.  Kabrick,  and  M.  Gold.  2003.  Tree 
establishment  in  floodplain  agroforestry  practices. 
Pages  102-115  in  Proceedings  of  the  8th  North 
American  agroforestry  conference.  Oregon  State 
University,  Corvallis. 

Erckmann,  W.  J.,  L.  D.  Beletsky,  G.  H.  Orians,  T. 
Johnson,  S.  Sharbaugh,  and  C.  D’ Antonio. 
1990.  Old  nests  as  cues  for  nest-site  selection:  an 
experimental  test  with  Red-winged  Blackbirds. 
Condor  92:113-117. 

Herkert,  J.  R.  1994.  The  effects  of  habitat  fragmen- 
tation on  Midwestern  grassland  bird  communities. 
Ecological  Applications  4:461-471. 

Joyner,  D.  E.  1978.  Use  of  an  old-field  habitat  by 
Bobolinks  and  Red-winged  Blackbirds.  Canadian 
Field-Naturalist  92:383-387. 

Kobal,  S.  N.,  N.  F.  Payne,  and  D.  R.  Ludwig.  1999. 
Habitat/area  relationships,  abundance,  and  com- 
position of  bird  communities  in  3 grassland  types. 
Transactions  of  the  Illinois  State  Academy  of  Sci- 
ence 92:109-131. 

Lapointe,  S.,  L.  Belanger,  J.  Giroux,  and  B.  Filion. 
2003.  Effects  of  plant  cover  for  nesting  ducks  on 
grassland  songbirds.  Canadian  Field-Naturalist 
1 17:167-172. 

McCoy,  T.  D.,  M.  R.  Ryan,  and  L.  W.  Burger,  Jr. 
2001.  Grassland  bird  conservation:  CPI  vs.  CP2 
plantings  in  Conservation  Reserve  Program  fields 


in  Missouri.  American  Midland  Naturalist  145:1- 
17. 

Moran,  M.  D.  2003.  Arguments  for  rejecting  the  se- 
quential Bonferroni  in  ecological  studies.  Oikos 
100:403-405. 

Murkin,  H.  R.,  E.  J.  Murkin,  and  J.  P.  Ball.  1997. 
Avian  habitat  selection  and  prairie  wetland  dy- 
namics: a 10-year  experiment.  Ecological  Appli- 
cations 7:1144-1159. 

Orians,  G.  H.  and  J.  F.  Wittenberger.  1991.  Spatial 
and  temporal  scales  in  habitat  selection.  American 
Naturalist  137:S29-S49. 

Payne,  N.  E,  S.  N.  Kobal,  and  D.  R.  Ludwig.  1998. 
Perch  use  by  7 grassland  bird  species  in  northern 
Illinois.  Transactions  of  the  Illinois  State  Acade- 
my of  Science  91:77-83. 

Pribil,  S.  and  J.  Picman.  1997.  The  importance  of  us- 
ing the  proper  methodology  and  spatial  scale  in 
the  study  of  habitat  selection  by  birds.  Canadian 
Journal  of  Zoology  75:1835-1844. 

Rice,  W.  R.  1989.  Analyzing  tables  of  statistical  tests. 
Evolution  43:223-225. 

Riffell,  S.  K.,  B.  E.  Keas,  and  T.  M.  Burton.  2001. 
Area  and  habitat  relationships  of  birds  in  Great 
Lakes  coastal  wet  meadows.  Wetlands  21:492- 
507. 

Rotenberry,  J.  T.  and  J.  A.  Wiens.  1980.  Habitat 
structure,  patchiness,  and  avian  communities  in 
North  American  steppe  vegetation:  a multivariate 
analysis.  Ecology  61:1228-1250. 

SAS  Institute,  Inc.  2003.  Version  9.1  for  Windows. 
SAS  Institute,  Inc.,  Cary,  North  Carolina. 

Scott,  P.  E.,  T.  L.  DeVault,  R.  A.  Bajema,  and  S.  L. 
Lima.  2002.  Grassland  vegetation  and  bird  abun- 
dances on  reclaimed  Midwestern  stripmines. 
Wildlife  Society  Bulletin  30:1006-1014. 

Shochat,  E.,  M.  A.  Patten,  D.  W.  Morris,  D.  L.  Re- 
inking, D.  H.  Wolfe,  and  S.  K.  Sherrod.  2005. 
Ecological  traps  in  isodars:  effects  of  tallgrass 
prairie  management  on  bird  nest  success.  Oikos 
111:159-169. 

Swengel,  S.  R.  1996.  Management  responses  of  three 
species  of  declining  sparrows  in  tallgrass  prairie. 
Bird  Conservation  International  6:241-253. 

Turner,  A.  M.  and  J.  P.  McCarty.  1998.  Resource 
availability,  breeding  site  selection,  and  reproduc- 
tive success  of  Red-winged  Blackbirds.  Oecologia 
113:140-146. 

Vickery,  P.  D.  and  M.  L.  Hunter.  1995.  Do  artificial 
song-perches  affect  habitat  use  by  grassland  birds 
in  Maine?  American  Midland  Naturalist  133:164- 
169. 

Wiens,  J.  A.  1969.  An  approach  to  the  study  of  eco- 
logical relationships  among  grassland  birds.  Or- 
nithological Monographs,  no.  8. 

Wood,  D.  R.,  L.  W.  Burger,  Jr.,  J.  L.  Bowman,  and 
C.  L.  Hardy.  2004.  Avian  community  response  to 
pine-grassland  restoration.  Wildlife  Society  Bul- 
letin 32:819-828. 

Yasukawa,  K.  and  W.  A.  Searcy.  1995.  Red- winged 
Blackbird  ( Agelaius  phoeniceus).  The  Birds  of 
North  America,  no.  184. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):399-410,  2006 


THE  USE  OF  SOUTHERN  APPALACHIAN  WETLANDS  BY 
BREEDING  BIRDS,  WITH  A FOCUS  ON  NEOTROPICAL 
MIGRATORY  SPECIES 

JASON  F.  BULLUCK1-2 3-4  AND  MATTHEW  P.  ROWE1 3 


ABSTRACT. — Although  loss  of  wetlands  in  southern  Appalachia  has  been  especially  severe,  no  avian  studies 
have  been  conducted  in  the  vestiges  of  these  ecosystems.  Our  research  assessed  avian  use  of  southern  Appala- 
chian wetlands  in  the  breeding  seasons  of  1999  through  2001.  Site  analyses  included  18  habitat  variables, 
including  total  wetland  area,  area  of  open  water,  beaver  or  livestock  evidence,  edge  type  (abrupt  or  gradual), 
and  percent  cover  of  nine  vegetation  types.  We  analyzed  avian  species  richness  and  abundance  at  the  community 
level  and  in  guilds  based  on  migratory  status  and  breeding  habitat  preference.  Measures  of  vegetation  and 
habitat — particularly  those  resulting  from  beaver  activities — and  gradual  edges  were  significantly  correlated  with 
guild-  and  community-level  variables.  Evidence  of  beaver  (i.e.,  forest  gaps  where  trees  had  been  felled,  ponds 
where  drainages  had  been  dammed;  hereafter  referred  to  simply  as  “beaver  evidence”)  was  significantly  cor- 
related with  greater  community-level  species  richness  and  abundance.  Both  beaver  evidence  and  gradual  edge 
were  positively  associated  with  greater  species  richness  and  abundance  of  Neotropical  migratory  birds  (NTMBs) 
overall,  as  well  as  with  the  late-successional  NTMB  guild.  Presence  of  gradual  edge  alone  also  was  significantly 
correlated  with  high  abundance  of  birds  in  the  early-successional  NTMB  guild.  Beaver  and  gradual  edge  may 
have  contributed  to  higher-quality  breeding  habitats  with  relatively  greater  overall  productivity  and  structural 
complexity  in  some  wetlands.  Received  24  November  2004,  accepted  22  March  2006. 


Wetlands  of  the  southern  Appalachians  are 
perhaps  the  rarest  and  most  threatened  in  the 
southeastern  U.S.  Weakley  and  Shafale  (1994) 
estimate  that  only  one-sixth  (about  2,000  ha) 
of  the  bogs  in  pre-European  settlement  south- 
ern Appalachia  remain  today.  Historically, 
post-glacial  southern  Appalachian  wetlands 
have  been  maintained  by  precipitation, 
groundwater  recharge,  and  natural  suppression 
of  woody  vegetation  (Weakley  and  Shafale 

1994,  Lee  and  Norden  1996);  humans,  how- 
ever, have  since  altered  the  woody  vegetation. 
Pleistocene  megafauna  (Weigl  and  Knowles 

1995,  Lee  and  Norden  1996),  including  elk 
( Cervus  elaphus)  and  American  bison  ( Bison 
bison ; Lee  and  Norden  1996,  but  see  Ward 
1990)  are  believed  to  have  maintained  these 
wetlands  in  early-successional  states  via 
browsing,  but  all  have  disappeared  concomi- 
tant with  human  settlement.  Native  American 
use  of  fire  also  may  have  suppressed  the  en- 
croachment of  woody  vegetation  (Lee  and 


1 Dept,  of  Biology,  Appalachian  State  Univ.,  Boone, 
NC  28608,  USA. 

2 Current  address:  ARCADIS,  114  Lovell  Rd., 
Knoxville,  TN  37934,  USA. 

3 Current  address:  Dept,  of  Biological  Sciences,  Sam 
Houston  State  Univ.,  Huntsville,  TX  77341,  USA. 

4 Corresponding  author;  e-mail: 
jbulluck@gmail.com 


Norden  1996)  into  southern  Appalachian  wet- 
lands (Delcourt  and  Delcourt  1997).  Today, 
fires  are  suppressed  and  quickly  extinguished 
when  they  do  occur  (Weakley  and  Shafale 
1994).  Widespread  loss  of  beaver  ( Castor 
canadensis ) via  the  fur  trade  of  the  18th  and 
19th  centuries  also  reduced  the  development 
(Snodgrass  1997)  and  maintenance  of  wet- 
lands throughout  the  landscape  (Webster  et  al. 
1975,  Naiman  et  al.  1988,  Weakley  and  Shaf- 
ale 1994,  Lawton  and  Jones  1995,  Lee  and 
Norden  1996).  Most  recently,  the  majority  of 
remaining  small  wetlands  in  southern  Appa- 
lachia have  been  converted  to  pasture,  devel- 
oped, or  manipulated  for  other  human  uses 
(Weakley  and  Shafale  1994). 

Today,  the  remaining  wetlands  of  southern 
Appalachia  are  considered  biological  hotspots 
(Murdock  1994);  until  now,  however,  no  study 
had  focused  on  the  breeding  avifauna  of  these 
ecosystems.  Southern  Appalachia’s  wetlands 
are  important  to  breeding  Neotropical  migra- 
tory birds  (NTMBs).  In  fact,  parts  of  the  re- 
gion harbor  the  greatest  species  richness  and 
abundance  of  NTMBs  in  North  America  (Si- 
mons et  al.  2000);  however,  the  region’s  pop- 
ulations of  NTMBs  are  declining  more  rapidly 
than  anywhere  else  in  North  America  (Rod- 
riguez 2002).  Species  preferring  open,  early- 
successional  habitats  or  late-successional  for- 


399 


400 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


ests  are  undergoing  the  most  rapid  declines 
(Robbins  et  al.  1989;  reviewed  in  Askins  et 
al.  1990).  Of  NTMBs  that  breed  in  southern 
Appalachia’s  early-successional  habitats,  76% 
are  declining  (Hunter  et  al.  2001,  Thompson 
and  DeGraaf  2001)  due  to  losses  of  early  suc- 
cessional  grasslands,  scrub-shrub,  open-cano- 
py woodlands,  and  small  canopy  gaps  (Hunter 
et  al.  2001);  some  of  North  America’s  greatest 
declines  in  early-successional  species  have 
been  reported  from  southern  Appalachia 
(Franzreb  and  Rosenberg  1997). 

Declines  among  NTMBs  that  breed  in  late- 
successional  habitats  are  due,  in  part,  to  forest 
fragmentation  resulting  from  agricultural,  res- 
idential, and  commercial  development  (Rob- 
bins et  al.  1989,  Askins  et  al.  1990,  Faaborg 
et  al.  1995).  Forest-interior  species  suffer  from 
increased  rates  of  brood  parasitism  (Britting- 
ham  and  Temple  1983,  Robbins  et  al.  1989) 
and  nest  predation  (Askins  et  al.  1990),  and 
from  increased  competition  with  other  bird 
species  (Askins  et  al.  1990,  Zannette  et  al. 
2000)  in  the  smaller  habitat  patches  that  result 
from  fragmentation.  Although  the  southern 
Appalachians  contain  approximately  80%  of 
the  primary  forests  in  the  eastern  U.S.  (Davis 
1993),  more  species  are  declining  in  the  re- 
gion (42%  of  forest-breeding  species)  than  in 
North  America  as  a whole  (27%;  Franzreb  and 
Rosenberg  1997). 

In  southern  Appalachia,  wetland  loss  has 
been  concurrent  with  declines  in  NTMB  pop- 
ulations, although  it  has  not  been  evaluated  as 
a contributing  factor  (Hunter  et  al.  1999).  In 
southern  Appalachian  wetlands,  habitat  suc- 
cession ranges  from  open,  early-successional 
grasslands  to  late-successional,  forested  bogs; 
thus,  these  wetlands  may  provide  important 
breeding  habitats  for  both  early-  and  late-suc- 
cessional breeding  species,  some  of  which  are 
undergoing  the  greatest  rates  of  population  de- 
cline. 

Considering  the  general  scarcity  of  southern 
Appalachian  wetlands  and  the  disproportion- 
ately high  rates  of  decline  among  NTMB  spe- 
cies in  that  region,  research  on  the  use  of 
southern  Appalachian  wetlands  by  breeding 
birds  is  overdue.  Herein,  we  report  the  results 
of  such  research,  focusing  specifically  on  the 
habitat  characteristics  that  make  certain  kinds 
of  wetlands  attractive  to  NTMBs  in  early-  and 


late-successional  habitat  guilds  of  breeding 
birds. 

METHODS 

Study  sites. — We  collected  data  at  57  south- 
ern Appalachian  wetlands  in  western  North 
Carolina  (n  = 44),  northeastern  Tennessee  ( n 
= 3),  and  southwestern  Virginia  ( n = 10). 
Wetland  elevations  ranged  from  442  to  1,254 
m.  The  total  wetland  area  in  our  study  was 
795  ha.  Individual  wetland  area  ranged  from 
0.40-95  ha  (mean  = 14  ha);  excluding  the 
four  largest  wetlands,  however,  mean  wetland 
size  was  only  0.64  ha.  Such  small  wetland  ar- 
eas are  typical  in  regions  of  high  topographic 
relief. 

All  wetland  sites  were  dominated  by  hydro- 
phytic  vegetation  and  other  hydrologic  fea- 
tures (i.e.,  hydric  soils,  periodic  to  permanent 
inundation  and/or  soil  saturation).  Forty-four 
of  our  sites  were  used  in  previous  botanical 
and  herpetofaunal  studies;  we  located  the  oth- 
ers by  using  natural  history  records  from  the 
North  Carolina  Natural  Heritage  Program  and 
the  North  Carolina  Museum  of  Natural  Sci- 
ences. All  wetlands  were  classified  as  one  of 
three  palustrine  system  types  (Cowardin  et  al. 
1979):  emergent  {n  = 23),  scrub-shrub  ( n — 
21),  or  forested  ( n = 13). 

Some  of  our  study  wetlands  were  low-pH, 
precipitation-fed  bogs,  wherein  peat-filled  de- 
pressions were  dominated  by  a lattice  of 
sphagnum  mats  and  standing  water.  In  these 
open  wetlands,  woody  vegetation  was  scarce, 
although  some  had  a sparse  shrub  layer  (e.g., 
Salix  spp.,  Alnus  spp.,  and  Acer  rubrum  sap- 
lings). Other  study  wetlands  were  groundwa- 
ter-sourced fens  characterized  by  thick  covers 
of  mosses,  lichens,  grasses,  and  forbs.  Most 
study  wetlands  were  located  in  floodplains 
and  characterized  by  a diverse,  structurally 
complex  vegetative  community.  These  flood- 
plain  wetlands  were  often  the  result  of  historic 
or  current  beaver  activity  and  may  have  been 
groundwater  and/or  surface-water  fed,  though 
detailed  hydrologic  characteristics  of  study 
sites  were  not  addressed. 

All  wetlands  were  owned  by  Appalachian 
State  University  (ASU;  n = 2),  the  Blue  Ridge 
Parkway  National  Park  (BRP;  n = 22),  The 
Nature  Conservancy  (TNC;  n = 6),  the  North 
Carolina  Department  of  Transportation 
(NCDOT;  n = 2),  the  U.S.  Department  of  Ag- 


Bulluck  and  Rowe  • NTMB  USE  OF  SOUTHERN  APPALACHIAN  WETLANDS 


401 


riculture  Forest  Service  (USFS;  n = 3),  or  pri- 
vate landowners  ( n = 22).  (Hereafter,  all  sites 
other  than  those  owned  by  private  landowners 
will  be  referred  to  as  “publicly  owned  sites,” 
including  TNC  sites,  although  we  recognize 
that  technically,  TNC  sites  are  “private.”) 

In  general,  publicly  owned  wetlands  are  ac- 
tively managed,  whereas  privately  owned  sites 
are  not.  Publicly  owned  wetlands  were  char- 
acterized by  fewer  land-use  disturbances  than 
those  that  were  privately  owned,  and  they 
were  managed  for  their  persistence  in  the 
landscape.  Privately  owned  sites  generally 
displayed  one  or  more  effects  of  land  use, 
such  as  logging,  grazing,  and  mowing,  or 
draining  for  agriculture,  residential  develop- 
ment, and/or  commercial  development. 

Small  southern  Appalachian  wetlands  are 
inherently  associated  with  edges,  and  we  clas- 
sified site  edges  as  either  abrupt  or  gradual. 
Our  qualitative  classification  of  edge  type  fol- 
lowed that  used  in  other  studies  of  edge-type 
effects  on  breeding  birds  (Suarez  et  al.  1997, 
Luck  et  al.  1999).  An  abrupt  edge  displayed 
a distinct,  drastic  change  in  vegetation  struc- 
ture between  two  vegetation  types.  Abrupt 
edges  ( n — 29  sites)  usually  resulted  from  per- 
sistent land  uses,  such  as  mowing  or  cattle 
grazing,  thus  creating  a sharp  edge  between 
grasses/forbs  and  forest.  In  some  sites,  beaver 
also  had  created  abrupt  edges.  For  example, 
sites  recently  flooded  by  beaver  dams  often 
had  no  transitional  vegetation  structure  be- 
tween the  new  pond  and  the  canopy-level  veg- 
etation (Snodgrass  1997). 

Twenty-eight  sites  had  a gradual  edge,  qual- 
itatively defined  as  a smooth  gradient  between 
vegetation  types  or  successional  stages  (Sua- 
rez et  al.  1997,  Luck  et  al.  1999).  Gradual 
edges  comprised  a complex  transition  between 
vegetation  types,  where  grasses,  forbs,  sap- 
lings, and  shrubs  were  intermixed.  Most  of  the 
beaver-impacted  wetlands  in  our  study  had 
gradual  edges,  primarily  because  there  had 
been  sufficient  time  since  beaver  invasion  for 
succession  to  occur;  gradual  edges  did  occur 
in  the  absence  of  beaver  evidence  wherever 
edges  were  not  maintained  by  anthropogenic 
disturbances. 

Presence/absence  of  beaver  evidence  was 
assessed  via  visual  observation.  Some  beaver- 
impacted  wetlands  were  inundated  hardwood 
forests.  Others  were  inundated  gaps  in  the 


canopy  that  had  resulted  from  tree-felling  and 
damming  activities;  these  wetlands  often  con- 
tained much  downed,  coarse  woody  debris 
and  many  exposed  stumps.  Some  beaver-im- 
pacted wetlands  had  been  abandoned,  as  evi- 
denced by  breached  dams  and  exposed  sedi- 
ments, which  supported  a variety  of  grasses, 
forbs,  and  shrub  species  (i.e.,  “beaver  mead- 
ows”). Overall,  beaver-impacted  wetlands 
were  characterized  by  a diversity  of  succes- 
sional seres  associated  with  beaver  coloniza- 
tion and  abandonment. 

Avian  censusing. — During  the  1999  field 
season,  we  conducted  a pilot  study  to  compare 
spot  mapping  and  50-m  fixed-radius  point 
counts.  Fixed-radius  point  counts  were  supe- 
rior for  this  study,  as  they  generated  more  bird 
detections  in  less  time  than  spot-mapping 
(Ralph  et  al.  1993),  allowing  us  to  increase 
sample  size  by  visiting  more  wetlands  in  2000 
and  2001.  Thus,  during  the  breeding  seasons 
of  2000  and  2001,  we  conducted  three  10-min, 
50-m  fixed-radius  point  counts  in  each  of  the 
57  wetlands  ( n = 33  sites  in  2000  and  n = 
24  sites  in  2001).  All  point  counts  were  con- 
ducted between  15  May  and  30  June,  from 
sunrise  to  10:00  EST,  on  days  when  neither 
precipitation  nor  wind  conditions  interfered 
with  bird  detections  (Ralph  et  al.  1995).  Dur- 
ing each  visit,  the  point  count  was  conducted 
from  the  center  of  the  core  wetland  area 
(Ralph  et  al.  1995)  and  always  at  the  same 
point  location  (Johnson  2001).  We  recorded 
all  birds  seen  or  heard  during  each  count 
(Ralph  et  al.  1995),  and  bird  detections  were 
categorized  as  <25.0,  25.1-50.0,  and  >50.0 
m from  the  point-count  center.  The  same  ob- 
server conducted  all  point  counts  in  all  3 
years. 

Although  point  counts — by  virtue  of  stan- 
dardized and  routinely  adopted  protocols 
(Ralph  et  al.  1995) — have  become  the  con- 
ventional technique  for  conducting  avian  cen- 
suses, differences  in  the  detectability  of  dif- 
ferent species  may  generate  inaccurate  counts 
(Thompson  2002).  Statistically  based  detect- 
ability adjustments  are  sometimes  used  to  at- 
tempt to  compensate  for  these  errors  (e.g., 
double-observer  approach,  Nichols  et  al. 
2000;  distance  sampling,  Rosenstock  et  al. 
2002;  double  sampling.  Bart  and  Earnst 
2002).  We  used  raw  data  for  our  analyses  be- 
cause our  sample  size  ( n = 57  wetlands  over 


402 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


3 years)  and  data  did  not  meet  all  the  as- 
sumptions necessary  for  use  of  distance-sam- 
pling methods  (Hutto  and  Young  2003).  In  ad- 
dition to  our  small  sample  size,  we  could  not 
be  certain  that  every  individual  present  was 
counted  only  once  or  that  precise  distances  for 
all  detections  were  estimated  accurately.  Thus, 
our  raw  data  were  used  to  assess  possible  re- 
lationships between  habitat  and  bird  commu- 
nities in  this  short-term  study. 

We  used  the  number  of  species  and  indi- 
viduals recorded  at  point  counts  to  calculate 
community-  and  guild-level  dependent  vari- 
ables for  statistical  analyses.  For  each  wet- 
land, we  calculated  community-level  species 
richness  as  the  total  number  of  species  ob- 
served across  all  three  visits.  Therefore,  spe- 
cies richness  assesses  all  species  observed  us- 
ing a wetland,  whether  or  not  they  were  breed- 
ing there;  some  birds  using  wetlands  for  for- 
aging (Pagen  et  al.  2000)  or  for  extraterritorial 
copulation  forays  (Norris  and  Stutchbury 
2001)  may  not  have  been  present  during  all 
census  visits.  For  each  wetland,  we  also  cal- 
culated community-level  avian  abundance  as 
the  mean  number  of  birds  observed  during  all 
three  visits. 

To  develop  guild-level  variables,  we  as- 
signed all  bird  species  to  guilds  based  upon 
classifications  used  by  the  Breeding  Bird  Sur- 
vey (Sauer  et  al.  2001).  We  focused  on  the 
NTMB  guild  (as  opposed  to  residents  and 
short-distance  migrants).  We  further  classified 
the  NTMBs  into  two  breeding-habitat  guilds: 
“late-successional”  (i.e.,  woodland)  and  “ear- 
ly-successional”  habitats.  All  early-succes- 
sional  NTMBs  nest  in  scrub,  except  the  East- 
ern Meadowlark  ( Sturnella  magna) — the  only 
“grassland”  nester  that  we  observed.  Because 
Eastern  Meadowlarks  represent  a unique  sub- 
guild of  early-successional  breeders,  and  be- 
cause we  observed  them  in  only  six  sites,  we 
excluded  this  species  from  our  analyses.  Thus, 
within  each  of  the  three  guilds  (i.e.,  NTMB 
and  two  habitat  guilds),  we  calculated  species 
richness  and  abundance,  which  we  used  as  de- 
pendent variables  in  statistical  analyses.  For 
each  wetland,  we  calculated  within-guild  spe- 
cies richness  as  the  total  number  of  species  in 
each  guild  observed  across  the  three  point 
counts.  We  calculated  within-guild  abundance 
as  the  mean  number  of  individuals  in  each 
guild  detected  across  all  three  visits. 


Vegetation  analyses. — At  each  site,  we  re- 
corded wetland  class  (Cowardin  et  al.  1979), 
presence  or  absence  of  livestock  evidence, 
presence  or  absence  of  beaver  evidence,  edge 
type,  and  ownership  status;  these  categorical 
variables  were  employed  as  independent  var- 
iables in  statistical  analyses  (Table  1).  For  a 
given  wetland,  vegetation  sampling  and  avian 
censuses  were  conducted  in  the  same  breeding 
season  (following  the  protocol  described  in 
Hamel  et  al.  1996).  At  each  wetland,  all  data 
were  collected  from  an  1 1.28-m-radius  circle 
surrounding  the  point-count  center  (see  table 
1 in  James  and  Shugart  1970). 

Percent  cover  of  several  classes  of  vegeta- 
tion structure  and  open  water  (Table  1)  were 
estimated  by  using  an  ocular  tube  (Hamel  et 
al.  1996).  In  each  of  the  four  cardinal  direc- 
tions, we  measured  2,  4,  6,  8,  and  10  m from 
the  point-count  center.  At  each  of  these  points 
we  looked  downward  and  upward  through  a 
5.08-cm  ocular  tube.  Presence  of  vegetation 
structural  layer(s)  observed  within  the  field  of 
view  of  the  ocular  tube  were  recorded  and 
used  to  calculate  the  percent  cover  of  vertical 
structural  layers  in  the  vegetation  plot. 

We  used  a vegetation  profile  board  to  assess 
horizontal  vegetation  structure  in  each  wet- 
land (Hamel  et  al.  1996).  This  method  entails 
using  a profile  board  (50.8  X 50.8  cm)  that  is 
divided  into  a grid  of  25  equally  sized  squares. 
The  board  was  placed  vertically  on  the 
ground,  10  m from,  and  facing,  the  point  cen- 
ter. We  recorded  number  of  squares  fully  vis- 
ible at  0,  2.5,  5,  and  7 m from  point  center,  in 
each  of  the  cardinal  directions.  A simple  cal- 
culation using  the  number  of  obstructed 
squares  (across  all  distances  and  directions) 
was  used  to  estimate  percent  horizontal  veg- 
etation density  in  each  wetland  (Hamel  et  al. 
1996). 

Data  analysis. — We  used  SAS  (SAS  Insti- 
tute, Inc.  2000)  to  conduct  stepwise  multiple 
linear  regressions  (SMLR)  with  the  PROC 
GLM  program  for  among-site  analyses  of 
wetland  use  by  breeding  birds  at  the  com- 
munity and  guild  levels.  Species  richness  and 
total  abundance  values  calculated  from  point- 
count  data  were  our  dependent  variables,  and 
vegetation  and  habitat  data  collected  from 
each  wetland  were  independent  variables.  Be- 
cause data  from  2000  and  2001  did  not  differ 
(r-tests),  we  pooled  data  from  both  years. 


Bulluck  and  Rowe  • NTMB  USE  OF  SOUTHERN  APPALACHIAN  WETLANDS 


403 


TABLE  1.  Description  of  18  independent  variables  measured  in  57  southern  Appalachian  wetlands  during 
2000  and  2001. 

Independent  variables 

Method  of  measurement 

Wetland  class 

Persistent-emergent,  scrub-shrub,  or  forested  (Cowardin  et  al.  1979) 

Livestock  evidence 

Presence  or  absence  of  recent  livestock  activity  (i.e.,  livestock, 
trampling,  and/or  manure) 

Edge  type 

Edge  nearest  the  point-count  center  was  gradual  or  abrupt 

Beaver  evidence 

Presence  or  absence  of  recent  beaver  activity  (i.e.,  actively  main- 
tained dams,  freshly  felled  trees,  and/or  recently  gnawed  stumps) 

Ownership  status 

Publicly  or  privately  owned 

Blue  Ridge  Parkway  ownership  status 

Under  the  jurisdiction  (or  not)  of  Blue  Ridge  Parkway  National  Park 

Size  of  wetland 

Publicly  owned  sites:  information  obtained  from  managers;  private- 
ly owned  sites:  estimated  (to  the  nearest  0. 1 ha)  from  1 :24,000 
USGS  topographic  maps 

Open  water 

Percent  cover  of  open  water3 

Stem  density  of  snags 

No.  snags  >10  cm  dbhb 

Stem  density  of  live  trees 

No.  trees  >10  cm  dbhb 

Basal  area  of  live  and  dead  trees  (cm2) 

Total  basal  area  of  trees  >10  cm  dbhb  (measured  with  a Biltmore 
stick;  Hamel  et  al.  1996) 

Canopy  cover 

Percent  canopy  cover3-6 

Midstory  cover 

Percent  cover  of  total  midstory  vegetation3-6 

Shrub  cover 

Percent  cover  of  shrub  vegetation3-6 

Ground  cover 

Percent  ground  cover3-6 

Forb  cover 

Percent  cover  of  forb  vegetation3  6 

Grass  cover 

Percent  cover  of  grass  vegetation3  6 

Vegetation  profile 

Estimated  horizontal  density  of  vegetation3-6 

a Vegetation  measures  made  using  the  ocular  tube  method  (Hamel  et  al.  1996). 
b Measurements  taken  within  1 1.28-m  circular  sample  plots. 


We  checked  all  dependent  and  independent 
variable  distributions  for  outliers  using  box 
plots  and  normal  probability  plots  (Tabach- 
nick  and  Fidell  1983,  Zar  1999).  Outlying  val- 
ues for  independent  variables  were  confirmed 
not  to  have  resulted  from  data  entry  errors, 
and  were  retained  for  final  regression  analy- 
ses. We  also  checked  all  variables  for  nor- 
mality using  residual  scatterplots  (Tabachnick 
and  Fidell  1983,  Zar  1999)  obtained  by  run- 
ning preliminary  multiple  regression  models 
for  every  dependent  variable  against  all  raw 
data  for  the  independent  variables  (SAS  Insti- 
tute, Inc.  2000).  Residual  scatterplots  for  all 
dependent  variables  were  normal,  and  regres- 
sion models  for  all  dependent  variables  were 
considered  valid. 

Prior  to  running  final  regression  analyses, 
we  conducted  a correlation  analysis  on  all  in- 
dependent variables  to  eliminate  redundancy 
in  habitat  measurements.  In  cases  where  cor- 
relations were  >0.50,  we  removed  one  of  the 
variables  before  running  final  regression  anal- 
yses. For  regression  models,  categorical  vari- 
ables, such  as  evidence  of  beaver  activity. 


edge  type,  and  wetland  type,  were  assigned 
absence/presence  values  of  0 or  1,  respective- 
ly. Only  parameters  significant  at  P < 0.05 
were  included  in  final  regression  models. 

RESULTS 

During  the  2000  and  2001  breeding  sea- 
sons, we  conducted  171  point  counts  in  the  57 
study  wetlands  and  detected  2,266  birds  rep- 
resenting 90  species  (see  Appendix  for  species 
observed). 

Community-level  analyses. — Presence  of 
gradual  edges,  beaver  evidence,  and  private 
ownership  collectively  explained  50%  of  the 
variation  in  community  species  richness  of 
NTMBs  (Table  2).  Beaver  evidence  also  ex- 
plained 16%  of  community  abundance,  and 
abundance  associated  with  wetlands  on  the 
BRP  was  lower  than  it  was  at  wetland  sites 
elsewhere  (Table  2). 

Guild-level  analyses. — Species  richness  of 
NTMBs  was  explained  by  the  presence  of 
gradual  edge  (42.5%  of  variation)  and  evi- 
dence of  beaver  activity  (7%;  Table  3).  Per- 
cent ground  cover  was  also  positively  corre- 


404 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE  2.  Significant  (P  < 0.05)  predictors  of  community-level 
wetlands  during  the  breeding  seasons  of  2000  and  200 1 . 

avian 

use  at  57  southern  Appalachian 

Community-level  parameter 

Predictor 

F 

SE 

Parameter  r2 

Model  R2 

Species  richness3 

Gradual  edge 

23.51 

0.96 

0.300***b 

Beaver  evidence 

10.15 

1.13 

0.111** 

0.41** 

Ownership  status 

9.66 

0.96 

0.091** 

0.50** 

Mean  avian  abundancec 

Beaver  evidence 

10.11 

4.64 

0.155** 

BRP  status 

7.05 

3.97 

0.098* 

0.25* 

a Total  number  of  species  detected  in  all  three  point  counts  in  each  wetland. 
b *P  < 0.05,  **  P < 0.01,  ***  P < 0.001. 

c Mean  number  of  individuals  observed  during  three  point-count  visits  to  each  study  wetland. 


lated  with  species  richness,  although  percent 
grass  cover  was  negatively  associated  with 
species  richness  (Table  3).  As  with  NTMB 
richness,  NTMB  abundance  was  most  strong- 
ly associated  with  gradual  edge  and  evidence 
of  beaver  activity;  collectively,  these  variables 
explained  37%  of  the  model  variation.  Also, 
though  to  a lesser  degree,  NTMB  abundance 
was  positively  associated  with  percent  cover 
of  canopy  vegetation  (Table  3). 

Basal  area  of  trees  at  our  sites  had  the  stron- 
gest negative  effect  on  species  richness  and 
abundance  of  early-successional  NTMBs:  it 
explained  16%  of  the  variation  in  both  rich- 
ness and  abundance  models  (Table  4).  Early- 
successional  NTMB  species  richness  and 
abundance  were  positively  correlated  with 
grazing  and  gradual  edge,  respectively  (Table 
4).  Late  successional  NTMB  species  richness 
and  abundance  both  were  positively  associat- 
ed with  gradual  edge,  basal  area,  and  evidence 
of  beaver  activity  (Table  4).  In  addition,  spe- 
cies richness  of  late-successional  NTMBs  was 
positively  associated  with  canopy  cover  (ac- 
counting for  19%  of  the  variation)  and  abun- 
dance was  positively  associated  with  midstory 
cover  (accounting  for  —5%  of  the  variation) 
(Table  4). 


DISCUSSION 

Although  our  vegetation  sampling  areas 
(1 1.28-m-radius  circular  plots)  did  not  corre- 
spond exactly  with  our  avian  census  areas  (50- 
m-radius  circular  plots),  the  wetland  sizes 
were  small,  in  which  case  our  quantitative 
vegetation  measurements  should  have  ade- 
quately represented  the  vegetation  of  most 
wetlands  overall;  only  the  largest  wetlands 
may  have  been  represented  inadequately  in 
our  11.28-m  vegetation  plots.  We  recognize 
that  this  spatial  inconsistency  may  have  driven 
the  effects  of  our  qualitative  habitat  variables 
(i.e.,  evidence  of  beaver  activity,  edge  type) 
more  than  the  continuous  variables  (e.g.,  per- 
cent cover  of  vegetation  types)  in  our  regres- 
sion models.  However,  relationships  between 
avian  community  structure  and  vegetation 
structure  should  not  be  disregarded. 

In  general,  many  of  our  results  support  ex- 
isting hypotheses  about  the  effects  of  land  use 
and  environmental  variables  on  NTMB  spe- 
cies richness  and  abundance.  At  the  commu- 
nity and  guild  levels,  species  richness  and 
abundance  were  associated  with  various  hab- 
itat characters  that  can  be  explained  by  the 
habitat  preferences  of  late-  and  early-succes- 


TABLE  3.  Significant  ( P < 0.05)  predictors  of  southern  Appalachian  wetland  use  by  the  Neotropical  mi- 
gratory bird  (NTMB)  guild  during  the  breeding  seasons  of  2000  and  2001. 


Guild-level  parameter 

Predictor 

F 

SE 

Parameter  r2 

Model  R 2 

NTMB  species  richness 

Gradual  edge 

40.59 

0.62 

0.425***3 

Grass  cover  (%) 

4.37 

0.01 

0.078** 

0.50** 

Beaver  evidence 

7.93 

0.83 

0.074** 

0.576** 

Ground  cover  (%) 

9.68 

1.03 

0.033* 

0.609* 

NTMB  abundance 

Gradual  edge 

21.53 

2.37 

0.284*** 

Beaver  evidence 

7.83 

2.80 

0.090** 

0.374** 

Canopy  cover  (%) 

4.53 

0.06 

0.063* 

0.436* 

a *P  < 0.05,  **  P < 0.01,  ***  P < 0.001. 


Bulluck  and  Rowe  • NTMB  USE  OF  SOUTHERN  APPALACHIAN  WETLANDS 


405 


TABLE  4.  Significant  ( P < 0.05)  predictors  of  southern  Appalachian  wetland  use  by  early-successional  (ES- 
NTMB)  and  late-successional  (LS-NTMB)  Neotropical  migratory  bird  guilds  during  the  breeding  seasons  of 
2000  and  2001. 

Guild-level  parameter 

Predictor 

F 

SE 

Parameter  r2 

Model  R2 

ES-NTMB  species  richness 

Basal  area 

10.14 

0.00 

0. 156**a 

Livestock  evidence 

11.47 

0.38 

0. 148** 

0.304** 

ES-NTMB  abundance 

Basal  area 

10.45 

0.01 

0.160** 

Gradual  edge 

10.84 

1.33 

0.141** 

0.300** 

LS-NTMB  species  richness 

Gradual  edge 

32.60 

0.66 

0.372*** 

Canopy  cover  (%) 

22.58 

0.02 

0.185*** 

0.557*** 

Beaver  evidence 

8.94 

0.73 

0.064** 

0.621** 

Basal  area 

4.68 

0.00 

0.031* 

0.652* 

LS-NTMB  abundance 

Basal  area 

40.45 

0.01 

0.424*** 

Beaver  evidence 

18.19 

1.80 

0.145*** 

0.569*** 

Midstory  cover  (%) 

6.40 

0.06 

0.046* 

0.615* 

Gradual  edge 

4.22 

1.62 

0.029* 

0.644* 

a * P < 0.05,  **  P < 0.01,  ***  P < 0.001. 


sional  NTMBs.  The  positive  association  be- 
tween private  ownership  and  species  richness, 
however,  was  unexpected.  Although  many  of 
the  publicly  owned  wetlands  we  studied  are 
managed,  in  part,  to  promote  biodiversity,  our 
results  show  that  private  wetlands  had  greater 
community-level  species  richness  than  sites 
held  in  public  trust.  This  may  reflect  land- 
scape-level  influences.  We  suspect  that  public- 
ly owned  sites  often  were  surrounded  by  less 
fragmented  landscapes  than  privately  owned 
sites,  which  often  were  embedded  in  land- 
scapes fragmented  by  various  land  uses.  The 
relatively  greater  number  of  small  habitat 
patches  surrounding  privately  owned  wetlands 
might  have  generated  a greater  diversity  of 
habitats  that  supported  a greater  variety  of 
birds  (Whitcomb  et  al.  1981). 

The  positive  effects  of  gradual  edges  on  the 
avian  community  overall,  and  on  NTMBs, 
were  also  unexpected.  Numerous  studies  have 
shown  that,  in  fragmented  forest  landscapes 
with  high  edge-to-interior  ratios,  area-sensi- 
tive NTMBs  experience  increased  predation 
due  to  greater  predator  abundance  (Temple 
and  Cary  1988,  Wilcove  and  Robinson  1990, 
Faaborg  et  al.  1995)  and  species  richness 
(Forsyth  and  Smith  1973,  Heske  1995,  Chal- 
foun  et  al.  2002),  as  well  as  greater  rates  of 
brood  parasitism  (Brittingham  and  Temple 
1983,  Johnson  and  Temple  1990).  However, 
the  differential  effects  of  gradual  versus 
abrupt  edges  on  NTMBs  have  received  far 
less  attention. 

Authors  of  previous  studies  have  reported 


greater  rates  of  nest  predation  along  abrupt 
edges  than  in  gradual  edges;  they  further  pro- 
pose that  the  more  developed  vegetation  struc- 
ture in  gradual  edges  provides  superior  nest 
concealment  (Ratti  and  Reese  1988)  and  min- 
imizes the  activity  and  efficiency  of  predators 
(Luck  et  al.  1999).  Gradual  edges  may  also 
provide  foraging  benefits.  Lopez  de  Casenave 
et  al.  (1998)  found  greater  avian  species  rich- 
ness and  abundance  in  “mature,”  or  gradual, 
edges  than  in  surrounding  habitats.  They  con- 
cluded that  complex,  mature  edges  provided 
greater  fruit  production  and  more  foraging 
niches  for  birds.  Along  with  these  findings, 
our  results  suggest  that  further  assessments  of 
parasitism,  predation,  and  foraging  efficiency 
in  abrupt  versus  gradual  edges  may  help  de- 
termine why  edge  structure  can  affect  avian 
community  structure. 

Overall  structure  of  wetland  vegetation  af- 
fected by  beaver  activity  also  may  have  influ- 
enced southern  Appalachian  bird  communi- 
ties. Grover  and  Baldassarre  (1995)  found  that 
wetlands  actively  maintained  by  beaver  har- 
bored greater  species  richness  of  breeding 
NTMBs  and  woodland  species  than  other  wet- 
lands, primarily  due  to  their  structurally  di- 
verse edges.  In  our  study,  beaver-impacted 
wetlands  were  characterized  by  gradual  edges 
more  often  than  by  abrupt  edges  ( P < 0.05, 
R2  = 0.302). 

Beaver  activity  is  also  thought  to  enhance 
avian  foraging  efficiency  by  creating  structur- 
ally diverse  habitats  with  a diversity  of  for- 
aging niches  (Medin  and  Clary  1990)  and  by 


406 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


increasing  the  productivity  of  insects — the 
dominant  component  of  NTMB  diets  (Reese 
and  Hair  1976).  Further  investigations  focus- 
ing on  differences  in  wetland  vegetation  struc- 
ture and  productivity  in  beaver-impacted  ver- 
sus other  wetlands  could  provide  more  con- 
clusive results  regarding  how  beaver  may  en- 
hance habitat  quality  for  nesting  NTMBs. 

From  a management  perspective,  results 
from  our  study  and  those  of  previous  studies 
suggest  new  approaches  to  managing  southern 
Appalachian  wetlands  to  promote  persistence 
of  native  birds.  “Gradualizing”  wetland  edges 
and  encouraging  beaver  could  be  especially 
beneficial  for  NTMBs.  Edges  are  inherent  re- 
sults of  current  land-use  practices,  and  al- 
though the  effects  of  edge  quantity  on  area- 
sensitive  songbirds  are  well-documented,  a 
better  understanding  of  how  edge  quality  af- 
fects these  species  may  help  to  refine  man- 
agement activities. 

Future  investigations  of  how  beaver  benefit 
songbirds  at  local  and  landscape  levels  also 
might  be  prudent.  Paradoxically,  populations 
of  avian  species  with  very  different  habitat  re- 
quirements are  in  decline,  including  those  that 
prefer  both  early-successional  grasslands  and 
late-successional  forests.  Some  researchers 
have  argued  that  landscapes  in  the  southeast- 
ern United  States  have  lost  their  heterogeneity 
and  are  now  dominated  by  homogeneous 
stands  of  mid-successional  forest  (Hunter  et 
al.  2001).  Prior  to  their  near  extirpation  over 
a century  ago,  the  estimated  60  million  beaver 
in  North  Carolina  alone  (McGrath  and  Sum- 
mer 1992)  would  have  generated  a remarkable 
mosaic  of  early-  to  late-successional  ponds, 
meadows,  and  forested  bogs.  The  physio- 
graphic diversity  of  these  sites,  coupled  with 
their  productivity,  may  have  benefited  bird 
species  with  a wide  range  of  resource  require- 
ments. Rather  than  treating  beaver  as  pests, 
public  land  managers  in  the  southern  Appa- 
lachians should  encourage  beaver  in  their  ef- 
forts to  restore  a heterogeneous  landscape  ca- 
pable of  supporting  a diverse  avifauna. 

ACKNOWLEDGMENTS 

The  U.S.  Fish  and  Wildlife  Service,  Explorer’s  Club, 
Wilderness  Society,  and  Appalachian  State  University 
Graduate  School  funded  this  research.  We  thank  N. 
Murdock,  D.  Lee,  C.  Haney,  and  D.  Herman  for  shar- 
ing their  expertise  and  knowledge  of  the  natural  his- 


tory, flora,  and  fauna  of  southern  Appalachian  wet- 
lands. This  study  could  not  have  been  conducted  with- 
out the  cooperation  of  The  Nature  Conservancy,  the 
National  Park  Service,  the  U.S.  Department  of  Agri- 
culture Forest  Service,  and  various  private  landowners 
who  allowed  us  access  into  highly  sensitive  areas  to 
collect  data.  We  would  also  like  to  thank  the  three 
anonymous  reviewers  of  this  manuscript  for  their  help- 
ful comments  and  suggestions. 

LITERATURE  CITED 

Askins,  R.  A.,  J.  F.  Lynch,  and  R.  Greenberg.  1990. 
Population  declines  in  migratory  birds  in  eastern 
North  America.  Current  Ornithology  7:1-57. 
Bart,  J.  and  S.  Earnst.  2002.  Double  sampling  to 
estimate  density  and  population  trends  in  birds. 
Auk  119:36-45. 

Brittingham,  M.  C.  and  S.  A.  Temple.  1983.  Have 
cowbirds  caused  forest  songbirds  to  decline?  Bio- 
science 33:31-35. 

Carter,  M.  F,  W.  C.  Hunter,  D.  N.  Pashley,  and  K. 
V.  Rosenberg.  2000.  Setting  conservation  priori- 
ties for  landbirds  in  the  United  States:  the  Partners 
in  Flight  approach.  Auk  117:541-548. 

Chalfoun,  A.  D„  F.  R.  Thompson,  and  M.  J.  Ratnas- 
wamy.  2002.  Nest  predators  and  fragmentation:  a 
review  and  meta-analysis.  Conservation  Biology 
16:306-318. 

Cowardin,  L.  M.,  V.  Carter,  F.  C.  Golet,  and  E.  T. 
LaRoe.  1979.  The  Cowardin  classification  of  wet- 
land and  deepwater  habitats  of  the  United  States. 
U.S.  Fish  and  Wildlife  Service  Report  FWS-OBS 
70-31,  Washington,  D.C. 

Davis,  M.  B.  1993.  Old  growth  in  the  east:  a survey. 

Cenozoic  Society,  Richmond,  Vermont. 
Delcourt,  H.  R.  and  P.  A.  Delcourt.  1997.  Pre-Co- 
lumbian Native  American  use  of  fire  on  southern 
Appalachian  landscapes.  Conservation  Biology 
11:1010-1014. 

Faaborg,  J.,  M.  Brittingham,  T.  Donovan,  and  J. 
Blake.  1995.  Habitat  fragmentation  in  the  tem- 
perate zone.  Pages  357-380  in  Ecology  and  man- 
agement of  Neotropical  migratory  birds  (T.  E. 
Martin  and  D.  M.  Finch,  Eds.).  Oxford  University 
Press,  New  York. 

Forsyth,  D.  J.  and  D.  A.  Smith.  1973.  Temporal  var- 
iation in  home  ranges  of  eastern  chipmunks  (Tam- 
ias  striatus)  in  a southeastern  Ontario  woodlot. 
American  Midland  Naturalist  90:107-117. 
Franzreb,  K.  E.  and  K.  V.  Rosenberg.  1997.  Are  for- 
est songbirds  declining?  Status  assessment  from 
the  southern  Appalachians  and  northeastern  for- 
ests. Transactions  of  the  62nd  North  American 
Wildlife  and  Natural  Resources  Conference  62: 
264-278. 

Grover,  A.  M.  and  G.  A.  Baldassarre.  1995.  Bird 
species  richness  within  beaver  ponds  in  south- 
central  New  York.  Wetlands  15:108—118. 

Hamel,  P.  B.,  W.  P.  Smith,  D.  J.  Twedlt,  J.  R.  Woehr, 
E.  Morris,  R.  B.  Hamilton,  and  R.  J.  Cooper. 
1996.  A land  manager’s  guide  to  point  counts  of 


Bulluck  and  Rowe  • NTMB  USE  OF  SOUTHERN  APPALACHIAN  WETLANDS 


407 


birds  in  the  Southeast.  General  Technical  Report 
SO- 120,  USDA  Forest  Service,  Southern  Re- 
search Station,  Asheville,  North  Carolina. 

Heske,  E.  J.  1995.  Mammalian  abundances  in  forest- 
farm  edges  versus  forest  interiors  in  southern  Il- 
linois: is  there  an  edge  effect?  Journal  of  Mam- 
mology  76:562-568. 

Hunter,  W.  C.,  D.  A.  Buehler,  R.  A.  Canterbury,  J. 

L.  Confer,  and  P.  B.  Hamel.  2001.  Conservation 
of  disturbance-dependent  birds  in  eastern  North 
America.  Wildlife  Society  Bulletin  29:440-455. 

Hunter,  W.  C.,  M.  F.  Carter,  D.  N.  Pashley,  and  K. 
Barker.  1993.  Neotropical  migratory  landbird 
species  and  habitats  of  special  concern  within  the 
southeast  region.  Pages  159-171  in  Status  and 
management  of  Neotropical  migratory  birds  (D. 

M.  Finch  and  P.  W.  Stangel,  Eds.).  General  Tech- 
nical Report  RM-229,  USDA  Forest  Service, 
Rocky  Mountain  Research  Station,  Fort  Collins, 
Colorado. 

Hunter,  W.  C.,  R.  Katz,  D.  N.  Pashley,  and  R.  P. 
Ford.  1999.  Partners  in  Flight  Bird  Conservation 
Plan  for  the  Southern  Blue  Ridge  (physiographic 
area  23).  http://www.blm.gov/wildlife/plan/pl_23_ 
10.pdf  (accessed  29  November  2003). 

Hutto,  R.  L.  and  J.  S.  Young.  2003.  On  the  design 
of  monitoring  programs  and  the  use  of  population 
indices:  a reply  to  Ellingson  and  Lukacs.  Wildlife 
Society  Bulletin  31:903-910. 

James,  F.  C.  and  H.  H.  Shugart,  Jr.  1970.  A quanti- 
tative method  of  habitat  description.  Audubon 
Field  Notes  24:727-736. 

Johnson,  D.  H.  2001.  Habitat  fragmentation  effects  on 
birds  in  grasslands  and  wetlands:  a critique  of  our 
knowledge.  Great  Plains  Research  11:211-230. 

Johnson,  R.  G.  and  S.  A.  Temple.  1990.  Nest  preda- 
tion and  brood  parasitism  of  tallgrass  prairie  birds. 
Journal  of  Wildlife  Management  54:106-1 1 1 . 

Lawton,  J.  H.  and  C.  G.  Jones.  1995.  Linking  species 
and  ecosystems:  organisms  as  ecosystem  engi- 
neers. Pages  141-150  in  Linking  species  and  eco- 
systems (C.  G.  Jones  and  J.  H.  Lawton,  Eds.). 
Chapman  and  Hall,  New  York. 

Lee,  D.  S.  and  A.  W.  Norden.  1996.  The  distribution, 
ecology  and  conservation  needs  of  bog  turtles, 
with  special  emphasis  on  Maryland.  Maryland 
Naturalist  40:7-46. 

LeGrand,  H.  E.,  Jr.,  S.  P.  Hall,  and  J.  T.  Finnegan. 
2001.  Natural  Heritage  Program  list  of  the  rare 
animal  species  of  North  Carolina.  North  Carolina 
Natural  Heritage  Program,  Division  of  Parks  and 
Recreation,  North  Carolina  Department  of  the  En- 
vironment and  Natural  Resources,  Raleigh. 

Lopez  de  Casenave,  J.,  J.  P.  Pelotto,  S.  M.  Caziani, 
M.  Mermoz,  and  J.  Protomastro.  1998.  Re- 
sponses of  avian  assemblages  to  a natural  edge  in 
a Chaco  semiarid  forest  in  Argentina.  Auk  115: 
425-435. 

Luck,  G.  W.,  H.  P.  Possingham,  and  D.  C.  Paton. 
1999.  Bird  responses  at  inherent  and  induced  edg- 


es in  the  Murray  Mallee,  South  Australia.  2.  Nest 
predation  as  an  edge  effect.  Emu  99:170-175. 

McGrath,  C.  and  P.  Summer.  1992.  Beaver:  Castor 
canadensis.  Wildlife  Profiles.  Division  of  Conser- 
vation Education,  North  Carolina  Wildlife  Re- 
sources Commission,  Raleigh. 

Medin,  D.  E.  and  W.  P.  Clary.  1990.  Bird  populations 
in  and  adjacent  to  a beaver  pond  ecosystem  in 
Idaho.  Research  Paper  INT-432,  USDA  Forest 
Service,  Intermountain  Research  Station,  Ogden, 
Utah. 

Murdock,  N.  A.  1994.  Rare  and  endangered  plants 
and  animals  of  southern  Appalachian  wetlands. 
Water,  Air  and  Soil  Pollution  77:385-405. 

Naiman,  R.  J.,  C.  A.  Johnston,  and  J.  C.  Kelley. 
1988.  Alteration  of  North  American  streams  by 
beaver.  Bioscience  38:753—762. 

Nichols,  J.  D..  J.  E.  Hines,  J.  R.  Sauer,  F.  W.  Fallon, 
J.  E.  Fallon,  and  P.  J.  Heglund.  2000.  A double- 
observer approach  for  estimating  detection  prob- 
ability and  abundance  from  point  counts.  Auk 
117:393-408. 

Norris,  D.  R.  and  B.  J.  M.  Stutchbury.  2001.  Extra- 
territorial movements  of  a forest  songbird  in  a 
fragmented  landscape.  Conservation  Biology  15: 
729-736. 

Pagen,  R.  W.,  F.  R.  Thompson,  III,  and  D.  E.  Bur- 
hans.  2000.  Breeding  and  post-breeding  habitat 
use  by  forest  migrant  songbirds  in  the  Missouri 
Ozarks.  Condor  102:738-747. 

Ralph,  C.  J.,  S.  Droege,  and  J.  R.  Sauer.  1995.  Man- 
aging and  monitoring  birds  using  point  counts: 
standards  and  applications.  Pages  161-168  in 
Monitoring  bird  populations  by  point  counts  (C. 
J.  Ralph,  J.  R.  Sauer,  and  S.  Droege,  Eds.).  Gen- 
eral Technical  Report  PSW-149.  USDA  Forest 
Service,  Pacific  Southwest  Research  Station,  and 
U.S.  Department  of  Agriculture,  Albany,  Califor- 
nia. 

Ralph,  C.  J.,  G.  R.  Geupel,  P.  Pyle,  T.  E.  Martin, 
and  D.  F.  SeSante.  1993.  Handbook  of  field 
methods  for  monitoring  landbirds.  General  Tech- 
nical Report  PSW-144,  USDA  Forest  Service,  Pa- 
cific Southwest  Research  Station,  Albany,  Cali- 
fornia. 

Ratti,  J.  T.  and  K.  P.  Reese.  1988.  Preliminary  test  of 
the  ecological  trap  hypothesis.  Journal  of  Wildlife 
Management  52:484-491. 

Reese,  K.  P.  and  J.  D.  Hair.  1976.  Avian  species  di- 
versity in  relation  to  beaver  pond  habitats  in  the 
Piedmont  region  of  South  Carolina.  Proceedings 
of  the  Annual  Conference  of  the  Southeast  Asso- 
ciation of  Fish  and  Wildlife  Agencies  30:437- 
447. 

Robbins,  C.  S.,  J.  R.  Sauer,  R.  S.  Greenberg,  and  S. 
Droege.  1989.  Population  declines  in  North 
American  birds  that  migrate  to  the  Neotropics. 
Proceedings  of  the  National  Academy  of  Sciences 
86:7658-7662. 

Rodriguez,  J.  P.  2002.  Range  contraction  in  declining 


408 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


North  American  bird  populations.  Ecological  Ap- 
plications 12:238-248. 

Rosenstock,  S.  S.,  D.  R.  Anderson,  K.  M.  Giesen,  T. 
Leukering,  and  M.  F.  Carter.  2002.  Landbird 
counting  techniques:  current  practices  and  an  al- 
ternative. Auk  119:46-53. 

SAS  Institute,  Inc.  2000.  SAS  ver.  8.1.  SAS  Institute, 
Inc.,  Cary,  North  Carolina. 

Sauer,  J.  R.,  J.  E.  Hines,  G.  Gough,  I.  Thomas,  and 
B.  G.  Peterjohn.  2001.  The  North  American 
Breeding  Bird  Survey:  results  and  analysis  1966- 
2000,  ver.  2001.2.  U.S.  Geological  Survey,  Patux- 
ent Wildlife  Research  Center,  Laurel,  Maryland. 
http://www.mbr-pwrc.usgs.gov/bbs/bbs00.html 
(accessed  12  July  2001). 

Simons,  T.  R.,  J.  L.  Lichstein,  K.  Weeks,  and  K.  E. 
Franzreb.  2000.  The  effects  of  landscape  pattern, 
core  areas,  and  the  forest  management  practices 
on  avian  communities  in  the  southern  Appala- 
chians. 1999  annual  report.  USD  A Forest  Service, 
Raleigh,  North  Carolina. 

Snodgrass,  J.  W.  1997.  Temporal  and  spatial  dynam- 
ics of  beaver-created  patches  as  influenced  by 
management  practices  in  a southeastern  North 
American  landscape.  Journal  of  Applied  Ecology 
34:1043-1056. 

Suarez,  A.  V.,  K.  S.  Pfennig,  and  S.  K.  Robinson. 
1997.  Nesting  success  of  a disturbance-  dependent 
songbird  on  different  kinds  of  edges.  Conserva- 
tion Biology  11:928-935. 

Tabachnick,  B.  G.  and  L.  S.  Fidell.  1983.  Using  mul- 
tivariate statistics.  Harper  and  Row,  New  York. 

Temple,  S.  A.  and  J.  R.  Cary.  1988.  Modeling  dy- 
namics of  habitat-interior  bird  populations  in  frag- 
mented landscapes.  Conservation  Biology  2:340- 
347. 

Thompson,  F.  R.,  Ill,  and  R.  M.  DeGraaf.  2001.  Con- 
servation approaches  for  woody,  early  succession- 


al  communities  in  the  eastern  United  States.  Wild- 
life Society  Bulletin  29:483-494. 

Thompson,  W.  L.  2002.  Towards  reliable  bird  surveys: 
accounting  for  individuals  present  but  not  detect- 
ed. Auk  119:18-25. 

Ward,  H.  T.  1990.  The  bull  in  the  North  Carolina  buf- 
falo. Southern  Indian  Studies  39:19-30. 

Weakley,  A.  S.  and  M.  P.  Shafale.  1994.  Non-allu- 
vial  wetlands  of  the  southern  Blue  Ridge:  diver- 
sity in  a threatened  ecosystem.  Water,  Air  and  Soil 
Pollution  77:359-383. 

Webster,  J.  R.,  J.  B.  Waide,  and  B.  C.  Patten.  1975. 
Nutrient  recycling  and  the  stability  of  ecosystems. 
Pages  1-27  in  Mineral  cycling  in  the  southeastern 
ecosystems  (F.  G.  Howell,  J.  B.  Gentry,  and  M. 
H.  Smith,  Eds.).  Energy  Research  and  Develop- 
ment Administration  Symposium  Series  CONF- 
740513,  Oak  Ridge,  Tennessee. 

Weigl,  P.  D.  and  T.  W.  Knowles.  1995.  Megaherbi- 
vores and  southern  Appalachian  grass  balds. 
Growth  and  Change  26:365-382. 

Whitcomb,  R.  E,  C.  S.  Robbins,  J.  F.  Lynch,  B.  L. 
Whitcomb,  M.  K.  Klimkiewicz,  and  D.  Bystrak. 
1981.  Effects  of  forest  fragmentation  on  avifauna 
of  the  eastern  deciduous  forest.  Pages  125-205  in 
Forest  island  dynamics  in  man-dominated  land- 
scapes (R.  L.  Burgess  and  D.  M.  Sharpe,  Eds.). 
Springer- Verlag,  New  York. 

Wilcove,  D.  S.  and  S.  K.  Robinson.  1990.  The  impact 
of  forest  fragmentation  on  bird  communities  in 
eastern  North  America.  Pages  319-331  in  Bio- 
geography and  ecology  of  forest  bird  communities 
(A.  Keast,  Ed.).  SPB  Academic  Publishing,  The 
Hague,  The  Netherlands. 

Zannette,  L.,  P.  Doyle,  and  S.  M.  Tremont.  2000. 
Food  shortage  in  small  fragments:  evidence  from 
an  area-sensitive  passerine.  Ecology  81:1 654 — 
1666. 

Zar,  J.  H.  1999.  Biostatistical  analysis.  Prentice  Hall, 
Upper  Saddle  River,  New  Jersey. 


Bulluck  and  Rowe  • NTMB  USE  OF  SOUTHERN  APPALACHIAN  WETLANDS 


409 


APPENDIX.  Occurrence  rates  of  bird  species  observed  in  57  southern  Appalachian  study  wetlands  during 
2000  and  2001. 

Common  name 

Scientific  name 

No.  sites  where 
observed  (%) 

Wood  Duck 

Aix  sponsa 

2 (3.51) 

Mallard 

Anas  platyrhynchos 

1 (1.75) 

Ruffed  Grouse 

Bonasa  umbellus 

1 (1.75) 

Northern  Bobwhiteab 

Colinus  virginianus 

1 (1.75) 

American  Bittern3 

Botaurus  lentiginosus 

1 (1.75) 

Green  Heron3 

Butorides  virescens 

1 (1.75) 

Turkey  Vulture 

Cathartes  aura 

4 (7.02) 

Red-tailed  Hawk 

Buteo  jamaicensis 

1 (1.75) 

Killdeer3 

Charadrius  vociferus 

1 (1.75) 

Mourning  Dove3 

Zenaida  macroura 

6 (10.53) 

Chimney  Swift3b 

Chaetura  pelagica 

1 (1.75) 

Ruby-throated  Hummingbird 

Archilochus  colubris 

11  (19.30) 

Belted  Kingfisher3 

Ceryle  alcyon 

10  (17.54) 

Red-bellied  Woodpecker 

Melanerpes  carolinus 

6 (10.53) 

Yellow-bellied  Sapsucker0 

Sphyrapicus  varius 

3 (5.26) 

Downy  Woodpeckerb 

Picoides  pubescens 

12  (21.05) 

Hairy  Woodpecker 

Picoides  villosus 

11  (19.29) 

Northern  Flicker3b 

Colaptes  auratus 

2 (3.51) 

Pileated  Woodpecker 

Dryocopus  pileatus 

3 (5.26) 

Eastern  Wood-Pewee3bc 

Contopus  virens 

12  (21.05) 

Acadian  Flycatcherb0 

Empidonax  virescens 

8 (14.04) 

Alder  Flycatcher0 

Empidonax  alnorum 

14  (24.56) 

Willow  Flycatcher3 

Empidonax  traillii 

9 (15.79) 

Least  Flycatcher3 

Empidonax  minimus 

9 (15.79) 

Eastern  Phoebeb 

Sayornis  phoebe 

21  (36.84) 

Great  Crested  Flycatcher 

Myiarchus  crinitus 

2 (3.51) 

White-eyed  Vireo 

Vireo  griseus 

4 (7.02) 

Blue-headed  Vireo0 

Vireo  solitarius 

16  (28.07) 

Red-eyed  Vireo 

Vireo  olivaceus 

35  (61.40) 

Blue  Jay3b 

Cyanocitta  cristata 

11  (19.30) 

American  Crow 

Corvus  brachyrhynchos 

21  (36.84) 

Tree  Swallow 

Tachycineta  bicolor 

1 (1.75) 

Bank  Swallow 

Riparia  riparia 

1 (1.75) 

Barn  Swallow 

Hirundo  rustic  a 

1 (1.75) 

Carolina  Chickadee3 

Poecile  carolinensis 

32  (56.14) 

Tufted  Timouse 

Baeolophus  bicolor 

28  (49.12) 

Red-breasted  Nuthatch0 

Sitta  canadensis 

2 (3.51) 

White-breasted  Nuthatch 

Sitta  carolinensis 

10  (17.54) 

Brown  Creeper0 

Certhia  americana 

1 (1.75) 

Carolina  Wrenb 

Thryothorus  ludovicianus 

17  (29.82) 

House  Wren 

Troglodytes  aedon 

9 (15.79) 

Winter  Wren0 

Troglodytes  troglodytes 

1 (1.75) 

Golden-crowned  Kinglet0 

Regulus  satrapa 

7 (12.28) 

Blue-gray  Gnatcatcher*5 

Polioptila  caerulea 

6 (10.53) 

Easten  Bluebird 

Sialia  sialis 

22  (38.60) 

Veery3 

Catharus  fuscescens 

4 (7.02) 

Wood  Thrush3b0 

Hylocichla  mustelina 

12  (21.05) 

American  Robinb 

Turdus  migratorius 

25  (43.86) 

Gray  Catbirdb  o 

Dumetella  carolinensis 

40  (70.18) 

Northern  Mockingbird3 

Mimus  polyglottos 

2 (3.51) 

Brown  Thrasher3 

Toxostoma  rufum 

8 (14.04) 

European  Starling3 

Sturnus  vulgaris 

9 (15.79) 

Cedar  Waxwing 

Bombycilla  cedrorum 

31  (54.39) 

Golden-winged  Warbler30 

Vermivora  chrysoptera 

3 (5.26) 

Northern  Parula0 

Parula  americana 

21  (36.84) 

Yellow  Warbler 

Dendroica  petechia 

8 (14.04) 

410 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


APPENDIX.  Continued. 

No.  sites  where 

Common  name 

Scientific  name 

observed  (%) 

Chestnut-sided  Warbler30 

Dendroica  pensylvanica 

17  (29.82) 

Black-throated  Blue  Warbler0 

Dendroica  caerulescens 

14  (24.56) 

Yellow-rumped  Warbler 

Dendrocia  coronata 

1 (1.75) 

Black-throated  Green  Warbler 

Dendroica  virens 

2 (3.51) 

Blackburnian  Warbler0 

Dendroica  fusca 

2 (3.51) 

Pine  Warbler 

Dendroica  pinus 

3 (5.26) 

Prairie  Warbler36 

Dendroica  discolor 

3 (5.26) 

Black-and-white  Warblerb 

Mniotilta  varia 

10  (17.54) 

American  Redstart 

Setophaga  ruticilla 

5 (8.77) 

Worm-eating  Warbler0 

Helmitheros  vermivorum 

2 (3.50) 

Swainson’s  Warbler60 

Limnothlypis  swainsonii 

1 (1.75) 

Ovenbirdb 

Seiurus  aurocapilla 

17  (29.82) 

Louisiana  Waterthrush0 

Seiurus  motacilla 

7 (12.28) 

Common  Yellowthroat3-6 

Geothlypis  trichas 

36  (63.16) 

Hooded  Warbler0 

Wilsonia  citrina 

22  (38.60) 

Canada  Warbler3  0 

Wilsonia  canadensis 

3 (5.26) 

Yellow-breasted  Chatb 

Icteria  virens 

7 (12.28) 

Scarlet  Tanager3  b0 

Piranga  olivacea 

15  (26.32) 

Eastern  Towhee3  6 

Pipilo  erythrophthalmus 

38  (66.67) 

Chipping  Sparrowb 

Spizella  passerina 

10  (17.54) 

Field  Sparrow3>b 

Spizella  pusilla 

13  (22.81) 

Song  Sparrow35 

Melospiza  melodia 

41  (71.93) 

White-throated  Sparrow3 

Zonotrichia  albicollis 

2 (3.51) 

Dark-eyed  Junco30 

Junco  hyemalis 

16  (28.07) 

Northern  Cardinal 

Cardinalis  cardinalis 

37  (64.91) 

Rose-breasted  Grosbeak3 

Pheucticus  ludovicianus 

1 (1.75) 

Indigo  Bunting35 

Passerina  cyanea 

42  (73.68) 

Red-winged  Blackbird3 

Agelaius  phoeniceus 

21  (36.84) 

Eastern  Meadowlark3-6 

Sturnella  magna 

5 (8.77) 

Common  Grackle 

Quiscalus  quiscula 

4 (7.02) 

Brown-headed  Cowbird3 

Molothrus  ater 

3 (5.26) 

House  Finch 

Carpodacus  mexicanus 

6 (10.53) 

American  Goldfinch 

Carduelis  tristis 

36  (63.16) 

House  Sparrow35 

Passer  domesticus 

1 (1.75) 

a Undergoing  significant  population  decline  throughout  the  species’  breeding  range  (Sauer  et  al.  2001). 

b Undergoing  a moderate  or  significant  population  decline  in  southern  Blue  Ridge  region  (Partners  in  Flight  physiographic  region  23;  Carter  et  al.  2000, 
Hunter  et  al.  1999)  or  in  the  Blue  Ridge  region  of  the  North  American  Breeding  Bird  Survey  (Sauer  et  al.  2001). 

0 Considered  a priority  species  in  the  southern  Blue  Ridge  region  (Partners  in  Flight  physiographic  region  23;  Carter  et  al.  2000,  Hunter  et  al.  1999)  or 
a species  of  local  concern  in  the  southern  Appalachians  (North  Carolina  Natural  Heritage  Program;  LeGrand  et  al.  2001,  Hunter  et  al.  1993,  D.  S.  Lee 
and  B.  Browning  unpubl.  data). 


Short  Communications 


The  Wilson  Journal  of  Ornithology  1 1 8(3):4 1 \ —4\3,  2006 


Breeding  Range  Extension  of  the  Northern  Saw-whet  Owl  in  Quebec 


Christophe  Buidin,1  Yann  Rochepault,1  Michel  Savard,2 3 4  and  Jean-Pierre  L.  Savard34 


ABSTRACT. — Although  the  breeding  range  of  the 
Northern  Saw-whet  Owl  ( Aegolius  acadicus ) is  re- 
stricted to  North  America,  the  northern  limits  of  its 
range  are  still  unclear.  In  Quebec,  the  most  northerly 
confirmed  breeding  records  had  come  from  the  Sag- 
uenay area  (Chicoutimi;  48°  25'  N,  71°  03'  W)  in  bal- 
sam fir-  ( Abies  balsamea ) white  birch  ( Betula  papyri- 
fera ) forest  and  on  the  Gaspe  Peninsula  (Amqui;  48° 
28' N,  67°  25 ' W)  in  balsam  fir-yellow  birch  (. B . al- 
leghaniensis ) forest.  Between  1998  and  2003, 
however,  we  documented  nine  Northern  Saw-whet 
Owl  nests  in  balsam  fir-black  spruce  ( Picea  marina) 
forest  in  boreal  Quebec  on  the  Mingan  Terraces.  These 
records  extend  the  species’  known  breeding  range 
northward  to  >50°  N.  Received  8 August  2005,  ac- 
cepted 24  March  2006. 


The  breeding  range  of  the  Northern  Saw- 
whet  Owl  {Aegolius  acadicus ) is  restricted  to 
North  America  (Cannings  1993),  and  includes 
most  of  the  southern  Canadian  forested  areas, 
the  mountainous  regions  of  the  United  States, 
and  the  mountains  of  Mexico  south  to  Oaxaca. 
The  northernmost  distribution  of  this  species 
occurs  along  the  Pacific  coast,  extending 
northward  from  British  Columbia  to  south- 
central  Alaska  (American  Ornithologists’ 
Union  1998).  However,  the  northern  limit  of 
its  range  remains  unclear  (Godfrey  1986,  Can- 
nings 1993).  In  Quebec,  Northern  Saw- whet 
Owls  breed  in  all  forested  areas  south  of  49° 
N,  with  the  exception  of  the  Abitibi  region 
(Cote  and  Bombardier  1996).  Previously,  the 
most  northerly  breeding  records  confirmed  in 
Quebec  came  from  the  Saguenay  area  (Chi- 
coutimi; 48°  25'  N,  71°  03'  W)  in  balsam  fir- 
( Abies  balsamea ) white  birch  {Betula  papyri- 
fera ) forest  and  on  the  Gaspe  Peninsula 


1 Assoc,  le  Balbuzard,  1 chemin  du  Grand  Ruisseau, 
Riviere-Saint-Jean,  QC  GOG  1N0,  Canada. 

2 Observatoire  d’oiseaux  de  Tadoussac,  302  rue  de 
la  Riviere,  Les  Bergeronnes,  QC  GOT  1G0,  Canada. 

3 Canadian  Wildlife  Service,  1141  Route  de  l’Eglise, 
P.O.  Box  10100,  Sainte-Foy,  QC  G1V  4H5,  Canada. 

4 Corresponding  author;  e-mail: 
jean-pierre.savard@ec.gc.ca 


(Amqui;  48°  28'  N,  67°  25'  W)  in  balsam  fir- 
yellow  birch  {B.  alleghaniensis ) forest  (Cote 
and  Bombardier  1996).  Seventeen  records, 
however,  in  the  1979-1998  regional  database 
housed  at  the  Etude  des  populations  d’oiseaux 
du  Quebec  indicated  that  Northern  Saw-whet 
Owls  breed  farther  north  in  the  Baie-Comeau 
area  (49°  13' N,  68°  09'  W)  than  what  was 
published  in  the  literature  as  their  confirmed 
breeding  range  in  Quebec  (Cote  and  Bombar- 
dier 1996). 

Between  1998  and  2003,  we  documented  a 
northerly  extension  of  the  known  breeding 
range  of  the  Northern  Saw-whet  Owl  in  bal- 
sam fir-black  spruce  {Picea  marina ) forest  in 
boreal  Quebec,  north  of  50°  N.  During  the 
1997-1998  winter,  we  had  erected  22  nest 
boxes  for  Boreal  Owls  {Aegolius  fune reus)  in 
the  Magpie  River  area  (50°  19'  N,  64°  27'  W) 
and,  during  the  1998-1999  winter,  we  erected 
51  nest  boxes  between  the  Manitou  River 
(50°  19'  N,  65°  14'  W)  and  Longue-Pointe-de- 
Mingan  (50°  17'  N,  64°  03'  W).  From  1998  to 
2003,  we  documented  9 Northern  Saw- whet 
Owl  nests  (Table  1),  as  well  as  15  Boreal  Owl 
and  11  American  Kestrel  {Falco  sparverius) 
nests,  in  the  nest  boxes.  On  1 1 June  1998,  we 
discovered  the  first  Northern  Saw-whet  Owl 
nest,  which  contained  a 1 -year-old  female 
brooding  four  young.  That  day,  we  banded  the 
female  at  her  nest,  located  at  Riviere-Saint- 
Jean  (50°  18'  N,  64°  22'  W);  on  29  February 
2000,  the  bird  was  recaptured  in  the  United 
States  at  Port  Elizabeth  on  Cape  May,  New 
Jersey  (39°  18'  N,  74°  58'  W)  (Patuxent  Bird 
Banding  Eaboratory,  Maryland).  In  1999,  we 
found  three  nest  boxes  occupied  by  Northern 
Saw-whet  Owls.  In  one  nest,  egg-laying  oc- 
curred in  early  April,  and  in  two  others  it  oc- 
curred at  the  beginning  of  May.  On  15  De- 
cember 1999,  we  captured  a hatching-year 
male  by  using  an  audio  lure  and,  on  24  June 
2000,  we  found  two  partially  hatched  clutch- 
es, indicating  that  egg-laying  had  occurred  be- 
tween 22  and  26  May.  No  breeding  attempts 


411 


412 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


TABLE 

1.  Nesting  records  for  Northern  Saw- whet  Owls  in 

the  Mingan  Region,  Quebec  (1998-2003). 

Year 

No.  eggs 

No.  fledged 

Location 

Latitude  (N) 

Longitude  (W) 

1998 

>4 

2 

Riviere-Saint-Jean 

50°20'31" 

64°26'38" 

1999 

>4 

4 

Riviere-Saint-Jean 

50°  18 '03" 

64°21'57" 

1999 

6 

5 

Longue-Pointe-de-Mingan 

50°16'24" 

64°08'44" 

1999 

4a 

2 

Longue-Pointe-de-Mingan 

50°16'25" 

64°08'45" 

2000 

3 

2 

Riviere-Saint-Jean 

50°  18 '03" 

64°21'55" 

2000 

3 

3 

Magpie  River 

50°19'12" 

64°28'07" 

200  lb 

— 

— 

— 

— 

— 

2002c 

>1 

>1 

Longue-Pointe-de-Mingan 

50°16'06" 

64°  12 '49" 

2002 

>1 

>1 

Longue-Pointe-de-Mingan 

50°  15 '40" 

64°09'41" 

2003 

6 

6 

Riviere-Saint-Jean 

50°  18 '03" 

64°21'55" 

a Two  eggs  abandoned. 
b No  nesting  attempts. 

c In  2002,  four  other  owl  nesting  attempts  were  recorded,  but  species  was  not  determined  (Association  Le  Balbuzard,  Riviere-Saint-Jean,  Quebec). 


were  recorded  in  2001.  During  a post-breed- 
ing check  of  nest  boxes  in  2002,  we  found  six 
Aegolius  nests,  including  two  Northern  Saw- 
whet  Owl  nests — identified  by  the  abandoned 
eggs  and  dead  nestlings  inside.  Finally,  on  23 
July  2003,  one  partially  hatched  Northern 
Saw-whet  Owl  clutch  (six  eggs)  was  recorded 
at  Riviere-Saint-Jean,  suggesting  that  egg-lay- 
ing likely  occurred  21-26  June;  on  24  August, 
three  young  had  fledged  and  three  were  still 
in  the  nest  box.  Overall,  the  Northern  Saw- 
whet  Owl  nests  we  found  contained  4.4  eggs 
± 1.5  SE  (range  = 3-6,  n — 5)  and  fledged 
3.4  young  ± 1.6  SE  (range  = 2-6,  n = 7). 
All  nest  boxes  were  located  in  forested  habi- 
tats within  5 km  of  the  St.  Lawrence  River. 


The  area  is  underlain  by  old  marine  deposits 
and  characterized  by  bogs,  conifer  forests 
(balsam  fir-black  spruce  and  balsam  fir-white 
birch),  and  igneous  rocky  hills  and  terraces 
rarely  >300  m in  elevation.  Egg-laying  dates 
ranged  from  early  April  to  late  June,  indicat- 
ing variable  breeding  conditions  between 
years. 

The  discovery  of  a Northern  Saw-whet  Owl 
nesting  population  on  the  north  shore  of  the 
St.  Lawrence  River  extends  the  species’ 
known  breeding  range  to  >50°  N latitude  (Fig. 
1).  We  have  no  data  indicating  that  this  rep- 
resents a recent  expansion  of  the  owl’s  range; 
more  likely,  our  observations  are  refinements 
of  what  is  known  about  the  limits  of  its  nor- 


FIG.  1.  Previous  northern  limit  of  known  breeding  range,  and  nest-site  locations,  of  Northern  Saw-whet 
Owls  in  the  Mingan  Region,  north  shore  of  the  St.  Lawrence  River,  Quebec  (1998-2003). 


SHORT  COMMUNICATIONS 


413 


mal  range.  The  Mingan  Terraces  were  thought 
to  be  inhabited  primarily  by  Boreal  Owls,  al- 
though, both  Boreal  and  Northern  Saw-whet 
owls  use  coastal  areas  and  even  nest  in  similar 
habitats.  Each  fall,  however,  southern  move- 
ments of  Northern  Saw-whet  Owls  are  ob- 
served along  the  north  shore  of  the  St.  Law- 
rence, whereas  southern  movements  by  Boreal 
Owls  occur  only  about  every  4 years  (Obser- 
vatoire  d’oiseaux  de  Tadoussac:  http://www. 
explos-nature.qc.ca/ootyindex_f.htm). 

In  North  America,  the  breeding  ranges  of 
Northern  Saw-whet  and  Boreal  owls  overlap 
broadly  in  western  mountain  ranges,  although 
Boreal  Owls  tend  to  occupy  the  higher  ele- 
vations (Palmer  1986,  Cannings  1993).  In 
some  years.  Northern  Saw-whet  Owls  estab- 
lish territories  adjacent  to  those  of  Boreal 
Owls  at  higher  elevations  in  British  Columbia 
(R.  J.  Cannings  pers.  comm.),  and  territorial 
overlap  between  the  two  species  has  been  doc- 
umented along  the  southern  edge  of  the  boreal 
forest  in  Minnesota  (Lane  and  McKeown 
1991).  Clearly,  the  cohabitation  of  these  close- 
ly related  species  in  Quebec  deserves  further 
study. 

ACKNOWLEDGMENTS 

We  thank  the  Canadian  Wildlife  Service,  Ministere 
des  Resources  naturelles  et  de  la  Faune,  Parc  national 
du  Saguenay,  Observatoire  d’oiseaux  de  Tadoussac, 
Explos-Nature,  the  Mingan  Archipelago  National  Park, 


and  the  Caisse  Populaire  Desjardins  de  Mingan-Anti- 
costi  for  their  financial  and  logistical  assistance  and/or 
data.  We  thank  M.  Pierre-Alain  Ravussin  for  his  advice 
at  the  beginning  of  our  nest-box  program.  We  thank 
R.  J.  Cannings,  J.  S.  Marks,  B.  Drolet,  and  an  anon- 
ymous reviewer  for  their  comments  and  M.  Melangon 
for  his  help  with  the  figure.  We  thank  the  volunteers 
that  participated  in  the  monitoring  of  nest  boxes:  S. 
Angel,  M.  Bourdon,  M.-H.  Lattaro,  L.  Lefebvre,  and 
V.  Vogel. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1998.  Check-list 
of  North  American  birds,  7th  ed.  American  Or- 
nithologists’ Union,  Washington,  D.C. 

Cannings,  R.  J.  1993.  Northern  Saw-whet  Owl  ( Ae - 
golius  acadicus ).  The  Birds  of  North  America,  no. 
42. 

Cote,  A.  and  M.  Bombardier.  1996.  Northern  Saw- 
whet  Owl.  Pages  618-621  in  The  breeding  birds 
of  Quebec:  atlas  of  the  breeding  birds  of  southern 
Quebec  (J.  Gauthier  and  Y.  Aubry,  Eds.).  Asso- 
ciation quebecoise  des  groupes  d’ornithologues. 
Province  of  Quebec  Society  for  the  Protection  of 
Birds,  Canadian  Wildlife  Service,  Environment 
Canada,  Quebec  Region,  Montreal. 

Godfrey,  W.  E.  1986.  Birds  of  Canada,  revised  ed. 
National  Museum  of  Natural  Sciences,  National 
Museums  of  Canada,  Ottawa,  Ontario. 

Lane,  B.  and  S.  McKeown.  1991.  Physical  interac- 
tions between  a male  Boreal  Owl  and  a male 
Northern  Saw- whet  Owl.  Loon  63:74-75. 
Palmer,  D.  A.  1986.  Habitat  selection,  movements, 
and  activity  of  Boreal  and  Saw-whet  owls.  M.Sc. 
thesis,  Colorado  State  University,  Fort  Collins. 


The  Wilson  Journal  of  Ornithology  1 18(3):413— 415,  2006 


Carolina  Wren  Nest  Successfully  Parasitized  by  House  Finch 

Douglas  R.  Wood13  and  William  A.  Carter1 2 3 


ABSTRACT. — We  report  the  first  observation  of 
a House  Finch  ( Carpodacus  mexicanus ) successful- 
ly parasitizing  a Carolina  Wren  ( Thryothorus  ludov- 
icianus)  nest.  On  24  May  2005,  we  found  a Carolina 
Wren  nest  in  south-central  Oklahoma  containing 
four  Carolina  Wren  eggs  and  two  House  Finch  eggs. 


1 Southeastern  Oklahoma  State  Univ.,  Dept,  of  Bi- 
ological Sciences,  PMB  4068,  1405  N.  4th  Ave.,  Du- 
rant, OK  74701-0609,  USA. 

2P.O.  Box  2209,  Ada,  OK  74821-2209,  USA. 

3 Corresponding  author;  e-mail:  dwood@sosu.edu 


The  House  Finch  eggs  hatched  and  nestlings  grew 
rapidly.  The  Carolina  Wren  eggs  hatched  but  the 
young  did  not  survive.  We  observed  a House  Finch 
fledgling  with  the  adult  Carolina  Wrens  the  day  after 
fledging.  Received  29  August  2005,  accepted  14 
March  2006. 


House  Finches  ( Carpodacus  mexicanus ) 
expanded  their  range  into  central  Oklahoma 
by  the  1990s  (Reinking  2004).  Typically, 


414 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


House  Finches  nest  near  human  habitation  and 
lay  an  average  of  four  eggs;  the  incubation 
period  is  13-14  days,  and  young  fledge  11- 
14  days  after  hatching.  This  species  has  been 
documented  as  an  occasional  interspecific 
brood  parasite;  however,  there  are  no  records 
of  House  Finches  successfully  parasitizing  an- 
other species  (i.e.,  a host  species  fledging 
House  Finch  young;  Shepardson  1915,  Hol- 
land 1923,  Woods  1968).  Therefore,  our  ob- 
servation of  a Carolina  Wren  ( Thryothorus  lu- 
dovicianus ) pair  successfully  fledging  two 
House  Finch  young  is  noteworthy. 

The  Carolina  Wren  is  a regular  breeding 
species  in  south-central  Oklahoma  (Reinking 
2004)  and  builds  a nest  of  various  materials 
in  a wide  variety  of  nest  sites.  Typically,  Car- 
olina Wrens  lay  four  eggs  that  hatch  in  ap- 
proximately 15  days  (Haggerty  and  Morton 
1995).  Brown-headed  Cowbirds  ( Molothrus 
ater ) occasionally  parasitize  Carolina  Wrens 
in  Oklahoma  (Bent  1948),  and  Carolina  Wrens 
have  successfully  incubated  cowbird  eggs  and 
fledged  cowbird  young  (Grzybowski  1995, 
Haggerty  and  Morton  1995). 

On  24  May  2005  at  16:15  CST,  we  flushed 
a Carolina  Wren  from  a nest  located  northeast 
of  Ada,  Pontotoc  County,  Oklahoma  (34°  49' 
N,  96°  36'  W).  The  nest  was  1.87  m above  the 
ground,  nestled  between  a branch  and  the  wall 
of  a chimney,  semi-domed,  and  constructed  of 
twigs,  leaves,  and  grass.  In  2003  and  2004, 
the  same  nest  site  was  used  by  a pair  of  Car- 
olina Wrens  that  were  banded  in  2003.  The 
nest  contained  four  Carolina  Wren  eggs  (mean 
size  = 19.5  X 15  mm)  and  two  House  Finch 
eggs  (23  X 16  mm  and  21  X 16  mm).  We 
determined  that  they  were  House  Finch  eggs 
based  on  size,  blue  color,  and  maculation  pat- 
tern (Baicich  and  Harrison  1997).  One  desic- 
cated Carolina  Wren  egg  was  found  just  out- 
side the  nest  and  was  not  present  the  follow- 
ing day. 

The  House  Finch  eggs  hatched  on  3 June 
and  two  Carolina  Wren  eggs  hatched  on  6 
June.  By  7 June,  a third  Carolina  Wren  egg 
had  hatched  and,  on  8 June,  only  two  House 
Finch  nestlings  and  one  unhatched  Carolina 
Wren  egg  remained  in  the  nest.  We  removed 
the  remaining  unhatched  wren  egg  and  deter- 
mined that  it  was  infertile;  we  found  no  em- 
bryo in  the  contents.  Prior  to  banding  the  nest- 
lings, we  definitively  identified  them  as  House 


Finches  based  on  size,  plumage,  bill  shape, 
and  general  morphology  (Hill  1993). 

We  observed  the  adult  wrens  feeding  in- 
sects and  insect  larvae  to  the  finch  nestlings. 
We  did  not  observe  adult  House  Finches  feed- 
ing the  nestlings,  although  adult  finches  used 
nearby  feeders  with  black  oil  sunflower  seeds. 
Typically,  House  Finch  nestlings  are  raised  on 
a diet  composed  of  seeds  (Beal  1907);  how- 
ever, our  observation  suggests  that  House 
Finch  nestlings  can  be  raised  on  a diet  of  pri- 
marily soft-bodied  insects  and  insect  larvae. 
On  13  June,  both  House  Finch  nestlings 
fledged  and  remained  within  10  m of  the  nest. 
We  observed  the  adult  wrens  feed  the  fledg- 
lings and  give  alarm  calls  when  we  ap- 
proached. On  14  June,  we  observed  the  adult 
wrens  foraging  and  feeding  one  House  Finch 
fledgling  50  m from  the  nest  site;  we  did  not 
observe  the  House  Finch  fledglings  after  that 
day. 

House  Finches  have  been  documented  as 
interspecific  brood  parasites  of  Black  Phoebe 
(Sayornis  nigricans ),  Cliff  Swallow  ( Petro - 
chelidon  pyrrhonota),  and  Hooded  Oriole  ( Ic- 
terus cucullatus ) (Shepardson  1915,  Holland 
1923);  to  our  knowledge,  however,  our  report 
is  the  first  to  document  House  Finch  nestlings 
fledging  from  a host  species’  nest.  Although 
House  Finches  intentionally  parasitize  and 
usurp  the  nests  of  other  species,  we  cannot 
exclude  the  possibility  that  egg  dumping  may 
be  an  alternate  explanation  for  our  observa- 
tion. Interspecific  egg  dumping  has  been  doc- 
umented for  a variety  of  passerines.  Wiens 
(1971)  reported  egg  dumping  by  a Grasshop- 
per Sparrow  ( Ammodramus  savannarum ) in  a 
Savannah  Sparrow  ( Passerculus  sandwichen- 
sis ) nest,  and  Sealy  (1989)  documented  egg 
dumping  by  a House  Wren  ( Troglodytes  ae- 
doh)  in  a Yellow  Warbler  (Dendroica  pete- 
chia) nest.  Hamilton  and  Orians  (1965)  spec- 
ulated that  egg  dumping  is  the  first  step  to- 
wards facultative  brood  parasitism  and,  even- 
tually, obligate  brood  parasitism. 

ACKNOWLEDGMENTS 

We  thank  D.  W.  Pogue,  M.  D.  Duggan,  and  three 
anonymous  reviewers  for  providing  comments  on  this 
manuscript. 

LITERATURE  CITED 

Baicich,  P.  J.  and  C.  J.  O.  Harrison.  1997.  A guide 
to  the  nests,  eggs,  and  nestlings  of  North  Ameri- 
can birds.  Academic  Press,  San  Diego,  California. 


SHORT  COMMUNICATIONS 


415 


Beal,  F.  E.  L.  1907.  Birds  of  California  in  relation  to 
fruit  industry.  U.S.  Department  of  Agriculture  Bi- 
ological Survey  Bulletin  30:13-17. 

Bent,  A.  C.  1948.  Thryothorus  ludovicianus  ludovi- 
cianus  (Latham),  Carolina  Wren.  Pages  205-216 
in  Life  histories  of  North  American  nuthatches, 
wrens,  thrashers,  and  their  allies.  U.S.  National 
Museum  Bulletin,  no.  195,  Smithsonian  Institute, 
Washington,  D.C. 

Grzybowski,  J.  A.  1995.  Carolina  Wrens  fledge 
Brown-headed  Cowbird  chick.  Bulletin  of  the 
Oklahoma  Ornithological  Society  28:6-7. 

Haggerty,  T.  M.  and  E.  S.  Morton.  1995.  Carolina 
Wren  ( Thryothorus  ludovicianus).  The  Birds  of 
North  America,  no.  188. 

Hamilton,  W.  J.  and  G.  H.  Orians.  1965.  Evolution 
of  brood  parasitism  in  altricial  birds.  Condor  67: 
361-382. 

Hill,  G.  E.  1993.  House  Finch  ( Carpodacus  mexican- 
us).  The  Birds  of  North  America,  no.  46. 


Holland,  H.  M.  1923.  Black  phoebes  and  house  finch- 
es in  joint  use  of  a nest.  Condor  25:131-132. 

Reinking,  D.  L.  (Ed.).  2004.  Oklahoma  breeding  bird 
atlas.  University  of  Oklahoma  Press,  Norman. 

Sealy,  S.  G.  1989.  Incidental  “egg  dumping”  by  the 
House  Wren  in  a Yellow  Warbler  nest.  Wilson 
Bulletin  101:491-493. 

Shepardson,  D.  I.  1915.  The  house  finch  as  a parasite. 
Condor  17:100-101. 

Wiens,  J.  A.  1971.  “Egg-dumping”  by  the  Grasshop- 
per Sparrow  in  a Savannah  Sparrow  nest.  Auk  88: 
185-186. 

Woods,  R.  S.  1968.  Carpodacus  mexicanus  frontalis 
(Say),  House  Finch.  Pages  290-314  in  Life  his- 
tories of  North  American  cardinals,  grosbeaks, 
buntings,  towhees,  finches,  sparrows,  and  allies 
(O.  L.  Austin,  Jr.,  Ed.).  U.S.  National  Museum 
Bulletin,  no.  237,  Smithsonian  Institution,  Wash- 
ington, D.C. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):4 1 5 — 4 1 8,  2006 


American  Coot  Parasitism  on  Least  Bitterns 

Brian  D.  Peer1 


ABSTRACT. — American  Coots  ( Fulica  americana) 
are  known  for  laying  eggs  in  the  nests  of  conspecifics, 
but  there  is  little  evidence  that  they  regularly  parasitize 
the  nests  of  other  species.  I found  13  Least  Bittern 
(. Ixobrychus  exilis ) nests,  2 of  which  were  parasitized 
by  coots.  These  are  the  first  records  of  coots  parasit- 
izing Least  Bitterns,  and  the  first  records  of  any  form 
of  brood  parasitism  on  Least  Bitterns.  Nests  of  Least 
Bitterns  also  were  parasitized  experimentally  with  a 
variety  of  nonmimetic  eggs  and  27%  were  rejected  (n 
= 1 1 nests).  This  indicates  that  Least  Bitterns  may 
possess  some  egg  recognition  abilities.  Received  15 
August  2005,  accepted  21  March  2006. 


Facultative  avian  brood  parasites  build 
nests  and  raise  their  own  young,  but  they  also 
lay  eggs  in  the  nests  of  conspecifics  (conspe- 
cific  brood  parasitism;  CBP)  and  sometimes 
in  the  nests  of  other  species  (interspecific 
brood  parasitism;  IBP).  CBP  has  been  docu- 
mented in  at  least  236  bird  species  (Yom-Tov 
200 1 ) and  appears  to  be  relatively  common  in 


1 Dept,  of  Biological  Sciences,  Western  Illinois 
Univ.,  Macomb,  IL  61455,  USA;  e-mail: 
BD-Peer@wiu.edu 


colonial  birds,  waterfowl,  and  cavity-nesters 
(MacWhirter  1989,  Rohwer  and  Freeman 
1989,  Yom-Tov  2001).  One  of  the  best-studied 
conspecific  brood  parasites  is  the  American 
Coot  ( Fulica  americana ; Arnold  1987;  Lyon 
1993a,  1993b,  2003).  CBP  appears  to  be  a rel- 
atively common  reproductive  strategy  among 
coots.  For  example,  Lyon  (1993a)  found  that 
13%  of  all  coot  eggs  over  a 4-year  period 
were  laid  parasitically  and  more  than  40%  of 
nests  were  parasitized  by  conspecifics.  The 
parasites  are  females  with  nesting  territories 
that  lay  parasitically  prior  to  laying  eggs  in 
their  own  nests,  and  floater  females  that  are 
unable  to  acquire  nesting  territories  of  their 
own  (Lyon  1993a). 

On  rare  occasions,  coots  have  been  known 
to  lay  eggs  in  the  nests  of  other  species.  To 
date,  three  host  species  have  been  recorded: 
Franklin’s  Gull,  (. Larus  pipixcan ; Burger  and 
Gochfeld  1994),  and  Cinnamon  Teal  {Anas  cy- 
anoptera)  and  Redhead  {Aythya  americana ) 
(Joyner  1973).  It  is  unknown  whether  any  of 
these  cases  of  parasitism  were  successful,  al- 
though coot  chicks  are  dependent  on  their  par- 


416 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  3,  September  2006 


TABLE 
Iowa,  2003- 

1.  Responses  of  Least  Bitterns 
-2004. 

to  natural  and  experimental  brood  parasitism 

in  Warren  County, 

Nest 

Host’s  clutch 
size  when 
parasitized 

Nesting  stage 
when 

parasitized 

Egg  type  added 

Accepted 
or  rejected 

03-3 

5 

Incubation 

Plaster  cowbird  egg 

Rejected 

03-16 

5 

Incubation 

Least  Bittern  egg  colored  black 

Accepted 

03-18 

6 

Unknown 

Two  naturally  laid  coot  eggs 

Accepted?3 

03-19 

6 

Incubation 

Wooden  egg  colored  black 

Rejected 

03-20 

3 

Laying 

Least  Bittern  egg  colored  black 

Accepted 

03-22 

4 

Unknown 

One  naturally  laid  coot  egg 

Accepted 

03-31 

5 

Laying 

One  coot  egg  placed  in  the  nest 

Accepted 

03-32 

6 

Incubation 

Wooden  egg  colored  black 

Accepted 

03-34 

6 

Incubation 

One  coot  egg  placed  in  the  nest 

Accepted 

04-49 

2 

Laying 

One  coot  egg  placed  in  the  nest 

Accepted 

04-55 

4 

Incubation 

Wooden  egg  colored  black 

Rejected 

One  of  two  coot  eggs  disappeared  from  this  nest  along  with  two  Least  Bittern  eggs. 


ents  for  food  and  typically  perish  without  their 
assistance  (Brisbin  et  al.  2002);  thus,  it  is  un- 
likely that  these  instances  of  parasitism  were 
successful  (B.  E.  Lyon  pers.  comm.).  I report 
the  first  records  of  American  Coot  parasitism 
on  Least  Bitterns  ( Ixobrychus  exilis).  I also 
experimentally  parasitized  Least  Bittern  nests 
to  determine  whether  bitterns  possess  defens- 
es, such  as  egg  rejection,  against  parasitism. 

METHODS 

This  study  was  conducted  in  a restored  wet- 
land in  Warren  County,  Iowa,  just  north  of 
Indianola  (41°  4' N,  93°  6' W),  in  2003  and 
2004.  The  dominant  vegetation  consisted  of 
cattails  ( Typha  spp.)  and  willows  ( Salix  spp.), 
and  water  depth  was  <1.5  m.  Nests  of  Least 
Bitterns,  American  Coots,  Pied-billed  Grebes 
( Podilymbus  podiceps ),  and  passerines  such  as 
Great-tailed  Grackles  ( Quiscalus  mexicanus ), 
Yellow-headed  Blackbirds  ( Xanthocephalus 
xanthocephalus ),  Red-winged  Blackbirds 
(Agelaius  phoeniceus ),  and  Marsh  Wrens 
( Cistothorus  palustris ) were  monitored  every 
1-3  days. 

I also  experimentally  parasitized  Least  Bit- 
tern nests  with  a variety  of  egg  types  during 
laying  and  incubation  to  determine  their  re- 
sponses to  parasitism.  These  eggs  included  (1) 
the  Least  Bittern’s  own  eggs  (31  X 24  mm; 
Baicich  and  Harrison  1997)  colored  black 
with  permanent-ink  markers  to  make  them 
nonmimetic,  (2)  real  coot  eggs  (49  X 34  mm; 
Baicich  and  Harrison  1997),  (3)  wooden  eggs 
colored  black  (34  X 22  mm),  and  (4)  plaster 


eggs  (21  X 16  mm)  made  to  look  like  those 
of  the  Brown-headed  Cowbird  ( Molothrus 
ater ; Table  1).  The  latter  two  egg  types  have 
been  used  in  similar  egg-recognition  experi- 
ments (Rothstein  1975,  Peer  and  Bollinger 
1998,  Peer  and  Sealy  2001).  Only  one  egg 
type  was  added  to  each  nest.  Experimentally 
parasitized  nests  were  checked  every  1-3  days 
to  determine  the  responses  of  Least  Bitterns. 
Eggs  were  considered  rejected  if  they  were 
missing  from  the  nest  after  it  was  parasitized. 

RESULTS 

Coots  parasitized  18.2%  ( n = 11)  of  Least 
Bittern  nests  in  2003  and  no  nests  ( n = 3)  in 
2004.  The  first  parasitized  nest  contained  six 
bittern  eggs  and  two  coot  eggs  when  found. 
Four  bittern  eggs  hatched,  and  two  bittern 
eggs  and  one  coot  egg  disappeared.  The  sec- 
ond parasitized  bittern  nest  was  found  con- 
taining four  young  bitterns  and  a coot  egg  that 
never  hatched.  Both  parasitized  nests  were  lo- 
cated near  the  water  level,  whereas  the  unpar- 
asitized bittern  nests  were  at  least  30-60  cm 
above  the  water  level.  Seven  Pied-billed 
Grebe  nests,  15  coot  nests,  and  1 unidentified 
duck  nest  also  were  monitored,  but  there  was 
no  evidence  of  parasitism  on  these  nests. 

The  single  artificial  cowbird  egg  that  was 
added  to  a bittern  nest  was  rejected  the  fol- 
lowing day,  as  were  two  of  three  black  wood- 
en eggs  (10  and  13  days;  Table  1).  None  of 
the  colored  bittern  eggs  was  rejected  ( n = 2) 
and  only  one  coot  egg  may  have  been  rejected 


SHORT  COMMUNICATIONS 


417 


within  8 days  after  it  was  found  (n  — 5;  Table 

1). 

DISCUSSION 

These  are  the  first  reported  instances  of 
American  Coot  parasitism  on  Least  Bitterns 
(see  Gibbs  et  al.  1992)  and  the  first  record  of 
any  form  of  brood  parasitism  on  Least  Bit- 
terns. The  Least  Bittern  is  likely  an  unsuitable 
host  for  the  coot  because  the  bittern’s  incu- 
bation period  is  17-20  days  (Gibbs  et  al. 
1992)  and  the  coot’s  is  23-27  days  (Brisbin 
et  al.  2002);  thus,  any  coot  eggs  laid  in  bittern 
nests  would  not  have  sufficient  time  to  devel- 
op and  hatch.  Indeed,  two  of  the  parasitic  coot 
eggs  did  not  hatch  and  the  fate  of  the  third 
egg  was  unclear  (see  discussion  below).  It  is 
also  unlikely  that  a coot  would  be  fed  properly 
or  receive  adequate  parental  care  from  a Least 
Bittern,  in  which  case  it  would  probably  die 
if  the  egg  did  hatch  (Brisbin  et  al.  2002). 

Why  would  coots  lay  their  eggs  in  a poten- 
tially unsuitable  host’s  nest?  It  is  possible  that 
the  coot  eggs  I observed  were  laid  by  floater 
females  (B.  E.  Lyon  pers.  comm.),  as  floater 
females  are  unable  to  obtain  their  own  nesting 
territories  and  presumably  attempt  to  make  the 
best  of  a bad  situation  by  practicing  CBP 
(Lyon  1993a).  Such  females  may  be  unable  to 
locate  and  successfully  parasitize  other  coots 
and  are  forced  to  parasitize  the  nests  of  un- 
suitable hosts  (e.g.,  bitterns).  Interestingly,  the 
two  parasitized  nests  that  I observed  were 
very  near  water  level — similar  to  the  floating 
platform  nests  used  by  coots.  The  coots  that 
parasitized  the  bittern  nests,  or  other  coots  in 
the  population,  also  may  have  been  practicing 
CBP.  Lyon  (1993a)  found  that  the  reproduc- 
tive success  of  floater  females  was  only  6% 
of  that  of  nesting  females,  and  only  3.6%  of 
parasitic  eggs  produced  by  floaters  produced 
young.  The  reasons  for  the  lower  reproductive 
success  of  floaters  were  the  anti-parasite  be- 
havior of  hosts  (rejected  38%  of  floater  eggs) 
and  the  timing  of  laying:  floaters  tended  to  lay 
late  in  the  host’s  nesting  cycle  (Lyon  1993a). 
CBP  in  general  is  not  a very  successful  strat- 
egy among  coots,  as  only  7.7%  of  all  parasitic 
eggs  produced  young  that  survived  (Lyon 
1993b);  however,  territorial  females  can  in- 
crease their  reproductive  success  by  laying 
eggs  in  the  nests  of  neighbors.  Brood  reduc- 
tion is  common  in  coots;  thus,  by  laying  eggs 


in  the  nests  of  conspecifics,  they  maximize 
their  reproductive  success  (Lyon  1993a). 

Least  Bitterns  rejected  some  of  the  foreign 
eggs  placed  into  their  nests.  One  of  the  natu- 
rally laid  coot  eggs  disappeared  from  a nest, 
but  it  is  unclear  whether  this  was  due  to  re- 
jection, partial  predation,  or  the  coot  chick 
hatching  and  leaving  the  nest.  Bitterns  reject- 
ed two  of  three  wooden  eggs  and  the  artificial 
cowbird  egg.  The  latter  may  have  been  so 
small  that  the  bitterns  viewed  it  as  debris  and 
removed  it  from  the  nest;  however,  the  wood- 
en eggs  were  approximately  the  same  size  as 
the  bittern  eggs,  indicating  that  bitterns  may 
possess  some  recognition  abilities.  Bitterns 
did  not  remove  any  of  their  own,  colored  eggs 
or  any  coot  eggs.  Egg  recognition  in  this  spe- 
cies deserves  further  study. 

ACKNOWLEDGMENTS 

I would  like  to  thank  B.  Lyon,  P.  Lowther,  and  an 
anonymous  reviewer  for  helpful  comments  on  the 
manuscript. 

LITERATURE  CITED 

Arnold,  T.  W.  1987.  Conspecific  egg  discrimination 
in  American  Coots.  Condor  89:675-676. 

Baicich,  P.  J.  and  C.  J.  O.  Harrison.  1997.  A guide 
to  the  nests,  eggs,  and  nestlings  of  North  Ameri- 
can Birds,  2nd  ed.  Academic  Press,  New  York. 
Brisbin,  I.  L.,  Jr.,  H.  D.  Pratt,  and  T.  B.  Mowbray. 
2002.  American  Coot  ( Fulica  americana)  and  Ha- 
waiian Coot  ( Fulica  alai).  The  Birds  of  North 
America,  no.  697. 

Burger,  J.  and  M.  Gochfeld.  1994.  Franklin’s  Gull 
(Larus  pipixcan).  The  Birds  of  North  America,  no. 
116. 

Gibbs,  J.  P.,  F.  A.  Reid,  and  S.  M.  Melvin.  1992.  Least 
Bittern  ( Ixobrychus  exilis).  The  Birds  of  North 
America,  no.  17. 

Joyner,  D.  E.  1973.  Interspecific  nest  parasitism  by 
ducks  and  coots  in  Utah.  Auk  90:692-693. 

Lyon,  B.  E.  1993a.  Conspecific  brood  parasitism  as  a 
flexible  female  reproductive  tactic  in  American 
Coots.  Animal  Behaviour  46:91 1-928. 

Lyon,  B.  E.  1993b.  Tactics  of  parasitic  American 
Coots:  host  choice  and  the  pattern  of  egg  disper- 
sion among  host  nests.  Behavioral  Ecology  and 
Sociobiology  33:87-100. 

Lyon,  B.  E.  2003.  Egg  recognition  and  counting  re- 
duce costs  of  avian  conspecific  brood  parasitism. 
Nature  422:495-499. 

MacWhirter,  R.  B.  1989.  On  the  rarity  of  intraspecific 
brood  parasitism.  Condor  91:485-492. 

Peer,  B.  D.  and  E.  K.  Bollinger.  1998.  Rejection  of 
cowbird  eggs  by  Mourning  Doves:  a manifesta- 
tion of  nest  usurpation?  Auk  115:1057-1062. 


418 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Peer,  B.  D.  and  S.  G.  Sealy.  2001.  Mechanism  of  egg 
recognition  in  the  Great-tailed  Grackle  ( Quiscalus 
mexicanus).  Bird  Behavior  14:71-73. 

Rohwer,  F.  C.  and  S.  Freeman.  1989.  The  distribution 
of  conspecific  nest  parasitism  in  birds.  Canadian 
Journal  of  Zoology  67:239-253. 


Rothstein,  S.  I.  1975.  An  experimental  and  teleonom- 
ic  investigation  of  avian  brood  parasitism.  Condor 
77:250-271. 

Yom-Tov,  Y.  2001.  An  updated  list  and  some  com- 
ments on  the  occurrence  of  intraspecific  nest  par- 
asitism in  birds.  Ibis  143:133-143. 


The  Wilson  Journal  of  Ornithology  1 18(3):418— 419,  2006 


Brown-headed  Cowbird’s  Fatal  Attempt  to  Parasitize  a 
Carolina  Chickadee  Nest 

David  A.  Zuwerink12  and  James  S.  Marshall1 2 


ABSTRACT. — On  5 June  2003,  a female  Brown- 
headed Cowbird  ( Molothrus  ater ) was  found  dead  in 
a Carolina  Chickadee  ( Poecile  carolinensis ) cavity 
nest  near  Bucyrus  in  Crawford  County,  Ohio.  The 
cowbird  had  little  room  in  the  cavity  and  likely  could 
not  remove  itself  after  laying  an  egg.  Carolina  Chick- 
adee nests  are  rarely  parasitized  by  brood  parasites, 
and  the  size  of  their  cavity  entrances  likely  limits  par- 
asitism by  Brown-headed  Cowbirds.  This  is  the  first 
known  instance  of  a Brown-headed  Cowbird  mortality 
after  laying  an  egg  in  the  cavity  nest  of  a host  species. 
Received  6 September  2005,  accepted  21  March  2006. 


More  than  220  avian  species  reportedly 
have  been  parasitized  by  Brown-headed  Cow- 
birds  ( Molothrus  ater: ; Lowther  1993).  Where- 
as the  Carolina  Chickadee  ( Poecile  carolinen- 
sis) is  an  uncommon  host  species,  there  are  a 
few  records  of  Brown-headed  Cowbirds  par- 
asitizing that  species  (Friedmann  1938,  Goertz 
1977).  The  closely  related  Black-capped 
Chickadee  ( P . atricapillus ) also  has  been  par- 
asitized, and  individuals  have  been  observed 
feeding  Brown-headed  Cowbird  fledglings 
(Lowther  1983).  Such  observations  suggest 
that  these  chickadee  species  are  capable  of 
raising  the  young  of  Brown-headed  Cowbirds, 
but  that  some  mechanism  may  be  limiting 
Brown-headed  Cowbirds  from  taking  advan- 
tage of  these  potential  host  species  more  of- 
ten. Cavity  nesting  seems  to  offer  some  pro- 


1  Dept,  of  Evolution,  Ecology,  and  Organismal  Bi- 
ology, 318  W.  12th  St.,  Ohio  State  Univ.,  Columbus, 
OH  43210,  USA. 

2 Corresponding  author;  e-mail: 
zuwerink.  1 @osu.edu 


tection  from  brood  parasites,  as  cavity  nesters 
have  been  found  to  have  low  levels  of  para- 
sitism (Strausberger  and  Ashley  1997).  Fe- 
male Carolina  Chickadees  cover  their  eggs 
during  the  egg-laying  stage  (Brewer  1961), 
which  also  may  offer  protection  against  par- 
asitism. Studies  have  revealed  lower  levels  of 
parasitism  among  some  host  species  because 
they  reject  cowbird  eggs  (Strausberger  and 
Ashley  1997)  or  because  they  do  not  provide 
adequate  nutrition  to  cowbird  young  (Mills 
1988). 

During  2003,  we  monitored  a pair  of  color- 
banded  Carolina  Chickadees  nesting  in  natural 
cavities  in  a 2.63-ha  woodlot  located  in  Craw- 
ford County,  Ohio  (40°  46'  N,  82°  58'  W).  The 
landscape  is  dominated  by  agriculture,  with 
woodlots  scattered  throughout  the  county.  On 
5 June  2003,  we  discovered  a Carolina  Chick- 
adee nest  cavity  from  which  most  of  a dead 
female  Brown-headed  Cowbird’s  tail  was  pro- 
truding. The  cowbird  appeared  to  have  died 
only  a day  or  two  before  we  found  the  nest 
and  appeared  cramped  in  the  cavity.  The  cav- 
ity entrance  dimensions  were  38  mm  high  X 
42  mm  wide,  similar  to  average  dimensions 
previously  reported  for  Carolina  Chickadee 
cavity  entrances  (Brewer  1961,  Albano  1992, 
Mostrom  et  al.  2002).  The  cavity  was  155  mm 
deep,  and  the  nest  was  made  with  grass,  hair, 
feathers,  and  plant  down.  We  did  not  measure 
the  female  cowbird,  but  her  size  appeared  to 
be  normal.  Inspection  of  the  nest  confirmed 
that  the  cowbird  had  laid  one  egg,  but  we 
found  no  chickadee  eggs  in  the  nest.  Given 
the  depth  of  the  nest  cavity,  we  can  only  as- 


SHORT  COMMUNICATIONS 


419 


sume  that  the  cowbird  died  after  laying  the 
egg  because  she  had  no  room  to  move  inside 
the  cavity  and  remove  herself  after  entering 
the  nest. 

The  chickadees’  cavity  appeared  to  have 
been  freshly  excavated  and  the  nest  inside  was 
intact.  The  cavity  was  located  in  a dead 
branch  (130  mm  in  diameter  at  the  cavity  en- 
trance, broken  but  still  barely  attached  to  the 
tree)  that  was  hanging  1.2  m above  ground, 
and  the  opening  was  oriented  north-northeast. 
The  nest  tree  was  located  about  22  m from  the 
northern  edge  of  the  woodlot.  Two  adult 
chickadees  were  heard  nearby,  but  if  they 
were  the  original  cavity  occupants,  it  appeared 
they  had  already  abandoned  the  nest.  This  was 
the  third  known  nesting  attempt  by  this  pair 
of  chickadees  in  2003.  The  first  nest  was  dis- 
covered on  18  April,  when  one  of  the  chick- 
adees was  observed  entering  a cavity.  On  24 
April,  their  nest  appeared  to  be  complete  and 
covered,  suggesting  they  had  laid  at  least  one 
egg.  On  28  April,  the  nest  was  gone  and  a few 
sticks  were  found  in  the  cavity.  A House  Wren 
( Troglodytes  aedon ) eventually  completed  a 
nest  and  laid  eggs  in  the  same  cavity.  On  4 
May,  again  the  chickadee  pair  was  observed 
building  a new  nest  in  a freshly  excavated 
cavity.  On  13  May,  the  nest  had  been  removed 
by  a House  Wren  and  sticks  were  placed  in 
the  cavity.  There  was  no  indication  that  the 
chickadees  had  laid  eggs  in  the  nest. 

The  small  entrances  of  chickadee  nest  cav- 
ities likely  prevent  most  Brown-headed  Cow- 
birds  from  even  attempting  to  parasitize  their 
nests.  Pribil  and  Pieman  (1997)  showed  that 
the  size  of  cavity  entrances  could  limit  a 
Brown-headed  Cowbird’s  ability  to  parasitize 
House  Wren  nests.  They  proposed  that  a 38- 
mm-diameter  hole  was  the  smallest  that  a 
Brown-headed  Cowbird  could  voluntarily 
exit;  however,  they  had  placed  the  cowbirds 
in  a nesting  box  (12  X 10  X 20  cm),  which 
provided  enough  room  for  the  birds  to  orient 
themselves  toward  the  exit  hole.  If  a cowbird 
is  cramped  in  a cavity — as  we  observed — it 
may  not  be  able  to  turn  and  face  the  cavity 
opening,  making  it  more  difficult  to  remove 
itself  from  the  cavity.  One  record  of  a para- 


sitized Black-capped  Chickadee  nest  indicated 
that  the  cavity  entrance  was  larger  than  nor- 
mal, allowing  intrusion  by  a cowbird  (Packard 
1936).  Whereas  some  cavities  may  permit  en- 
try by  Brown-headed  Cowbirds,  most  cow- 
birds  may  not  attempt  to  parasitize  such  nests 
because  of  the  difficulty  in  removing  them- 
selves from  the  nests  after  they  have  com- 
pletely entered  the  cavities.  This  is  the  first 
reported  instance  of  a Brown-headed  Cowbird 
mortality  after  egg-laying  in  the  nest  of  a cav- 
ity-nesting species. 

ACKNOWLEDGMENTS 

This  observation  was  made  during  research  funded 
by  the  Ohio  Department  of  Natural  Resources,  Divi- 
sion of  Wildlife,  and  by  the  Columbus  Zoo  and  Aquar- 
ium. We  also  would  like  to  thank  P.  E.  Lowther,  B.  M. 
Strausberger,  and  A.  M.  Mostrom  for  their  comments, 
which  improved  this  manuscript. 

LITERATURE  CITED 

Albano,  D.  J.  1992.  Nesting  mortality  of  Carolina 
Chickadees  breeding  in  natural  cavities.  Condor 
94:371-382. 

Brewer,  R.  1961.  Comparative  notes  on  the  life  his- 
tory of  the  Carolina  Chickadee.  Wilson  Bulletin 
73:348-373. 

Friedmann,  H.  1938.  Additional  hosts  of  the  parasitic 
cowbirds.  Auk  55:41-50. 

Goertz,  J.  W.  1977.  Additional  records  of  Brown- 
headed Cowbird  nest  parasitism  in  Louisiana.  Auk 
94:386-389. 

Lowther,  P.  E.  1983.  Chickadee,  thrasher,  and  other 
cowbird  hosts  from  northwest  Iowa.  Journal  of 
Field  Ornithology  54:414-417. 

Lowther,  P.  E.  1993.  Brown-headed  Cowbird  (Mol- 
othrus  ater).  The  Birds  of  North  America,  no.  47. 
Mills,  A.  M.  1988.  Unsuitability  of  Tree  Swallows  as 
hosts  to  Brown-headed  Cowbirds.  Journal  of  Field 
Ornithology  59:331-333. 

Mostrom,  A.  M.,  R.  L.  Curry,  and  B.  Lohr.  2002. 
Carolina  Chickadee  ( Poecile  carolinensis ).  The 
Birds  of  North  America,  no.  636. 

Packard,  F.  M.  1936.  A Black-capped  Chickadee  vic- 
timized by  the  Eastern  Cowbird.  Bird-Banding  7: 
129-130. 

Pribil,  S.  and  J.  Picman.  1997.  Parasitism  of  House 
Wren  nests  by  Brown-headed  Cowbirds:  why  is  it 
so  rare?  Canadian  Journal  of  Zoology  75:302- 
307. 

Strausberger,  B.  M.  and  M.  V.  Ashley.  1997.  Com- 
munity-wide patterns  of  a host  “generalist” 
brood-parasitic  cowbird.  Oecologia  112:254-262. 


420 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


The  Wilson  Journal  of  Ornithology  1 18(3):420-422,  2006 


Likely  Predation  of  Adult  Glossy  Ibis  by  Great  Black-backed  Gulls 


Christina  E.  Donehower1 


ABSTRACT. — Great  Black-backed  Gulls  ( Larus 
marinus ) are  known  to  prey  upon  a wide  range  of  bird 
species,  particularly  adults,  young,  and  eggs  of  sea- 
birds and  waterfowl.  Here,  I provide  the  first  account 
of  Great  Black-backed  Gulls  pursuing  and  attacking, 
in  flight,  a medium-sized  wading  bird,  the  Glossy  Ibis 
( Plegadis  falcinellus ).  I recorded  two  observations  at 
Stratton  Island,  Maine,  the  northernmost  breeding  site 
for  the  Glossy  Ibis  in  North  America.  Received  12  Sep- 
tember 2005,  accepted  21  March  2006. 


Great  Black-backed  Gulls  ( Larus  marinus ) 
are  well-known  predators  of  colonial  water- 
birds.  Many  studies  have  attributed  heavy 
losses  of  seabird  and  waterfowl  eggs  and 
young  to  this  species  (Hatch  1970,  Menden- 
hall and  Milne  1985,  Mawhinney  and  Dia- 
mond 1999,  Whittam  and  Leonard  1999,  Mas- 
saro  et  al.  2000),  particularly  following  human 
disturbance  (Johnson  1938,  Kury  and  Goch- 
feld  1975,  Ahlund  and  Gotmark  1989,  Mikola 
et  al.  1994).  Great  Black-backed  Gulls  have 
also  been  observed  attacking  and  killing  adult 
waterfowl  (reviewed  in  Ryan  1990),  seabirds 
(Robinson  1930;  Snyder  1960;  Harris  1965, 
1980;  Pierotti  1983;  Russell  and  Montevecchi 
1996;  reviewed  in  Good  1998),  migrating  pas- 
serines (reviewed  in  Macdonald  and  Mason 
1973),  and  even  other  gulls  (Corkhill  1971; 
reviewed  in  Good  1998).  Large  birds  may  be 
seized  or  struck  on  the  wing  (Snyder  1960, 
Harris  1980,  Burger  and  Gochfeld  1984,  Ryan 
1990),  harassed  and  pursued  on  the  water 
(Addy  1945,  Sobkowiak  1986,  Ryan  1990),  or 
surprised  on  land  (Robinson  1930,  Snyder 
1960).  Here,  I describe  the  first  observation  of 
Great  Black-backed  Gulls  (length  71-79  cm, 
wingspan  152-167  cm,  mass  1,300-2,000  g; 
Good  1998)  attacking  adult  Glossy  Ibis  ( Ple- 
gadis falcinellus ),  a medium-sized  wading 


1 Dept,  of  Natural  Resource  Sciences,  Macdonald 
Campus,  McGill  Univ.,  21111  Lakeshore  Rd.,  Ste- 
Anne-de-Bellevue,  QC  H9X  3V9,  Canada;  e-mail: 
christina.donehower@mail.mcgill.ca 


bird  (length  48-66  cm,  wingspan  92  cm,  mass 
500-800  g;  Davis  and  Kricher  2000). 

On  15  June  2005,  I observed  two  aerial 
chases  in  which  Great  Black-backed  Gulls 
pursued  and  struck  Glossy  Ibis  in  flight.  Both 
events  were  recorded  on  a handheld  camcord- 
er ( Sony  Handy  cam  Vision  with  200  X digital 
zoom)  and  later  reviewed.  All  video  was  taken 
from  a 6-m-high  observation  tower  on  Strat- 
ton Island  (43°  31'  N,  70°  19'  W),  a 12-ha  Na- 
tional Audubon  Society  waterbird  sanctuary 
located  2.4  km  south  of  Prout’s  Neck,  Saco 
Bay,  Maine  (see  Kress  1998  and  Chase  1994 
for  a detailed  site  description  and  history). 
The  island  supports  approximately  100  breed- 
ing pairs  of  Glossy  Ibis  (C.  S.  Hall  pers. 
comm.)  and  represents  the  northernmost  nest- 
ing colony  for  this  species  in  North  America 
(Davis  and  Kricher  2000).  Although  gulls  do 
not  breed  on  Stratton  Island  (National  Audu- 
bon Society  gull  control  measures  include  nest 
destruction  and  shooting  of  gulls  seen  entering 
the  island’s  tern  colony),  more  than  400  Her- 
ring (L.  argentatus ) and  Great  Black-backed 
gulls  reside  on  Stratton  and  nearby  Bluff  Is- 
land— an  active,  unmanaged  gull  colony  less 
than  400  m away  (CED  unpubl.  data). 

Event  1. — At  15:30  EDT,  I observed  a Great 
Black-backed  Gull  adult  in  breeding  plumage 
chasing  an  adult  Glossy  Ibis  above  the  tree 
line  of  the  wading  bird  colony.  The  ibis  flew 
erratically,  climbing  high  and  then  low,  bank- 
ing and  trying  to  elude  the  gull.  The  aerial 
chase  continued  for  about  1 min,  at  which 
point  a second  Great  Black-backed  Gull  adult 
in  breeding  plumage  joined  in  the  pursuit.  At 
15:32,  the  latter  gull  struck  the  ibis  with  its 
bill,  hitting  it  with  such  force  that  the  ibis 
plummeted  to  the  ground  and  out  of  view.  I 
was  unable  to  determine  whether  one  or  both 
gulls  further  pursued  the  ibis. 

Event  2. — At  16:01,  I again  saw  an  adult 
Great  Black-backed  Gull  pursuing  an  ibis  in 
flight.  At  16:06,  a second  adult  Great  Black- 
backed  Gull  again  joined  in  the  chase  and 


SHORT  COMMUNICATIONS 


421 


struck  the  ibis  10-15  sec  later,  hitting  it  on 
the  back  near  the  rump  and  tearing  off  a small 
section  of  skin  and  feathers  with  its  bill.  The 
ibis  tumbled  out  of  the  air  and  fell  into  the 
vegetation.  The  latter  gull  immediately  fol- 
lowed the  ibis  into  the  vegetation.  Although 
my  view  was  partially  obscured  by  the  vege- 
tation, it  was  clear  that  for  the  next  2-3  min, 
the  gull  was  trying  to  gain  control  of  the 
struggling  ibis.  At  one  point,  the  gull  could  be 
seen  grasping  the  ibis’  neck  in  its  bill.  At 
16:07,  the  gull  flew  away,  abandoning  the  ibis 
in  the  vegetation. 

Following  the  gull’s  departure,  Audubon 
staff  and  I retrieved  and  inspected  the  ibis.  It 
was  alive  but  appeared  exhausted,  with  droop- 
ing wings  and  little  reaction  to  approaching 
humans.  There  were  no  visible  injuries  other 
than  the  small  surface  wound  inflicted  during 
the  chase.  We  placed  the  bird  in  a box  and 
released  it  several  hours  later. 

While  this  is  the  first  account  of  Great 
Black-backed  Gulls  attacking  adult  Glossy 
Ibis,  such  attacks  may  be  fairly  common  at 
this  site  but  seldom  observed.  I have  observed 
gulls  feeding  on  fresh  ibis  carcasses  on  several 
occasions  but  never  witnessed  the  kill.  Addi- 
tionally, during  an  annual  wading  bird  and 
seabird  census  in  late  May,  I found  remains 
of  24  adult  ibis.  All  carcasses  had  been 
cleaned  of  flesh  and  viscera,  but  they  retained 
wings  and  sometimes  the  head/neck  or  legs, 
indicating  gull  predation  (there  are  no  mam- 
malian predators  on  Stratton,  and  raptors  sel- 
dom visit  the  site).  Perhaps  aerial  pursuit  is 
not  the  usual  means  of  capture,  and/or  the 
events  are  easily  missed  due  to  the  dense  veg- 
etation and  trees  favored  by  nesting  ibis.  Au- 
dubon personnel  have  also  seen  gulls  occa- 
sionally take  ibis  fledglings  from  the  air  and 
noticed  fledgling  remains  in  the  wading  bird 
colony,  but  they  have  never  conducted  sys- 
tematic observations  to  quantify  predation 
rates  (C.  S.  Hall  pers.  comm.,  S.  Sanborn  pers. 
comm.). 

In  contrast,  Great  Black-backed  Gull  dep- 
redation of  other  species  nesting  on  Stratton 
(e.g.,  adult  and  duckling  Common  Eiders  [So- 
materia  mollissima ] and  tern  [Sterna  spp.] 
eggs  and  chicks)  is  frequently  observed  (CED 
unpubl.  data).  In  the  breeding  seasons  of 
2004-2005,  few  (if  any)  ducklings  survived 
to  fledging  as  a result  of  opportunistic,  group 


attacks  by  gulls  (CED  unpubl.  data).  Some  at- 
tacks involved  more  than  20  gulls  simulta- 
neously descending  on  a creche,  fighting  and 
plunge-diving  to  consume  ducklings.  Existing 
gull  control  practices  to  enhance  tern  resto- 
ration (nest  destruction  and  shooting  of  tern 
predators)  seem  to  have  little  benefit  for  eiders 
(and  perhaps  ibis),  as  predatory  gulls  continue 
to  congregate  in  large  numbers  around  crech- 
ing  and  nesting  areas. 

For  a small  ibis  colony  of  100  breeding 
pairs,  the  presumed  number  of  Great  Black- 
backed  Gull  kills  reported  here  seems  consid- 
erable and  warrants  further  investigation.  In  a 
recent  review,  Davis  and  Kricher  (2000)  found 
no  reports  of  predation  on  adult  Glossy  Ibis, 
though  they  described  the  Glossy  Ibis  as  “an 
understudied  species”  and  suggested  that  Per- 
egrine Falcons  ( Falco  peregrinus ) likely  take 
adults  at  some  colonies.  It  appears,  then,  that 
this  level  of  adult  mortality  is  unprecedented 
and,  if  continued,  could  lead  to  colony  extinc- 
tion. Additional  study  is  needed  to  determine 
whether  a few  “specialist”  gulls  prey  on  ibis 
at  Stratton  Island,  and,  if  so,  whether  they 
prey  on  weak,  sick,  or  otherwise  unfit  indi- 
viduals. 

ACKNOWLEDGMENTS 

I thank  D.  M.  Bird,  S.  W.  Kress,  C.  S.  Hall,  and  R. 
D.  Titman  for  supporting  my  graduate  work.  Staff  and 
volunteers  of  the  National  Audubon  Society’s  Seabird 
Restoration  Program  provided  assistance,  field  camp 
facilities,  and  logistical  support  on  Stratton  Island. 
This  observation  was  recorded  during  a gull  predation 
study  funded  by  the  Cornell  Lab  of  Ornithology,  the 
Garden  Club  of  America  (Frances  M.  Peacock  Schol- 
arship for  Native  Bird  Habitat),  and  the  Avian  Science 
and  Conservation  Centre  of  McGill  University.  M.  A. 
Gahbauer,  M.  A.  Hudson,  and  three  anonymous  re- 
viewers provided  helpful  comments  on  earlier  drafts  of 
this  manuscript. 

LITERATURE  CITED 

Addy,  C.  E.  1945.  Great  Black-backed  Gull  kills  adult 
Black  Duck.  Auk  62:142-143. 

Ahlund,  M.  and  F.  Gotmark.  1989.  Gull  predation  on 
eider  ducklings  Somateria  mollissima : effects  of 
human  disturbance.  Biological  Conservation  48: 
115-127. 

Burger,  J.  and  M.  Gochfeld.  1984.  Great  Black- 
backed  Gull  predation  on  kittiwake  fledglings  in 
Norway.  Bird  Study  31:149-151. 

Chase,  G.  P.  1994.  Stratton’s  islands  of  Saco  Bay:  an 
interwoven  history.  Mendocino  Lithographers, 
Fort  Bragg,  California. 


422 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Corkhill,  P.  1971.  Cannibalistic  Great  Black-backed 
Gulls.  British  Birds  64:30-32. 

Davis,  W.  E.,  Jr.,  and  J.  Kricher.  2000.  Glossy  Ibis 
(Plegadis  falcinellus).  The  Birds  of  North  Amer- 
ica, no.  545. 

Good,  T.  P.  1998.  Great  Black-backed  Gull  (Lams 
marinus).  The  Birds  of  North  America,  no.  330. 

Harris,  M.  P.  1965.  The  food  of  some  Larus  gulls. 
Ibis  107:43-53. 

Harris,  M.  P.  1980.  Breeding  performance  of  puffins 
Fratercula  arctica  in  relation  to  nest  density,  lay- 
ing date  and  year.  Ibis  122:193-209. 

Hatch,  J.  J.  1970.  Predation  and  piracy  by  gulls  at  a 
ternery  in  Maine.  Auk  87:244-254. 

Johnson,  R.  A.  1938.  Predation  of  gulls  in  murre  col- 
onies. Wilson  Bulletin  185:161-170. 

Kress,  S.  W.  1998.  Applying  research  for  effective 
management:  case  studies  in  seabird  restoration. 
Pages  141-154  in  Avian  conservation:  research 
and  management  (J.  M.  Marzluff  and  R.  Salla- 
banks,  Eds.).  Island  Press,  Washington,  D.C. 

Kury,  C.  R.  and  M.  Gochfeld.  1975.  Human  inter- 
ference and  gull  predation  in  cormorant  colonies. 
Biological  Conservation  8:23-34. 

Macdonald,  S.  M.  and  C.  F.  Mason.  1973.  Predation 
of  migrant  birds  by  gulls.  British  Birds  66:361- 
363. 

Massaro,  M.,  J.  W.  Chardine,  I.  L.  Jones,  and  G.  J. 
Robertson.  2000.  Delayed  capelin  ( Mallotus  vil- 
losus)  availability  influences  predatory  behaviour 
of  large  gulls  on  Black-legged  Kitti wakes  (Rissa 
tridactyla),  causing  a reduction  in  kittiwake  breed- 
ing success.  Canadian  Journal  of  Zoology  78: 
1588-1596. 


Mawhinney,  K.  and  A.  W.  Diamond.  1999.  Using  ra- 
dio-transmitters to  improve  estimates  of  gull  pre- 
dation on  Common  Eider  ducklings.  Condor  101: 
824-831. 

Mendenhall,  V.  M.  and  H.  Milne.  1985.  Factors  af- 
fecting duckling  survival  of  eiders  Somateria  mol- 
lissima  in  northeast  Scotland.  Ibis  127:148-158. 

Mikola,  J.,  M.  Miettinen,  E.  Lehikoinen,  and  K.  Leh- 
tila.  1994.  The  effects  of  disturbance  caused  by 
boating  on  survival  and  behaviour  of  Velvet  Sco- 
ter Melanitta  fusca  ducklings.  Biological  Conser- 
vation 67:1 19-124. 

Pierotti,  R.  1983.  Gull-puffin  interactions  on  Great 
Island,  Newfoundland.  Biological  Conservation 
26:1-14. 

Robinson,  H.  W.  1930.  Departure  and  landing  of  Manx 
Shearwaters.  British  Birds  23:224-225. 

Russell,  J.  and  W.  A.  Montevecchi.  1996.  Predation 
on  adult  puffins  Fratercula  arctica  by  Great 
Black-backed  Gulls  Larus  marinus  at  a New- 
foundland colony.  Ibis  138:791-794. 

Ryan,  R.  A.  1990.  Predation  by  Great  Black-backed 
Gulls  on  banded  waterfowl.  North  American  Bird 
Bander  15:10-12. 

Snyder,  F.  1960.  Great  Black-backed  Gulls  killing 
Dovekies.  Auk  77:476-477. 

Sobkowiak,  S.  1986.  Greater  Black-backed  Gull  and 
Bald  Eagle  predation  on  American  Coots.  M.Sc. 
thesis,  McGill  University,  Montreal,  Quebec. 

Whitt  am,  R.  M.  and  M.  L.  Leonard.  1999.  Predation 
and  breeding  success  in  Roseate  Terns  ( Sterna 
dougallii).  Canadian  Journal  of  Zoology  77:851- 
856. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):422-A23,  2006 


Tailless  Whipscorpion  ( Phrynus  longipes)  Feeds  on  Antillean  Crested 
Hummingbird  ( Orthorhyncus  cristatus ) 

Jennifer  L.  Owen13  and  James  C.  Cokendolpher1 2 3 


ABSTRACT. — A tailless  whipscorpion  (Phrynus 
longipes ) was  observed  feeding  on  an  Antillean  Crest- 
ed Hummingbird  (Orthorhyncus  cristatus)  atop  a large 


1 Dept,  of  Range,  Wildlife,  and  Fisheries  Manage- 
ment, Texas  Tech  Univ.,  Box  42125,  Lubbock,  TX 
79409-2125,  USA;  and  Texas  Parks  and  Wildlife, 
Bentsen  Rio  Grande  Valley  State  Park,  World  Birding 
Center,  2800  South  Bentsen  Palm  Dr.,  Mission,  TX 
78572,  USA. 

2 Invertebrates,  Natural  Science  Research  Lab.,  Mu- 
seum of  Texas  Tech  Univ.,  Lubbock.  TX  79409,  USA. 

3 Corresponding  author;  e-mail: 
jennifer.owen@tpwd. state. tx. us 


boulder  on  the  island  of  Virgin  Gorda  in  the  British 
Virgin  Islands.  This  is  the  first  record  of  any  avian 
species  serving  as  prey  for  an  amblypygid.  Received 
13  June  2005,  accepted  21  March  2006. 


Whip  spiders  (tailless  whipscorpions),  or 
amblypygids,  are  members  of  the  class  Arach- 
nida,  order  Amblypygi.  Phrynus  longipes  is 
the  largest  amblypygid  on  many  Caribbean  is- 
lands, including  the  U.S.  and  British  Virgin 
Islands  (Lazell  2005).  The  average  body 
length  of  P.  longipes  is  —35  mm  and  the  an- 


SHORT  COMMUNICATIONS 


423 


tenniform  legs  can  reach  an  additional  34  mm 
(Quintero  1981).  Amblypygids  have  no  ven- 
om glands;  instead,  they  use  their  sharp  rap- 
toral  pedipalps  (first  pair  of  appendages)  to 
capture  prey.  They  are  generally  nocturnal  and 
are  considered  mostly  “sit  and  wait”  preda- 
tors, feeding  on  prey  items  found  around  their 
home  territory  in  the  caves  and  crevices  be- 
tween and  under  large  rocks,  and  on  trees 
(Weygoldt  2000).  Although  the  diet  of  P.  lon- 
gipes  consists  primarily  of  arthropods,  espe- 
cially insects,  it  has  been  recorded  to  prey 
upon  vertebrates,  such  as  Anolis  lizards  (Wey- 
goldt 2000)  and  Eleuthrodactylus  frogs  (Rea- 
gan and  Waide  1996).  There  are  no  previous 
records  of  avian  species  serving  as  prey  for 
any  amblypygid. 

Antillean  Crested  Hummingbirds  ( Ortho - 
rhyncus  cristatus ) are  diurnal  and  inhabit  the 
Lesser  Antilles,  including  the  British  Virgin 
Islands  (Lazell  2005).  The  main  cause  of  mor- 
tality for  hummingbirds  is  predation  of  their 
eggs  and  nestlings;  predation  on  adult  hum- 
mingbirds is  relatively  rare  (Miller  and  Glass 
1985).  Thirteen  cases  of  adult  hummingbird 
predation  have  been  documented  worldwide, 
with  only  two  events  involving  an  invertebrate 
predator;  the  Chinese  praying  mantis  ( Ten - 
odera  aridifolia)  was  the  predator  in  both  cas- 
es (Miller  and  Glass  1985).  Like  amblypygids, 
the  Chinese  praying  mantis  is  a “sit  and  wait” 
predator. 

At  22:00  EST  on  20  October  2004,  J.  Egel- 
hoff  observed  an  adult  P.  longipes  (body  —30 
mm  long)  feeding  on  an  adult  Antillean  Crest- 
ed Hummingbird  (—80  mm  long),  1 m above 
ground,  atop  a large  boulder  behind  the  Little 
Secrets  Nature  Gallery  in  Spanish  Town,  Vir- 
gin Gorda,  British  Virgin  Islands  (18°  26.68' 


N,  64°  26.38'  W).  The  P.  longipes  was  hold- 
ing the  hummingbird  with  its  raptoral  pedi- 
palps and  was  feeding  on  the  hummingbird’s 
body;  it  continued  to  feed  for  2 hr.  At  the  time 
of  observation,  the  hummingbird  was  no  lon- 
ger alive,  and  due  to  the  mutilation  caused  by 
the  feeding  amblypygid,  we  were  unable  to 
obtain  information  on  the  hummingbird’s 
weight,  sex,  or  breeding  status.  The  ambly- 
pygid is  now  part  of  the  living  exhibit  at  the 
Little  Secrets  Nature  Gallery. 

Although  it  is  unknown  how  the  P.  longipes 
acquired  its  avian  prey,  our  observation  is  the 
first  record  of  an  amblypygid  feeding  on  a 
hummingbird,  or  any  other  avian  species. 

ACKNOWLEDGMENTS 

We  thank  Jim  Egelhoff  and  the  Little  Secrets  Nature 
Gallery  on  Virgin  Gorda,  British  Virgin  Islands,  for 
locating  and  photographing  the  predation  incident,  and 
to  Dr.  J.  Lazell,  The  Conservation  Agency,  and  the 
Falconwood  Foundation  for  supporting  research  in  the 
British  Virgin  Islands.  We  also  thank  Texas  Tech  Uni- 
versity for  financial  support.  This  is  manuscript  T-9- 
1054  of  the  College  of  Agricultural  Sciences  and  Nat- 
ural Resources,  Texas  Tech  University.  We  thank  J.  M. 
Wunderle  and  two  anonymous  reviewers  for  the  help- 
ful comments  that  improved  this  manuscript. 

LITERATURE  CITED 

Lazell,  J.  2005.  Island:  fact  and  theory  in  nature.  Uni- 
versity of  California  Press,  Berkeley. 

Miller,  R.  S.  and  C.  L.  Glass.  1985.  Survivorship  in 
hummingbirds:  is  predation  important?  Auk  102: 
175-178. 

Quintero,  D.  1981.  The  amblypygid  genus  Phrynus  in 
the  Americas.  Journal  of  Arachnology  9:117-166. 
Reagan,  D.  P.  and  R.  B.  Waide  (Eds.).  1996.  The  food 
web  of  a tropical  rainforest.  University  of  Chicago 
Press,  Chicago,  Illinois. 

Weygoldt,  P.  2000.  Whip  spiders  (Chelicerata:  Am- 
blypygi):  their  biology,  morphology  and  system- 
atics.  Apollo  Books,  Stenstrup,  Denmark. 


424 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


The  Wilson  Journal  of  Ornithology  1 18(3):424^126,  2006 


Polydactyly  in  a Vaux’s  Swift 


Walter  H.  Sakai1 


ABSTRACT. — I report  on  polydactyly  in  a Vaux’s 
Swift  ( Chaetura  vauxi).  An  extra,  asymmetrically  lo- 
cated toe  was  found  on  each  foot  of  one  swift.  A check 
of  329  swifts  from  several  museums  produced  no  other 
examples  of  polydactyly  in  this  species.  A review  of 
the  literature  and  a query  over  the  Internet,  however, 
produced  10  other  examples  of  polydactyly  in  wild 
birds.  Received  5 August  2005,  accepted  27  February 
2006. 


Polydactyly  is  a relatively  common  malfor- 
mation phenomenon  in  vertebrates.  It  has  been 
well  documented  in  humans  and  domestic  an- 
imals such  as  cats,  dogs,  mice,  and  chickens 
(Clark  et  al.  2000);  however,  it  is  an  uncom- 
mon phenomenon  and  rarely  reported  in  wild 
birds.  A group  of  eight  Vaux’s  Swifts  ( Chae- 
tura vauxi , family  Apodidae)  was  brought  to 
me  from  the  California  Wildlife  Center,  an  an- 
imal rehabilitation  center  in  the  Santa  Monica 
Mountains  in  Malibu,  California.  On  29  April 
2002,  the  swifts  were  found  dead  along  Cross 
Creek  Road  (34°  02'  35"  N,  1 1 8°  41 ' 02"  W) 


1 Life  Sciences  Dept.,  Santa  Monica  College,  1900 
Pico  Blvd.,  Santa  Monica,  CA  90405-1628,  USA; 
e-mail:  sakai_walter@smc.edu 


near  Malibu  Creek,  Malibu,  Los  Angeles 
County,  California. 

As  I was  preparing  the  birds  as  study  skins 
and  examining  the  swifts’  pamprodactyl-type 
feet  (Proctor  and  Lynch  1993),  I found  that 
seven  of  the  birds  were  normal  and  one  had 
an  extra,  asymmetrically  located  toe  on  each 
foot.  On  both  feet,  digit  one  (the  hallux)  was 
located  11  mm  below  the  joint  of  the  tibi- 
otarsus  and  tarsometatarsus.  The  tarsometatar- 
si  were  13.5  mm  long.  On  the  left  foot,  the 
extra  digit  was  located  on  the  tarsometatarsus 
6 mm  from  the  joint  of  the  tibiotarsus  and 
tarsometatarsus  (Fig.  1A)  and  was  6 mm  long. 
In  addition,  digit  one  and  the  extra  toe  of  the 
left  foot  were  joined  by  a webbing  of  tissue; 
thus,  the  nails  touched.  The  extra  digit  on  the 
right  foot  was  located  at  the  joint  of  the  ti- 
biotarsus and  the  tarsometatarsus  (Fig.  IB) 
and  was  10  mm  long. 

A survey  of  the  literature  and  a query  to 
museum  bird  curators  and  collection  managers 
via  the  “AVECOL”  listserve  produced  reports 
of  10  birds  with  polydactyly.  Extra  toes  were 
reported  for  Mallard  {Anas  platyrhynchos’,  Na- 
pier 1963),  Common  (currently  Wilson’s) 


FIG.  1.  Left  (A)  and  right  (B)  feet  with  extra  toe  of  a Vaux’s  Swift  {Chaetura  vauxi)  collected  29  April 
2002  along  Cross  Creek  Road  near  Malibu  Creek,  Malibu,  Los  Angeles  County,  California. 


SHORT  COMMUNICATIONS 


425 


Snipe  ( Capella  gallinago  [currently  Gallinago 
delicata ];  Fogarty  1969),  Sooty  Tern  ( Sterna 
fuscata ; Austin  1969),  Long-billed  Curlew 
( Numenius  americanus;  Forsythe  1972),  Ring- 
billed Gull  ( Larus  delawarensis;  Ryder  and 
Chamberlain  1972),  Common  Nighthawk 
( Chordeiles  minor ; Chandler  1992),  Common 
Loon  ( Gavia  immer ; R.  Y.  McGowan  pers. 
comm.),  Common  Swift  ( Apus  apus\  Gory 
1992),  Common  (currently  Eurasian)  Kestrel 
(Falco  tinnunculus’,  Trinkaus  et  al.  1999),  and 
Eastern  Screech-Owl  ( Otus  [currently  Mega- 
scops] asio’,  Albers  et  al.  2001).  An  uncon- 
firmed case  of  polydactyly  in  Anna’s  Hum- 
mingbird ( Calypte  anna ) was  reported  from 
the  San  Francisco  Bay  Area,  California  (W.  H. 
Baltosser  pers.  comm.) 

I also  checked  Vaux’s  Swifts  in  the  collec- 
tions of  two  nearby  museums:  75  specimens 
at  the  Los  Angeles  County  Museum  of  Nat- 
ural History  (LACMNH),  Los  Angeles,  Cali- 
fornia, and  157  specimens  at  the  Western 
Foundation  of  Vertebrate  Zoology  (WFVZ), 
Camarillo,  California,  all  of  which  were  nor- 
mal. The  73  Vaux’s  and  Chimney  Swifts 
( Chaetura  pelagica ) in  the  collection  at  Del- 
aware Museum  of  Natural  History,  Wilming- 
ton, Delaware,  also  were  reported  as  normal 
(J.  L.  Woods  pers.  comm.).  C.  M.  Dardia 
(pers.  comm.)  reported  that  all  24  Vaux’s 
Swifts  in  the  collection  at  Cornell  Museum  of 
Vertebrates,  Ithaca,  New  York,  were  normal. 

The  causes  of  polydactyly  among  vertebrate 
groups  have  included  UV-B  radiation  (Blau- 
stein  et  al.  1997),  parasites  (Johnson  et  al. 
2001),  parasites  and  pesticides  in  amphibians 
(Kiesecker  2002),  nuclear  radiation  in  humans 
(Lazjuk  et  al.  1998),  and  congenital  defects  in 
humans  (Castilla  et  al.  1996).  Extensive  tera- 
tological  studies  have  been  conducted  on  Do- 
mestic Chicken  ( Gallus  domesticus),  and  sev- 
eral breeds  normally  have  five  toes  (Warren 
1941,  1944).  Unfortunately,  the  life  history  of 
the  Vaux’s  Swift  with  polydactyly  is  un- 
known. The  individual  in  question  appeared 
healthy  and  its  weight  (12.8  g)  did  not  differ 
from  that  of  the  other  seven  individuals  (mean 
= 12.67  ± 0.62;  Z-test,  P = 0.71)  found  with 
it,  although  it  was  lower  than  the  mean  (17.1 
± 1.3  SD,  n = 72)  weight  of  birds  reported 
by  Dunning  (1984). 

The  Vaux’s  Swift  specimen  with  polydac- 
tyly (Santa  Monica  College  [SMC]  SMC 


1 100)  was  prepared  as  a wet  specimen,  and 
the  other  seven  specimens  (SMC  1049,  1051, 
1052,  1053,  1056,  1057,  and  1058)  were  pre- 
pared as  study  skins.  All  eight  specimens  were 
then  transferred  to  the  LACMNH’s  Ornithol- 
ogy Collection  (wet  specimen:  LACM 

113615;  skins:  112233,  112234,  112230, 
11232,  11231,  11229,  and  11228). 

ACKNOWLEDGMENTS 

I thank  the  various  museum  ornithology  curators 
and  collection  managers  who  responded  with  both  pos- 
itive and  negative  reports,  and  for  suggesting  possible 
specimens.  Thanks  to  L.  Matsui,  who  brought  the 
specimens  to  me  from  the  California  Wildlife  Center 
where  she  volunteers.  Thanks  to  R.  A.  Cobb  and  K. 
L.  Garrett  for  suggestions  on  preservation  of  the  spec- 
imen. Thanks  to  K.  L.  Garrett  and  R.  Corado  for  access 
to  the  swifts  at  the  LACMNH  and  the  WFVZ,  respec- 
tively. J.  L.  Woods  provided  information  on  swifts  at 
the  Delaware  Museum  of  Natural  History,  and  C.  M. 
Dardia  provided  information  on  Vaux’s  Swifts  at  the 
Cornell  Museum  of  Vertebrates.  L.  S.  Hall  provided 
useful  comments.  Photographs  were  taken  by  J.  Smar- 
gis.  I would  like  to  thank  E.  L.  Bull  and  two  anony- 
mous reviewers  for  their  useful  and  helpful  comments 
on  this  paper. 

LITERATURE  CITED 

Albers,  P.  H.,  D.  J.  Hoffman,  and  I.  L.  Brisbin,  Jr. 
2001.  Unusual  leg  malformations  in  screech  owls 
from  a South  Carolina  superfund  site.  Journal  of 
Toxicology  and  Environmental  Health,  Part  A 63: 
89-99. 

Austin,  O.  L.,  Jr.  1969.  Extra  toes  on  a Sooty  Tern 
chick.  Auk  86:352. 

Blaustein,  A.  R.,  J.  M.  Kiesecker,  D.  P.  Chivers,  and 
R.  G.  Anthony.  1997.  Ambient  UV-B  radiation 
causes  deformities  in  amphibian  embryos.  Pro- 
ceedings of  the  National  Academy  of  Sciences 
USA  94:13735-13737. 

Castilla,  E.  E.,  R.  L.  Da  Fonseca,  M.  Da  G.  Dutra, 
E.  Bermejo,  L.  Cuevas,  and  M.  Martinez-Frias. 
1996.  Epidemiological  analysis  of  rare  polydacty- 
lies.  American  Journal  of  Medical  Genetics  65: 
295-303. 

Chandler,  R.  M.  1992.  Polydactyly  in  a Common 
Nighthawk.  Kansas  Ornithological  Society  Bulle- 
tin 43:17. 

Clark,  R.  M.,  P.  C.  Marker,  and  D.  M.  Kingsley. 
2000.  A novel  candidate  gene  for  mouse  and  hu- 
man preaxial  polydactyly  with  altered  expression 
in  limbs  of  hemimelic  extra-toes  mutant  mice. 
Genomics  67:19-27. 

Dunning,  J.  B.,  Jr.  1984.  Body  weight  of  686  species 
of  North  American  birds.  Western  Bird  Banding 
Association  Monograph,  no.  1. 

Fogarty,  M.  J.  1969.  Extra  toes  on  the  halluces  of  a 
Common  Snipe.  Auk  86:132. 


426 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


Forsythe,  D.  M.  1972.  Long-billed  Curlew  with  su- 
pernumerary hallux.  Auk  89:457. 

Gory,  G.  1992.  Un  cas  de  monstruousite  chez  le  Mar- 
tinet Noir  ( Apus  apus  L.).  Bulletin  de  la  Societe 
d’etude  des  Sciences  Naturelles  de  Nimes  et  du 
Gard  59:139.  [In  French] 

Johnson,  R T.  J.,  K.  B.  Lunde,  R.  W.  Haight,  J.  Bow- 
erman,  and  A.  R.  Blaustein.  2001.  Ribeiroia  on- 
datrae  (Trematoda:  Digenea)  infection  induces  se- 
vere limb  malformations  in  western  toads  ( Bufo 
boreas ).  Canadian  Journal  of  Zoology  79:370- 
379. 

Kiesecker,  J.  M.  2002.  Synergism  between  trematode 
infection  and  pesticide  exposure  magnifies  am- 
phibian limb  deformities  in  nature.  Proceedings  of 
the  National  Academy  of  Sciences  USA  99:9900- 
9904. 

Lazjuk,  G.,  Y.  Satow,  D.  Nikolaev,  and  I.  Novikova. 
1998.  Genetic  consequences  of  the  Chernobyl  ac- 


cident for  Belarus  Republic.  Gijutsu-to-Ningen 
283:26-32. 

Napier,  A.  1963.  Congenital  malformations  of  the  feet 
in  Mallard  ducklings.  Wildlife  Trust  Annual  Re- 
port 14:170-171. 

Proctor,  N.  S.  and  P.  J.  Lynch.  1993.  Manual  of  or- 
nithology: avian  structure  and  function.  Yale  Uni- 
versity Press,  New  Haven,  Connecticut. 

Ryder,  J.  P.  and  D.  J.  Chamberlain.  1972.  Congenital 
foot  abnormality  in  the  Ring-billed  Gull.  Wilson 
Bulletin  84:342-344. 

Trinkaus,  von  K.,  F.  Mueller,  and  E.  F.  Kaleta. 
1999.  Polydactyly  in  a kestrel  ( Falco  tinnunculus 
tinnunculus  Linne,  1758):  a case  study.  Zeitschrift 
fur  Jagdwissenschaft  45:66-72. 

Warren,  D.  C.  1941.  A new  type  of  polydactyly  in 
the  fowl.  Journal  of  Heredity  32:1-5. 

Warren,  D.  C.  1944.  Inheritance  of  polydactylism  in 
the  fowl.  Genetics  29:217-231. 


The  Wilson  Journal  of  Ornithology  1 1 8(3):427^t29,  2006 


Once  Upon  a 'Time  in  Tmerican  Ornithology 


James  Little  Baillie,  whose  parents  had  em- 
igrated from  Great  Britain  to  Canada,  was 
born  on  4 July  1904  in  Toronto,  Ontario.  The 
fifth  of  11  children,  he  went  to  work  at  the 
age  of  13  after  completing  elementary  school. 
When  he  was  16,  Baillie  began  bird  watching, 
and,  just  two  years  later  in  1922,  he  was  ap- 
pointed as  technical  assistant  for  the  ornithol- 
ogy department  of  the  Royal  Ontario  Museum 
(ROM)  of  Zoology.  From  1927  to  1931,  he 
attended  high  school  night  classes,  although 
he  never  earned  enough  credits  to  graduate. 
Nonetheless,  his  enthusiasm  and  profound 
knowledge  of  birds  eventually  resulted  in  his 
promotion  to  assistant  curator  of  ornithology 
at  ROM,  a position  in  which  he  served  for 
nearly  50  years. 

Recognizing  the  value  of  public  awareness 
in  conservation  endeavors,  for  39  years  Bail- 
lie  wrote  a weekly  column,  Birdland,  for  the 
Toronto  Evening  Telegram.  He  liked  work- 
ing with  youth  and  mentored  countless  be- 
ginning ornithologists,  including  ecologist 
Robert  MacArthur  and  artist  Robert  Bate- 
man. Today,  Bailie’s  conservation  and  public 
education  legacies  continue  through  the 
James  L.  Baillie  Memorial  Fund  for  Bird  Re- 
search and  Preservation  (see  http://www. 
bsc-eoc.org/organization/jlbmf.html),  which 
provides  funding  opportunities  for  Canadian 
students  interested  in  field  studies  and  pro- 
jects that  improve  our  understanding  and 
conservation  of  birds.  In  1935,  Baillie  was 
elected  a member  of  the  American  Ornithol- 
ogists’ Union — only  the  eighth  Canadian  to 
be  so  honored. 

Above  all  else,  however,  Baillie  was  a ded- 
icated museum  man.  He  published  reports  of 
numerous  museum  expeditions  and  actively 
sought  to  enhance  ROM’s  bird  collection.  In 
a 1970  tribute  to  Baillie,  C.  H.  D.  Clarke 


wrote,  “Jim  had  a rare  sense  of  the  museum 
collection  as  . . . documents  that  would  never 
cease  yielding  new  information.  . . . The  fact 
that  the  whole  history  of  environmental  pol- 
lution in  Sweden  has  been  read  from  the  mo- 
lecular analyses  of  piths  from  the  feathers  of 
birds  in  the  Swedish  National  Museum,  the 
dates  being  the  dates  on  labels,  fitted  precisely 
Jim’s  concept  of  the  specimen  as  a storehouse 
of  information  yet  undreamed  of.”  In  fact, 
Baillie’s  dedication  to  the  museum  concept 
drove  him  to  what  he  felt  was  the  most  re- 
warding accomplishment  of  his  entire  career: 
acquiring  Great  Auk  ( Pinguinus  impennis ) 
and  Labrador  Duck  ( Camptorhynchus  labra- 
dorius ) specimens  for  the  ROM.  Although  he 
was  proud  that  the  ROM  already  held  108 
specimens  of  the  Passenger  Pigeon  {Ectopistes 
migratorius ) — “the  largest  collection  of  them 
in  existence,”  he  wrote  to  a friend — he  was 
distraught  that  not  one  Canadian  museum  pos- 
sessed a mounted  specimen  of  the  Great  Auk. 
The  other  species  that  had  once  inhabited 
parts  of  Canada — the  Labrador  Duck — was 
represented  in  Canadian  museums  by  only 
two  specimens. 

As  Baillie  searched  for  possible  specimens 
of  the  Great  Auk  and  Labrador  Duck,  he  ap- 
pealed to  his  weekly  newspaper  readership 
and  his  network  of  patrons  for  funding.  In 
1964,  his  resolve  and  efforts  were  finally  re- 
warded (see  Fig.  1).  The  reference  for  the 
quotes  that  follow  is  Anglin,  L.  1987.  Birder 
Extraordinaire:  The  life  and  legacy  of  James 
L.  Baillie.  Toronto  Ornithological  Club,  To- 
ronto, Ontario.  Thanks  to  Lise  Anglin  and  the 
book’s  publishers — Toronto  Ornithological 
Club  and  Long  Point  Bird  Observatory — for 
providing  quotations  and  permission  to  quote 
from  the  book.— ALEXANDER  T.  CRINGAN; 
e-mail:  alexc @ lamar.colostate.edu 


On  July  22  1964,  [Baillie’s]  son-in-law  drove  [Bailie]  to  New  York  with  Helen 
[Baillie’s  second  wife]  and  Florence  [his  daughter]  to  negotiate  the  deal  with  Dr. 
R.  S.  Palmer  of  the  American  Museum  of  Natural  History. 

On  July  26,  [they]  made  the  return  trip  to  Toronto  with  two  more  inanimate 
passengers  aboard — one  Great  Auk  and  one  Labrador  Duck.  Jim  was  nervous  dur- 
ing the  drive  lest  an  accident  might  result  in  damage  to  the  glass  case  or  the  birds. 
However  there  was  no  mishap. 


427 


428 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


FIG.  1.  James  L.  Baillie  contemplating  the  Great  Auk  specimen  he  procured  in  1964  for  the  Royal  Ontario 
Museum  in  Toronto.  This  specimen  is  widely  believed  to  have  been  the  one  previously  owned  by  John  James 
Audubon  (see  pages  154-160,  “Audubon’s  Auk.  bird  no.  20,”  in  Fuller,  E.  1999.  The  Great  Auk.  Harry  N. 
Abrams,  Inc.  Publishers,  New  York).  Anxious  to  see  and  paint  a Great  Auk  and  other  sea  birds  of  northern 
latitudes,  John  James  Audubon  embarked  on  a voyage  to  Labrador  in  summer  1833.  Poor  weather,  however, 
precluded  the  expedition  from  ever  reaching  locations  where  Audubon  could  observe  Great  Auks.  Thus,  he  had 
to  acquire  a mounted  specimen  from  which  to  make  his  painting  for  Birds  of  North  America.  As  reported  by 
an  officer  of  the  Toronto  Ornithological  Club,  “It  is  strongly  suspected  that  the  ROM’s  Great  Auk  was  indeed 
Audubon’s  specimen.”  He  went  on,  however,  to  mention  at  least  one  source  that  brought  this  belief  into  question: 
“ . . . although  everything  collected  was  consistent  with  that  specimen  being  Audubon’s  (nothing  glaring  dis- 
proving that  possibility),  the  chain  of  ownership  was  not  complete  enough  to  provide  ‘absolute  proof’  of  this, 
but  it  is  very  likely  that  this  indeed  is  the  case.”  According  to  Fuller  (1999),  when  Audubon’s  Great  Auk  was 
restored  and  remounted  in  1921,  the  renovator  discovered  that  it  was  stuffed  with  old  German  newspapers,  thus 
dispelling  the  prevailing  notion  that  Audubon’s  auk  was  American  in  origin.  Rather,  the  German  association 
indicates  an  Icelandic  origin. 


Against  somewhat  unexpected  odds,  he  had  achieved  a goal  seen  by  many  as 
unattainable.  On  19  May  1970,  just  days  before  his  death,  Jim  wrote  from  the 
Toronto  General  Hospital,  “With  a staff  of  three  or  four,  we  . . . acquired  a Great 
Auk,  a long-extinct  Canadian  bird  previously  represented  in  Canadian  collections 
only  by  bones.  The  fact  that  the  specimen  turned  out  to  be  John  James  Audubon’s 
very  own  specimen,  from  which  he  made  his  famous  painting,  was  an  unexpected 
bonus.  Happily,  at  the  same  time,  from  the  same  U.S.  ladies’  college  [Vassar],  we 
acquired  another  Canadian  we  did  not  previously  possess — a drake  Labrador  Duck. 
Previous  Canadian-held  Labrador  ducks  exist  only  in  Dalhousie  and  McGill  Uni- 
versities. . . . The  possession  of  these  two  treasures  is  an  accepted  criterion  of  the 
value  of  a museum’s  collection,  in  ornithological  circles.  . . . Both  ours  are  magnif- 
icent birds  in  first-class  condition,  mounted  in  hermetically  sealed  cases.” 


ONCE  UPON  A TIME  IN  AMERICAN  ORNITHOLOGY 


429 


EPILOGUE:  Pinguinus,  the  Great  Auk’s 
genus  name,  reflects  the  species’  widely  used 
common  name:  “penguin.”  Although  the  der- 
ivation of  Pinguinus  is  uncertain,  possibilities 
include  “pen-winged”  or  “pinioned,”  from 
the  Welsh  terms  for  white  (pen)  and  head 
(gwyn),  or  the  Latin  word  for  fat  (penquis ).  It 
was  after  Europeans  discovered  Pinguinus  im- 
pennis  in  the  northern  Atlantic  that  explorers 
found  members  of  the  similar-looking — but 
very  different — Spheniscidae  family  (pen- 
guins) in  the  Southern  hemisphere  (Monte  - 
vecchi,  W.  A.  and  D.  A.  Kirk.  1996.  Great 
Auk.  Birds  of  North  America,  no.  260).  Al- 
though the  Great  Auk  inhabited  much  of  the 
northern  Atlantic,  there  is  evidence  that  pre- 
historic people  had  extirpated  the  species  from 
many  parts  of  its  original  range.  Climate 
changes  also  may  have  factored  into  the  spe- 
cies’ range  contractions. 


Human  exploitation  of  this  flightless  spe- 
cies for  its  meat,  eggs,  oil,  and  down  contin- 
ued right  up  until  the  early  19th  century,  by 
which  time  the  northern  Atlantic  “penguin” 
had  become  quite  rare.  Another  significant 
blow  to  the  population  came  in  1830,  when 
an  underwater  volcanic  eruption  occurred  near 
Iceland,  causing  tremors  and  massive  waves 
that  washed  away  the  Island  of  Geirfuglas- 
ter — one  of  the  species’  last  important  breed- 
ing sites.  The  largest-known  nesting  colony  of 
Great  Auks,  however,  was  found  on  Funk  Is- 
land (historically  known  as  Penguin  Island), 
located  off  the  coast  of  Newfoundland;  in 
1841,  the  last  of  Funk  Island’s  auks  was 
killed.  In  1844,  the  species  disappeared  alto- 
gether when  two  Great  Auks  found  on  Eldey 
Island  near  Iceland  were  beaten  to  death  and 
sold  for  use  as  stuffed  specimens. — CYNTHIA 
P.  MELCHER;  e-mail:  wjo@usgs.gov 


The  Wilson  Journal  of  Ornithology  1 18(3):430— 435,  2006 


Ornithological  Literature 


HANDBOOK  OF  THE  BIRDS  OF  THE 
WORLD,  VOLUME  9:  COTINGAS  TO  PIP- 
ITS AND  WAGTAILS.  Edited  by  Josep  del 
Hoyo,  Andrew  Elliott,  and  David  Christie. 
Lynx  Edicions,  Barcelona,  Spain.  2004:  864 
pp.,  78  color  plates,  440  photographs,  809 
maps.  ISBN:  84-87334-69-5.  $245.00 
(cloth). — Volume  9 in  the  landmark  series. 
Handbook  of  the  Birds  of  the  World,  con- 
cludes the  suboscines  with  cotingas,  mana- 
kins,  tyrant  flycatchers,  New  Zealand  wrens, 
scrub-birds,  and  lyrebirds,  and  begins  the  os- 
cines  with  larks,  swallows,  pipits,  and  wag- 
tails. This  volume  follows  the  format  proven 
in  earlier  volumes  of  the  series,  with  a chapter 
for  each  family — lavishly  illustrated  with  col- 
or photographs — followed  by  the  species  ac- 
counts. The  chapters  include  discussions  of 
the  family’s  systematics,  morphological  as- 
pects, habitats,  general  habits,  vocalizations, 
foods  and  foraging,  breeding,  movements,  re- 
lationship with  humans,  and  status  and  con- 
servation, and  they  wrap  up  with  a general 
bibliography.  The  species  accounts  are  illus- 
trated with  color  plates  that  often  include  sub- 
species and  both  sexes.  The  accounts  are  or- 
ganized by  the  same  section  headings  as  those 
in  the  family  chapters — with  the  substitution 
of  taxonomy,  subspecies,  and  distribution  for 
systematics,  and  the  addition  of  descriptive 
notes. 

As  in  previous  volumes,  the  photographs  in 
volume  9 are  superlative:  they  capture  court- 
ship displays,  bathing,  agonistic  behaviors, 
roosting  birds,  nests,  recently  rediscovered 
species,  birds  in  their  habitats,  and  “birds  be- 
ing birds.”  Those  who  see  the  photographs  in 
this  volume  will  be  left  with  the  impression 
that  all  one  needs  is  a camera,  and  then  mag- 
ically pipits  will  pose  for  the  camera  while 
carrying  insects  in  their  bills  and  rare  rain- 
forest birds  will  display  in  plain  view  (and,  of 
course,  in  good  weather).  Anyone  who  has 
ever  tried  to  photograph  wild  birds  (especially 
those  where  the  subject  is  actually  doing 
something)  will  recognize  the  difficulty  in- 
volved in  taking  photographs  of  high  technical 
quality  with  a pleasing  composition.  The  few 


photographs  of  birds  in  the  hand  are  of  ex- 
ceptionally rare  species,  making  them  worth  a 
second  look.  The  informative  photo  captions 
provide  information  that  is  not  covered  else- 
where in  the  text. 

The  Foreword  by  Richard  C.  Banks  covers 
the  topic  of  ornithological  nomenclature.  He 
begins  with  an  overview  of  the  history  and 
development  of  ornithological  nomenclature, 
which  leads  to  a discussion  on  its  state  today, 
including  the  current  International  Code  of 
Zoological  Nomenclature.  He  recounts  the  de- 
velopment of  the  trinomial  for  subspecies  in 
what  was  originally  a binomial  system,  the 
purpose  and  use  of  a superspecies  or  subge- 
nus, and  the  availability  of  names,  holotypes, 
and  syntypes.  He  goes  on  to  discuss  the  rel- 
atively recent  practice  of  naming  new  species 
with  a photograph  serving  as  the  “type,”  the 
difficulties  that  this  presents  to  nomenclatur- 
ists,  and  why  naming  new  species  inadver- 
tently is  problematic.  The  use  of  real  life  ex- 
amples brings  to  light  the  difficulty  of  naming 
bird  species.  Banks  also  covers  the  issue  of 
prevailing  usage,  which  is  contrary  to  the 
principle  of  priority.  The  section  concludes 
with  a summary  of  the  number  of  new  species 
described  from  the  years  1920  to  2000,  30- 
56%  of  which  were  estimated  to  be  truly  new 
species — depending  on  the  years  considered. 

Another  Foreword  (that  somehow  did  not 
make  the  Contents)  by  John  Fitzpatrick  entails 
a formal  description  of  a new  tribe  of  tyrant- 
flycatchers.  According  to  the  volume’s  Intro- 
duction, John  Fitzpatrick  realized  that  one  of 
the  subdivisions  he  intended  to  recognize  in 
the  Tyrannidae  had  not  been  named  formally, 
and  he  remedies  this  by  naming  the  tribe, 
Contopini,  in  the  volume’s  introductory  ma- 
terial. 

Some  of  the  common  names  used  in  this 
volume  were  surprising.  Rock  Wren  (. Xenicus 
gilviventris ) was  used  for  a member  of  Acan- 
thisittidae,  which  brings  up  the  question  of 
what  future  editors  will  call  the  Rock  Wren 
{Salpinctes  obsoletus ) when  they  get  to  the 
volume  that  includes  Troglodytidae.  I was  also 
intrigued  to  see  the  use  of  Collared  Sand  Mar- 


430 


ORNITHOLOGICAL  LITERATURE 


431 


tin  ( Riparia  riparia ) as  the  common  name  for 
Bank  Swallow.  I was  familiar  with  the  use  of 
Sand  Martin,  but  the  modifier  was  new  to  me. 
The  resolution  of  taxonomic  tangles,  such  as 
that  of  the  Yellow  Wagtail  ( Motacilla  flava) 
complex,  is  outside  the  true  purpose  of  this 
work;  accordingly,  the  editors  treat  Yellow 
Wagtail  as  one  species,  but  the  taxonomy  sec- 
tion provides  a good  description  of  recent 
DNA  work  on  this  complex. 

As  in  all  previous  volumes  of  this  series,  the 
References  section  is  split  into  two  parts:  Ref- 
erences of  Scientific  Descriptions  and  the  Gen- 
eral List  of  References.  The  former  lacks  the 
titles  of  publications  listed  but  does  include  sci- 
entific name(s),  whereas  the  latter  includes  the 
titles  of  listed  publications.  I am  uncertain  why 
the  two  were  not  merged  and  one  standard  ci- 
tation used,  but  because  this  is  Volume  9,  it  is 
likely  too  late  for  questions.  Regardless,  this 
book  is  highly  recommended. — MARY  GUS- 
TAFSON, Rio  Grande  Joint  Venture,  Texas 
Parks  and  Wildlife  Department,  Mission,  Tex- 
as; e-mail:  mary.gustafson@tpwd. state. tx. us 


A BIRDER’S  GUIDE  TO  MICHIGAN.  By 
Allen  T.  Chattier  and  Jerry  Ziamo.  American 
Birding  Association,  Colorado  Springs,  Colo- 
rado. 2004:  660  pp.,  284  maps,  6 photographs. 
ISBN:  1-878788-13-2.  $28.95  (paper).— In  his 
Foreword  to  A Birder's  Guide  to  Michigan, 
Allen  Chattier  and  Jerry  Ziamo’s  exhaustive 
guide  to  birding  in  the  Great  Lakes  State,  re- 
nowned bird-tour  leader  Jon  Dunn  describes 
his  first  trip  to  the  state  on  a cross-country 
birding  adventure.  In  June  1971,  he  and  his 
four  friends  visited  the  jack-pine  country  near 
Mio  to  search  for  Kirtland’s  Warbler,  which, 
as  most  birders  know,  breeds  exclusively  in 
the  north-central  Lower  Peninsula  (LP).  After 
successfully  seeing  the  warbler,  he  and  his 
group  left  the  following  day  for  the  eastern 
coast.  Dunn’s  trip  was  typical  of  many  bird- 
ers’ experiences  with  birding  in  Michigan — to 
see  Kirtland’s  Warbler  and  leave  a day  or  two 
later.  With  the  publication  of  this  book,  how- 
ever, more  adventurous  birders  will  decide  to 
make  Michigan  the  destination  of  longer  trips 
to  see  its  31  other  warbler  species,  as  well  as 
all  the  other  species  this  unique  northern  state 
has  to  offer. 


Four  years  in  the  making,  this  guide  is  by 
far  the  most  thorough  state-wide  guide  avail- 
able for  Michigan.  The  book  includes  266 
birding  sites  in  67  of  the  state’s  83  counties, 
including  all  15  counties  located  in  the  Upper 
Peninsula  (UP).  Indicative  of  the  authors’ 
knowledge  of  Michigan,  they  wrote  or  con- 
tributed to  166  of  the  site  descriptions.  Vir- 
tually all  the  site  descriptions  for  the  South- 
eastern LP  section  were  authored  exclusively 
by  Chartier,  and  Ziarno  wrote  nearly  all  those 
included  in  the  book  for  the  Northeastern  LP 
and  UP  sections.  Forty-three  other  birders 
from  across  the  state  authored  the  remaining 
site  descriptions.  Also  contributing  their  tal- 
ents to  this  guide  were  the  24  birders  who 
reviewed  and  checked  the  text  and  mileages, 
and  another  12  that  reviewed  the  bar  graphs 
depicting  each  species’  status  in  Michigan. 

Visitors  planning  their  first  trip  to  Michigan 
will  benefit  from  the  introductory  sections  on 
topography,  vegetation,  bird  habitats,  and  cli- 
mate— now  standard  information  included  in 
all  state  birding  guides  recently  published  by 
the  American  Birding  Association  (ABA).  A 
section  entitled  “The  Michigan  Birding  Year’’ 
gives  an  overview  of  bird  activity  that  one  can 
expect  in  each  month  of  a given  season,  sup- 
plementing the  excellent  status  and  occurrence 
bar  graphs  for  Michigan’s  303  annually  oc- 
curring species  and  the  list  of  casual  and  ac- 
cidental bird  species.  In  addition,  the  guide 
lists  Michigan’s  mammals,  amphibians,  rep- 
tiles, butterflies,  damselflies  and  dragonflies, 
and  orchids  and  other  plant  species  referenced 
in  the  book,  and  it  provides  weather  data  for 
selected  cities.  The  authors  also  discuss  Mich- 
igan’s few  potential  hazards  to  birders,  from 
the  prevalent  (e.g.,  biting  insects  and  weather) 
to  the  least  likely  (e.g.,  black  bears,  moose, 
and  massasauga  rattlesnakes).  Finally,  the 
book  lists  contact  information  for  Michigan 
tourism  councils,  birding-related  telephone 
hotlines,  internet  chat  groups,  websites,  festi- 
vals, and  parks  and  conservation  organiza- 
tions. 

The  birding  site  descriptions  are  organized 
into  six  regions  of  the  state;  Southeastern  LP, 
Northeastern  LP,  Northwestern  LP,  South- 
western LP,  Eastern  UP,  and  Western  UP.  Pre- 
ceding each  of  these  sections  is  a map  illus- 
trating the  region’s  major  birding  areas  and 
the  alpha-numerical  identifiers  used  for  bird- 


432 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


ing  sites  in  that  area.  For  instance,  the  regional 
map  of  the  Southeastern  LP  indicates  that  the 
“St.  Clair  Marshes”  is  birding  area  #10,  for 
which  sites  SE67  to  SE71  are  listed.  After 
paging  to  the  site  description  for  SE67,  the 
user  will  find  a more  detailed  map  showing 
the  locations  of  all  five  sites  in  the  St.  Clair 
Marshes  area.  For  a given  site,  the  authors 
have  included  seasonal  ratings  of  the  site’s 
birding  quality,  as  well  as  the  latitude/longi- 
tude reference  and  the  page  number  and  grid 
location  where  one  would  find  that  site  in  the 
Delorme  Atlas.  The  directions  for  getting  to 
site  SE67 — Metro  Beach  Metropark,  one  of 
the  most  popular  migrant  traps  among  Detroit- 
area  birders — advise  the  reader  that  taking  I- 
94  East  actually  entails  traveling  north  from 
Detroit.  This  is  one  example  of  the  detail  and 
thought  that  went  into  the  directions  to  all 
sites  included  in  the  book.  The  authors  also 
advise  visitors  to  call  ahead  for  the  park’s 
hours  of  operation,  warns  that  the  park  is  pop- 
ular with  non-birders,  and  that  birders  should 
check  South  Beach  at  Metro  Beach  first,  be- 
fore the  non-birders  arrive. 

In  another  location  at  Metro  Beach — Pt. 
Rosa  Marsh — I was  surprised  to  learn  that  as 
many  as  500  Common  Loons  have  been  tal- 
lied in  one  day  during  spring  migration.  The 
text  also  mentions  that  the  bushes  behind  the 
nature  center  are  a reliable  place  to  find  the 
elusive  Connecticut  Warbler,  and  that  the 
Meadow  Area  should  be  checked  for  Red- 
headed Woodpecker,  Orchard  Oriole,  and  Yel- 
low-breasted Chat — all  uncommon  in  Michi- 
gan. Rarities  that  have  made  appearances  here, 
such  as  Magnificent  Frigatebird,  Great  White 
Heron,  and  Heerman’s  Gull,  are  mentioned  as 
well. 

Birding  areas  in  the  Northeast  LP  include 
groups  of  five  to  eight  sites,  each  being  close 
to  a state  highway  or  expressway;  thus,  each 
can  be  regarded  as  the  equivalent  of  a “bird- 
ing trail,”  such  as  those  promoted  in  Texas  or 
Minnesota.  Tawas  Point — a park  at  the  north 
end  of  Saginaw  Bay  on  Lake  Huron — is  one 
of  the  state’s  premier  migrant  traps  and  de- 
serves at  least  one  full  day  of  birding.  Men- 
tioned by  Jon  Dunn  as  “indeed  my  favorite 
place  to  bird  in  all  of  North  America,”  Tawas 
Point  truly  measures  up  to  such  high  praise. 
As  one  of  the  few  extremely  fortunate  birders 
to  have  been  with  Dunn  in  May  1996  to  see 


the  only  White-collared  Swift  recorded  in  the 
Midwest,  I can  personally  attest  to  the  magic 
that  can  happen  at  Tawas  Point.  Now  that  the 
park’s  greatness  is  no  longer  a secret,  Ziamo’s 
description  of  this  location  and  other  nearby 
sites  will  make  birding  in  the  Tawas  area 
obligatory  for  those  also  taking  a Kirtland’s 
Warbler  tour  in  the  nearby  Mio  area.  The  site 
description  mentions  the  park’s  seasonal  high- 
lights, including  Common  Loons  and  diving 
ducks  in  the  early  spring  and  late  fall,  and 
nesting  Piping  Plovers,  as  well  as  all  the  best 
nearby  places  for  observing  up  to  24  warbler 
species  and  many  other  passerines  in  a single 
day.  It  also  suggests  checking  the  pier  behind 
the  Holiday  Inn  for  waterfowl  and  along 
Brownell  Road  near  Tuttle  Marsh  to  listen  for 
Kirtland’s  Warbler — locations  of  which  I was 
unaware. 

An  even  more  famous  birding  destination 
in  Michigan — Whitefish  Point  Bird  Observa- 
tory (WPBO)  in  the  Eastern  UP — has  nine 
pages  devoted  to  it.  Along  with  an  enticing 
list  of  casual  and  accidental  sightings  from 
“the  point,”  the  authors  provide  a thorough 
history  of  WPBO  and  what  can  be  expected 
there  on  a seasonal  basis.  The  site  description 
also  includes  tables  listing  the  site’s  mean  ear- 
ly, late,  and  peak  dates  of  migration,  as  well 
as  seasonal  averages  and  minimum  and  max- 
imum counts  for  spring  and  fall  waterbird 
counts,  spring  raptor  counts,  and  owl  banding 
conducted  at  this  intensively  studied  migrant 
hotspot.  The  last  weekends  of  April  and  May, 
when  experienced  Michigan  birders  flock  to 
the  area,  are  recommended  as  especially  good 
birding  times  for  first-time  visitors.  Tradition- 
ally, Memorial  Day  weekend  is  considered  the 
beginning  of  tourist  season  in  the  UP;  thus, 
readers  are  rightly  warned  to  check  on  the 
opening  and  closing  times  of  restaurants  in  the 
nearby  town  of  Paradise  to  avoid  the  possi- 
bility of  going  hungry.  WPBO  visitors  also 
are  cautioned  that,  “even  in  Mid-May,  tem- 
peratures can  be  low  enough  to  require  winter 
clothes.”  As  one  who  has  shivered  through 
numerous  early  mornings  of  waterbird  watch- 
ing in  the  area,  I would  take  this  one  step 
farther  by  suggesting  that  one  bring  along 
some  winter  clothing  at  any  time  of  the  year 
for  birding  along  Lake  Superior. 

The  Western  UP,  up  to  a 1 2-hour  drive  from 
Detroit,  receives  much  less  coverage  from 


ORNITHOLOGICAL  LITERATURE 


433 


birders  than  the  Eastern  UP;  thus,  Michigan’s 
county  listers,  and  anyone  else  with  a sense 
of  adventure,  will  appreciate  the  guide’s  in- 
clusion of  33  sites  west  of  Luce  and  Mackinac 
counties.  One  of  the  lesser-known  birding 
sites  listed  is  the  Garden  Peninsula,  which 
projects  south  into  Lake  Michigan  towards 
Wisconsin’s  Door  Peninsula.  On  Garden  Pen- 
insula, the  State  Forest  campground  at  Portage 
Bay  is  an  excellent  spot  for  both  passerines 
and  shorebirds  in  the  fall;  however,  this  is  not 
mentioned  in  the  site  description,  illustrating 
that  there  are  many  birding  spots  yet  to  be 
discovered  in  the  UP,  especially  the  western 
portion.  I look  forward  to  making  another  La- 
bor Day  weekend  trip  there  soon,  and  I’ll  be 
sure  that  my  itinerary  includes  two  other  plac- 
es described  for  that  area — the  Mead  Planta- 
tion and  the  Nahma  Marsh  Trail.  With  the 
Stonington  Peninsula  being  so  close  to  the 
Garden  Peninsula,  I’ll  have  to  visit  there  as 
well.  The  guide  makes  Peninsula  Point  Park 
sound  like  an  excellent  migrant  trap  and,  con- 
sidering how  little  old-growth  forest  is  left  in 
the  state,  the  hemlock  stand  at  Squaw  Creek 
also  sounds  intriguing. 

At  660  pages  long,  this  is  a very  thick  bird- 
ing guide,  and  it  can  be  difficult  to  make  it  lie 
open.  The  back  cover,  however,  extends  an  ad- 
ditional 4.5  inches  for  use  as  a bookmark.  In- 
side the  back  cover  is  a handy  state  map  de- 
noting the  state’s  birding  regions  and  selected 
birding  sites.  On  the  map,  sites  are  labeled 
according  to  the  page  numbers  where  their  de- 
scriptions are  located.  The  facing  page  has  a 
map  key,  which  lists  all  the  birding  sites  and 
their  page  numbers  for  each  of  the  state’s  six 
regions. 

I saw  only  a few  errors  in  this  guide.  One 
pertained  to  a birding  site  near  where  I live  in 
Genesee  County  (in  the  Southeastern  LP);  the 
site  was  mislabeled  as  being  presented  on 
page  42  and  occurring  in  adjoining  Livingston 
County.  After  checking  the  text,  however,  I 
found  that  there  was  no  birding  site  in  Liv- 
ingston County,  and  page  42  actually  de- 
scribes the  site  labeled  as  occurring  on  page 
43 — Gratiot-Saginaw  State  Game  Area,  locat- 
ed about  thirty  miles  to  the  northwest  of  Liv- 
ingston County.  Clare  County  is  misspelled  on 
the  state  map  on  the  inside  back  cover.  I also 
noticed  that  there  are  two  different  area  codes 
listed  in  the  site  description  for  Metro  Beach 


Metropark’s  phone  number.  Noted  in  the 
guide’s  introduction  is  a request  to  send  any 
comments  and  corrections  to  ABA’s  website 
for  use  in  future  editions  of  the  guide. 

In  conclusion,  all  Michigan  birders,  and 
anyone  else  planning  a birding  trip  to  that 
state,  should  own  a copy  of  A Birder’s  Guide 
to  Michigan.  There  is  no  other  guide  like  it 
for  the  state,  and  its  detail  and  completeness 
are  impressive.  Thanks  to  Chartier  and  Ziarno 
for  providing  such  a useful  tool  to  promote 
more  complete  birding  coverage  of  Michigan 
and  for  giving  out-of-state  birders  such  a user- 
friendly  guide  for  discovering  all  that  Michigan 
has  to  offer.— JEFF  A.  BUECKING,  Michigan 
Rare  Birds  Committee,  1225  Dauner  Rd.,  Fen- 
ton, Michigan;  e-mail:  jbuecking@juno.com 


A FIELD  GUIDE  TO  THE  BIRDS  OF 
THE  GAMBIA  AND  SENEGAL.  By  Clive 
Barlow  and  Tim  Wacher.  Yale  University 
Press,  New  Haven,  Connecticut.  2006:  400 
pp.,  48  color  plates.  ISBN:  0-300-11574-1. 
$40.00  (paper). — This  comprehensive  guide 
has  been  very  popular  with  birders  for  its  in- 
clusion of  many  tropical  African  birds.  It  was 
first  published  in  1997  in  the  United  Kingdom 
by  Christopher  Helm,  then  reprinted  with 
amendments  in  1999,  and  now  it  has  been  re- 
leased again  in  paperback  by  Yale  University. 
It  is  the  first  field  guide  to  the  birds  of  Gambia 
and  Senegal,  and  includes  other  areas  of  West 
Africa  popular  with  birders  from  around  the 
world. 

Clive  Barlow  has  lived  in  the  Gambia  area 
since  1985,  and  has  become  very  familiar  with 
the  region’s  bird  fauna.  He  presently  runs 
birdwatching  safaris  and  is  very  active  with 
the  conservation  of  Gambian  birds  through  ef- 
forts in  the  Kiang  West  National  Park  and 
Tanji  Bird  Reserve  conservation  areas.  Tim 
Wacher,  a mammalian  ecologist,  resided  in 
Gambia  for  five  years,  where  he  assembled  a 
database  of  bird  records  from  which  came 
most  of  the  distributional  information  for  this 
book. 

This  400-page  guide  provides  full  accounts 
of  more  than  600  bird  species  and  depicts 
nearly  all  of  them  in  the  48  color  plates  clus- 
tered at  the  forefront  of  this  attractive  volume. 
The  end-boards  depict  maps  of  both  Senegal 


434 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  3,  September  2006 


and  Gambia,  and  the  nine-page  introduction 
provides  short,  but  useful,  discussions  on  the 
region’s  geography,  climate,  vegetation,  and 
major  habitats.  The  habitat  descriptions  in- 
clude marine,  coastal,  estuarine,  mangrove, 
freshwater  riverbank,  and  other  wetland  hab- 
itats, as  well  as  farmlands  and  villages,  hotel 
gardens,  Guinea  savanna,  Sudan  savanna,  and 
dry  Sahel  of  northern  Senegal.  The  habitat 
section  is  followed  by  a short  section  on  the 
Sejegambian  avifauna,  which  boasts  over  660 
species,  about  a third  of  which  are  migrants 
from  the  Palearctic  region.  Additionally,  there 
are  descriptions  and  locator  maps  of  the  pro- 
tected areas  in  Gambia  and  Senegal,  a short 
discussion  that  will  aid  the  reader  in  using  this 
book,  and  illustrations  of  avian  plumage  to- 
pography that  should  be  useful  in  understand- 
ing the  keys  and  descriptions  throughout  the 
text. 

High-quality  plates  are  an  important  feature 
in  any  field  guide,  and  the  present  volume 
meets  that  criterion  nicely.  The  48  plates, 
however,  provide  rather  small  images,  which 
reduces  the  size  of  key  characteristics  used  for 
identification.  The  plates  also  lack  arrows 
pointing  out  key  identification  characteristics. 
Nonetheless,  they  are  of  excellent  quality  and 
will  prove  highly  useful  for  anyone  visiting 
Gambia  and  Senegal  or  surrounding  areas. 

Each  species  account  includes  the  species’ 
common  and  scientific  names,  relevant  plate 
numbers,  and  a comprehensive  section  on 
identification.  Comments  on  similar  or  con- 
fusing species  are  followed  by  remarks  on 
flight  characteristics,  habits,  voice,  status  and 
distribution,  and  reproduction,  as  well  as 
when  migrant  species  typically  appear.  Occa- 
sional vignettes  illustrate  such  things  as  the 
differences  in  the  nests  of  weaver  birds,  char- 
acteristic patterns  of  gull  flights,  and  aerial 
song-flight  displays  among  Cisticola  species. 
Most  field  guides  provide  range  maps  for  each 
species,  but  this  guide  provides  none.  This 
omission  may  be  due  to  the  fact  that  nearly 
one-third  of  the  species  are  migratory,  but 
range  maps  would  have  been  very  useful  for 
resident  species.  Following  the  species  ac- 
counts, this  guide  provides  a listing  of  three 
conservation  organizations  and  their  member- 
ship information,  a bibliography  of  cited  ref- 
erences, and  an  index  of  English  and  scientific 
names  that  will  allow  those  familiar  with  the 


region’s  avifauna  to  easily  locate  species  ac- 
counts and  plates. 

Overall,  the  authors  certainly  should  be 
commended  for  producing  such  a compact 
and  badly  needed  field  guide  for  Gambia  and 
Senegal.  I found  it  reasonably  priced  and  a 
welcome  resource  for  those  planning  to  visit 
the  area  and  enjoy  its  diversity  and  abundance 
of  resident  and  migratory  species. — HARLAN 
D.  WALLEY,  Department  of  Biology,  North- 
ern Illinois  University,  DeKalb;  e-mail:  hdw@ 
niu.edu 


BIRDS  OF  TROPICAL  AMERICA:  A 
WATCHER’S  INTRODUCTION  TO  BE- 
HAVIOR, BREEDING,  AND  DIVERSITY. 
By  Steven  Hilty.  University  of  Texas  Press, 
Austin.  2005:  312  pp.,  12  black-and-white  il- 
lustrations. ISBN:  0-292-70673-1.  $19.95  (pa- 
per).— This  title  was  originally  published  by 
Chapters  Publishing  of  Shelburne,  Vermont, 
as  part  of  their  The  Curious  Naturalist  series, 
and  then  it  was  reprinted  in  2005  by  the  Uni- 
versity of  Texas  Press  with  an  updated  sug- 
gested reading  list  and  epilogue.  After  being 
out  of  print  for  several  years,  this  particularly 
well-written  book  is  finally  back  in  print  and 
readily  available  to  interested  readers. 

Steven  Hilty  discusses  issues  of  tropical  or- 
nithology in  a readable  and  engaging  manner. 
He  has  organized  the  book  in  a series  of  twen- 
ty stand-alone  essays,  each  of  which  focuses 
on  a theme  related  to  Neotropical  birds.  The 
essays  are  as  varied  as  tropical  habitats  and 
the  birds  they  support.  Not  only  do  they  ed- 
ucate and  entertain  the  reader,  they  provide 
some  insight  as  to  why  tropical  habitats  and 
birds  are  so  different  from  those  of  northern 
latitudes.  The  text  is  enhanced  by  black-and- 
white  illustrations  of  tropical  birds  in  their 
habitats. 

Initial  chapters  cover  avian  community 
structure  and  diversity  of  Neotropical  rain  for- 
ests, biogeography  of  the  Amazon  River  ba- 
sin, and  how  the  most  recent  Ice  Age  affected 
bird  distribution,  migration,  and  mixed-spe- 
cies flocks.  Subsequent  essays  cover  ant 
swarms  and  the  bird  species  that  follow  them; 
avian  coloration;  fruit,  frugivory,  and  avian 
dispersal  of  seeds;  displays  performed  by 
manakins  and  cotingas;  hummingbird  forag- 


ORNITHOLOGICAL  LITERATURE 


435 


ing  strategies;  hummingbirds,  flycatchers,  vul- 
tures, and  caciques  that  inhabit  high  altitudes; 
vocal  production  and  sound  characteristics; 
ecology  of  island  specialists  in  the  Amazon 
River  basin;  and  seasonality  in  the  tropics.  I 
particularly  enjoyed  Hilty’s  explanations  of 
commonly  observed  behaviors,  including  the 
song  flight  of  the  Blue-black  Grassquit  ( Vol - 
atinia  jacarinci). 


This  book  is  recommended  for  all  those 
interested  in  tropical  birds  and  birding.  It 
would  make  an  interesting  collection  of  read- 
ings for  an  ornithology  class  or  a good  read 
for  your  next  tropical  birding  trip. — MARY 
GUSTAFSON,  Rio  Grande  Joint  Venture, 
Texas  Parks  and  Wildlife  Department,  Mis- 
sion, Texas;  e-mail:  mary.gustafson@tpwd. 
state. tx. us 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY 


Editor  JAMES  A.  SEDGWICK 
U.S.  Geological  Survey 
Fort  Collins  Science  Center 
2150  Centre  Ave.,  Bldg.  C. 

Fort  Collins,  CO  80256-8118,  USA 
E-mail:  wjo@usgs.gov 


Managing  Editor  CYNTHIA  MELCHER 

Copy  Editors  ALISON  GOFFREDI 

JULIETTE  WILSON 


Editorial  Board  KATHY  G.  BEAL 
CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 
KARL  E.  MILLER 


Review  Editor  MARY  GUSTAFSON 

Texas  Parks  and  Wildlife  Dept. 

2800  S.  Bentsen  Palm  Dr. 

Mission,  TX  78572,  USA 
E-mail:  WilsonBookReview@aol.com 


GUIDELINES  FOR  AUTHORS 

Consult  the  detailed  “Guidelines  for  Authors”  found  on  the  Wilson  Ornithological  Society  Web  site  (http:// 
www.ummz.lsa.umich.edu/birds/wilsonbull.html).  Beginning  in  2007,  Clait  E.  Brain  will  become  the  new  editor 
of  The  Wilson  Journal  of  Ornithology.  As  of  1 July  2006,  all  manuscript  submissions  and  revisions  should  be 
sent  to  Clait  E.  Brain,  Editor,  The  Wilson  Journal  of  Ornithology,  5572  North  Ventana  Vista  Rd.,  Tucson,  AZ 
85750-7204,  USA.  The  New  Wilson  Journal  of  Ornithology  office  and  fax  telephone  number  will  be  (520)  529- 
0365,  and  the  E-mail  address  will  be  TWilsonJO@comcast.net. 

NOTICE  OF  CHANGE  OF  ADDRESS 

If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornithological 
Societies  of  North  America,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  %The  Museum  of  Zoology,  The 
Univ.  of  Michigan,  Ann  Arbor,  MI  48109.  Persons  having  business  with  any  of  the  officers  may  address  them 
at  their  various  addresses  given  on  the  inside  of  the  front  cover,  and  all  matters  pertaining  to  the  journal  should 
be  sent  directly  to  the  Editor. 


MEMBERSHIP  INQUIRIES 

Membership  inquiries  should  be  sent  to  James  L.  Ingold,  Dept,  of  Biological  Sciences,  Louisiana  State  Univ., 
Shreveport,  LA  71115;  e-mail:  jingold@pilot.lsus.edu 

THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 

The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  in  the  Univ.  of 
Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  Univ.  of  Michigan  in  1930.  Until  1947 
the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  ornithological  magazines  from 
members  and  friends  of  the  Society.  Two  members  have  generously  established  a fund  for  the  purchase  of  new 
books;  members  and  friends  are  invited  to  maintain  the  fund  by  regular  contribution.  The  fund  will  be  admin- 
istered by  the  Library  Committee.  Terry  L.  Root,  Univ.  of  Michigan,  is  Chairman  of  the  Committee.  The  Library 
currently  receives  over  200  periodicals  as  gifts  and  in  exchange  for  The  Wilson  Journal  of  Ornithology.  For 
information  on  the  Library  and  our  holdings,  see  the  Society’s  web  page  at  http://www.ummz.lsa.umich.edu/ 
birds/wos.html.  With  the  usual  exception  of  rare  books,  any  item  in  the  Library  may  be  borrowed  by  members 
of  the  Society  and  will  be  sent  prepaid  (by  the  Univ.  of  Michigan)  to  any  address  in  the  United  States,  its 
possessions,  or  Canada.  Return  postage  is  paid  by  the  borrower.  Inquiries  and  requests  by  borrowers,  as  well  as 
gifts  of  books,  pamphlets,  reprints,  and  magazines,  should  be  addressed  to:  Josselyn  Van  Tyne  Memorial  Library, 
Museum  of  Zoology,  The  Univ.  of  Michigan,  1109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079,  USA.  Contri- 
butions to  the  New  Book  Fund  should  be  sent  to  the  Treasurer. 


This  issue  of  The  Wilson  Journal  of  Ornithology  was  published  on  22  September  2006. 


438 


Continued  from  outside  back  cover 


418  Brown-headed  Cowbirds  fatal  attempt  to  parasitize  a Carolina  Chickadee  nest 
David  A.  Zuwerink  and  James  S.  Marshall 

420  Likely  predation  of  adult  Glossy  Ibis  by  Great  Black-backed  Gulls 
Christina  E.  Donehower 

422  Tailless  whipscorpion  (Phrynus  longipes ) feeds  on  Antillean  Crested  Hummingbird  ( Orthorhyncus 
cristatus ) 

Jennifer  L.  Owen  and  James  C.  Cokendolpher 

424  Polydactyly  in  a Vaux’s  Swift 
Walter  H.  Sakai 

427  Once  Upon  a Time  in  American  Ornithology 
430  Ornithological  Literature 


The  Wilson  Journal  of  Ornithology 

(formerly  The  Wilson  Bulletin) 

Volume  118,  Number  3 CONTENTS  September  2006 

Major  Articles 

28 1 Nest-site  selection  and  productivity  of  American  Dippers  in  the  Oregon  Coast  Range 
John  P.  Loegering  and  Robert  G.  Anthony 

295  Upland  bird  communities  on  Santo,  Vanuatu,  Southwest  Pacific 
Andrew  W Kratter,  Jeremy  J.  Kirchman,  and  David  W.  Steadman 

309  A description  of  the  first  Micro nesian  Honeyeater  (. Myzomela  rubratra  sajfordi ) nests  found  on  Saipan, 
Mariana  Islands 

Thalia  Sachtleben,  Jennifer  L.  Reidy,  and  Julie  A.  Savidge 

316  Within-pair  interactions  and  parental  behavior  of  Cerulean  Warblers  breeding  in  eastern  Ontario 
Jennifer  J.  Barg,  Jason  Jones,  M.  Katharine  Girvan,  and  Raleigh  J.  Robertson 

326  Comparative  spring  migration  arrival  dates  in  the  two  morphs  of  White- throated  Sparrow 
Sarah  S.  A.  Caldwell  and  Alexander  M.  Mills 

333  Can  supplemental  foraging  perches  enhance  habitat  for  endangered  San  Clemente  Loggerhead  Shrikes? 
Suellen  Lynn,  John  A.  Martin,  and  David  K Garcelon 

341  Do  American  Robins  acquire  songs  by  both  imitating  and  inventing? 

Steven  L.  Johnson 

353  Effects  of  mowing  and  burning  on  shrubland  and  grassland  birds  on  Nantucket  Island,  Massachusetts 
Benjamin  Zuckerberg  and  Peter  D.  Vickery 

364  Spatial  behavior  of  European  Robins  during  migratory  stopovers:  a telemetry  study 
Nikita  Chernetsov  and  Andrey  Mukhin 

374  Age-related  timing  and  patterns  of  prebasic  body  molt  in  wood  warblers  (Parulidae) 

Christine  A.  Debruyne,  Janice  M.  Hughes,  and  David  J.  T.  Hussell 

380  Foraging  ecology  of  Bald  Eagles  at  an  urban  landfill 

Kyle  H.  Elliott,  Jason  Dujfe,  Sandi  L.  Lee,  Pierre  Mineau,  and  John  E.  Elliott 

391  Territory  selection  by  upland  Red-winged  Blackbirds  in  experimental  restoration  plots 
Maria  A.  Furey  and  Dirk  E.  Burhans 

399  The  use  of  southern  Appalachian  wetlands  by  breeding  birds,  with  a focus  on  Neotropical  migratory 
species 

Jason  E Bulluck  and  Matthew  P.  Rowe 

Short  Communications 

411  Breeding  range  extension  of  the  Northern  Saw-whet  Owl  in  Quebec 

Christophe  Buidin,  Vann  Rochepault,  Michel  Savard,  and  Jean-Pierre  L.  Savard 

413  Carolina  Wren  nest  successfully  parasitized  by  House  Finch 
Douglas  R.  Wood  and  William  A.  Carter 

415  American  Coot  parasitism  on  Least  Bitterns 
Brian  D.  Peer 


Continued  on  inside  back  cover 


bCThe  Wilson  Journal 

of  Ornithology 

Volume  118,  Number  4,  December  2006 


UCT 

yBRAB' 


• a : ;oK' 

• ' ■ - »••••-•  -•  ' ' - 

■>••  ' "’v  *?/  ' * vJr:v»::,x 


: •:.  V: 

•r\% ,.i/<^'4rS2y-y .'  •:•  . 


\S& 


Published  by  the 
Wilson  Ornithological  Society 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  ornithologist. 


President — Doris  J.  Watt,  Dept,  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  IN  46556,  USA;  e-mail: 
dwatt@sai  ntmary  s . edu 

First  Vice-President — James  D.  Rising,  Dept,  of  Zoology,  Univ.  of  Toronto,  Toronto,  ON  M5S  3G5, 
Canada;  e-mail:  rising@zoo.utoronto.ca 

Second  Vice-President — E.  Dale  Kennedy,  Biology  Dept.,  Albion  College,  Albion,  MI  49224,  USA; 
e-mail:  dkennedy@albion.edu 

Editor — James  A.  Sedgwick,  U.S.  Geological  Survey,  Fort  Collins  Science  Center,  2150  Centre  Ave., 
Bldg.  C,  Fort  Collins,  CO  80526,  USA;  e-mail:  wjo@usgs.gov 

Secretary — John  A.  Smallwood,  Dept,  of  Biology  and  Molecular  Biology,  Montclair  State  University, 
Montclair,  NJ  07043,  USA;  e-mail:  smallwoodj@montclair.edu 

Treasurer — -Melinda  M.  Clark,  52684  Highland  Dr.,  South  Bend,  IN  46635,  USA;  e-mail:  MClark@tcservices.biz 

Elected  Council  Members— Mary  Bomberger  Brown,  Robert  L.  Curry,  and  James  R.  Hill,  III  (terms 
expire  2007);  Kathy  G.  Beal,  Daniel  Klem,  Jr.,  and  Douglas  W.  White  (terms  expire  2008);  Carla  J. 
Dove,  Greg  H.  Farley,  and  Mia  R.  Revels  (terms  expire  2009). 

Membership  dues  per  calendar  year  are:  Active,  $21.00;  Student,  $15.00;  Family,  $25.00;  Sustaining, 
$30.00;  Life  memberships,  $500  (payable  in  four  installments). 

The  Wilson  Journal  of  Ornithology  is  sent  to  all  members  not  in  arrears  for  dues. 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 
(formerly  The  Wilson  Bulletin) 

THE  WILSON  JOURNAL  OF  ORNITHOLOGY  (ISSN  1559-4491)  is  published  quarterly  in  March,  June, 
September,  and  December  by  the  Wilson  Ornithological  Society,  810  East  10th  St.,  Lawrence,  KS  66044-8897.  The 
subscription  price,  both  in  the  United  States  and  elsewhere,  is  $40.00  per  year.  Periodicals  postage  paid  at  Lawrence,  KS. 
POSTMASTER:  Send  address  changes  to  OSNA,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

All  articles  and  communications  for  publications  should  be  addressed  to  the  Editor.  Exchanges  should  be  addressed 
to  The  Josselyn  Van  Tyne  Memorial  Library,  Museum  of  Zoology,  Ann  Arbor,  Michigan  48 1 09. 

Subscriptions,  changes  of  address,  and  claims  for  undelivered  copies  should  be  sent  to  OSNA,  5400  Bosque  Blvd., 
Ste.  680,  Waco,  TX  76710.  Phone:  (254)  399-9636;  e-mail:  business@osnabirds.org.  Back  issues  or  single  copies  are 
available  for  $12.00  each.  Most  back  issues  of  the  journal  are  available  and  may  be  ordered  from  OSNA.  Special  prices 
will  be  quoted  for  quantity  orders.  All  issues  of  the  journal  published  before  2000  are  accessible  on  a free  Web  site  at  the 
Univ.  of  New  Mexico  library  (http://elibrary.unm.edu/sora/).  The  site  is  fully  searchable,  and  full-text  reproductions  of  all 
papers  (including  illustrations)  are  available  as  either  PDF  or  DjVu  files. 


© Copyright  2006  by  the  Wilson  Ornithological  Society 
Printed  by  Allen  Press,  Inc.,  Lawrence,  Kansas  66044,  U.S. A. 


COVER:  Wilson’s  Plover  ( Charadrius  wilsonia).  Illustration  by  Robin  Corcoran. 


® This  paper  meets  the  requirements  of  ANSI/NISO  Z39.48-1992  (Permanence  of  Paper). 


FRONTISPIECE.  Male  American  Restarts  ( Setophaga  ruticilla)  in  second-  (above)  and  after-second-year  (below) 
plumage.  Staicer  et  al.  (p.  439)  found  that  singing  behavior  changes  with  male  pairing  status;  although  a larger 
proportion  of  second-year  males  were  unpaired  than  after-second-year  males,  the  authors  found  no  evidence  that 
male  age  affected  singing  behavior.  Original  painting  (gouache  water  color  and  acrylic,  on  paper)  by  Barry  Kent 
Mac  Kay. 


Jt>e  Wilson  Journal 

of  Ornithology 


Published  by  the  Wilson  Ornithological  Society 


VOL.  118,  NO.  4 December  2006 PAGES  439-610 

The  Wilson  Journal  of  Ornithology  1 1 8(4):439 — 45 1 , 2006 

SINGING  BEHAVIOR  VARIES  WITH  BREEDING  STATUS  OF 
AMERICAN  REDSTARTS  ( SETOPHAGA  RUTICILLA) 


CYNTHIA  A.  STAICER,1 2  VICTORIA  INGALLS,24  AND  THOMAS  W.  SHERRY3 4 


ABSTRACT. — We  examined  the  relationship  between  singing  behavior  and  breeding  status  in  the  American 
Redstart  ( Setophaga  ruticilla ) by  analyzing  song  rates,  singing  mode  (Repeat  or  Serial),  and  variability  of  song 
delivery  in  relation  to  the  age  and  breeding  status  of  129  males  in  the  Hubbard  Brook  Experimental  Forest, 
New  Hampshire.  Unpaired  males  spent  most  of  their  time  (>90%)  after  dawn  singing  in  Repeat  mode,  whereas 
paired  males  sang  sporadically,  in  Serial  as  well  as  Repeat  mode  (51%  of  their  singing  time).  Males  who  lost 
their  mates  sang  in  Repeat  mode  at  rates  indistinguishable  from  males  who  had  not  yet  obtained  a mate.  Overall, 
unpaired  males  sang  in  Repeat  mode  at  significantly  higher  and  less  variable  rates  than  did  paired  males. 
Although  a larger  proportion  of  second-year  males  were  unpaired  than  after-second-year  males,  we  found  no 
evidence  that  age  affected  singing  behavior. 

We  also  assessed  the  effect  of  pairing  status  on  male  detectability  in  song-based  monitoring  surveys  (e.g., 
point  counts),  and  we  suggest  a field  protocol  for  identifying  unpaired  males.  Simulations  of  5-min  field  samples, 
obtained  from  continuous  samples  >3  hr  in  duration,  suggest  that  human  listeners  would  be  twice  as  likely  to 
detect  unpaired  males  as  paired  males.  This  result  suggests  that  surveys  based  on  aural  detections  may  be  biased 
in  favor  of  unpaired  males.  In  our  population,  >90%  of  males  who  sang  >40  Repeat  songs  in  5 min  were 
unpaired.  Unpaired  males  were  >3  times  as  likely  as  paired  males  to  sing  only  Repeat  songs  in  a given  5-min 
period.  These  results  suggest  that  it  may  be  possible  to  identify  unpaired  male  American  Redstarts  by  their  high 
singing  rates  of  exclusively  Repeat  songs.  Received  23  May  2005,  accepted  30  March  2006. 


Recent  interest  in  the  song  rates  of  male 
passerines  has  focused  on  the  information 
contained  in  a male’s  singing,  especially  that 
available  to  females  for  assessing  prospective 
mates  (e.g.,  Hoi-Leitner  et  al.  1995).  Many 
studies  have  found  that  females  prefer  males 
with  a higher  song  rate  (Gottlander  1987,  Ala- 


1 Biology  Dept.,  Dalhousie  Univ.,  Halifax,  NS  B3H 
4J1,  Canada. 

2 Dept,  of  Biology,  Marist  College,  Poughkeepsie, 
NY  12601,  USA. 

3 Dept,  of  Ecology  and  Evolutionary  Biology,  Tu- 
lane  Univ.,  New  Orleans,  LA  70118,  USA. 

4 Corresponding  author;  e-mail: 
victoria,  ingalls  @ marist.edu 


talo  et  al.  1990,  Westcott  1992,  Gentner  and 
Hulse  2000,  Nolan  and  Hill  2004),  perhaps 
because  song  rate  is  correlated  with  male 
health  (Saino  et  al.  1997,  Smith  and  Moore 
2003),  dominance  in  winter  flocks  (Otter  et  al. 
1997),  food  abundance  before  female  arrival 
(Nystrom  1 997),  time  on  territory  since  arrival 
(Arvidsson  and  Neergaard  1991),  territory 
quality  (Radesater  and  Jakobsson  1989),  egg 
size  (Smith  and  Moore  2003),  feeding  rate  of 
older  chicks  by  the  male  (Hofstad  et  al.  2002), 
and  subsequent  nest  success  (Hoi-Leitner  et 
al.  1995).  Thus,  song  rate  appears  to  be  an 
honest  signal  of  male  quality  in  many  species. 

Song  rate  also  may  be  an  honest  signal  of 


439 


440 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


pairing  status,  since  unpaired  males  typically 
sing  more  than  their  paired,  nesting  neighbors 
(Hayes  et  al.  1986,  Ratti  and  Siikamaki  1993, 
Staicer  1996b,  Gil  et  al.  1999,  Amrhein  et  al. 
2004),  and  males  who  lose  their  mates  in- 
crease their  song  output  (Johnson  1983,  Han- 
ski  and  Laurila  1993).  Field  experiments  have 
shown  clear  effects  of  pairing  status  on  male 
song,  with  an  increase  in  singing  after  female 
removal  and  a decrease  to  pre-removal  levels 
after  female  return  (Krebs  et  al.  1981,  Cuthill 
and  Hindmarsh  1985,  Staicer  1996b).  If  fe- 
males can  use  these  differences  in  singing  be- 
havior and  song  rates  to  locate  unpaired  males 
in  a population,  then  perhaps  male  singing  be- 
havior contains  sufficient  information  for  hu- 
mans to  distinguish  paired  and  unpaired  males 
when  monitoring  songbird  populations. 

Typically,  songbird  monitoring  techniques 
involve  counts  of  singing  males  to  obtain  an 
estimate  of  the  number  of  breeding  pairs  at  a 
site  (e.g.,  Ralph  et  al.  1995),  but,  if  some  pro- 
portion of  singing  males  remains  unpaired, 
these  estimates  may  be  biased  and  confound 
comparisons  among  sites  (Rappole  1995). 
Males  that  remain  unmated  throughout  the 
breeding  season  are  not  uncommon  in  many 
socially  monogamous  species  (Breitwisch 
1989,  Marra  and  Holmes  1997).  For  example, 
in  populations  of  the  American  Redstart  (Se- 
tophaga  ruticilla ) — a Neotropical  migrant 
species  (Parulidae) — over  half  the  yearling 
males  remain  unmated  due  to  polygyny  (pre- 
dominantly in  older  males)  and,  possibly,  to 
disproportionate  female  mortality  at  various 
times  of  the  year  (Secunda  and  Sherry  1991, 
Sherry  and  Holmes  1997).  Moreover,  in  other 
parulids  habitat  fragmentation  has  been  asso- 
ciated with  edge-  and  patch-size-related  ex- 
cesses of  unmated  males  (Faaborg  et  al.  1995, 
Faaborg  2002),  possibly  in  relation  to  altered 
habitat  quality  or  dispersal  behavior.  The  re- 
sulting variability  in  male  mating  opportuni- 
ties could  influence  life-history  evolution. 
These  considerations  illustrate  why  precise 
determination  of  mating  status  is  important, 
and  song  behavior  provides  a diagnostic  tool 
(e.g.,  Gibbs  and  Faaborg  1990).  Song  behav- 
ior, and  its  interpretation,  is  also  crucial  for 
monitoring  populations  of  migratory  species 
like  the  American  Redstart  even  if  populations 
of  many  such  species  are  not  as  imminently 
threatened  as  once  thought  (Faaborg  2002). 


Few  researchers  have  quantified  the  differ- 
ences in  male  song  rates  with  respect  to  mat- 
ing status  or  breeding  stage  (e.g.,  Searcy  et  al. 
1991,  Nemeth  1996),  nor  have  most  research- 
ers considered  how  song  rate  may  bias  pop- 
ulation estimates  (Best  1981,  Hayes  et  al. 
1986,  Gibbs  and  Wenny  1993,  McShea  and 
Rappole  1997).  If  unpaired  males  could  be 
distinguished  from  paired  males  by  their  sing- 
ing behavior,  then  more  accurate  estimates  of 
population  density  and  habitat  quality  could 
be  obtained.  Although  the  American  Red- 
start— a species  in  which  many  males  often 
fail  to  obtain  a mate — has  been  the  subject  of 
many  studies  (reviewed  in  Sherry  and  Holmes 
1997),  the  species’  song  rate  has  not  been  ex- 
amined. 

Most  of  the  closely  related  Dendroica,  Ver- 
mivora,  Mniotilta,  Parula,  and  Setophaga  spe- 
cies have  two  categories  of  song  and  they  use 
these  in  different  social  contexts,  suggesting  a 
functional  dichotomy  (e.g.,  Ficken  and  Ficken 
1965;  Morse  1970;  Kroodsma  1981;  Lemon 
et  al.  1985;  Spector  1992;  Staicer  1989;  Wea- 
ry et  al.  1994;  Staicer  1996a,b;  Staicer  et  al. 
1996).  In  Repeat  mode,  which  is  more  com- 
mon early  in  the  season  before  pairing,  males 
sing  one  song  type  in  repetitive  fashion;  in 
Serial  mode,  which  is  more  common  later  in 
the  season,  they  alternate  among  two  or  more 
other  song  types  (Lemon  et  al.  1985,  1987). 
Thus,  any  study  involving  song  use  in  this 
species  must  consider  song  modes. 

The  delayed  plumage  maturation  of  Amer- 
ican Redstarts  has  received  much  interest 
(e.g..  Sherry  and  Holmes  1989,  Lozano  et  al. 
1996,  Perreault  et  al.  1997).  Yearling  adult 
male  American  Redstarts,  in  their  second  cal- 
endar year  of  life  (SY),  are  distinguishable  by 
plumage  from  older  males  (after-second-year, 
ASY),  making  it  easy  to  assess  the  effect  of 
age  on  singing  behavior.  Most  males  that  re- 
main unpaired  are  SY  (Lemon  et  al.  1987), 
but  whether  this  can  be  explained  by  song  is 
unclear  (e.g.,  Morris  and  Lemon  1988). 

The  primary  goal  of  our  study  was  to  ex- 
amine differences  in  the  singing  behavior  of 
paired  and  unpaired  male  American  Redstarts 
with  respect  to  song  rates,  regularity  of  song 
delivery,  and  use  of  song  mode.  In  addition, 
we  wanted  to  see  whether  (1)  the  breeding 
stage  of  females  would  influence  the  singing 
behavior  of  their  mates  and  (2)  whether  SY 


Staicer  et  al.  • SINGING  BEHAVIOR  AND  BREEDING  STATUS  IN  REDSTARTS 


441 


versus  ASY  males  differ  with  respect  to  sing- 
ing behavior.  Such  information  should  be  use- 
ful to  those  interested  in  monitoring  breeding 
populations  of  American  Redstarts  and  for 
stimulating  similar  investigations  of  related 
species. 

METHODS 

Study  area  and  subjects. — Our  main  study 
area  was  a 140-ha  stand  of  old,  second- 
growth,  northern  hardwood  forest  dominated 
by  yellow  birch  ( Betula  alleghaniensis ),  sugar 
maple  ( Acer  saccharum),  and  American  beech 
(Fagus  grandifolia ) in  the  Hubbard  Brook  Ex- 
perimental Forest,  White  Mountains,  New 
Hampshire  (Holmes  and  Sturges  1975).  Sub- 
jects were  male  American  Redstarts,  for 
which  breeding  data  were  being  collected  as 
part  of  a long-term  population  study  that  was 
independent  of  our  vocal  behavior  study. 
Males  defended  contiguous  territories  across 
the  study  area,  except  where  eastern  hemlock 
( Tsuga  canadensis ) and  other  conifers  domi- 
nated. Additional  observations  were  made  in 
adjacent  experimental,  regenerating  clear-cuts 
dominated  by  dense  stands  of  paper  birch  ( B . 
papyrifera). 

Classification  of  breeding  stages. — For 
paired  individuals,  we  classified  breeding 
stages  as  early  association  (the  first  hours  dur- 
ing which  a female  was  on  territory,  or  briefly 
visiting  and  then  moving  on  to  another  terri- 
tory, up  to  the  first  day  the  male  had  pair 
bonded  with  a female),  nest  prospecting  (fe- 
male associating  with  the  male  and  visiting 
various  tree  crotches),  nest  building,  egg  lay- 
ing, incubation,  dependence  (when  adults 
were  feeding  nestlings  or  fledglings),  or  lost 
mate  (some  nesting  females  disappeared  from 
the  territories  of  seven  males,  usually  coincid- 
ing with  nest  predation).  Information  on  the 
presence,  behavior,  and  pairing  and  breeding 
status  of  males  was  updated  every  few  days 
by  another  team  of  observers  who  banded 
birds,  mapped  territories,  and  monitored  nests. 

Extensive  song  sampling. — To  document 
what  songs  birds  were  singing  and  at  what 
rates,  we  recorded  singing  males  for  short  pe- 
riods throughout  the  breeding  season.  We  at- 
tempted to  record  each  singing  male  in  a giv- 
en area  for  at  least  5 min.  Samples  were  well 
distributed  across  the  study  area,  breeding  sea- 
son, and  hours  of  the  morning.  It  took  7 days 


to  cover  the  entire  study  area;  thus,  we  visited 
different  sections  on  consecutive  observation 
days,  repeating  the  cycle  every  7—10  days. 
These  extensive  samples  composed  our  main 
data  set  for  examining  the  relationship  be- 
tween singing  behavior  and  breeding  stage; 
they  did  not  reveal,  however,  whether  birds 
were  singing  at  a given  time  of  day,  because 
we  only  recorded  males  that  were  already 
singing. 

A total  of  129  different  males  were  record- 
ed over  parts  of  three  breeding  seasons  (23 
May-19  June  1991,  13  May-26  June  1992, 
and  8-23  June  1993).  We  recorded  10  males 
in  2 consecutive  years  and  one  male  in  all  3 
years.  Any  males  that  were  not  uniquely  col- 
or-banded were  identified  by  individual  plum- 
age; chest  markings  vary  among  males,  and 
drawings  were  made  for  those  without  bands. 
We  used  sonograms  to  confirm  the  identities 
of  males.  Individuals  have  fairly  unique  rep- 
ertoires and  the  songs  of  each  male  have 
unique  features,  making  sonograms  the  equiv- 
alent of  fingerprints.  We  determined  the  age 
of  males  (SY  versus  ASY)  by  plumage  col- 
oration (e.g..  Sherry  and  Holmes  1997). 

We  made  recordings  between  03:33  and 
15:45  EST,  mostly  between  sunrise  (—04:15) 
and  1 1 :00,  when  songbird  population  surveys 
are  typically  conducted.  We  recorded  songs  on 
Type  IV  metal  tape  using  a Marantz  PMD-222 
monaural  cassette  recorder  and  a Dan  Gibson 
parabolic  microphone.  Using  Sound-Edit  soft- 
ware on  a Macintosh  computer,  we  made  a 
sonogram  of  each  song  type  in  each  recording 
and  compared  sonograms  to  document  reper- 
toires and  verify  subject  identity.  Once  the 
sonograms  from  all  recordings  had  been  ex- 
amined, Repeat-  and  Serial-mode  songs  were 
identified  for  each  subject.  Typical  songs  re- 
corded from  the  study  population  are  present- 
ed in  Sherry  and  Holmes  (1997). 

Intensive  song  sampling. — To  assess  how 
singing  behavior  changed  throughout  the 
morning  hours,  and  to  provide  data  for  mod- 
eling detectability,  we  studied  a subset  of  nine 
(five  paired,  four  unpaired)  focal  males  more 
intensively.  Males  were  selected  for  ease  of 
study  (territories  accessible  at  dawn)  and  to 
encompass  a range  of  breeding  stages.  On 
mornings  in  early-  to  mid-breeding  season, 
starting  with  a focal  male’s  first  song  at  dawn, 
we  followed  each  male  for  210  min  continu- 


442 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


ously.  To  facilitate  maintaining  contact  with 
the  focal  male,  we  mapped  his  territory 
boundaries  and  studied  both  his  song  reper- 
toire and  that  of  his  neighbors  prior  to  the 
sampling  date.  We  made  sonograms  of  the  Se- 
rial and  Repeat  songs  of  the  focal  male  and 
his  neighbors,  and  learned  to  recognize  them 
by  ear.  For  each  song  the  focal  male  sang,  we 
noted  the  singing  mode  and  time  the  song  be- 
gan (measured  to  the  nearest  second  with  a 
stopwatch).  The  first  30  min  of  song  was  re- 
corded on  magnetic  tape,  and  for  the  remain- 
ing 180  min,  time  of  song  and  singing  mode 
were  tallied  on  data  sheets. 

Detectability. — We  used  the  intensive  sam- 
ples to  obtain  an  estimate  of  detectability  for 
paired  and  unpaired  males.  Samples  were  di- 
vided into  5-min  intervals;  we  considered  a 
male  “detected”  if  he  sang  at  least  one  song 
(in  either  Repeat  or  Serial  mode)  during  a giv- 
en 5-min  interval.  We  compared  the  propor- 
tion of  intervals  in  which  the  5 paired  and  4 
unpaired  males  sang.  Median  values  were 
used  as  estimates  of  the  detectability  of  paired 
and  unpaired  males. 

Calculations  for  song  rate  and  song  ca- 
dence.— For  each  extensive  sample,  we  cal- 
culated song  rate  (number  of  songs/min)  and 
cadence  (the  time  between  the  beginnings  of 
successive  songs;  Reynard  1963).  The  time 
from  the  start  of  one  song  to  the  beginning  of 
the  next  consecutive  song  was  measured  with 
a stopwatch;  the  median  value  per  sample  was 
used  for  all  analyses.  Cadence  is  essentially  a 
measure  of  the  male’s  singing  “rhythm.”  To 
quantify  the  variability  of  this  rhythm,  we 
used  the  coefficient  of  variation  (CV)  of  the 
cadence  (corrected  for  small  samples;  Sokal 
and  Rohlf  1995)  expressed  as  a percentage, 
and  hereafter  referred  to  as  cadence  CV;  a 
higher  cadence  CV  indicates  a more  irregular 
delivery  of  songs.  Whereas  song  rate  and  ca- 
dence should  be  negatively  correlated  (i.e.,  as 
song  rate  increases,  time  between  songs  nec- 
essarily decreases),  song  rate  and  cadence  CV 
need  not  be.  Additional  information  associat- 
ed with  each  sample  included  sample  dura- 
tion, date  and  time  of  day,  and  the  male’s 
identity,  age,  pairing  status  (paired  or  un- 
paired), breeding  stage  (if  paired),  and  singing 
mode  (Repeat  or  Serial). 

Statistical  analyses. — We  used  nonparamet- 
ric  tests  to  determine  whether  pairing  status. 


breeding  stage,  or  time  of  day  affected  song 
rate  or  cadence  CV.  Data  were  not  normally 
distributed  and  sample  sizes  for  some  groups 
were  small,  so  we  report  medians  instead  of 
means  as  a measure  of  central  tendency.  Mul- 
tiple samples  of  the  same  male  were  averaged 
so  that  each  male  contributed  a single  datum 
to  a given  group.  We  used  Mann-Whitney 
U- tests  to  compare  two  groups  of  males,  and 
all  tests  were  two-tailed  unless  otherwise  not- 
ed. To  determine  the  significance  of  Mann- 
Whitney  U- tests  involving  multiple  compari- 
sons, we  used  a sequential  Bonferroni  test  ( k 
comparisons  by  the  Dunn-Sidak  method)  and 
an  experiment-wise  a = 0.05  (Sokal  and 
Rohlf  1995).  We  report  the  significance  level 
of  each  test;  if  the  Bonferroni  revealed  signif- 
icance, we  also  report  the  Bonferroni-adjusted 
critical  value  (Padj).  We  also  calculated  Spear- 
man’s rank  correlations  to  examine  the  rela- 
tionship between  song  rate  and  cadence  CV. 

RESULTS 

Song  modes. — The  total  singing  time  cap- 
tured in  our  514  samples  of  129  males  was 
27.5  hr  (median  sample  duration  = 3.2  min). 
In  few  samples  (<2%),  males  switched  sing- 
ing modes;  for  these,  we  separated  the  Serial 
song  bouts  from  the  Repeat  bouts  before  anal- 
ysis. 

The  dawn  chorus  was  a period  of  intense 
singing  of  Serial-mode  songs.  Males  sang  in 
Serial  mode  at  greater  rates  at  dawn  (14.4 
songs/min,  n = 17  males)  than  they  did  later 
in  the  day  (10.3  songs/min;  n = 76  males; 
Mann-Whitney  U-test:  P < 0.001).  For  a sub- 
set of  eight  paired  males,  we  recorded  Serial 
mode  sequences  during  their  dawn  singing 
bouts  as  well  as  during  later  morning  bouts  on 
the  same  day.  These  males  sang  in  Serial 
mode  at  higher  rates  at  dawn  (15.3  songs/min) 
than  they  did  later  in  the  morning  (9.7  songs/ 
min;  one-tailed  Wilcoxon  Matched  Pairs  test: 
P = 0.006).  Because  of  the  robust  difference 
between  dawn  and  daytime  song  rates,  sub- 
sequent analyses  include  only  recordings  ob- 
tained after  sunrise  (i.e.,  daytime  songs). 

Post-sunrise  use  of  song  modes  varied  with 
pairing  status  and  nesting  stage.  When  multi- 
ple samples  from  the  same  male  in  the  same 
breeding  stage  were  averaged,  Repeat  mode 
comprised  68%  of  the  225  resulting  samples. 
Unpaired  males  sang  in  Repeat  mode  in  91% 


Staicer  et  al.  • SINGING  BEHAVIOR  AND  BREEDING  STATUS  IN  REDSTARTS 


443 


of  69  samples  and  males  who  lost  their  mate 
sang  in  Repeat  in  100%  of  7 samples.  In  the 
early  association  stage,  males  sang  in  Repeat 
mode  in  93%  of  15  samples  and  in  100%  of 
7 samples  during  the  nest  prospecting  stage. 
Once  males  were  nesting,  their  use  of  Repeat 
mode  declined.  Paired  males  sang  in  Repeat 
mode  in  51%  of  71  samples  during  the  nest- 
building period,  54%  of  13  samples  during  the 
egg-laying  period,  36%  of  31  samples  during 
the  incubation  period,  and  67%  of  12  samples 
during  the  dependence  period.  Overall,  use  of 
song  mode  after  sunrise  was  dependent  on 
pairing  status:  paired  males  sang  in  Repeat 
mode  in  only  51%  of  134  samples  compared 
to  unpaired  males  or  males  who  had  lost  their 
mates;  these  males  sang  in  repeat  mode  in 
92%  of  76  samples  (Chi-square  test  of  inde- 
pendence: x2  — 26.95,  df  = 1,  P < 0.001). 

After  dawn  song  rates  and  cadence  CV. — 
Unpaired  males  sang  in  Repeat  mode  at  sig- 
nificantly higher  rates  (8.0  songs/min,  n — 68 
males)  than  did  paired  males  (6.3  songs/min, 
n = 82  males;  Mann- Whitney  U-test  and  Bon- 
ferroni  adjustment:  P = 0.001,  Padj  = 0.013; 
Fig.  1A).  Unpaired  males  also  sang  in  Repeat 
mode  with  a significantly  less  variable  ca- 
dence (cadence  CV  = 25.3%)  than  did  paired 
males  (37.8%;  Mann-Whitney  U-test  and 
Bonferroni  adjustment:  P = 0.001  > Padj  ~ 
0.013;  Fig.  IB). 

Only  6 (8.7%)  of  the  unpaired  males  we 
recorded  sang  in  Serial  mode  after  dawn,  and 
they  did  so  only  on  1 day  of  observation  for 
a brief  period  (median  duration  of  recording 
= 1.0  min)  in  the  first  few  days  after  arrival. 
Their  Serial  song  rates  were  not  significantly 
different  (1 1.6  songs/min)  than  those  of  paired 
males  (10.1  songs/min,  n = 69  males;  Mann- 
Whitney  U-test:  P = 0.82;  Fig.  1A).  Further- 
more, when  unpaired  males  sang  in  serial 
mode  after  sunrise,  their  cadence  CV  was  sim- 
ilar to  that  of  paired  males  (Mann-Whitney 
U-test:  P = 0.61;  Fig.  IB). 

Overall,  males  sang  in  Serial  mode  at  sig- 
nificantly higher  rates  than  they  sang  in  Re- 
peat mode,  regardless  of  pairing  status  (Mann- 
Whitney  U-test  and  Bonferroni  adjustment  for 
paired  males:  P = 0.010,  Padj  = 0.017;  for 
unpaired  males:  P = 0.017,  Padj  = 0.025). 
Paired  males  sang  in  Serial  mode  with  a lower 
cadence  CV  (29.0%;  Mann-Whitney  U-test 
and  Bonferroni  adjustment:  P = 0.012,  Padj  = 


Repeat  Serial 


UN  p UN  p 
(68)  (82)  (6)  (69) 

Pairing  status 

FIG.  1 . Effects  of  pairing  status  on  (A)  song  rate 
and  (B)  variability  of  song  delivery  (cadence  CV)  for 
male  American  Redstarts  at  Hubbard  Brook  Experi- 
mental Forest,  New  Hampshire,  1991-1993.  Repeat 
and  Serial  mode  sequences  of  paired  (P)  and  unpaired 
(UN)  males  were  recorded  after  04:15  EST.  Higher  ca- 
dence CV  values  indicate  more  variation  in  timing  be- 
tween songs.  Sample  sizes  in  parentheses  indicate 
number  of  males;  for  a given  status,  multiple  samples 
per  male  were  averaged,  so  that  each  male  contributed 
a single  datum.  Samples  from  males  in  the  early  as- 
sociation stage  (early  stages  of  pairing  or  unpaired 
males  who  were  visited  briefly  by  unpaired  females) 
could  not  be  classified  unambiguously  and  were  ex- 
cluded from  this  analysis.  Box  plots  show  the  medians 
(horizontal  center  lines),  interquartile  ranges  (between 
the  upper  and  lower  edges  of  the  box,  within  which 
50%  of  the  data  lie),  values  within  ±1.5  times  the 
interquartile  range  (bars  extending  from  box  edges), 
and  outliers  (open  circles).  Unpaired  males  sang  in  Re- 
peat mode  significantly  faster  and  with  a more  regular 
cadence  than  paired  males  (Mann-Whitney  C-test; 
Bonferroni  adjustment  for  both  comparisons:  P adj  ~ 
0.013).  See  text  for  additional  results  and  statistical 
tests. 


444 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


0.017)  than  they  sang  in  Repeat  mode  (Fig. 
IB).  Unpaired  males  sang  in  Repeat  mode 
with  a similar  cadence  CV  as  did  paired  males 
singing  in  Serial  mode  (CV  = 27.0%;  Mann- 
Whitney  U-test:  P = 0.36). 

Cadence  CV  was  negatively  correlated  with 
song  rate  for  combined  Repeat-  and  Serial- 
mode samples  (Spearman’s  rank  correlation:  r 
= —0.41,  n = 219,  P < 0.001).  Results  were 
similar  for  Serial  mode  when  samples  were 
analyzed  separately  (r  = —0.46,  n = 75,  P < 
0.001).  For  Repeat-mode  samples,  the  nega- 
tive correlation  between  cadence  CV  and  song 
rate  was  strong  for  paired  males  (r  = -0.61, 
n = 76,  P < 0.001)  and  weak  for  unpaired 
males  (r  = -0.24,  n = 68,  P = 0.050);  thus, 
unpaired  males  sang  in  Repeat  mode  with  a 
more  regular  rhythm  than  paired  males,  re- 
gardless of  song  rate. 

Rates  of  Repeat  mode  song  also  changed 
with  breeding  stage  (Fig.  2A).  Males  who  lost 
their  mate  sang  at  rates  similar  to  those  who 
had  not  yet  paired  (8.3  versus  8.0  songs/min; 
Mann-Whitney  U-test:  P = 0.90).  Males  sang 
at  greater  rates  before  pairing  than  did  males 
whose  mates  were  nest  prospecting  (5.0 
songs/min;  Mann-Whitney  U-test  and  Bonfer- 
roni  adjustment:  P = 0.006,  Padj  — 0.010), 
nest  building  (6.6  songs/min;  P = 0.001,  Padj 
= 0.007),  incubating  (6.1  songs/min;  P = 
0.009,  Padj  = 0.013),  or  feeding  dependent 
young  (4.2  songs/min;  P = 0.002,  Padj  — 
0.009).  Repeat-song  rates  of  unpaired  males 
did  not  differ  significantly  from  those  of  males 
in  early  stages  of  pairing  (early  association 
stage,  6.5  songs/min,  P = 0.16),  or  in  the  egg- 
laying  stage  (6.9  songs/min;  P = 0.11;  Mann- 
Whitney  U-tests). 

Cadence  CV  of  Repeat  songs  also  changed 
with  breeding  stage  (Fig.  2B).  Again,  the  ca- 
dence CV  of  males  who  lost  their  mates 
(22.5%)  was  similar  to  that  of  males  who  had 
not  yet  paired  (25.1%;  Mann-Whitney 
U-test:  P = 0.79).  Before  pairing,  males  sang 
with  a significantly  more  regular  rhythm  than 
did  males  who  were  beginning  to  associate 
with  a female  (37.0%;  Mann-Whitney  U-test 
and  Bonferroni  adjustment:  P = 0.008,  Padj  = 
0.0 1 3)  or  paired  males  whose  mates  were  nest 
prospecting  (46.3%;  Mann-Whitney  U-test 
and  Bonferroni  adjustment:  P = 0.001  » P adj  ~ 
0.007),  nest  building  (38.7%;  P = 0.001,  Padj 
= 0.009),  or  feeding  dependent  young 


150 


8 100 
c 
<D 
T5 

O 50 
> 

O 


B 


aiHO* 


s /vyvv-^  jf 


v # A 


FIG.  2.  Effects  of  breeding  stage  on  (A)  song  rate 
and  (B)  variability  of  song  delivery  (cadence  CV)  for 
Repeat-mode  sequences  for  male  American  Redstarts 
at  Hubbard  Brook  Experimental  Forest,  New  Hamp- 
shire, 1991-1993.  Breeding  stage:  lost  mate,  before 
pairing,  early  association,  nest  prospecting,  nest  build- 
ing, egg  laying,  incubation,  and  dependence  (feeding 
nestlings  or  fledglings).  Sample  sizes  in  parentheses 
indicate  number  of  males;  often  a given  male  contrib- 
uted data  to  more  than  one  stage,  but  within  each  stage, 
all  data  were  independent  (i.e.,  multiple  samples  per 
male  were  averaged  to  obtain  a single  datum).  See  text 
for  explanations  of  statistical  tests  and  the  Figure  1 
caption  for  an  explanation  of  the  box  plots. 


(68.5%;  P = 0.002;  P^  = 0.010).  Cadence 
CV  of  unpaired  males  did  not  differ  from  that 
of  males  whose  mates  were  in  the  egg-laying 
stage  (CV  = 22.4%;  Mann-Whitney  U-test:  P 
— 0.88)  or  incubating  (35.9%;  P = 0.09). 
Thus,  although  song  rates  of  unpaired  males 
and  males  in  the  early  association  stage  did 
not  differ,  the  latter  sang  with  a less  regular 
rhythm.  Conversely,  although  song  rates  of 
unpaired  males  were  significantly  greater  than 
those  of  paired  males  whose  mates  were  in- 
cubating, both  groups  sang  with  a similarly 
regular  rhythm. 

Age  and  song  rate. — We  found  no  signifi- 


Staicer  et  al.  • SINGING  BEHAVIOR  AND  BREEDING  STATUS  IN  REDSTARTS 


445 


cant  age  effects  on  song  rate  (SY  versus  ASY 
males).  Unpaired  SY  and  ASY  males  sang  in 
Repeat  mode  at  similar  rates  (8.6  versus  8.0 
songs/min,  n = 32  versus  n — 28,  respective- 
ly; Mann-Whitney  U-test:  P — 0.24).  Paired 
SY  and  ASY  males  also  sang  in  Repeat  mode 
at  similar  rates  (4.9  versus  5.7  songs/min,  n 
= 17  versus  49,  respectively;  Mann-Whitney 
U-test:  P = 0.10).  Only  3 of  the  36  unpaired 
SY  males  that  we  observed  sang  in  Serial 
mode  after  the  dawn  bout.  For  paired  SY  and 
ASY  males  singing  in  Serial  mode,  song  rates 
were  similar  (10.6  and  10.2  songs/min,  n — 
1 1 and  43,  respectively;  Mann-Whitney 
U-test:  P = 0.76).  Thus,  song  rate  was  not 
affected  by  male  age,  regardless  of  pairing  sta- 
tus. The  similarity  in  singing  behaviors  of  SY 
and  ASY  males  can  be  seen  in  the  3.5-hr  sam- 
ples of  the  nine  focal  males  (Fig.  3). 

Temporal  patterns  in  song  activity  and 
pairing  status. — Obvious  differences  between 
paired  and  unpaired  males  with  regard  to  their 
singing  behaviors  are  illustrated  by  3.5-hr  song 
counts  for  the  nine  intensively  sampled  males 
(Fig.  3).  Typical  of  breeding  males,  the  five 
paired  males  (Fig.  3A)  sang  a large  number  of 
Serial  mode  songs  at  rapid  rates  during  their 
dawn  bouts.  Around  sunrise,  however,  paired 
males  usually  stopped  singing  and  for  the  rest 
of  the  morning  sang  sporadic,  but  typically  dis- 
tinct (not  mixed),  bouts  of  Repeat-  or  Serial- 
mode  songs.  During  the  incubation  stage,  some 
males  (e.g.,  10  June;  Fig.  3 A)  sang  little  on  their 
territory  after  their  dawn  bouts,  whereas  others 
(e.g.,  16  June;  Fig.  3 A)  sang  during  most  of  the 
5-min  periods  after  sunrise.  Temporal  patterns 
in  Serial-  and  Repeat-mode  song  activity  were 
similar  for  the  five  paired  males  (two  SY  and 
three  ASY  males). 

In  contrast,  the  four  males  who  lacked  es- 
tablished pair  bonds  (Fig.  3B)  sang  only  in  Re- 
peat mode  after  sunrise,  and  did  so  more  fre- 
quently and  at  higher  rates  than  paired  males. 
A male’s  time  on  territory  rather  than  date  or 
pairing  status  seemed  to  influence  whether  he 
sang  Serial  mode  in  the  dawn  chorus.  The  two 
unpaired  males  that  did  not  sing  in  serial  mode 
during  a dawn  bout,  but  sang  only  in  Repeat 
mode  before  04:00,  were  late  arrivals  in  the 
study  area  (28  May  and  10  June;  Fig.  3B).  Al- 
though these  SY  and  ASY  males  were  ob- 
served at  different  times  of  season,  both  had 
been  singing  for  only  a few  days  on  territories 


that  were  adjacent  to  contiguous  clusters  of  es- 
tablished territories.  The  other  two  unpaired 
males  (13  June  and  15  June;  Fig.  3B),  which 
had  defended  territories  within  a contiguous 
cluster  of  ASY  males  for  >10  days  by  the  time 
they  were  recorded,  sang  dawn  Serial  bouts 
like  those  of  their  paired  neighbors  but  then 
switched  at  sunrise  to  Repeat  mode  and  steadi- 
ly sang  in  that  mode  through  the  morning.  The 
male  who  attracted  a mate  during  the  obser- 
vation period  (13  June;  Fig.  3B)  sang  only  in 
Repeat  mode  but  at  a rate  that  decreased 
through  the  morning.  On  the  previous  days,  no 
female  was  present;  after  the  sample  date,  he 
remained  paired  and  commenced  nesting.  The 
male  who  lost  his  mate  after  her  nest  was  dep- 
redated (15  June;  Fig.  3 A)  sang  only  in  Repeat 
mode  after  sunrise,  but  at  a slightly  lower  rate 
and  with  less  regularity  than  did  the  males  who 
had  not  yet  paired. 

Confounding  factors. — To  test  whether  time 
of  day  or  time  of  season  influenced  Repeat- 
song  rates,  we  calculated  Spearman’s  rank 
correlation  coefficients.  Song  rates  of  un- 
paired males  were  negatively  correlated  with 
time  of  day  ( n = 70,  r = —0.350,  P = 0.010). 
For  paired  males,  however,  there  was  no  sig- 
nificant relationship  between  song  rate  and 
time  of  day  ( n — 54,  r - —0.10)  or  time  of 
season  (n  = 54,  r = —0.05),  and,  for  unpaired 
males,  there  was  no  correlation  between  song 
rate  and  time  of  season  (n  = 70,  r — —0.12; 
all  P > 0.10). 

Sampling  duration  was  another  potentially 
confounding  factor.  Although  Repeat-song 
rates  of  paired  and  unpaired  males  differed 
significantly,  data  for  the  two  groups  did  over- 
lap to  some  extent  (Fig.  1).  Overlap  between 
paired  and  unpaired  males,  however,  de- 
creased as  sample  duration  increased  (Fig.  4). 
In  samples  lasting  >5  min.  Repeat  song  rates 
of  paired  and  unpaired  males  overlapped  little. 
In  samples  of  >5-min  duration,  82%  of  27 
unpaired  males,  but  only  7%  of  paired  males, 
sang  >8  Repeat  songs/min.  In  samples  of  10- 
to  1 5-min  duration,  the  median  for  the  first  5 
min  was  similar  to  the  median  for  the  entire 
sample. 

Detectability. — Data  for  the  nine  intensively 
sampled  males  (Fig.  3)  were  split  into  5-min 
intervals  and  each  was  examined  for  occur- 
rence of  song.  Only  intervals  after  the  dawn 
chorus  were  used  (median  = 37,  range  = 35— 


Song  rate  (songs/min) 


446 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


16 

12 

8 

4 

0 

16 

12 

8 

4 

0 

16 

12 

8 

4 

0 

16 

12 

8 

4 

0 

16 

12 

8 

4 

0 


B 


■ Serial 
□ Repeat 


, Incubation  (SY)  16  June 

^aI/Aa 


0400  0500  0600  0700 

Time  of  day  (hours  EST) 


FIG.  3.  Singing  activity  of  nine  American  Redstart  males  in  various  breeding  stages  at  Hubbard  Brook 
Experimental  Forest,  New  Hampshire,  1992-1993.  SY  = yearlings,  ASY  = older  adults.  Areas  under  curves 
show  median  number  of  Serial  (black)  and  Repeat  (white)  songs  that  the  subject  sang  per  minute  for  each  5- 
min  period,  from  his  first  songs  at  dawn  until  3 hr  after  sunrise.  Sunrise  varied  from  04:10  (28  May)  to  04:05 
EST  (15  June),  as  indicated  by  arrows  on  the  x-axis.  Subjects  were  (A)  five  paired  males  and  (B)  two  unpaired 
males  within  a few  days  of  territory  establishment,  one  male  who  first  attracted  a mate  during  the  observation 
period,  and  one  male  whose  mate  had  disappeared  when  her  nest  was  depredated.  Note  the  larger  output  of 
Repeat-mode  songs  from  males  who  lacked  an  established  pair  bond  (B). 


Staicer  et  al.  • SINGING  BEHAVIOR  AND  BREEDING  STATUS  IN  REDSTARTS 


447 


Song  rate  (songs/min) 


FIG.  4.  Repeat-song  rates  of  paired  versus  un- 
paired male  American  Redstarts  using  samples  of  three 
durations  at  Hubbard  Brook  Experimental  Forest,  New 
Hampshire,  1991-1993.  For  each  duration,  a given 
male  was  entered  into  the  analysis  only  once.  (A)  Sam- 
ples of  short  duration  (0.3-2.9  min)  for  37  paired  and 
34  unpaired  males.  (B)  Samples  of  medium  duration 
(3. 0-4. 9 min)  for  24  paired  and  31  unpaired  males. 
(C)  Samples  of  long  duration  (5-15  min)  for  27  paired 
and  27  unpaired  males.  Note  that  as  sample  duration 
increased,  the  amount  of  overlap  between  the  two  sam- 
ples decreased. 


38  intervals  per  male).  Unpaired  males  ( n = 
4)  sang  in  99%  (median;  range  = 92-100%) 
of  the  5-min  intervals,  whereas  paired  males 
(n  = 5)  sang  in  only  49%  (median;  range  = 
16-74%)  of  the  5-min  intervals.  Detectability 
was  defined  as  the  proportion  of  5-min  inter- 
vals in  which  a bird  sang  one  or  more  songs. 
Detectability  of  unpaired  males  (0.99)  was 
significantly  greater  than  the  detectability  of 
paired  males  (0.49;  Mann- Whitney  t/-test:  P 
= 0.014). 


DISCUSSION 

Singing  behavior  and  breeding  status. — We 
identified  three  ways  in  which  the  singing  be- 
havior of  unpaired  male  American  Redstarts 
differed  significantly  from  that  of  paired 
males:  (1)  after  sunrise,  unpaired  males  sang 
in  Repeat  mode  almost  exclusively,  whereas 
paired  males  sang  in  both  modes;  (2)  unpaired 
males  sang  Repeat  songs  at  a significantly 
faster  rate  than  did  paired  males;  and  (3)  un- 


paired males  sang  with  a more  regular  ca- 
dence than  did  paired  males.  We  also  docu- 
mented variation  in  song  rates  and  regularity 
of  cadence  in  relation  to  breeding  stage  of 
paired  males. 

After  the  dawn  bout  ended,  use  of  Serial 
mode  varied  with  pairing  status  and  breeding 
stage.  In  almost  all  cases  in  which  we  heard 
Serial  mode  after  dawn,  it  was  delivered  by  a 
paired  male.  Use  of  Serial  mode  after  dawn 
may  reflect  the  presence  of  nests  or  young 
(see  also  Ficken  and  Ficken  1965,  Lemon  et 
al.  1985),  and  males  seem  to  have  the  greatest 
propensity  to  use  Serial  mode  (or  the  equiv- 
alent song  category  in  other  species)  when 
their  mates  are  incubating  (this  study;  Staicer 
1989,  1996b;  but  see  Lemon  et  al.  1987). 

Breeding  stage  also  affected  Repeat-song 
rates  and  cadence.  As  males  began  to  pair, 
they  continued  singing  primarily  in  Repeat 
mode,  but  cadence  became  more  irregular. 
Lowest  rates  of  singing  in  Repeat  mode  were 
found  in  males  whose  mates  were  building 
nests  and  males  who  were  feeding  nestlings 
or  fledglings.  Slower  song  rates  and  more  ir- 
regular cadences  have  been  associated  with 
the  activities  of  foraging  and  associating  with 
females  (e.g.,  Nolan  1978,  Gil  et  al.  1999). 
Although  we  had  relatively  few  song  samples 
from  the  egg-laying  stage,  these  males  some- 
times sang  for  brief  periods  at  rates  that  over- 
lapped those  of  unpaired  males.  Our  males, 
however,  were  silent  while  following  their 
mates;  thus,  we  found  no  evidence  that  song 
functions  to  guard  females  during  their  fertile 
period  (see  also  Titus  et  al.  1997).  Males  sang 
in  Repeat  mode  least  often  when  their  mates 
were  incubating,  a pattern  shared  with  other 
parulid  species  (Staicer  1989,  1996b;  but  see 
Lemon  et  al.  1987). 

Time  of  season  did  not  appear  to  alter  these 
singing  patterns.  Pairing  and  nesting  were 
asynchronous  in  our  population  due  to  differ- 
ent arrival  times  of  males  and  high  rates  of 
nest  predation,  after  which  females  sometimes 
disappeared  or,  rarely,  changed  mates.  Thus, 
at  any  given  time,  neighboring  males  often 
were  in  different  breeding  stages.  Males  who 
lost  their  mates  sang  at  high  rates,  similar  to 
males  before  they  were  paired.  This  change  in 
behavior  has  been  noted  for  other  wood-war- 
blers  (Nolan  1978,  Kroodsma  et  al.  1989, 
Spector  1991,  Staicer  1996b)  and  other  groups 


448 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  11H,  No.  4,  December  2006 


of  passerines  (e.g.,  Wasserman  1977,  Krebs  et 
al.  1981). 

To  determine  whether  females  select  males 
with  higher  song  rates,  additional  data,  such 
as  pairing  order,  male  condition  or  quality,  and 
territory  quality  must  be  obtained  (e.g.,  Hoi- 
Leitner  et  al.  1995,  Nystrom  1997).  If  vocal 
behavior  is  important  in  mate  choice,  how- 
ever, we  might  expect  to  find  differences  be- 
tween SY  and  ASY  male  American  Redstarts. 
We  found  no  evidence  that  age  affects  song 
rate  or  singing  mode  when  pairing  status  was 
taken  into  account.  Although  age  influences 
competitive  ability  (Sherry  and  Holmes 
1989),  pairing  success  (Morris  and  Lemon 
1988),  and  extra-pair  fertilizations  (Perreault 
et  al.  1997),  these  effects  appear  to  be  caused 
by  the  later  arrival  of  yearlings  rather  than  age 
effects  on  song  behavior  (Lozano  et  al.  1996; 
TWS  unpubl.  data). 

Implications  for  population  monitoring. — 
Few  researchers  have  examined  the  possibility 
of  distinguishing  unpaired  from  paired  males 
based  on  their  song  behaviors,  despite  the  po- 
tential utility  of  such  information  in  popula- 
tion monitoring.  Our  results  suggest  that  a 
considerable  amount  of  potentially  useful  in- 
formation is  available  in  the  singing  behavior 
of  male  American  Redstarts.  Unpaired  males 
sang  at  steadier  and  higher  rates,  took  fewer 
and  shorter  breaks  from  singing  (usually  <5 
min),  and  typically  sang  only  in  Repeat  mode 
after  sunrise.  After  the  dawn  chorus.  Serial 
mode  was  heard  from  paired  males  almost  ex- 
clusively; typically,  if  a male  sang  in  Serial 
mode,  he  was  paired.  A trained  ear  can  easily 
distinguish  Repeat  from  Serial  mode.  In  Re- 
peat mode,  the  same  song  type  is  repeated, 
whereas  in  Serial  mode,  males  rapidly  alter- 
nate between  2—5  noticeably  different  songs 
(e.g..  Lemon  et  al.  1985). 

In  5-min  samples  from  a large  number  of 
males,  the  Repeat-song  rates  of  unpaired  and 
paired  males  overlapped  little.  We  further  as- 
sessed the  information  available  in  a 5-min 
sample  by  combining  estimates  of  detectabil- 
ity (whether  a male  sang  any  songs  in  the  5- 
min  period)  with  the  likelihood  that  a male 
already  detected  was  singing  in  Repeat  mode. 
The  probability  that  a singing  male  sang  in 
Repeat  instead  of  Serial  mode  differed  for 
paired  (0.51)  versus  unpaired  (0.92)  males. 
Detectability  also  differed  for  paired  (0.49) 


and  unpaired  (0.99)  males.  The  chances  that 
a paired  male  would  sing  any  Repeat  songs 
within  a 5-min  interval  was  only  0.25  (0.51 
X 0.49).  In  contrast,  the  chances  that  an  un- 
paired male  would  sing  in  Repeat  mode  within 
a 5-min  period  was  0.91  (0.92  X 0.99).  Thus, 
unpaired  males  were  3.6  times  (0.91/0.25) 
more  likely  to  sing  in  Repeat  mode  in  a given 
interval  than  were  paired  males. 

Our  results  suggest  that  unpaired  males 
should  be  distinguishable  from  paired  males 
in  field  surveys.  When  conducting  point 
counts,  an  observer  could  listen  to  a singing 
male  for  a prescribed  period  of  time,  note 
whether  he  is  repeating  the  same  song  (Repeat 
mode)  or  alternating  songs  (Serial  mode),  and 
tally  the  number  of  Repeat  songs  he  sings  per 
minute  or  the  number  of  seconds  that  lapse 
between  successive  songs.  In  our  study  pop- 
ulation, a critical  song  rate  of  8.0  Repeat 
songs  per  min  for  5 min  (>40  songs  total) 
would  identify  the  male  as  “unpaired”  with 
reasonable  certainty.  If  a male  sang  in  Serial 
mode  during  the  same  5-min  period,  we  could 
be  reasonably  certain  that  he  was  “paired.” 

The  presence  of  unpaired  males  can  con- 
found estimates  of  the  numbers  of  breeding 
birds.  Unpaired  males  are  common  in  Amer- 
ican Redstart  populations,  with  yearlings 
forming  the  bulk  of  males  that  are  unsuccess- 
ful in  obtaining  mates  (Sherry  and  Holmes 
1997).  Our  data  show  that  unpaired  males  are 
about  twice  as  likely  as  paired  males  to  be 
detected  during  brief  listening  intervals  (e.g., 
5 min).  Similar  results  have  been  reported  for 
several  other  species  (Best  1981,  Mayfield 
1981,  Gibbs  and  Wenny  1993). 

The  utility  of  such  a protocol  for  detection 
of  trends  over  time  (or  space)  is  demonstrated 
in  the  following  hypothetical  case.  Assume 
that  100  males  are  within  earshot,  5-min 
counts  are  conducted,  and  the  listener  always 
detects  and  correctly  identifies  a given  song. 
If,  in  year  1 (or  habitat  A),  all  males  are 
paired,  only  49  males  would  be  reported  (us- 
ing our  calculated  detection  probability  = 
0.49).  If  only  half  of  the  100  total  males  are 
paired  in  year  2 (or  habitat  B),  then  only  —25 
(50  X 0.49)  of  the  paired  males  would  be  de- 
tected while  nearly  all  of  the  unpaired  males 
(50)  would  be  detected  (using  our  calculated 
detection  probability  = 0.99),  for  a total  of 
~75  males  reported.  Based  on  the  data,  we 


Staicer  et  al.  • SINGING  BEHAVIOR  AND  BREEDING  STATUS  IN  REDSTARTS 


449 


would  erroneously  conclude  that  the  popula- 
tion increased  from  year  1 to  year  2 (or  that 
the  population  in  habitat  B was  larger  than 
that  in  habitat  A). 

Correcting  the  data  by  removing  unpaired 
males  from  the  total  detected  and  taking  into 
account  the  lower  detectability  of  paired 
males  provides  a very  different  picture  of  pop- 
ulation status.  Assume  we  use  the  protocol 
whereby,  for  a given  male,  detecting  >40 
songs  per  5-min  sample  indicates  that  he  is 
unpaired,  and  10%  of  males  are  misclassified 
(based  on  the  type  of  overlap  illustrated  in 
Fig.  4C).  In  year  1 (or  in  habitat  A),  we  would 
correctly  classify  44  (and  misclassify  5)  of  the 
49  paired  males  that  were  detected,  and  then 
double  this  number  for  a total  estimate  of  88 
breeding  pairs.  In  year  2 (or  in  habitat  B),  22 
of  the  25  paired  males  detected  would  be  cor- 
rectly classified  as  paired  and  5 of  the  detected 
unpaired  males  would  be  misclassified  as 
paired,  for  a total  of  27  paired  males  (22  + 
5)  detected.  Correcting  for  the  0.49  detection 
rate  of  paired  males  yields  a total  estimate  of 
—54  pairs  in  year  2 (or  in  habitat  B).  Both 
corrected  estimates  fall  within  10%  of  the  ac- 
tual number  of  breeding  pairs.  The  large  pop- 
ulation decline  from  year  1 to  year  2 becomes 
visible  (or  the  lower  population  density  in 
habitat  B becomes  obvious).  Thus,  the  infor- 
mation about  the  relationship  between  pairing 
status  and  song  rates  in  this  species,  and  per- 
haps others,  can  potentially  be  used  to  obtain 
more  accurate  population  estimates. 

ACKNOWLEDGMENTS 

This  study  was  supported  by  a Carnes  Award  and 
an  American  Ornithologists’  Union  Research  Grant  to 
CAS  and  VI,  and  National  Science  Foundation  Grants 
to  Dartmouth  College  (R.  T.  Holmes)  and  Tulane  Uni- 
versity (TWS).  Marist  College  provided  three  Summer 
Research  Grants  and  a sabbatical  leave  (VI).  Record- 
ing and  analysis  equipment  and  software  were  provid- 
ed by  Marist  College  (including  equipment  loans  from 
the  School  of  Communication  and  the  Arts),  and  by 
Dalhousie  University  (research  development  grant  to 
CAS).  We  thank  the  U.S.  Department  of  Agriculture 
Forest  Service  for  allowing  us  to  conduct  the  study  at 
the  Hubbard  Brook  Experimental  Forest.  We  also 
thank  the  many  field  assistants  (including  J.  Clark,  J. 
Crews,  M.  L.  Deinlein,  R.  Heins,  J.  I.  Lovette,  P.  P. 
Marra,  and  T.  S.  Wilkinson)  for  monitoring  nests,  G. 
Drake  for  technical  assistance,  and  R.  T.  Holmes  for 
logistical  support  and  encouragement  at  Hubbard 
Brook.  We  wish  to  extend  special  thanks  to  A.  Molloy, 


who  encouraged  this  research  and  helped  us  obtain  the 
equipment  necessary  to  process  and  analyze  sono- 
grams. We  also  thank  J.  P.  Saunders  and  M.  G.  Tan- 
nenbaum  for  proofreading  the  manuscript.  D.  Nelson, 
J.  C.  Nordby,  D.  A.  Spector,  and  four  anonymous  re- 
viewers provided  useful  comments  on  previous  drafts. 

LITERATURE  CITED 

Alatalo,  R.  V.,  C.  Glynn,  and  A.  Lundberg.  1990. 
Singing  rate  and  female  attraction  in  the  Pied  Fly- 
catcher: an  experiment.  Animal  Behaviour  39: 
601-603. 

Amrhein,  V.,  H.  P.  Kunc,  and  M.  Naguib.  2004.  Sea- 
sonal patterns  of  singing  activity  vary  with  time 
of  day  in  the  Nightingale  ( Luscinia  megarhyn- 
chos).  Auk  121:110-117. 

Arvidsson,  B.  L.  and  R.  Neergaard.  1991.  Mate 
choice  in  Willow  Warbler:  a field  experiment.  Be- 
havioral Ecology  and  Sociobiology  29:225-229. 
Best,  L.  B.  1981.  Seasonal  changes  in  detection  of 
individual  bird  species.  Studies  in  Avian  Biology 
6:252-261. 

Breitwisch,  R.  1989.  Mortality  patterns,  sex  ratio,  and 
parental  investment  in  monogamous  birds.  Cur- 
rent Ornithology  6:1-50. 

Cuthill,  I.  and  A.  Hindmarsh.  1985.  Increase  in  star- 
ling song  activity  with  removal  of  mate.  Animal 
Behaviour  33:326-328. 

Faaborg,  J.  2002.  Saving  migrant  birds:  developing 
strategies  for  the  future.  University  of  Texas  Press, 
Austin. 

Faaborg,  J.,  M.  Brittingham,  T.  Donovan,  and  J. 
Blake.  1995.  Habitat  fragmentation  in  the  tem- 
perate zone.  Pages  357-380  in  Ecology  and  man- 
agement of  Neotropical  migratory  birds:  a synthe- 
sis and  review  of  critical  issues  (T.  E.  Martin  and 
D.  M.  Finch,  Eds.).  Oxford  University  Press,  New 
York. 

Ficken,  M.  S.  and  R.  W.  Ficken.  1965.  Comparative 
ethology  of  the  Chestnut-sided  Warbler,  Yellow 
Warbler,  and  American  Redstart.  Wilson  Bulletin 
77:363-375. 

Gentner,  T.  Q.  and  S.  H.  Hulse.  2000.  Female  Eu- 
ropean Starling  preference  and  choice  for  varia- 
tion in  conspecific  male  song.  Animal  Behaviour 
59:443-458. 

Gibbs,  J.  P.  and  J.  Faaborg.  1990.  Estimating  the  vi- 
ability of  Ovenbird  and  Kentucky  Warbler  popu- 
lations occupying  forest  fragments  in  central  Mis- 
souri, USA.  Conservation  Biology  4:192-196. 
Gibbs,  J.  P.  and  D.  G.  Wenny.  1993.  Song  output  as 
a population  estimator:  effect  of  male  pairing  sta- 
tus. Journal  of  Field  Ornithology  64:316-322. 
Gil,  D.,  J.  A.  Graves,  and  P.  J.  B.  Slater.  1999.  Sea- 
sonal patterns  of  singing  in  the  Willow  Warbler: 
evidence  against  the  fertility  announcement  hy- 
pothesis. Animal  Behaviour  58:995-1000. 
Gottlander,  K.  1987.  Variation  in  the  song  rate  of  the 
male  Pied  Flycatcher  ( Ficedula  hypoleuca ):  caus- 
es and  consequences.  Animal  Behaviour  35:1037- 
1043. 


450 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Hanski,  I.  K.  and  A.  Laurila.  1993.  Variation  in  song 
rate  during  the  breeding  cycle  of  the  Chaffinch, 
Fringilla  coelebs.  Ethology  93:161-169. 

Hayes,  J.  P.,  J.  R.  Probst,  and  D.  Rakstad.  1986. 
Effect  of  mating  status  and  time  of  day  on  Kirt- 
land’s  Warbler  song  rates.  Condor  88:386-388. 

Hofstad,  E.,  Y.  Espmark,  A.  Moksnes,  T.  Haugan, 
and  M.  Ingebrigtsen.  2002.  The  relationship  be- 
tween song  performance  and  male  quality  in 
Snow  Buntings  ( Plectrophenax  nivalis).  Canadian 
Journal  of  Zoology  80:524-531. 

Hoi-Leitner,  M..  H.  Nechtelberger,  and  H.  Hoi. 
1995.  Song  rate  as  a signal  for  nest-site  quality  in 
Blackcaps  ( Sylvia  atricapilla).  Behavioral  Ecolo- 
gy and  Sociobiology  37:399-405. 

Holmes,  R.  T.  and  F.  W.  Sturges.  1975.  Bird  com- 
munity dynamics  and  energetics  in  a northern 
hardwoods  ecosystem.  Journal  of  Animal  Ecology 
44:175-200. 

Johnson,  L.  S.  1983.  Effect  of  mate  loss  on  song  per- 
formance in  the  Plain  Titmouse.  Condor  85:378- 
380. 

Krebs,  J.  R.,  M.  Avery,  and  R.  J.  Cowie.  1981.  Effect 
of  removal  of  mate  on  the  singing  behaviour  of 
Great  Tits.  Animal  Behaviour  29:635-637. 

Kroodsma,  D.  E.  1981.  Geographic  variation  and 
functions  of  song  types  in  warblers  (Parulidae). 
Auk  98:743-751. 

Kroodsma,  D.  E.,  R.  C.  Bereson,  B.  E.  Byers,  and 
E.  Minear.  1989.  Use  of  song  types  by  the  Chest- 
nut-sided  Warbler:  evidence  for  both  intrasexual 
and  intersexual  functions.  Canadian  Journal  of  Zo- 
ology 67:447-456. 

Lemon,  R.  E.,  R.  Cotter,  R.  C.  MacNally,  and  S. 
Monette.  1985.  Song  repertoires  and  song  shar- 
ing by  American  Redstarts.  Condor  87:457-470. 

Lemon,  R.  E.,  S.  Monette,  and  D.  Roff.  1987.  Song 
repertoires  of  American  warblers  (Parulinae):  hon- 
est advertising  or  assessment?  Ethology  74:265- 
284. 

Lozano,  G.  A.,  S.  Perreault,  and  R.  E.  Lemon.  1996. 
Age,  arrival  date  and  reproductive  success  of  male 
American  Redstarts  Setophaga  ruticilla.  Journal 
of  Avian  Biology  27:164-170. 

Marra,  P.  P.  and  R.  T.  Holmes.  1997.  Avian  removal 
experiments:  do  they  test  for  habitat  saturation  or 
female  availability?  Ecology  78:947-952. 

Mayfield,  H.  F.  1981.  Problems  in  estimating  popu- 
lation size  through  counts  of  singing  males.  Stud- 
ies in  Avian  Biology  6:220-224. 

McShea,  W.  J.  and  J.  H.  Rappole.  1997.  Variable  song 
rates  in  three  species  of  passerines  and  implica- 
tions for  estimating  bird  populations.  Journal  of 
Field  Ornithology  68:367-375. 

Morris,  M.  M.  J.  and  R.  E.  Lemon.  1988.  Mate  choice 
in  American  Redstarts:  by  territory  or  quality?  Ca- 
nadian Journal  of  Zoology  66:2255-2261. 

Morse,  D.  H.  1970.  Territorial  and  courtship  songs  of 
birds.  Nature  226:659-661. 

Nemeth,  E.  1996.  Different  singing  styles  in  mated  and 


unmated  Reed  Buntings  Emberiza  schoeniclus. 
Ibis  138:172-176. 

Nolan,  V.,  Jr.  1978.  The  ecology  and  behavior  of  the 
Prairie  Warbler  Dendroica  discolor.  Ornithologi- 
cal Monographs  no.  26. 

Nolan,  P.  M.  and  G.  E.  Hill.  2004.  Female  choice  for 
song  characteristics  in  the  House  Finch.  Animal 
Behaviour  67:403-410. 

Nystrom,  K.  K.  G.  1997.  Food  density,  song  rate,  and 
body  condition  in  territory-establishing  Willow 
Warblers  ( Phylloscopus  trochilus).  Canadian  Jour- 
nal of  Zoology  75:47-58. 

Otter,  K.,  B.  Chruszcz,  and  L.  Ratcliffe.  1997. 
Honest  advertisement  and  song  output  during  the 
dawn  chorus  of  Black-capped  Chickadees.  Behav- 
ioral Ecology  8:167-173. 

Perreault,  S.,  R.  E.  Lemon,  and  U.  Kuhnlein.  1997. 
Pattern  and  correlates  of  extrapair  paternity  in 
American  Redstarts  ( Setophaga  ruticilla).  Behav- 
ioral Ecology  8:612-621. 

Radesater,  T.  and  S.  Jakobsson.  1989.  Song  rate  cor- 
relations of  replacement  Willow  Warblers  Phyl- 
loscopus trochillus.  Ornis  Scandinavia  20:71-73. 

Ralph,  C.  J.,  S.  Droege,  and  J.  R.  Sauer.  1995.  Man- 
aging and  monitoring  birds  using  point  counts: 
standards  and  applications.  Pages  161-168  in 
Monitoring  bird  populations  by  point  counts  (C. 
J.  Ralph,  J.  R.  Sauer,  and  S.  Droege,  Eds.).  Gen- 
eral Technical  Report  PSW-GTR-144,  Pacific 
Southwest  Research  Station,  USDA  Forest  Ser- 
vice, Albany,  California. 

Rappole,  J.  H.  1995.  The  ecology  of  migrant  birds:  a 
Neotropical  perspective.  Smithsonian  Institution 
Press,  Washington,  D.C. 

Ratti,  O.  and  P.  Siikamaki.  1993.  Female  attraction 
behaviour  of  radio  tagged  polyterritorial  Pied  Fly- 
catcher males.  Behaviour  127:279-288. 

Reynard,  G.  B.  1963.  The  cadence  of  bird  song.  Liv- 
ing Bird  2:139-146. 

Saino,  N.,  P.  Galeottl  R.  Sacchi,  and  A.  P.  Mpller. 
1997.  Song  and  immunological  condition  in  male 
Barn  Swallows  ( Hirundo  rustica).  Behavioral 
Ecology  8:364-371. 

Searcy,  W.  A.,  D.  Eriksson,  and  A.  Lundberg.  1991. 
Deceptive  behavior  in  Pied  Flycatchers.  Behav- 
ioral Ecology  and  Sociobiology  29:167-175. 

Secunda,  R.  C.  and  T.  W.  Sherry.  1991.  Polyterrito- 
rial polygyny  in  the  American  Redstart.  Wilson 
Bulletin  103:190-203. 

Sherry,  T.  W.  and  R.  T.  Holmes.  1989.  Age-specific 
social  dominance  affects  habitat  use  by  breeding 
American  Redstarts  ( Setophaga  ruticilla):  a re- 
moval experiment.  Behavioral  Ecology  and  Socio- 
biology 25:327-333. 

Sherry,  T.  W.  and  R.  T.  Holmes.  1997.  American 
Redstart  ( Setophaga  ruticilla).  The  Birds  of  North 
America,  no.  277. 

Smith,  R.  J.  and  F.  R.  Moore.  2003.  Arrival  fat  and 
reproductive  performance  in  a long-distance  pas- 
serine migrant.  Oecologia  134:325-331. 


Staicer  et  al.  • SINGING  BEHAVIOR  AND  BREEDING  STATUS  IN  REDSTARTS 


451 


Sokal,  R.  R.  and  E J.  Rohlf.  1995.  Biometry,  3rd  ed. 
Freeman,  New  York. 

Spector,  D.  A.  1991.  The  singing  behaviour  of  Yellow 
Warblers.  Behaviour  117:29-52. 

Spector,  D.  A.  1992.  Wood-warbler  song  systems:  a 
review  of  paruline  singing  behaviors.  Current  Or- 
nithology 9:199-238. 

Staicer,  C.  A.  1989.  Characteristics,  use,  and  signifi- 
cance of  two  singing  behaviors  in  Grace’s  Warbler 
{Dendroica  graciae).  Auk  106:49—63. 

Staicer,  C.  A.  1996a.  Acoustical  features  of  song  cat- 
egories of  the  Adelaide’s  Warbler  ( Dendroica  ade- 
laidae ).  Auk  113:771-783. 

Staicer,  C.  A.  1996b.  Honest  advertisement  of  pairing 
status:  evidence  from  a tropical  resident  wood- 
warbler.  Animal  Behaviour  51:375-390. 

Staicer,  C.  A.,  D.  A.  Spector,  and  A.  G.  Horn.  1996. 
The  dawn  chorus  and  other  diel  patterns  of  acous- 


tic signaling.  Pages  426-453  in  Ecology  and  evo- 
lution of  acoustic  communication  in  birds  (D.  E. 
Kroodsma  and  E.  H.  Miller,  Eds.).  Cornell  Uni- 
versity Press,  New  York. 

Titus,  R.  C.,  C.  R.  Chandler,  E.  D.  Ketterson,  and 
V.  Nolan,  Jr.  1997.  Song  rates  of  Dark-Eyed  Jun- 
cos  do  not  increase  when  females  are  fertile.  Be- 
havioral Ecology  and  Sociobiology  41:165-169. 

Wasserman,  F.  R.  1977.  Mate  attraction  function  of 
song  in  the  White-throated  Sparrow.  Animal  Be- 
haviour 79: 125-127. 

Weary,  D.  M.,  R.  E.  Lemon,  and  S.  Perreault.  1994. 
Different  responses  to  different  song  types  in 
American  Redstarts  ( Setophaga  ruticilla).  Auk 
111:730-734. 

Westcott,  D.  1992.  Inter-  and  intra-sexual  selection: 
the  role  of  song  in  a lek  mating  system.  Animal 
Behaviour  44:695-703. 


The  Wilson  Journal  of  Ornithology  1 18(4):452 — 460,  2006 


INVESTMENT  IN  NEST  DEFENSE  BY  NORTHERN  FLICKERS: 
EFFECTS  OF  AGE  AND  SEX 

RYAN  J.  FISHER123  AND  KAREN  L.  WIEBE1 2 3 


ABSTRACT. — At  early  breeding  stages,  male  woodpeckers  invest  heavily  in  nest  construction  and  defense, 
but  parental  contributions  to  brood  defense  among  Picidae  are  not  well  known.  We  studied  the  Northern  Flicker 
( Colaptes  auratus)  to  determine  whether  sex,  age,  brood  size,  body  size,  or  body  condition  influenced  defense 
behavior.  When  presented  with  a model  predator  (red  squirrel,  Tamiasciurus  hudsonicus ) during  the  brood- 
rearing period,  parents  exhibited  a range  of  behaviors,  such  as  blocking  the  nest  hole,  diving  at  the  model,  and 
striking  the  model;  however,  defense  scores  did  not  differ  between  males  and  females  aged  1,  2,  or  3+  years 
old.  Although  we  predicted  that  defense  level  would  be  positively  correlated  with  brood  size,  we  found  no  such 
relationship.  Adult  body  size  and  condition  also  were  not  related  to  defense  intensity.  We  conclude  that  the  sexes 
may  exhibit  similar  levels  of  defense  because  they  have  similar  apparent  annual  survival  rates  and  males  are 
only  slightly  larger  than  females.  If  flickers  optimize  clutch  size  according  to  the  number  of  offspring  they  can 
rear,  then  there  may  be  no  relationship  between  defense  and  brood  size.  Received  20  September  2005,  accepted 
6 July  2006. 


Although  nest  defense  may  deter  predators, 
it  may  place  the  parent  bird  at  considerable 
risk  while  requiring  significant  energy  expen- 
diture (Blancher  and  Roberstson  1982,  Nealen 
and  Breitwisch  1997,  Olendorf  and  Robinson 
2000).  For  many  birds,  the  intensity  of  nest 
defense  may  increase  (1)  as  the  breeding  sea- 
son and  reproductive  value  of  the  brood  in- 
creases (see  Montgomerie  and  Weatherhead 
1988  for  a review),  (2)  as  the  potential  for 
renesting  declines  (Andersson  et  al.  1980), 
and  (3)  with  clutch  or  brood  size  (Olendorf 
and  Robinson  2000).  Moreover,  the  intensity 
of  defense  may  depend  on  the  sex  of  the  par- 
ent defending  the  nest  (Breitwisch  1988, 
Sproat  and  Ritchison  1993,  Nealen  and  Breit- 
wisch 1997). 

Age  may  be  correlated  with  the  level  of 
nest  defense  for  several  reasons,  but  this  has 
rarely  been  tested  (Veen  et  al.  2000).  Older 
birds  have  a lower  probability  of  future  repro- 
duction; thus,  they  should  invest  more  in 
broods  than  younger  individuals  (Hatch 
1997).  In  addition,  it  is  often  difficult  to  sep- 
arate the  effects  of  age  from  experience  with 
predators  because  they  are  often  directly  cor- 
related. Similar  to  older  birds,  birds  with  more 


1 Dept,  of  Biology,  Univ.  of  Saskatchewan,  1 12  Sci- 
ence Place,  Saskatoon,  SK  S7N  5E2,  Canada. 

2 Current  address:  Dept,  of  Biology,  Univ.  of  Regi- 
na, 3737  Wascana  Pkwy..  Regina.  SK  S4S  0A2,  Can- 
ada. 

3 Corresponding  author;  e-mail: 
fisherry@uregina.ca 


experience  also  may  be  willing  to  defend  their 
nests  more  aggressively  (Veen  et  al.  2000). 

Levels  of  defense  also  may  vary  between 
the  sexes  (e.g.,  Breitwisch  1988,  Sproat  and 
Ritchison  1993,  Tryjanowski  and  Golawski 
2004)  because  of  intersexual  differences  in  fu- 
ture survival  and  body  size  (Montgomerie  and 
Weatherhead  1988).  The  sex  with  the  lower 
survival  rate  and,  consequently,  the  lower 
probability  of  future  breeding,  should  defend 
broods  more  vigorously  than  its  partner 
(Montgomerie  and  Weatherhead  1988).  Mor- 
tality is  usually  female  biased  in  many  bird 
species,  likely  as  a result  of  high  reproductive 
costs  (Promislow  et  al.  1992).  Generally,  the 
larger  sex  defends  the  nest  more  aggressively, 
perhaps  because  the  risk  of  injury  is  lower  or 
because  larger  birds  are  able  to  mount  strong 
attacks  (Tryjanowski  and  Golawski  2004).  Be- 
cause healthy  birds  may  have  relatively  great- 
er energy  reserves,  they  may  take  more  risks 
when  defending  their  nests  than  birds  in  poor- 
er condition  (Martin  and  Horn  1993).  For  ex- 
ample, females  may  be  in  poorer  condition  af- 
ter incubation  and  defend  the  nest  less  ag- 
gressively than  the  male  (Sproat  and  Ritchison 
1993). 

Cavity  nesters  may  rely  more  on  the  inac- 
cessible or  cryptic  nature  of  their  nest  than  on 
active  nest  defense  (Weidinger  2002);  how- 
ever, there  have  been  few  studies  of  wood- 
pecker behavioral  responses  to  predators  at 
the  nest  site.  Wiebe  (2004)  examined  respons- 
es of  the  Northern  Flicker  ( Colaptes  auratus) 


452 


Fisher  and  Wiebe  • NORTHERN  FLICKER  NEST  DEFENSE 


453 


to  the  European  Starling  ( Sturnus  vulgaris ) — 
a kleptoparasite  of  cavity  nests  (Kappes 
1997) — but  found  no  sex-  or  age-related  dif- 
ferences in  cavity  defense.  Ingold  (1994)  also 
described  aggressive  interactions  between 
starlings  and  flickers,  but  did  not  examine  sex 
or  age  differences  in  these  behaviors.  Law- 
rence (1967)  described  woodpeckers  defend- 
ing their  nests  from  inside  their  cavities,  en- 
gaging in  alarm  vocalizations  and  diving  at- 
tacks; she  also  reported  a male  Northern 
Flicker  that  delivered  a blow  with  its  beak  to 
a squirrel  entering  a nest  hole,  effectively  de- 
terring the  squirrel  from  entering. 

In  this  study,  we  presented  a model  predator 
(red  squirrel,  Tamiasciurus  hudsonicus)  at 
nest  sites  of  Northern  Flickers  to  examine 
adult  nest-defense  behavior  in  relation  to  age, 
sex,  brood  size,  body  size,  and  body  condi- 
tion. Because  flickers  are  relatively  short-lived 
and  their  probability  of  survival  is  indepen- 
dent of  age  (Fisher  and  Wiebe  2006a),  we  pre- 
dicted that  there  would  be  no  differences  in 
defense  between  young  and  older  birds.  Sim- 
ilarly, mark-recapture  models  suggest  only  a 
2%  difference  in  annual  survival  rate  between 
the  sexes  (Fisher  and  Wiebe  2006a),  and  the 
sexes  invest  about  equally  in  nestling  provi- 
sioning (Moore  1995,  Wiebe  and  Elchuk 
2003).  Thus,  we  predicted  that  male  and  fe- 
male flickers  would  defend  their  broods  with 
similar  intensity.  We  also  predicted  that  indi- 
viduals in  better  condition  and  with  larger 
broods  would  defend  their  nests  more  aggres- 
sively. 

METHODS 

Study  site  and  study  species. — Our  study 
site  was  near  Riske  Creek,  British  Columbia 
(51°  52'  N,  122°  21'  W),  and  encompassed  ap- 
proximately 100  km2;  90-120  pairs  of  flickers 
nest  there  each  year  (Fisher  and  Wiebe 
2006a).  Habitats  on  the  site  are  patchy  and 
variable.  Flickers  prefer  grasslands  for  forag- 
ing (Elchuck  and  Wiebe  2003)  and  patches  of 
quaking  aspen  ( Populus  tremuloides ) and 
lodgepole  pine  ( Pinus  contorta ) for  nesting 
(Martin  and  Eadie  1999).  Continuous  forests 
of  Douglas-fir  ( Pseudotsuga  menziesii ) and 
hybrid  spruce  ( Picea  engelmannii  X P.  glau- 
ca ) also  occur. 

Flickers  migrate  to  the  area  in  mid-April 
and  begin  egg-laying  in  early-  to  mid-May 


(mean  clutch-initiation  date  = 13  May,  range 
= 26  April-2  July;  Moore  1995,  KLW  un- 
publ.  data).  Each  year  since  1998,  the  area  has 
been  surveyed  in  spring  (22  April- 15  May, 
1998-2005)  for  finding  newly  excavated  cav- 
ities and  to  check  old  cavities  for  new  breed- 
ing pairs  (flickers  tend  to  reuse  old  cavities 
more  often  than  other  woodpeckers;  Moore 
1995,  Aitken  et  al.  2002,  Wiebe  et  al.  2006). 
Tape-recorded  territorial  playback  calls  also 
were  used  to  locate  flicker  territories  and  nest 
sites.  Average  clutch  size  in  this  area  is  eight 
eggs  and  mean  number  of  young  fledged  per 
successful  nest  is  six  (Wiebe  2003).  Once  a 
clutch  was  complete,  we  cut  a small  door  into 
the  side  of  the  nest  tree  for  access  to  adults, 
eggs,  and  nestlings  (see  Wiebe  2001).  Flickers 
seem  to  tolerate  the  doors  and  readily  re-use 
such  cavities  (Fisher  and  Wiebe  2006a).  Ap- 
proximately 18%  of  monitored  nests  are  dep- 
redated annually  by  mammalian  predators, 
mainly  red  squirrels  (Fisher  and  Wiebe 
2006b). 

We  captured  flickers  by  flushing  individuals 
from  the  nest  cavity  into  a small  net  placed 
over  the  cavity  entrance  (Fisher  and  Wiebe 
2006b).  Three  colored  plastic  and  one  alumi- 
num band  were  attached  to  each  individual  to 
aid  in  individual  identification  (>95%  of  the 
known  annual  breeding  population  is  color 
banded  and  individually  identifiable).  During 
banding,  we  used  molt  criteria  to  determine 
the  birds’  ages  (up  to  4 years  old;  Pyle  et  al. 
1997).  We  developed  an  index  of  flicker  body 
size  (i.e.,  score  on  the  first  axis  of  a principle 
components  analysis  based  on  six  measures: 
bill  depth,  and  lengths  of  the  wing,  bill,  tail, 
tarsus,  and  ninth  primary)  and  body  condition 
(i.e.,  residuals  of  a regression  of  body  mass 
on  body  size);  because  of  sexual  size  dimor- 
phism, we  made  separate  calculations  for 
males  and  females  (see  Wiebe  and  Swift 
2001).  A year-specific  estimate  of  body  con- 
dition was  made  only  for  individuals  that  were 
trapped  and  weighed  in  2003  and  2004;  thus, 
only  individuals  captured  during  2003  or  2004 
were  included  in  analyses  with  body  condition 
as  a covariate  (see  below).  We  assumed  that 
body  size  (i.e.,  the  structural  size  of  an  indi- 
vidual and  not  body  mass)  did  not  change 
from  year  to  year. 

Model  presentations. — Birds  with  altricial 
young  generally  defend  their  nests  most 


454 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  4,  December  2006 


strongly  during  the  nestling  stage  and  as  nest- 
lings age  (Montgomerie  and  Weatherhead 
1988).  We  measured  nest  defense  when  nest- 
lings were  10-15  days  old  to  control  for  ef- 
fects of  nest  stage  and  nestling  age  on  defense 
behavior.  At  each  nest,  we  tested  nest  defense 
once  with  a predator  (taxidermic  model  of  a 
red  squirrel)  and  once  with  a control  (taxider- 
mic model  of  a Yellow-headed  Blackbird, 
Xanthocephalus  xanthocephalus,  or  a Cedar 
Waxwing,  Bombycilla  cedrorum).  The  same 
individuals  were  tested  only  once  with  each 
model  during  the  2-year  study  to  avoid  poten- 
tial habituation  of  parents  to  the  models 
(Knight  and  Temple  1986a,  1986c).  Blackbird 
and  waxwing  models  were  used  as  controls 
because  they  are  both  common  in  the  study 
area  and  neither  poses  a threat  to  flicker 
broods  (Wiebe  2004).  In  2004,  during  60%  of 
control  trials  we  used  the  waxwing  because 
the  blackbird  model  was  irreparably  damaged 
from  transportation  to  and  from  trials. 

Predator  and  control  trials  were  conducted 
randomly  at  a given  nest,  with  1-5  days  be- 
tween trials  (i.e.,  one  trial  = one  model  pre- 
sentation). Because  the  perceived  threat  from 
a predator  could  vary  with  distance  between 
the  predator  and  the  nest  (Ratti  2000),  we  fas- 
tened the  models  at  a fixed  distance  (1  m be- 
low the  cavity  entrance)  with  a bungee  cord 
tied  to  the  tree  trunk.  The  model  squirrel  was 
attached  to  a small,  flat  board  base  that  was 
then  attached  to  the  tree  trunk.  Control  models 
were  mounted  in  an  upright,  perched  position 
on  a natural  branch,  which  was  then  attached 
to  the  tree  trunk.  During  a given  trial,  terri- 
torial “chatter”  calls  of  squirrels  or  songs  of 
Yellow-headed  Blackbirds  or  Cedar  Wax- 
wings  were  played  at  the  base  of  the  nest  tree 
to  increase  model  detectability  (Ghalambor 
and  Martin  2002).  After  models  were  placed 
at  the  nest,  we  retreated  to  a concealed  posi- 
tion >15  m away  to  record  responses  of  the 
returning  parents. 

The  first  variable  we  recorded  was  response 
time  of  the  adult  (i.e.,  sec  between  when  we 
had  set  up  the  model  and  were  hidden,  to 
when  the  parent  returned  and  we  judged  it  was 
within  10  m of  the  nest  and  in  sight  of  the 
model).  Ten  meters  from  the  nest  was  usually 
the  maximum  distance  from  which  we  could 
observe  a bird  responding,  because  of  dense 
foliage  around  some  nests.  We  were  confident 


that  the  flicker  was  responding  to  the  model 
at  distances  ^10  m from  the  nest  once  we 
judged  that  it  could  see  the  model.  If  parents 
did  not  return  to  within  10  m and  in  sight  of 
the  model  in  1 hr,  then  these  trials  were  re- 
moved from  all  analyses.  After  an  adult(s)  re- 
turned within  <10  m,  we  recorded  its  behav- 
ior for  5 min  (if  both  parents  returned  simul- 
taneously, we  treated  them  as  individual  re- 
sponses). Flickers  respond  to  models  with 
slow,  deliberate  movements  (Wiebe  2004),  so 
the  5-min  period  should  have  provided  a rep- 
resentative sample  of  behavior.  We  quantified 
defense  levels  based  on  four  behaviors  re- 
corded during  the  5-min  period:  (1)  number 
of  alarm  calls  {peak  and  wicka  calls;  Moore 
1995);  (2)  the  closest  distance  that  the  re- 
sponding parent  approached  the  model  (m;  a 
visual  estimate);  (3)  whether  or  not  the  parent 
dived  at  or  hit  the  model  (dichotomous  vari- 
able); and  (4)  time  (sec)  an  individual  spent 
inside  the  cavity  during  each  trial  (flickers  en- 
tered cavities  and  then  peered  back  out,  usu- 
ally with  their  beaks  protruding  from  the  cav- 
ity entrances).  Time  spent  in  the  cavity  should 
reflect  investment  in  nest  defense  because 
blocking  the  entrance  prevents  predation  of 
the  nest  (Cordero  and  Senar  1990).  Assessing 
the  risk  a parent  incurs  by  blocking  the  cavity 
entrance  is  difficult.  This  defensive  strategy 
may  be  safer  than  others  because  most  of  the 
parent’s  body  is  inside  the  cavity  (Cordero  and 
Senar  1990);  conversely,  there  are  no  avenues 
of  escape  for  the  parent. 

Statistical  analyses. — Response  time  was 
square-root  transformed  to  meet  assumptions 
of  normality,  and  we  analyzed  it  separately 
from  other  defense  variables  because  it  was 
unlikely  to  have  been  influenced  by  model 
type  (parents  presumably  had  not  had  time  to 
see  the  model  before  returning).  We  used  an 
ANCOVA  to  test  whether  age,  sex,  brood 
size,  and/or  body  condition  affected  response 
time  to  the  predator  model  (we  assumed  that 
the  structural  size  of  an  individual  would  not 
influence  response  time).  Because  data  trans- 
formations of  the  other  four  defense  variables 
did  not  result  in  normality,  we  used  non-para- 
metric  tests  for  subsequent  analyses.  Statisti- 
cal significance  was  set  at  P < 0.05. 

With  respect  to  the  four  nest-defense  vari- 
ables, there  was  no  difference  between  control 
model  types  (blackbird  versus  waxwing; 


Fisher  and  Wiehe  • NORTHERN  FLICKER  NEST  DEFENSE 


455 


Mann- Whitney  U and  Fisher  Exact  tests:  all 
P > 0.47).  Similar  tests  also  showed  that  there 
were  no  significant  differences  between  years 
in  terms  of  responses  to  control  and  predator 
models  (all  P > 0.12).  Therefore,  we  pooled 
all  responses  (for  years  and  control  models) 
in  subsequent  analyses. 

We  first  analyzed  each  defense  variable  sin- 
gly to  determine  which  differed  significantly 
between  control  and  squirrel  models,  without 
any  other  effects.  This  allowed  us  to  eliminate 
model  type  as  a variable  if  it  was  non-signif- 
icant, thus  simplifying  subsequent  models  in- 
volving age  class,  sex,  brood  size,  body  size, 
and  body  condition.  We  used  paired  tests 
(Wilcoxon’s  signed-rank  tests)  to  analyze  min- 
imum distance  to  the  model,  time  in  the  cav- 
ity, and  number  of  alarm  calls  to  account  for 
both  predator  and  control  trials  taking  place  at 
the  same  nest.  This  approach  may  have  been 
more  stringent  than  necessary  because  it  was 
not  necessarily  the  same  individual  that  re- 
sponded to  each  trial;  however,  independent 
test  results  were  consistent  with  those  of  the 
paired  tests.  We  used  a Fisher’s  exact  test  to 
compare  the  frequency  of  diving  at  the  squir- 
rel versus  the  control  models.  All  means  pre- 
sented are  ± SD. 

After  separate  analysis  of  each  defense  be- 
havior (see  results),  we  constructed  an  overall 
defense  score  based  on  the  three  variables  that 
differed  significantly  between  control  and 
predator  models.  This  score  was  used  in  sub- 
sequent analyses  involving  the  relationship 
between  various  parental  attributes  and 
strength  of  response  to  the  squirrel  model.  A 
score  of  1 indicated  the  bird  returned  to  the 
nest  and  was  judged  to  be  within  sight  of  the 
model  but  did  not  dive  at  the  model  or  enter 
the  cavity,  and  always  remained  >2  m away 
from  the  model  (there  is  a low  probability  that 
a squirrel  could  contact  the  parent  at  a dis- 
tance of  2 m).  A score  of  2 indicates  that  the 
parent  approached  <1  m from  the  predator 
model  but  otherwise  performed  no  other  nest- 
defense  behaviors.  In  developing  score  2,  we 
assumed  that  a squirrel  might  be  able  to  phys- 
ically contact  a flicker  <1  m away  and  that 
parents  approaching  within  1 m were  placing 
themselves  at  a greater  risk  than  those  in  score 
category  1.  Responses  in  category  2 included 
perching  on  the  cavity  lip  from  the  outside  or 
on  a branch  within  1 m of  the  model.  A score 


TABLE  1.  Sample  sizes  of  Northern  Flickers  re- 
sponding to  a model  predator  (red  squirrel)  or  control 
(Yellow-headed  Blackbird  or  Cedar  Waxwing)  placed 
at  their  nests  during  the  brood-rearing  stage  at  Riske 
Creek,  British  Columbia  (2003  and  2004  data  pooled). 
Totals  include  instances  in  which  both  parents  re- 
sponded to  the  models,  plus  those  in  which  only  one 
parent  responded;  thus,  sample  sizes  are  larger  than  the 
total  number  of  trials  conducted  for  each  model  type. 


Model  type 
(total  no.  trials) 

Sex 

Age 

n 

Control  (91) 

Male 

1 year 

15 

2 years 

17 

3+  years 

25 

Female 

1 year 

19 

2 years 

16 

3+  years 

15 

Predator  (94) 

Male 

1 year 

17 

2 years 

19 

3+  years 

24 

Female 

1 year 

20 

2 years 

14 

3+  years 

13 

of  3 indicates  that  the  parent  entered  the  cavity 
and  blocked  it  from  the  inside.  Finally,  a score 
of  4 indicates  that  birds  dived  at  or  hit  the  mod- 
el, indicating  the  riskiest  and  most  energetically 
expensive  behavior  to  a defending  adult. 

For  statistical  analyses  involving  age,  we 
categorized  males  or  females  as  either  1,  2,  or 
3+  years  old,  such  that  there  was  at  least  a 
sample  size  of  13  in  each  age  category  (Table 
1).  A further  subdivision  of  age  was  not  pos- 
sible to  analyze  statistically,  as  it  would  have 
resulted  in  some  categories  with  a sample  size 
<5.  We  used  a Kruskal- Wallis  test  to  examine 
whether  the  median  defense  scores  of  birds  in 
the  six  different  age-sex  classes  differed.  To 
analyze  the  effect  of  brood  size  on  defense 
score  (a  categorical  variable),  we  used  Spear- 
man’s rank  correlations.  Body  size  and  con- 
dition met  assumptions  of  normality;  there- 
fore, we  could  use  parametric  tests  (two-factor 
ANOVA)  to  assess  the  relationship  between 
defense  score  and  sex  on  body  size  and  con- 
dition (dependent  variables). 

RESULTS 

We  conducted  9 1 control  trials  and  94  pred- 
ator trials  at  94  Northern  Flicker  nests  in  2003 
and  2004.  Control  trials  were  not  conducted 
at  three  nests  because  nestlings  were  >15 
days  old  by  the  time  the  second  model  could 


456 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  2.  Effects  of  sex,  age  class  (1,  2,  and  3 + 
years  old),  brood  size,  and  body  condition  of  flicker 
parents  on  their  response  time  (see  description  in  text) 
to  a model  nest  predator  presented  at  the  nest  during 
the  brood-rearing  stage  at  Riske  Creek,  British  Colum- 
bia, 2003  and  2004.  No  predictor  was  significant  ac- 
cording to  a 2-factor  ANCOVA  (n  = 84  individuals) 
using  Type  III  sums  of  squares. 


Effect 

ss 

df 

F 

P 

Sex 

231.67 

1 

1.18 

0.29 

Age 

181.15 

2 

0.44 

0.65 

Sex  X age 

438.81 

2 

1.06 

0.35 

Brood  size 

16.50 

1 

0.56 

0.46 

Body  condition 

589.02 

1 

2.84 

0.10 

Sex  X brood  size 

211.50 

1 

1.02 

0.32 

Age  X brood  size 

92.43 

2 

0.22 

0.80 

be  presented.  Parents  occasionally  returned  to- 
gether to  defend  the  nest  ( 1 6 out  of  9 1 control 
and  13  out  of  94  predator  trials)  and  responses 
by  these  individuals  were  considered  to  be  in- 
dependent trials  (i.e.,  two  parents  responding 
increased  sample  size  by  two).  Sample  sizes 
of  responding  parents  of  both  age  classes  and 
sexes  varied  according  to  model  type  (Table 
1). 

Response  time  and  defense  behaviors. — The 
mean  overall  response  time  to  the  predator 
model  was  1,090  ± 876  sec  (n  = 107).  There 
was  a weak  trend  (P  = 0.10)  that  birds  in 
better  condition  responded  to  the  predator 
model  more  quickly,  but  there  was  no  effect 
of  age,  sex,  brood  size,  or  body  condition,  and 
there  were  no  interactions  (Table  2). 

Flickers  dived  significantly  more  at  the 
predator  model  (26%  of  trials)  than  at  the  con- 
trol (2%  of  trials;  Fisher’s  exact  test:  P < 
0.001).  Parents  also  approached  the  predator 
model  more  closely  (3  m ± 4)  than  the  control 
model  (5  m ± 4;  Wilcoxon’s  signed-rank  test: 
Z = —4.98,  P < 0.001).  During  the  5-min 
trials,  flickers  spent  significantly  more  time  in 
their  cavities  when  responding  to  the  predator 
model  than  to  the  control  model  (16%  ± 33 
versus  5%  ± 20,  respectively;  Wilcoxon’s 
signed-rank  test:  Z = —2.35,  P < 0.001).  Par- 
ents gave  wicka  and  peah  alarm  calls  in  36% 
of  the  trials,  but  there  was  no  effect  of  model 
type  on  the  number  of  alarm  calls  (mean  num- 
ber of  alarm  calls  = 11  ±32  and  18  ± 37  in 
response  to  predator  and  control  models,  re- 


Male  Female 

Sex  and  age  class 

FIG.  1.  Nest-defense  scores  of  parent  flickers  did 
not  differ  by  sex  and  age  categories  when  responding 
to  a model  predator  (red  squirrel)  placed  at  their  nest 
during  the  brood-rearing  stage  in  Riske  Creek,  British 
Columbia,  2003  and  2004.  Bold  horizontal  lines  rep- 
resent median  defense  scores,  boxes  represent  25th  and 
75th  percentiles,  and  error  bars  represent  10th  and  90th 
percentiles.  Because  several  birds  within  each  age  and 
sex  category  received  the  same  defense  score,  some 
10th,  25th,  75th,  and  90th  percentiles  overlap;  thus, 
symbols  for  each  age  and  sex  class  are  not  necessarily 
apparent. 


spectively;  Wilcoxon’s  signed-rank  test:  Z = 
-1.41,  P = 0.16). 

Traits  of  the  parent  and  brood. — The  me- 
dian defense  score  for  males  ^3  years  of  age 
was  marginally  higher  that  than  of  any  other 
age-sex  category  (Kruskal- Wallis  test:  x2  = 
6.63,  df  = 3,  P = 0.085;  Fig.  1).  Brood  sizes 
of  parents  tested  with  the  squirrel  model 
ranged  from  2 to  9,  but  there  were  no  signif- 
icant correlations  between  brood  size  and 
nest-defense  score  for  the  six  age-sex  classes 
when  considered  separately  (Spearman’s  rank 
correlations:  all  P > 0.28,  but  two-year  old 
males  showed  a marginally  significant  trend 
of  defending  smaller  broods  more  aggressive- 
ly, r = —0.45,  P =0.060).  Similarly,  with  all 
ages  and  sexes  combined,  there  was  no  effect 
of  brood  size  on  defense  score  (Spearman’s 
rank  correlation:  r = 0.02,  P = 0.83).  In  an- 
other analysis,  we  categorized  brood  sizes  as 
small  (<6  chicks,  n = 45)  versus  large  (>7 
chicks,  n = 62).  Approximately  30%  of  in- 
dividuals with  large  broods  exhibited  the  most 
intense  defensive  behavior  (score  = 4), 
whereas  22%  of  individuals  with  small  broods 
had  score  4;  however,  the  overall  frequency  of 


Fisher  and  Wiebe  • NORTHERN  FLICKER  NEST  DEFENSE 


457 


90- 


A 


88- 


f 


l 


i 


dition  and  defense  score  (two-factor  ANOVA: 
F = 1.48,  df  =3,  P = 0.84)  for  either  sex  (F 
= 2.13,  df  = 1,  P = 0.15;  Fig.  2)  or  a sex  X 
defense  score  interaction  (F  = 1.48,  df  =3,  P 
= 0.23;  Fig.  2). 


0 

N 

'</) 

■o 

o 

CO 


86 


84- 


82- 


l 


o 


i 


FIG.  2.  Mean  and  95%  Cl  of  (A)  body  size  and 
(B)  body  condition  for  male  (filled  circles)  and  female 
(open  circles)  Northern  Rickers  performing  four  levels 
of  nest  defense  (1=  least,  4 = greatest;  see  text  for 
description  of  defense  scores)  in  response  to  a model 
predator  placed  at  nests  during  the  brood-rearing  stage 
at  Riske  Creek,  British  Columbia,  2003  and  2004. 
Body  size  differed  between  the  sexes,  but  defense 
scores  did  not  vary  with  body  size  or  condition. 


defense  scores  was  not  associated  with  brood 
size  (x2  = 2.48,  df  = 3,  P = 0.48). 

As  expected,  adult  body  size  was  signifi- 
cantly associated  with  sex  (males  were  struc- 
turally larger  than  females;  two-factor  ANO- 
VA: F = 345.67,  df  = 1,  P < 0.001),  but  there 
was  no  relationship  between  body  size  and  de- 
fense score  (F  = 0.33,  df  = 3,  P = 0.80;  Fig. 
2),  nor  was  there  a sex  X defense  score  inter- 
action (F  = 0.41,  df  = 3,  P = 0.75).  Similarly, 
there  was  no  relationship  between  body  con- 


DISCUSSION 

Relationship  between  sex  and  nest  de- 
fense.— Although  a model  predator  may  not 
elicit  the  same  intensity  of  nest  defense  as  a 
real  predator,  the  fact  that  flickers  responded 
to  it  more  intensely  than  to  the  control  model 
suggests  that  they  did  perceive  danger.  Con- 
sistent with  initial  predictions,  we  found  no 
differences  between  nest  defense  of  male  and 
female  flickers.  Although  many  studies  have 
revealed  sex-related  differences  in  nest  de- 
fense among  birds  (Gill  and  Sealy  1996,  Caw- 
thom  et  al.  1998,  Pavel  and  Bures  2001,  Grig- 
gio  et  al.  2003),  others  have  not,  including 
studies  on  the  American  Goldfinch  ( Carduelis 
tristis;  Knight  and  Temple  1986b)  and  Red- 
backed  Shrike  ( Lanius  collurio;  Tryjanowski 
and  Golawski  2004).  Adult  male  and  female 
American  Goldfinches  may  exhibit  equal  de- 
fense responses  because  they  are  monoga- 
mous and  both  sexes  are  required  to  raise  the 
young  (Knight  and  Temple  1986b).  Tryja- 
nowski and  Golawski  (2004)  suggested  that 
net  costs  and  benefits  of  nest  defense  by  male 
and  female  Red-backed  Shrikes  were  equal 
because  males  were  larger  than  females,  but 
females  had  greater  confidence  of  parenthood. 
For  flickers,  the  sex-related  differences  in  sur- 
vival (male  survival  is  2%  lower  than  that  of 
females;  Fisher  and  Wiebe  2006b),  body  size 
(males  are  —3%  larger  than  females;  Moore 
1995,  Wiebe  2000),  and  investment  in  the  cur- 
rent brood  (Moore  1995,  Wiebe  and  Elchuk 
2003)  are  likely  too  small  to  alter  the  costs 
and  benefits  of  sex-related  nest  defense. 
Among  cavity  nesters,  male  Eastern  Screech- 
owls  (Otus  [currently  Megascops ] asio ) de- 
fend nestlings  more  aggressively  than  females 
(Sproat  and  Ritchison  1993),  as  do  male  Great 
Tits  ( Parus  major,  Currio  and  Onnebrink 
1995)  and  male  Tree  Swallows  ( Tachycineta 
bicolon ; Winkler  1992). 

Age  and  nest  defense. — In  general,  we 
found  no  significant  association  between  age 
and  nest  defense,  although  males  ^3  years  old 
tended  to  engage  in  more  risky  defense  be- 
havior (attributed  to  their  greater  tendency  to 


458 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


block  the  cavity  entrance)  than  the  other 
groups.  Blocking  the  cavity  entrance  may  be 
used  by  cavity  nesters  to  prevent  usurpation 
of  cavities  (Cordero  and  Senar  1990).  With 
the  head  and  bill  in  striking  position  at  the 
entrance  hole,  it  also  may  be  an  effective  strat- 
egy for  fending  off  an  attack  while  minimiz- 
ing risk  to  the  rest  of  the  parent’s  body.  The 
lack  of  strong  age  or  sex  effects  on  any  de- 
fense behavior  suggests  that  individuals  of 
different  ages  perceive  the  overall  costs  and 
benefits  of  nest  defense  in  a similar  way. 

According  to  economic  models  of  nest  de- 
fense (Montgomerie  and  Weatherhead  1988), 
an  older  bird  should  defend  its  current  brood 
more  aggressively  than  a younger  bird  be- 
cause it  has  a lower  future  reproductive  po- 
tential; however,  we  found  no  evidence  for 
this  in  flickers.  Winkler  (1992)  explained  that 
age-independent  survival  probabilities  pre- 
cluded an  effect  of  age  on  nest  defense  by 
Tree  Swallows.  Similarly,  the  annual  apparent 
survival  rates  (42%)  for  flickers  do  not  vary 
with  age,  and  the  birds  are  relatively  short- 
lived (Fisher  and  Wiebe  2006b),  so  it  is  prob- 
ably not  surprising  that  age  has  little  influence 
on  defense  intensity. 

Although  future  reproductive  potential  is 
one  component  that  could  lead  to  age-depen- 
dent nest  defense,  experience  also  may  be  a 
key  factor  if  defense  is  learned  and  becomes 
less  risky  for  the  adult  over  time  (Montgom- 
erie and  Weatherhead  1988).  We  could  not 
separate  age  from  experience  in  our  study  and 
it  is  impossible  to  know  the  previous  experi- 
ence that  a wild  bird  may  have  had  with  a 
predator. 

Effects  of  body  size  and  condition  on  nest 
defense . — It  was  surprising  that  neither  body 
size  nor  condition  were  positively  associated 
with  our  measures  of  flicker  nest  defense.  Al- 
though sexual-size  dimorphism  is  often  cited 
as  contributing  to  differences  in  nest  defense 
between  the  sexes  (Tryjanowski  and  Golawski 
2004),  effects  of  body-size  differences  within 
the  sexes  have  rarely  been  tested  (Hamer  and 
Furness  1993,  Radford  and  Blakey  2000).  If 
large  and  small  birds  are  both  effective  nest 
defenders  for  different  reasons — for  example, 
if  small  individuals  have  greater  maneuver- 
ability and  large  individuals  are  more  power- 
ful— then  overall  costs  and  benefits  may  be 
similar  for  each  (Montgomerie  and  Weather- 


head 1988).  The  few  studies  that  have  tested 
for  within-sex  effects  of  body  condition  have 
been  equivocal  at  best,  ranging  from  no  effect 
(Radford  and  Blakey  2000)  to  a sex-specific 
effect  (Winkler  1992,  Hamer  and  Furness 
1993).  There  is  little  direct  evidence  that  body 
condition  affects  the  intensity  of  active  de- 
fense in  any  species,  but  good  nutrient  re- 
serves may  allow  a parent  to  reduce  foraging 
time  away  from  the  nest  and  be  more  attentive 
to  the  nest  site  during  incubation  and  brooding 
(Slagsvold  and  Lifjeld  1989,  Wiebe  and  Mar- 
tin 1997);  in  turn,  these  factors  would  result 
in  greater  nesting  success  (Chastel  et  al. 
1995).  We  found  some  evidence  that  birds  in 
better  body  condition  responded  more  quickly 
to  the  predator  model,  which  may  provide 
support  for  this  hypothesis.  Flicker  condition 
was  measured  in  the  late  stages  of  incubation 
or  early  stages  of  brooding  when  parents 
could  be  captured;  thus,  they  may  not  have 
been  in  exactly  the  same  condition  at  the  time 
of  our  defense  trials  (about  10-15  days  later). 
However,  if  relative  rankings  of  body  condi- 
tion among  individuals  remain  similar,  we 
should  have  been  able  to  detect  a pattern. 

Effects  of  brood  size  on  nest  defense. — We 
predicted  that  male  and  female  flickers  with 
larger  broods  should  defend  them  more  ag- 
gressively than  flickers  with  smaller  broods, 
but  brood  size  was  not  correlated  with  any  of 
the  defense  behaviors  that  we  measured.  Try- 
janowski and  Golawski  (2004)  suggested  that 
brood  size  manipulation  experiments  are 
needed  to  adequately  test  for  effects  of  brood 
size  on  nest  defense.  However,  even  some  ex- 
perimental studies  have  failed  to  reveal  any 
differences  in  nest  defense  as  a result  of  brood 
size  (Tolonen  and  Korpimaki  1995).  If  parents 
optimize  their  clutch  size  according  to  their 
ability  to  raise  all  their  young,  then  large  and 
small  broods  may  represent  equal  value  to  the 
defending  adults,  in  which  case  brood  size 
may  not  be  expected  to  influence  nest  defense 
(Tolonen  and  Korpimaki  1995,  Dawson  and 
Bortolotti  2003). 

In  summary,  anecdotal  data  from  the  liter- 
ature (Lawrence  1967)  and  video-tape  evi- 
dence from  our  own  study  site  (KLW  unpubl. 
data)  indicates  that  the  defense  behaviors  we 
observed  may  successfully  protect  cavity 
nests  from  live  predators,  such  as  red  squir- 
rels. Individual  flickers  varied  in  their  re- 


Fisher  and  Wiebe  • NORTHHRN  FLICKER  NEST  DEFENSE 


459 


sponses,  but  we  were  unable  to  find  strong 
correlates  of  that  variation  associated  with 
common  traits  of  those  individuals  or  their 
broods. 

ACKNOWLEDGMENTS 

We  sincerely  thank  C.  L.  Galatiuk,  H.  J.  Kalyn,  J. 
R M.  Johnston,  and  K.  M.  Warner,  who  helped  with 
the  model  presentations.  We  would  also  like  to  thank 
K.  Martin,  who  allowed  us  to  conduct  model  presen- 
tations on  parts  of  her  study  area.  D.  J.  Ingold,  J.  J. 
Kappes,  Jr.,  and  one  anonymous  reviewer  improved 
earlier  drafts  of  this  manuscript.  This  project  was  fund- 
ed through  National  Sciences  and  Engineering  Re- 
search Council  of  Canada  and  Southern  Interior  Blue- 
bird Trail  Society  scholarships  to  RJF,  and  through  a 
National  Sciences  and  Engineering  Research  Council 
of  Canada  operating  grant  to  KLW.  We  would  also  like 
to  thank  the  Borror  Acoustics  Laboratory  for  providing 
squirrel  and  blackbird  vocalizations  used  in  the  model 
presentations. 

LITERATURE  CITED 

Aitken,  K.  E.  H„  K.  L.  Wiebe,  and  K.  Martin.  2002. 
Nest-site  reuse  patterns  for  a cavity  nesting  bird 
community  in  interior  British  Columbia.  Auk  1 19: 
391-402. 

Andersson,  M.,  C.  G.  Wiklund,  and  H.  Rundgren. 
1980.  Parental  defence  of  offspring:  a model  and 
an  example.  Animal  Behaviour  28:536-542. 
Blancher,  P.  J.  and  R.  J.  Robertson.  1982.  Kingbird 
aggression:  does  it  deter  predation?  Animal  Be- 
haviour 30:929-945. 

Breitwisch,  R.  1988.  Sex  differences  in  defence  of 
eggs  and  nestlings  by  Northern  Mockingbirds,  Mi- 
mus  polyglottos.  Animal  Behaviour  36:62-72. 
Cawthorn,  J.  M.,  D.  L.  Morris,  E.  D.  Ketterson, 
and  V.  Nolan.  1998.  Influence  of  experimentally 
elevated  testosterone  on  nest  defence  in  Dark- 
eyed Juncos.  Animal  Behaviour  56:617-621. 
Chastel,  O.,  H.  Weimerskirch,  and  P.  Jouventin. 
1995.  Influence  of  body  condition  on  reproductive 
decision  and  reproductive  success  in  the  Blue  Pe- 
trel. Auk  112:964-972. 

Cordero,  P.  J.  and  J.  C.  Senar.  1990.  Interspecific 
nest  defence  in  European  Sparrows:  different 
strategies  to  deal  with  a different  species  of  op- 
ponent? Ornis  Scandinavica  21:71-73. 

Curio,  E.  and  H.  Onnebrink.  1995.  Brood  defense  and 
brood  size  in  the  Great  Tit  ( Parus  major):  a test 
of  a model  of  unshared  parental  investment.  Be- 
havioral Ecology  6:235-241. 

Dawson,  R.  D.  and  G.  R.  Bortolotti.  2003.  Parental 
effort  of  American  Kestrels:  the  role  of  variation 
in  brood  size.  Canadian  Journal  of  Zoology  81: 
852-860. 

Elchuk,  C.  L.  and  K.  L.  Wiebe.  2003.  Ephemeral  food 
resources  and  high  conspecific  densities  as  factors 
explaining  lack  of  feeding  territories  in  Northern 
Flickers  ( Colaptes  auratus).  Auk  120:187-193. 


Fisher,  R.  J.  and  K.  L.  Wiebe.  2006a.  Effects  of  sex 
and  age  on  survival  of  Northern  Flickers:  a six- 
year  field  study.  Condor  108:193-200. 

Fisher,  R.  J.  and  K.  L.  Wiebe.  2006b.  Nest  site  attri- 
butes and  temporal  patterns  of  Northern  Flicker 
nest  loss:  effects  of  predation  and  competition. 
Oecologia  147:744-753. 

Ghalambor,  C.  K.  and  T.  E.  Martin.  2002.  Compar- 
ative manipulation  of  predation  risk  in  incubating 
birds  reveals  variability  in  the  plasticity  of  re- 
sponses. Behavioral  Ecology  13:101-108. 

Gill,  S.  A.  and  S.  G.  Sealy.  1996.  Nest  defence  by 
Yellow  Warblers:  recognition  of  a brood  parasite 
and  an  avian  nest  predator.  Behaviour  133:263- 
282. 

Griggio,  M.,  G.  Matessi,  and  A.  Pilastro.  2003. 
Male  Rock  Sparrow  ( Petronia  petronia)  nest  de- 
fence correlates  with  female  ornament  size.  Ethol- 
ogy 109:659-669. 

Hamer,  K.  C.  and  R.  W.  Furness.  1993.  Parental  in- 
vestment and  brood  defense  by  male  and  female 
Great  Skuas  Catharacta  skua:  the  influence  of 
food-supply,  laying  date,  body  size  and  body  con- 
dition. Journal  of  Zoology  230:7-18. 

Hatch,  M.  I.  1997.  Variation  in  Song  Sparrow  nest 
defense:  individual  consistency  and  relationship  to 
nest  success.  Condor  99:282-289. 

Ingold,  D.  J.  1994.  Influence  of  nest-site  competition 
between  European  starlings  and  woodpeckers. 
Wilson  Bulletin  106:227-241. 

Kappes,  J.  J.,  Jr.  1997.  Defining  cavity-associated  in- 
teractions between  Red-cockaded  Woodpeckers 
and  other  cavity-dependent  species:  interspecific 
competition  or  cavity  kleptoparasitism?  Auk  1 14: 
778-780. 

Knight,  R.  L.  and  S.  A.  Temple.  1986a.  Methodolog- 
ical problems  in  studies  of  avian  nest  defence.  An- 
imal Behaviour  34:561-566. 

Knight,  R.  L.  and  S.  A.  Temple.  1986b.  Nest  defence 
in  the  American  Goldfinch.  Animal  Behaviour  34: 
887-897. 

Knight,  R.  L.  and  S.  A.  Temple.  1986c.  Why  does 
intensity  of  avian  nest  defense  increase  during  the 
nesting  cycle?  Auk  103:318-327. 

Lawrence,  L.  D.  K.  1967.  A comparative  life-history 
study  of  four  species  of  woodpeckers.  Allen  Press, 
Lawrence,  Kansas. 

Martin,  K.  and  J.  M.  Eadie.  1999.  Nest  webs:  a com- 
munity-wide approach  to  the  management  and 
conservation  of  cavity-nesting  forest  birds.  Forest 
Ecology  and  Management  15:243-257. 

Martin,  K.  and  A.  G.  Horn.  1993.  Clutch  defense  by 
male  and  female  Willow  Ptarmigan  Lagopus  la- 
gopus.  Ornis  Scandinavica  24:261-266. 

Montgomerie,  R.  D.  and  P.  J.  Weatherhead.  1988. 
Risks  and  rewards  of  nest  defence  by  parent  birds. 
Quarterly  Review  of  Biology  63:167-187. 

Moore,  W.  S.  1995.  Northern  Flicker  ( Colaptes  aura- 
tus). The  Birds  of  North  America,  no.  166. 

Nealen,  P.  M.  and  R.  Breitwisch.  1997.  Northern 


460 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Cardinal  sexes  defend  nests  equally.  Wilson  Bul- 
letin 109:269-278. 

Olendorf,  R.  and  S.  K.  Robinson.  2000.  Effective- 
ness of  nest  defence  in  the  Acadian  Flycatcher 
Empidonax  virescens.  Ibis  142:365-371. 

Pavel,  V.  and  S.  Bures.  2001.  Offspring  age  and  nest 
defence:  test  of  the  feedback  hypothesis  in  the 
Meadow  Pipit.  Animal  Behaviour  61:297-303. 

Promislow,  D.  E.,  R.  Montgomerie,  and  T.  E.  Mar- 
tin. 1992.  Mortality  costs  of  sexual  dimorphism 
in  birds.  Proceedings  of  the  Royal  Society  of  Lon- 
don, Series  B 250:143-150. 

Pyle,  R,  N.  G.  Howell,  D.  F.  DeSante,  R.  P.  Yunick, 
and  M.  Gustafson.  1997.  Identification  guide  to 
North  American  birds,  part  I.  Columbidae  to  Plo- 
ceidae.  Slate  Creek  Press,  Bolinas,  California. 

Radford,  A.  N.  and  J.  K.  Blakey.  2000.  Intensity  of 
nest  defence  is  related  to  offspring  sex  ratio  in  the 
Great  Tit  Parus  major.  Proceedings  of  the  Royal 
Society  of  London,  Series  B 267:535-538. 

Ratti,  O.  2000.  Characteristics  and  level  of  aggression 
by  female  Pied  Flycatchers  at  different  distances 
from  the  nest  hole.  Ornis  Fennica  77:1 1-16. 

Slagsvold,  T.  and  J.  T.  Lifjeld.  1989.  Constraints  on 
hatching  asynchrony  and  egg  size  in  Pied  Fly- 
catchers. Journal  of  Animal  Ecology  58:837-849. 

Sproat,  T.  M.  and  G.  Ritchison.  1993.  The  nest  de- 
fense behavior  of  Eastern  Screech-owls:  effects  of 
nest  stage,  sex,  nest  type  and  predator  location. 
Condor  95:288-296. 

Tolonen,  P.  and  E.  Korpimaki.  1995.  Parental  effort 
of  Kestrels  ( Falco  tinnunculus ) in  nest  defense: 
effects  of  laying  time,  brood  size,  and  varying  sur- 
vival prospects  of  offspring.  Behavioral  Ecology 
6:435-441. 

Tryjanowski,  P.  and  A.  Golawski.  2004.  Sex  differ- 
ences in  nest  defence  by  the  Red-backed  Shrike 
Lanius  collurio:  effects  of  offspring  age,  brood 
size,  and  stage  of  breeding  season.  Journal  of 
Ethology  22:13-16. 


Veen,  T.,  D.  S.  Richardson,  K.  Blaakmeer,  and  J. 
Komdeur.  2000.  Experimental  evidence  for  innate 
predator  recognition  in  the  Seychelles  Warbler. 
Proceedings  of  the  Royal  Society  of  London,  Se- 
ries B 267:2253-2258. 

Weidinger,  K.  2002.  Interactive  effects  of  conceal- 
ment, parental  behaviour  and  predators  on  the  sur- 
vival of  open  passerine  nests.  Journal  of  Animal 
Ecology  71:424-437. 

Wiebe,  K.  L.  2000.  Assortative  mating  by  color  in  a 
population  of  hybrid  Northern  Flickers.  Auk  117: 
525-529. 

Wiebe,  K.  L.  2001.  Microclimate  of  tree  cavity  nests: 
is  it  important  for  reproductive  success  in  North- 
ern Flickers?  Auk  118:412-421. 

Wiebe,  K.  L.  2003.  Delayed  timing  as  a strategy  to 
avoid  nest-site  competition:  testing  a model  using 
data  from  starlings  and  flickers.  Oikos  100:291- 
298. 

Wiebe,  K.  L.  2004.  Innate  and  learned  components  of 
defence  by  flickers  against  a novel  nest  competi- 
tor, the  European  Starling.  Ethology  110:1-13. 

Wiebe,  K.  L.  and  C.  L.  Elchuk.  2003.  Correlates  of 
parental  care  in  Northern  Flickers  Colaptes  aura- 
tus : do  the  sexes  contribute  equally  while  provi- 
sioning young?  Ardea  91:91-101. 

Wiebe,  K.  L.,  W.  D.  Koenig,  and  K.  Martin.  2006. 
Evolution  of  clutch  size  in  cavity-excavating 
birds:  the  nest  site  limitation  hypothesis  revisited. 
American  Naturalist  167:343-353. 

Wiebe,  K.  L.  and  K.  Martin.  1997.  Effects  of  pre- 
dation, body  condition  and  temperature  on  incu- 
bation rhythms  of  White-tailed  Ptarmigan  Lago- 
pus  leucurus.  Wildlife  Biology  3:219-227. 

Wiebe,  K.  L.  and  T.  L.  Swift.  2001.  Clutch  size  rel- 
ative to  tree  cavity  size  in  Northern  Flickers.  Jour- 
nal of  Avian  Biology  32:167-173. 

Winkler,  D.  W.  1992.  Causes  and  consequences  of 
variation  in  parental  defense  behavior  by  Tree 
Swallows.  Condor  94:502-520. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):46 1—470,  2006 


BLACK-THROATED  BLUE  WARBLER  AND  VEERY  ABUNDANCE 
IN  RELATION  TO  UNDERSTORY  COMPOSITION  IN 
NORTHERN  MICHIGAN  FORESTS 

LAURA  J.  KEARNS,134  EMILY  D.  SILVERMAN,1  AND  KIMBERLY  R.  HALL2 3 4 


ABSTRACT. — Balsam  fir  ( Abies  balsamea)  understory  may  be  an  important  predictor  of  Black-throated  Blue 
Warbler  ( Dendroica  caerulescens ) and  Veery  {Catharus  fuscescens)  distributions  in  northern  hardwood  forests 
that  are  heavily  browsed  by  white-tailed  deer  ( Odocoileus  virginianus).  We  examined  the  abundance  and  age 
ratios  of  Black-throated  Blue  Warblers,  and  the  abundance  of  Veerys,  in  16  plots  of  hardwood  forest  with 
different  understory  composition  within  a heavily  browsed  region  of  the  Hiawatha  National  Forest  in  Michigan’s 
eastern  Upper  Peninsula.  Four  of  these  36-ha  plots  had  minimal  understory  and  12  had  dense  understory  with 
variable  amounts  of  balsam  fir.  Black-throated  Blue  Warbler  abundance  was  significantly  greater  in  plots  with 
an  average  of  27%  balsam  fir  understory  cover  than  in  plots  dominated  by  deciduous  understory;  no  Black- 
throated  Blue  Warblers  were  detected  on  the  minimal  understory  plots.  Age  ratios  did  not  differ  significantly 
relative  to  balsam  fir  understory  density.  Veery  abundance  also  did  not  vary  with  balsam  fir  understory  density, 
but  it  increased  with  overall  understory  density.  In  forests  such  as  these,  where  deer  are  abundant  but  rarely 
browse  balsam  fir,  active  management  of  balsam  fir  understory  could  provide  key  habitat  for  sustaining  popu- 
lations of  Black-throated  Blue  Warblers  and  Veerys.  We  recommend  that  managers  consider  the  presence  of 
balsam  firs  in  the  understory  when  planning  forest  harvests  in  deer-impacted  areas,  so  that  they  leave  some 
balsam  fir  and  stagger  the  cutting  of  stands  with  balsam  fir  over  time  to  create  and  maintain  heterogeneous 
understory  structure.  Received  2 September  2005,  accepted  16  May  2006. 


Identifying  key  habitat  characteristics  that 
predict  songbird  distributions  represents  an 
important  step  towards  incorporating  song- 
birds into  forest  management  plans  (Martin 
1992,  Donovan  et  al.  2002).  In  the  eastern 
United  States,  browsing  of  understory  vege- 
tation by  white-tailed  deer  ( Odocoileus  virgi- 
nianus) produces  forests  that  differ  in  terms 
of  their  structural  characteristics  and  plant 
species  compositions  from  those  in  less  im- 
pacted areas  (reviewed  by  Rooney  and  Waller 
2003,  Cote  et  al.  2004),  and  these  changes  can 
affect  the  abundance  of  understory-dependent 
songbirds  (Casey  and  Hein  1983,  deCalesta 
1994,  McShea  and  Rappole  2000).  Browsing 
impacts,  however,  are  likely  to  differ  across 
species’  ranges  because  of  variation  in  the 
plant  community,  the  landscape  context,  and, 
in  the  Great  Lakes  region,  the  degree  to  which 


1 School  of  Natural  Resources  and  Environment, 
Univ.  of  Michigan,  440  Church  St.,  Ann  Arbor,  MI 
48109,  USA. 

2 Dept,  of  Forestry  and  Dept,  of  Fisheries  and  Wild- 
life, Michigan  State  Univ.,  126  Natural  Resources 
Building,  East  Lansing,  MI  48824,  USA. 

3 Current  address:  Smithsonian  Conservation  and 
Research  Center,  1500  Remount  Rd.,  Front  Royal,  VA 
22630,  USA. 

4 Corresponding  author;  e-mail: 
laurajkeams@yahoo.com 


the  understory  is  protected  from  deer  by  snow. 
Therefore,  predicting  the  abundance  of  under- 
story-dependent  birds  is  best  approached  us- 
ing habitat  indicators  based  on  local  infor- 
mation, a key  element  of  which  may  be  the 
distribution  of  browse-resistant  plants. 

We  investigated  the  relationship  between 
understory  characteristics  and  the  abundance 
of  two  forest  songbird  species,  the  Black- 
throated  Blue  Warbler  ( Dendroica  caerules- 
cens; BTBW)  and  the  Veery  ( Catharus  fus- 
cescens),  in  managed  northern  hardwood  for- 
ests in  the  eastern  Upper  Peninsula  of  Mich- 
igan, where  the  overabundance  of  deer  is  a 
conservation  concern  (The  Nature  Conservan- 
cy 2000,  Rooney  and  Waller  2003,  Kraft  et  al. 
2004).  Our  sites  were  dominated  by  sugar  ma- 
ple {Acer  saccharum)  and  located  near  conif- 
erous forest  “deeryards” — areas  that  provide 
winter  habitat  for  high  densities  of  deer  (Van 
Deelen  et  al.  1998).  At  similar  Great  Lakes 
forest  sites,  browsing  has  decreased  understo- 
ry density  and  reduced  structural  complexity, 
especially  for  sugar  maple  seedlings  and  sap- 
lings (Alverson  et  al.  1988,  Kraft  et  al.  2004). 

Veerys  and  BTBWs  are  likely  to  be  suscep- 
tible to  browsing  impacts  because  they  nest 
and  forage  in  the  understory  (Holmes  1994, 
Moskoff  1995).  Both  species  are  also  of  con- 


461 


462 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


servation  concern  in  northern  forests  (U.S. 
Fish  and  Wildlife  Service  2002;  Matteson  et 
al.  in  press).  BTBWs  have  been  studied  inten- 
sively in  New  Hampshire,  where  population 
density  is  positively  associated  with  shrub  and 
sapling  density  (Steele  1992,  1993;  Holmes  et 
al.  1996),  and  the  density  of  deciduous  leaves 
in  the  shrub  layer  is  a key  predictor  of  terri- 
tory quality  (Rodenhouse  et  al.  2003).  Less  is 
known  about  key  habitat  features  for  Veerys 
but,  in  Michigan,  they  are  typically  found  in 
mesic  to  wet  forest  with  dense  understory  and 
a conifer  component  (Winnett-Murray  1991). 

We  hypothesized  that  the  density  of  under- 
story balsam  fir  (Abies  balsamea ),  a species 
rarely  browsed  by  deer  in  our  region  (Borg- 
mann  et  al.  1999),  may  better  predict  BTBW 
abundance  than  deciduous  species  in  Great 
Lakes  forests.  Our  previous  work  in  Michigan 
hardwood  forests  near  deeryards  revealed  that 
100-m-radius  point-count  locations  with  abun- 
dant balsam  fir  had  higher  relative  abundances 
of  BTBWs  than  locations  with  dense,  decid- 
uous-dominated  understory  (Hall  2002).  In 
this  paper,  we  considered  a management-rel- 
evant scale  (36-ha  stand)  and  compared 
BTBW  and  Veery  abundance  between  plots 
that  varied  in  their  proportion  of  balsam  fir 
understory.  We  also  predicted  that  areas  with 
more  balsam  fir  would  have  a higher  ratio  of 
older  to  yearling  BTBWs,  thus  indicating  hab- 
itat preference  (Holmes  et  al.  1996,  Hunt 
1996). 

METHODS 

Study  area. — We  collected  data  in  16  stands 
of  mature,  relatively  even-aged  hardwood  for- 
est within  a section  (—15  X 7 km2)  of  the 
southeastern  Hiawatha  National  Forest  in 
Mackinac  County,  Michigan,  between  46°  09' 
06"  N to  46°  05'  1 8"  N and  84°  52' 23"  W to 
84°  40'  50"  W (Fig.  1).  All  plots  were  located 
within  the  St.  Ignace  subsection  of  the  Nia- 
garan  Escarpment,  an  area  characterized  by 
shallow  morainal  soils  and  occasional  glacial 
erratics  (Albert  1995).  Sugar  maple  was  the 
dominant  overstory  tree  on  the  study  plots, 
but  often  was  co-dominant  with  American 
beech  ( Fagus  grandifolia ) and,  to  a lesser  ex- 
tent, aspen  ( Populus  spp.),  paper  birch  ( Betula 
papyrifera),  and  American  basswood  (Tilia 
americana );  rarely,  balsam  fir  and  white  pine 
(Pinus  strobus ) were  also  co-dominant.  Typi- 


cal understory  species  included  sugar  maple, 
hop-hornbeam  ( Ostrya  virginiana),  and  bal- 
sam fir;  occasionally  we  found  seedlings  and 
saplings  of  other  canopy  species  and  white 
spruce  (Picea  glauca ),  white  ash  (Fraxinus 
americana ),  and  black  cherry  ( Prunus  seroti- 
na ).  The  study  area  receives  an  annual  average 
of  1.5-2  m of  snow  (Albert  1995),  which  ap- 
pears to  protect  many  plants  from  being  com- 
pletely removed  by  overwintering  deer  that 
seek  shelter  in  the  nearby  deeryards  and  enter 
these  stands  to  forage. 

We  chose  site  locations  using  a 2002  GIS 
database  of  forest  management  units  in  the  Hi- 
awatha National  Forest  within  the  Niagaran 
Escarpment  (U.S.  Department  of  Agriculture 
Forest  Service  unpubl.  data).  We  used  Arc- 
View  (Environmental  Systems  Research  Insti- 
tute 2002)  to  select  hardwood  management 
units  large  enough  to  accommodate  a square 
36-ha  plot,  then  visited  those  units  in  random 
order  for  the  purpose  of  selecting  our  16  sites, 
with  four  in  each  of  the  following  understory 
categories:  (1)  minimal  understory  vegetation, 
(2)  deciduous-dominated  understory  vegeta- 
tion with  sparse  balsam  fir,  (3)  understory 
vegetation  with  moderate  balsam  fir  density, 
and  (4)  understory  vegetation  with  high  bal- 
sam fir  density  (Fig.  1).  The  initial  assignment 
of  sites  to  understory  categories  was  based  on 
visual  estimates  conducted  in  May,  prior  to 
the  standardized  collection  of  vegetation  data 
(see  below).  The  dark  vegetated  areas  (Fig.  1) 
were  dominated  by  coniferous  overstory  and 
comprised  the  habitat  type  typical  of  deer- 
yards in  this  region  (Van  Deelen  et  al.  1998). 
The  36-ha  plot  size  was  small  enough  so  that 
sites  were  internally  similar  (e.g.,  within  the 
same  management  unit,  with  similar  canopy 
cover  and  understory  density,  and  with  few 
old  logging  roads  or  other  openings),  yet  large 
enough  to  encompass  a wide  range  in  the 
number  of  BTBW  territories  (typically  1-4  ha 
in  size;  Holmes  1994,  Hall  2002). 

Vegetation  sampling. — We  measured  under- 
story composition  using  a modified  method 
from  Mueller-Dombois  and  Ellenberg  (1974). 
Within  each  plot,  we  established  three  paral- 
lel, 600-m  transects  spaced  200  m apart,  and 
randomly  oriented  the  transects  east-west  or 
north-south.  We  then  divided  each  transect 
into  100-m  segments  and  randomly  chose  a 
16-m2  quadrat  within  each  segment,  for  a total 


Kearns  et  al.  • SONGBIRDS  AND  FOREST  UNDERSTORY 


463 


Michigan 


Hiawatha 

National 

Forest 


Study  area 


□ Minimal  ^Sparse  ■ Moderate  ■High 
understory  balsam  fir  balsam  fir  balsam  fir 


Kilometers 


FIG.  1 . Distribution  of  four  understory  vegetation  plot  types  in  the  Hiawatha  National  Forest,  Mackinac 
County,  Michigan,  summers  2002  and  2003.  Digital  orthophoto  taken  before  leaf-out  in  March  2001  shows 
conifer  stands  in  dark  gray,  hardwood  stands  in  light  gray,  and  water  in  near  black.  Squares  represent  36-ha 
plots  ( n = 16):  white  = minimal  understory,  light  gray  = deciduous-dominated  understory  with  sparse  stem 
densities  of  balsam  fir,  dark  gray  = understory  with  moderate  balsam  fir  density,  black  = understory  with  high 
balsam  fir  density.  Points  in  each  plot  ( n = 9)  represent  approximate  locations  of  100-m  radius  avian  point 
counts  and  11.3-m  overstory  sampling  subplots.  (Sources:  Environmental  Systems  Research  Institute  2002  Pro- 
jection: UTM  Zone  16  N,  Datum:  NAD  1927;  U.S.  Department  of  Agriculture  Forest  Service  unpubl.  data.) 


of  18  quadrats  per  plot.  For  each  quadrat,  we 
calculated  the  total  stem  count  and  average 
percent  cover  of  woody  understory  plant  spe- 
cies within  six  height  categories — five  0.5-m 
categories  (ranging  from  0.5  to  3 m)  and  a 3- 
to  5-m  category — based  on  estimates  from 
four  4-m2  sub-quadrats.  Using  a spherical  den- 
siometer,  we  also  measured  the  canopy  cover 
in  each  quadrat.  Following  a modification  of 
James  and  Shugart’s  (1970)  vegetation  sam- 
pling method,  in  each  36-ha  plot  we  estab- 
lished nine  points  spaced  200  m apart  on  a 3 
X 3 grid  (Fig.  1);  within  an  11.3-m  radius  of 
each  point,  we  counted  the  number  of  trees  in 
two  size  categories  (small:  7.5-22.5  cm  in  di- 
ameter at  breast  height  [dbh],  large:  >22.5  cm 
dbh).  We  sampled  all  vegetation  between  late 


July  and  September,  prior  to  leaf  fall,  in  2002 
and  2003. 

We  calculated  mean  stem  density,  percent 
cover,  and  height  for  both  balsam  fir  and  de- 
ciduous understory  species  from  the  18  quad- 
rats in  each  plot.  We  calculated  the  standard 
deviation  of  percent  cover  as  a measure  of  un- 
derstory patchiness.  We  used  the  standard  de- 
viation of  height  as  a measure  of  understory 
vertical  structure.  We  also  determined  mean 
density  of  small  and  large  trees  in  the  1 1.3-m 
point  samples. 

Bird  sampling. — In  2002,  we  measured  the 
abundance  of  territorial  male  BTBWs  by  tar- 
get-netting and  color-banding  birds.  An  ob- 
server (LJK)  first  surveyed  each  plot  during 
late  May-early  June  by  walking  the  three  tran- 


464  THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


sects  and  using  song  playbacks  to  detect  and 
record  the  locations  of  BTBWs;  Wolf  et  al. 
(1995)  estimated  that  BTBW  song  is  detect- 
able up  to  120  m from  an  observer.  Plots  were 
revisited  up  to  10  more  times  between  late 
May  and  late  July,  depending  on  the  density 
of  male  BTBWs  and  how  catchable  they  were. 
During  these  visits,  two  or  three  observers 
once  again  searched  the  plots  for  male 
BTBWs  by  walking  transects  and  using  song 
playbacks;  nearly  all  males  within  each  plot 
were  captured  and  color-banded  by  targeted 
mist-netting  (song  playback  and  model  bird). 
We  banded  each  bird  with  a federal  aluminum 
leg  band  and  two  colored  plastic  leg  bands. 
During  banding,  we  determined  age  as  older 
(after  second  year;  ASY)  or  yearling  (second 
year;  SY)  on  the  basis  of  plumage  character- 
istics (Pyle  1997).  Experienced  observers 
(KRH,  LJK)  aged  three  uncatchable  birds  by 
using  binoculars  to  study  their  plumage  char- 
acteristics (Graves  1997a).  Between  late  May 
and  early  June  2003,  we  systematically  resur- 
veyed all  plots  using  song  playback  to  deter- 
mine 2003  abundance. 

From  early  June  to  mid-July  2002,  we  con- 
ducted 10-min  point  counts  (100-m  fixed  ra- 
dius) of  singing  males  to  estimate  the  relative 
abundances  of  Veerys  and  BTBWs  (as  a sec- 
ond measure)  in  each  plot  (Ralph  et  al.  1993). 
For  each  bird,  we  recorded  its  location  within 
one  of  three  distance  categories  (0-25,  25-50, 
50-100  m)  and  time  to  detection  (0-3,  3-5, 
5-10  min).  Weather  permitting,  LJK  surveyed 
one  plot  per  day,  starting  the  count  within  30 
min  of  sunrise.  After  randomly  selecting  a 
starting  point  from  one  of  the  nine  points 
within  a given  plot  (Fig.  1),  the  observer  con- 
ducted the  count  following  the  most  efficient 
route.  We  minimized  the  potential  for  double- 
counting birds  that  moved  between  survey 
points  by  eliminating  individual  detections  in 
similar  locations  on  adjacent  counts.  Since 
BTBWs  often  move  quickly  across  large  ter- 
ritories (e.g.,  >200  m in  diameter;  Hall  2002), 
double-counting  birds  during  point  counts  was 
a particular  concern.  Thus,  our  BTBW  anal- 
yses focused  on  the  banding  data,  whereas  we 
used  the  point  count  data  only  as  an  additional 
measure  of  BTBW  abundance  and  to  verify 
that  we  had  banded  all  birds  in  locations 
where  they  were  detected  during  point  counts. 

Statistical  analyses. — We  performed  Prin- 


TABLE 1.  Eigenvectors  of  the  first  three  principal 
components  for  13  vegetation  variables  measured  in 
36-ha  plots  (n  = 16)  in  the  Hiawatha  National  Forest, 
Michigan,  summer  2002.  The  standard  deviation  (SD) 
of  percent  cover  for  the  18  16-m2  quadrats  in  each  plot 
was  a measure  of  vegetation  patchiness;  the  SD  of  av- 
erage height  was  a measure  of  vertical  structure. 

Eigenvectors 

Variable 

PCAl 

PCA2 

PCA3 

Canopy  cover 

-0.14 

-0.32 

-0.07 

Large-tree  density 

-0.30 

-0.33 

0.20 

Small-tree  density 

0.36 

0.06 

-0.02 

Balsam  fir 

Stem  density 

0.37 

0.03 

-0.04 

Percent  cover 

0.38 

0.03 

0.01 

Cover  SD 

0.37 

0.06 

-0.14 

Height 

0.31 

0.09 

-0.00 

Height  SD 

0.36 

0.11 

-0.01 

Deciduous  spp. 

Stem  density 

-0.19 

0.44 

-0.10 

Percent  cover 

-0.21 

0.51 

-0.10 

Cover  SD 

-0.21 

0.47 

0.04 

Height 

0.02 

0.24 

0.61 

Height  SD 

0.04 

0.16 

0.65 

ciple  Components  Analysis  (PCA)  using  the 
correlation  matrix  for  13  vegetation  variables 
to  explore  the  relationship  between  vegetation 
characteristics  in  the  16  plots  and  to  evaluate 
our  visual  estimates  of  plot  characteristics.  We 
investigated  the  relationships  of  BTBW  abun- 
dance and  age  ratio  (percent  older  birds),  and 
Veery  abundance,  to  plot  characteristics  by 
comparing  the  bird  variables  among  plot  types 
(Kruskal- Wallis  test,  a = 0.05;  Zar  1999)  and 
by  correlating  abundance  with  plot  scores  for 
principal  components  with  eigenvalues  >1. 
Statistical  analyses  were  conducted  in  S-Plus 
6.1  (Insightful  Corporation  2002).  Means  are 
presented  ± SE. 

RESULTS 

Vegetation. — Principle  components  analysis 
identified  three  axes  that  accounted  for  84% 
of  the  variation  in  vegetation  measurements. 
The  first  principle  component,  which  account- 
ed for  50%  of  the  variation  (eigenvalue  = 
6.5),  positively  weighted  all  balsam  fir  vari- 
ables and  small-tree  density,  and  negatively 
weighted  deciduous  understory  and  large-tree 
density  (Table  1).  This  component  distin- 
guished the  eight  plots  classified  by  visual  es- 


Kearns  et  al.  • SONGBIRDS  AND  FOREST  UNDERSTORY 


465 


rs 

< 


FIG.  2.  Principal  components  analysis  (PCA)  showing  variation  in  vegetation  composition  and  structure 
among  36-ha  plots  (n  = 16)  in  the  Hiawatha  National  Forest,  Michigan,  summer  2002.  (A)  Plot-type  distribution: 
triangles  = minimal  understory  plots,  squares  = deciduous-dominated  understory  with  sparse  stem  densities  of 
balsam  fir,  open  circles  = understory  with  moderate  densities  of  balsam  fir,  and  closed  circles  = understory 
with  high  densities  of  balsam  fir.  (B)  Pattern  of  variables  along  PCA  axes.  Axes  1 and  2 accounted  for  50% 
and  18%,  respectively,  of  the  variation  among  plots.  The  first  component  positively  loads  balsam  fir  variables 
and  the  second  positively  loads  stem  density,  percent  cover,  and  patchiness  of  deciduous  vegetation,  thus  sep- 
arating plots  containing  minimal  understory  from  deciduous-dominated  understory;  plots  containing  moderate 
and  high  stem  densities  of  balsam  fir  were  not  clearly  separated. 


timation  as  containing  moderate  to  high  den- 
sities of  balsam  fir  in  the  understory  from  the 
four  minimal  understory  and  four  deciduous- 
dominated  understory  plots  (Fig.  2A).  Stem 
density  of  balsam  fir  in  the  understory  and 
small-tree  overstory  were  highly  correlated 
(Fig.  2B).  The  second  principle  component, 
accounting  for  18%  of  the  variation  (eigen- 
value = 2.3),  positively  weighted  deciduous 
understory  stem  density,  cover,  and  patchiness 
and  negatively  weighted  large-tree  density 
(Table  1,  Fig.  2B).  This  component  distin- 
guished the  four  minimal  understory  plots 


from  the  four  deciduous,  sparse  balsam  fir  un- 
derstory plots.  The  third  principle  component 
described  16%  of  the  variation  (eigenvalue  = 
2.0)  and  positively  weighted  deciduous  un- 
derstory height  and  vertical  structure  (Table 
1);  this  component  was  not  clearly  associated 
with  the  four  understory  plot  types. 

Based  on  the  results  of  the  PCA,  we  rede- 
fined the  understory  categories  of  plots,  re- 
ducing the  number  to  three  categories:  mini- 
mal understory  ( n = 4),  deciduous-dominated 
understory  ( n = 4),  and  balsam  fir-dominated 
understory  ( n = 8).  Compared  to  balsam  fir- 


466 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  2.  Mean  vegetation  and  avian  measurements  (SE)  for  plot  types  after  redefinition  by  principle 
components  analysis:  minimal  understory  (n  = 4),  deciduous-dominated  understory  ( n = 4),  and  balsam  fir- 
dominated  understory  ( n = 8)  in  the  Hiawatha  National  Forest,  Michigan,  summers  of  2002  and  2003.  Vegetation 
variables  included  measures  with  the  largest  loadings  for  the  first  three  principle  components  and  densities  of 
overstory  trees;  plot  types  were  subsequently  defined  by  the  PC  A results.  Deciduous-  and  balsam  fir-dominated 
plots  had  similar  total  understory  cover  but  differed  with  respect  to  composition;  minimal  understory  plots 
contained  more  large  (>22.5  cm  in  diameter  at  breast  height)  trees.  There  were  significant  differences  in  the 
abundances  of  Black-throated  Blue  Warblers  (BTBW)  and  Veerys  by  plot  type  (Kruskal-Wallis  test,  P < 0.05); 
between-plot  differences  in  the  ratio  of  older  to  younger  male  BTBWs  were  not  significant  (Kruskal-Wallis  test, 
P = 0.49). 

Variable 

Plot  type 

Minimal 

understory 

Deciduous-dominated 

understory 

Balsam  fir-dominated 
understory 

Large-tree  density  (stems/ha) 

240  (9) 

162  (10) 

128  (7) 

Small-tree  density  (stems/ha) 

283  (25) 

306  (14) 

487  (24) 

Balsam  fir  understory 

Cover  (%) 

0.0  (0) 

2.5  (2.5) 

26.9  (2.5) 

Height  (m) 

0.50  (0.50) 

0.60  (0.35) 

1.51  (0.07) 

Height  SD 

0.09  (0.09) 

0.28  (0.19) 

0.83  (0.14) 

Deciduous  species  understory 

Cover  (%) 

12.0  (1.9) 

36.0  (5.7) 

12.6  (2.1) 

Height  (m) 

1.27  (0.14) 

1.33  (0.15) 

1.25  (0.10) 

Height  SD 

0.91  (0.17) 

0.98  (0.09) 

0.94  (0.16) 

Black-throated  Blue  Warbler 

Abundance  (2002  banding) 

0.0  (0) 

3.5  (0.6) 

7.1  (1.0) 

Abundance  (2002  point  counts) 

0.0  (0) 

3.8  (0.9) 

5.2  (0.5) 

Abundance  (2003  survey) 

0.0  (0) 

3.3  (1.0) 

6.4  (1.0) 

Age  ratio  (%  older) 

NA 

58.8  (21.2) 

77.8  (7.2) 

Veery 

Abundance  (2002  point  counts) 

1.3  (0.5) 

6.5  (1.3) 

4.2  (0.9) 

and  deciduous-dominated  understories,  mini- 
mal understory  plots  were  characterized  by 
sparse  understory  cover,  all  of  which  was  de- 
ciduous (Table  2).  Plots  containing  deciduous- 
dominated  understory  had  a moderate  amount 
of  understory  cover  but  sparse  balsam  fir  un- 
derstory cover  (2.5%  ± 2.5),  whereas  balsam 
fir-dominated  plots  contained  moderate  under- 
story cover,  of  which  26.9%  ± 2.5  was  balsam 
fir  (Table  2).  Deciduous  stems  typically  fell  in 
the  shortest  height  category:  in  the  12  plots 
with  the  densest  understory  (deciduous-  and 
balsam  fir-dominated),  66%  ± 4 of  the  stems 
were  0.5—1  m tall,  whereas  only  15%  ± 2 and 
19%  ± 3 fell  in  the  1-2  m and  >2  m cate- 
gories, respectively.  In  contrast,  40%  ± 3 of 
the  balsam  firs  were  0.5—1  m tall;  a similar 
percentage  were  1—2  m tall  (41%  ± 3),  and  a 
lower  percentage  (18%  ± 3)  fell  in  the  >2-m 
height  category.  Finally,  there  were  fewer 
large  trees  in  the  twelve  plots  with  dense  un- 


derstory, and  more  small  trees  in  the  balsam 
fir-dominated  plots  (Table  2). 

Birds. — Sixty-seven  BTBWs  were  banded 
in  12  plots  and  3 additional  males  were  re- 
peatedly observed  and  counted,  resulting  in  2- 
12  males  per  36-ha  plot.  The  three  measures 
of  BTBW  abundance  (2002  banding  and  point 
counts,  and  2003  repeat  surveys)  were  highly 
correlated  (r  = 0.90-0.92,  n = 16)  and  the 
results  of  our  analyses  using  each  of  these 
measures  were  identical.  BTBW  abundance 
differed  between  plot  types  (Kruskal-Wallis 
test,  k 3,  ^minimal  — 4,  Wdeciduous  4,  8, 

P < 0.01  for  all  three  abundance  measures). 
On  average,  there  were  1.4  to  3.6  more 
BTBWs  per  36  ha  (low  estimate:  2002  point 
counts;  high  estimate:  2002  banding  data)  on 
plots  averaging  27%  balsam  fir  understory 
cover  than  on  plots  with  sparse  balsam  fir  (Ta- 
ble 2).  The  positive  relationship  between  bal- 
sam fir  and  BTBW  abundance  was  apparent 


Kearns  et  al.  • SONGBIRDS  AND  FOREST  UNDERSTORY 


467 


A 

B 

12- 

• 12- 

• 

• 

• 

io- 

io- 

£ 

“ 8 

8- 

CO 

• • 

• • 

p 6 

• 6- 

• 

0) 

MB 

E 4 

■ #4- 

• ■ 

3 

z 

■ 

■ 

2 

■ 2- 

■ 

O' 

aaa  o- 

AA  A A 

4 -2  0 2 4 

-2-10123 

C 

D 

10- 

■ • io- 

• ■ 

</»  8 ' 

8- 

>x 

cD 

■ 

■ 

CD 

> 6- 

6- 

*4— 

o 

■ • 

• ■ 

CD 

_Q  4- 

■ M • 4- 

• • • ■ 

E 

3 

• 

• 

2 2- 

AA  • • 2‘ 

AA  •• 

A 

A 

o- 

A O' 

A 

4 -2  0 2 4 

-2-10  1 2 3 

PCA1  scores 

PCA2  scores 

FIG.  3.  Relationships  between  Black-throated 
Blue  Warbler  (BTBW)  and  Veery  abundances  and  the 
scores  for  vegetation  characteristics  summarized  by 
principal  components  analysis  (PCA)  for  36-ha  plots 
(n  = 16)  in  the  Hiawatha  National  Forest,  Michigan. 
BTBW  abundance  (based  on  banding  data)  in  2002 
versus  scores  for  (A)  PCA  1 and  (B)  PCA  2;  Veery 
relative  abundance  (based  on  point  counts)  in  2002 
versus  scores  for  (C)  PCA  1 and  (D)  PCA  2.  Triangles 
= minimal  understory  plots,  squares  = deciduous- 
dominated  plots,  and  circles  = balsam  fir-dominated 
plots.  For  the  plots  that  contained  dense  understory  (n 
= 12),  BTBW  abundance  increased  significantly  with 
increasing  values  of  PCA  1 (r  = 0.68),  and  decreased 
significantly  with  increasing  PCA  2 (r  =-0.65).  Veery 
abundance  was  not  linearly  related  to  the  PCA  scores. 


when  BTBW  abundance  was  compared  to  the 
first  principal  component  (r  = 0.68,  n = 16, 
P = 0.004;  Fig.  3A).  Excluding  plots  with 
minimal  understory,  BTBW  abundance 
showed  a negative  association  with  deciduous 
understory  (r  =—0.65,  n = 12 , P = 0.021; 
Fig.  3B).  There  was  no  relationship  between 
BTBW  abundance  and  the  height  of  decidu- 
ous understory,  as  measured  by  the  third  prin- 


ciple component  (r  =—0.25,  n = 16,  P = 
0.35). 

Overall,  74%  (52  of  70)  of  the  BTBWs 
were  older  males  in  2002.  The  BTBW  age  ra- 
tio (%  older)  did  not  differ  significantly  be- 
tween plot  types  (Kruskal- Wallis  test:  k = 2, 
X2  = 0.47,  P = 0.49;  Table  2)  and  showed  no 
pattern  of  relationship  with  any  of  the  prin- 
cipal components  ( n = 12,  P > 0.25  for  all 
three  correlations). 

Veery  relative  abundance  differed  signifi- 
cantly by  plot  type  (Kruskal- Wallis  test:  k = 
3,  x2  ~ 9.12,  P = 0.010)  and  there  were  no 
significant  differences  among  the  plot  types  in 
detection  probabilities  by  distance  or  time 
(distance:  x2  — 3.41,  P = 0.065;  time:  x2  = 
2.14,  P = 0.14;  n = 65).  Veery  abundance 
was  somewhat  greater  in  plots  with  abundant 
deciduous  understory  than  it  was  in  balsam 
fir-dominated  plots  and  there  were  few  Veerys 
in  minimal  understory  plots  (Table  2).  Veery 
abundance  did  not  show  any  relationship  to 
the  three  principle  components  (n  = 16,  P > 
0.20  for  all  three  correlations;  Fig.  3C,  D). 
Thus,  Veery  abundance  increased  with  under- 
story cover,  but  did  not  show  a pattern  with 
respect  to  understory  type  (Table  2). 

DISCUSSION 

In  maple-dominated,  managed  stands  in  the 
Hiawatha  National  Forest  that  experience  high 
winter  deer  densities,  Black-throated  Blue 
Warbler  abundance  was  significantly  greater 
in  areas  with  a dense  understory  of  balsam  fir 
than  in  areas  with  a dense  understory  of  de- 
ciduous trees.  Previous  studies  have  shown 
that  BTBWs  breed  in  both  pure  stands  of 
northern  hardwoods  and  mixed  stands  of  hard- 
wood-conifer, and  exhibit  little  preference  for 
particular  understory  species  if  dense  cover 
exists  (Steele  1993,  Holmes  1994,  DeGraaf  et 
al.  1998,  Steffes  1999).  In  New  Hampshire, 
BTBWs  often  nest  in  hobblebush  ( Viburnum 
alnifolium ),  a shade-tolerant  deciduous  shrub, 
probably  because  it  is  abundant  and  provides 
structural  characteristics  and  branch  heights 
suitable  for  nesting  (Holway  1991,  Holmes 
1994).  Hobblebush  and  shrubs  with  similar 
characteristics  (e.g.,  Rhododendron  spp.)  used 
by  nesting  BTBW  in  other  parts  of  the  spe- 
cies’ range  (Holmes  1994)  do  not  occur  in 
most  Great  Lakes  forests,  and  we  suggest  that 
at  sites  like  ours,  where  most  of  the  understo- 


468 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


ry  comprises  regenerating  tree  species,  balsam 
fir  can  play  a role  similar  to  that  of  hobble- 
bush,  particularly  in  areas  where  deer  brows- 
ing reduces  the  abundance  and  heights  of  de- 
ciduous species.  Therefore,  the  proportion  of 
balsam  fir  in  the  understory,  which  ranged  in 
our  study  from  0-40%  cover  in  plots  with  3— 
53%  total  understory  cover,  can  be  a useful 
tool  for  predicting  the  occurrence  of  BTBWs 
in  managed,  maple-dominated  stands. 

BTBW  densities  in  our  study  area,  which  is 
near  the  western  edge  of  the  species’  range, 
were  low  compared  to  those  in  more  central 
parts  of  their  range  (e.g..  New  Hampshire,  the 
Appalachians);  this  result  agrees  with  esti- 
mates from  Breeding  Bird  Survey  data 
(Holmes  1994)  and  work  by  Graves  (1997b). 
Densities  averaged  0.16  ± 0.02  males/ha  in 
plots  where  BTBWs  were  present  ( n = 12, 
maximum  = 0.3),  versus  0.8-0. 9 males/ha  in 
New  Hampshire  forest  with  a dense  shrub  lay- 
er (Holmes  1994).  The  presence  of  balsam  fir 
and  some  short  (<1  m)  deciduous  understory 
(presumably  present  due  to  snow  protection) 
appears  to  allow  BTBWs,  Veerys,  and  other 
understory-dependent  species  to  persist  in 
these  heavily  deer-impacted  hardwood  forests. 
For  both  bird  species,  the  peak  relative  abun- 
dance values  were  similar  to  high  values  ob- 
served in  Michigan  forests  with  much  less  ev- 
idence of  browsing  by  deer  (Hall  2002).  Our 
results  indicate  that  if  local  forest  managers 
rely  on  studies  of  how  deer  impact  bird  hab- 
itats in  other  regions,  especially  those  with 
hardwood-dominated  understory  (e.g.,  de- 
Calesta  1994,  McShea  and  Rappole  2000), 
they  will  underestimate  habitat  values  for  un- 
derstory-dependent species  at  sites  similar  to 
ours. 

On  balsam  fir-dominated  understory  plots 
with  abundant  BTBWs,  not  only  were  balsam 
fir  stem  densities  greater,  balsam  firs  also  were 
taller  than  other  understory  species  (Table  2). 
In  particular,  many  (41%)  balsam  firs  were  1- 
2 m tall,  whereas  most  (66%)  of  the  understo- 
ry maples  were  <1  m tall  and  only  15%  were 
in  the  1-2  m category;  taller  deciduous  stems 
typically  showed  evidence  of  being  repeatedly 
browsed  (i.e.,  many  short  remnants  of  branch- 
es persisted  along  the  main  stem).  We  suggest 
that  this  difference  in  height  distribution  is 
likely  an  important  driver  of  the  positive 
BTBW  response  to  balsam  fir  at  these  sites. 


In  addition  to  nesting  in  both  balsam  fir  and 
deciduous  cover  <1  m tall,  BTBWs  often 
nested  in  the  lower  branches  of  balsam  firs 
that  were  1-2  m high  (LJK  and  KRH  pers. 
obs.).  Furthermore,  habitats  providing  a great- 
er proportion  of  taller,  more  structurally  com- 
plex saplings  may  provide  more  cover  and 
foraging  substrate  for  recently  fledged  young 
and  adult  BTBWs  (Kolozsvary  2002;  LJK, 
KRH  pers.  obs.)  Although  height  differences 
in  deciduous  understory  explained  a substan- 
tial percentage  of  the  vegetation  variability  in 
our  study  area  (Table  1),  this  was  not  the  focus 
of  our  sampling  design.  Typically,  height  of 
deciduous  understory  is  strongly  linked  to 
both  the  intensity  of  deer  browsing  and  time 
since  the  last  selection  cut  or  forest  thinning, 
and  further  research  focused  on  height  would 
likely  improve  our  understanding  of  habitat 
use  by  BTBWs  in  these  forests. 

Holmes  et  al.  (1996)  found  that  areas  with 
more  understory  had  greater  densities  of 
BTBWs  and  greater  proportions  of  older  birds. 
The  age-ratio  pattern  in  our  plots  indicated  that 
older  birds  preferred  areas  with  more  balsam 
firs;  however,  the  ASY:SY  age  ratio  was  not 
significantly  greater  in  balsam  fir-dominated 
plots,  although  these  plots  had  the  greatest  den- 
sities of  BTBWs.  In  plots  where  we  found 
BTBWs,  74%  were  older  males;  this  is  at  the 
high  end  of  the  range  (50-79%)  observed  by 
Holmes  et  al.  (1996)  in  New  Hampshire,  and 
is  greater  than  ratios  reported  by  Graves 
(1997b)  for  birds  in  northern  Michigan  and 
Ontario  (50-60%).  It  is  possible  that  the  rela- 
tive scarcity  of  yearling  birds  on  our  study  sites 
precluded  detection  of  an  association  between 
age  and  understory  characteristics.  Return  rates 
also  indicated  a preference  for  abundant  balsam 
fir  in  the  understory  (mean  return  rates  were 
26%  in  balsam  fir-dominated  plots  and  1 1 % in 
deciduous-dominated  plots,  a non-significant 
difference),  but  these  values  were  based  on 
only  one  year  of  data  collected  during  a single 
survey  per  site. 

Veery  abundance  did  not  increase  as  balsam 
fir  understory  increased,  but  Veerys  were 
more  abundant  in  plots  with  dense  understory 
than  in  those  with  minimal  understory.  Veerys 
use  a broader  range  of  nest  sites  than  BTBWs, 
including  on  the  ground,  on  downed  branches 
or  logs,  and  in  understory  vegetation  (Mos- 
koff  1995;  KRH  unpubl.  data).  In  a study  by 


Kearns  et  al.  • SONGBIRDS  AND  FOREST  UNDERSTORY 


469 


Heckscher  (2004),  Veerys  generally  built  their 
nests  where  dense  vegetation  was  < 1 .5  m tall 
and  there  was  sparse  vegetation  between  2.5 
and  3 m high;  this  is  consistent  with  our  ob- 
servation that  Veerys  were  more  common  in 
sites  with  dense  understory.  We  observed  that 
Veerys  commonly  nested  in  taller  firs  (2-4 
m),  indicating  that  an  abundance  of  taller  bal- 
sam firs  may  be  important  in  some  stands,  but 
balsam  fir  density  alone  does  not  appear  to 
reliably  predict  the  relative  abundance  of 
Veerys.  The  fact  that  a few  Veerys  were  found 
at  sites  with  little  understory  also  suggests  that 
factors  we  did  not  measure,  such  as  presence 
of  coarse  woody  debris,  may  be  useful  pre- 
dictors of  Veery  abundance  in  Great  Lakes 
hardwood  forests. 

Our  results  indicate  that  stem  density  of 
balsam  fir  understory  predicted  BTBW  abun- 
dance in  deer-browsed  forests  of  northern 
Michigan.  The  density  of  small  trees,  which 
covaried  with  balsam  fir  and  total  understory 
density  (because  both  variables  reflect  time 
since  the  last  thinning  or  selective  harvest), 
also  predicted  BTBW  abundance.  Balsam  fir 
is  a conspicuous  plant  that  is  easily  mapped 
and  quantified  from  aerial  photographs  taken 
in  spring,  which  could  make  it  a useful,  prac- 
tical indicator  of  BTBW  habitat.  Managers 
seeking  to  determine  the  spatial  and  temporal 
pattern  of  harvest  activities  in  hardwood  forest 
(currently,  harvest  methods  for  hardwood 
stands  in  the  Hiawatha  National  Forest  focus 
on  selection  cutting)  could  rank  sites  based  on 
the  prevalence  of  balsam  fir  and  then  stagger 
the  times  at  which  sites  containing  high  den- 
sities of  balsam  fir  would  be  harvested.  We 
recommend  that  small  balsam  firs  be  left  in 
the  understory  when  overstory  trees  are  re- 
moved, especially  in  areas  most  impacted  by 
deer.  Ideally,  these  activities  would  be  paired 
with  avian  population  monitoring  to  verify  the 
effectiveness  of  using  balsam  fir  density  as  an 
indicator  of  BTBW  abundance,  and  to  identify 
relationships  between  other  songbirds  and  this 
plant  species. 

ACKNOWLEDGMENTS 

We  thank  the  following  individuals  and  organiza- 
tions that  helped  make  this  project  possible.  J.  A.  Wit- 
ter, T.  L.  Root,  S.  J.  Sjogren,  and  J.  A.  Craves  were 
essential  in  enabling  us  to  complete  this  project.  Those 
who  helped  with  fieldwork  included  K.  S.  Sheldon,  S. 


K.  Fruchey,  H.  A.  Petrillo,  L.  K.  Peterson,  D.  Y.  Sa- 
saki, G.  R.  Kearns,  L.  A.  Jacobs,  B.  J.  Dantzer,  J.  E. 
Law,  G.  L.  Norwood,  and  J.  J.  Segula.  C.  A.  Geddes, 
S.  J.  Brines,  and  R.  L.  Pickert  assisted  with  GIS  anal- 
yses. M.  A.  Kolozsvary  collaborated  in  grant-writing; 
R.  Bowman,  B.  A.  Hahn,  M.  E.  McPhee,  and  L.  A. 
Riopelle  reviewed  drafts  of  this  manuscript.  Project 
funding  was  provided  by  the  Doris  Duke  Charitable 
Foundation,  U.S.  Fish  and  Wildlife  Service,  U.S.  De- 
partment of  Agriculture  Mclntire-Stennis  Research 
Program,  Rackham  Graduate  School  at  the  University 
of  Michigan,  and  the  U.S.  Department  of  Agriculture 
Forest  Service,  Hiawatha  National  Forest. 

LITERATURE  CITED 

Albert,  D.  A.  1995.  Regional  landscape  ecosystems 
of  Michigan,  Minnesota,  and  Wisconsin:  working 
map  and  classification.  General  Technical  Report 
NC-GTR-178,  USDA  Forest  Service,  North  Cen- 
tral Forest  Experiment  Station,  St.  Paul,  Minne- 
sota. 

Alverson,  W.  S.,  D.  M.  Waller,  and  S.  L.  Solheim. 
1988.  Forests  too  deer:  edge  effects  in  northern 
Wisconsin.  Conservation  Biology  2:348-358. 
Borgmann,  K.  L.,  D.  M.  Waller,  and  T.  P.  Rooney. 
1999.  Does  balsam  fir  {Abies  balsamea ) facilitate 
the  recruitment  of  eastern  hemlock?  American 
Midland  Naturalist  141:391-397. 

Casey,  D.  and  D.  Hein.  1983.  Effects  of  heavy  brows- 
ing on  a bird  community  in  deciduous  forest.  Jour- 
nal of  Wildlife  Management  47:829-836. 

Cote,  S.  D.,  T.  P.  Rooney,  J.-P.  Tremblay,  C.  Dus- 
sault,  and  D.  W.  Waller.  2004.  Ecological  im- 
pacts of  deer  overabundance.  Annual  Review  of 
Ecology  and  Systematics  35:113-47. 
deCalesta,  D.  S.  1994.  Effect  of  white-tailed  deer  on 
songbirds  within  managed  forests  in  Pennsylva- 
nia. Journal  of  Wildlife  Management  58:711-718. 
DeGraaf,  R.  M.,  J.  B.  Hestbeck,  and  M.  Yamasaki. 
1998.  Associations  between  breeding  bird  abun- 
dance and  stand  structure  in  the  White  Mountains, 
New  Hampshire  and  Maine,  USA.  Forest  Ecology 
and  Management  103:217-233. 

Donovan,  T.  M.,  C.  J.  Beardmore,  D.  N.  Bonter,  J. 
D.  Brawn,  R.  J.  Cooper,  J.  A.  Fitzgerald,  R. 
Ford,  S.  A.  Gauthreaux,  T.  L.  George,  W.  C. 
Hunter,  et  al.  2002.  Priority  research  needs  for 
the  conservation  of  Neotropical  migrant  landbirds. 
Journal  of  Field  Ornithology  73:329-39. 
Environmental  Systems  Research  Institute.  2002. 
ArcView  GIS,  ver.  3.3.  Environmental  Systems 
Research  Institute,  Redlands,  California. 

Graves,  G.  R.  1997a.  Age  determination  of  free-living 
male  Black-throated  Blue  Warblers  during  the 
breeding  season.  Journal  of  Field  Ornithology  68: 
443-449. 

Graves,  G.  R.  1997b.  Geographic  dines  of  age  ratios 
of  Black- throated  Blue  Warblers  ( Dendroica  cae- 
rulescens ).  Ecology  78:2524-2531. 

Hall,  K.  R.  2002.  An  assessment  of  habitat  quality  of 
heavily-  and  less-browsed  Michigan  forests  for  a 


470 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


shrub-nesting  songbird.  Ph.D.  dissertation.  Uni- 
versity of  Michigan,  Ann  Arbor. 

Heckscher,  C.  M.  2004.  Veery  nest  sites  in  Mid-At- 
lantic Piedmont  forest:  vegetative  physiognomy 
and  use  of  alien  shrubs.  American  Midland  Nat- 
uralist 151:326-337. 

Holmes,  R.  T.  1994.  Black- throated  Blue  Warbler 
(Dendroica  caerulescens).  The  Birds  of  North 
America,  no.  87. 

Holmes,  R.  T.,  P.  P.  Marra,  and  T.  W.  Sherry.  1996. 
Habitat-specific  demography  of  breeding  Black- 
throated  Blue  Warblers  ( Dendroica  caerulescens ): 
implications  for  population  dynamics.  Journal  of 
Animal  Ecology  65:183-195. 

Holway,  D.  A.  1991.  Nest-site  selection  and  the  im- 
portance of  nest  concealment  in  the  Black-throat- 
ed Blue  Warbler.  Condor  93:575-581. 

Hunt,  P.  1996.  Habitat  selection  by  American  Red- 
starts along  a successional  gradient  in  northern 
hardwoods  forest:  evaluation  of  habitat  quality. 
Auk  113:875-888. 

Insightful  Corporation.  2002.  S-Plus  for  Windows, 
Student  Version,  ver.  6.1.  Insightful  Corporation, 
Seattle,  Washington. 

James,  L C.  and  H.  H.  Shugart,  Jr.  1970.  A quanti- 
tative method  of  habitat  description.  Audubon 
Field  Notes  24:727-736. 

Kolozsvary,  M.  L.  2002.  Foraging  strategies  of 
Black-throated  Blue  Warblers:  effects  of  habitat 
alteration  due  to  browsing  by  white-tailed  deer. 
M.Sc.  thesis.  University  of  Michigan,  Ann  Arbor. 

Kraft,  L.  S.,  T.  R.  Crow,  D.  S.  Buckley,  E.  A. 
Nauertz,  and  J.  C.  Zasada.  2004.  Effects  of  har- 
vesting and  deer  browsing  on  attributes  of  under- 
story plants  in  northern  hardwood  forests.  Upper 
Michigan,  USA.  Forest  Ecology  and  Management 
199:219-30. 

Martin,  T.  E.  1992.  Breeding  productivity  consider- 
ations: what  are  the  appropriate  habitat  features 
for  management?  Pages  455-73  in  Ecology  and 
conservation  of  Neotropical  migrant  landbirds  (J. 
M.  Hagan  and  D.  W.  Johnston,  Eds.).  Smithsonian 
Institution  Press,  Washington,  D.C. 

Matteson,  S.,  A.  Paulios,  G.  Bartelt,  G.  Butcher, 
D.  Sample,  J.  Fitzgerald,  G.  Niemi,  J.  Hanowski, 
and  R.  Howe.  In  press.  Partners  in  Flight  Bird 
Conservation  Plan  for  the  Boreal  Hardwood  Tran- 
sition (Bird  Conservation  Region  12,  Physio- 
graphic Area  20).  Wisconsin  Department  of  Nat- 
ural Resources  in  cooperation  with  Partners  in 
Flight,  Madison,  Wisconsin. 

McShea,  W.  J.  and  J.  H.  Rappole.  2000.  Managing 
the  abundance  and  diversity  of  breeding  bird  pop- 
ulations through  manipulation  of  deer  populations. 
Conservation  Biology  14:1 161-1  170. 

Moskoff,  W.  1995.  Veery  ( Catharus  fuscescens).  The 
Birds  of  North  America,  no.  142. 


Mueller-Dombois,  D.  and  H.  Ellenberg.  1974.  Aims 
and  methods  of  vegetation  ecology.  John  Wiley 
and  Sons,  New  York. 

Pyle,  P.  1997.  Identification  guide  to  North  American 
birds.  Slate  Creek  Press,  Bolinas,  California. 

Ralph,  C.  J.,  G.  R.  Geupel,  P.  Pyle,  T.  E.  Martin, 
and  D.  F.  Desante.  1993.  Handbook  of  field 
methods  for  monitoring  landbirds.  General  Tech- 
nical Report  PSW-GTR-144,  USDA  Forest  Ser- 
vice, Pacific  Southwest  Research  Station,  Albany, 
California. 

Rodenhouse,  N.  L.,  T.  S.  Sillett,  P.  J.  Doran,  and  R. 
T.  Holmes.  2003.  Multiple  density-dependence 
mechanisms  regulate  a migratory  bird  population 
during  the  breeding  season.  Proceedings  of  the 
Royal  Society  of  London  B 270:2105-2110. 

Rooney,  T.  P.  and  D.  M.  Waller.  2003.  Direct  and 
indirect  effects  of  white-tailed  deer  in  forest  eco- 
systems. Forest  Ecology  and  Management  181: 
165-76. 

Steele,  B.  B.  1992.  Habitat  selection  by  breeding 
Black-throated  Blue  Warblers  at  two  spatial 
scales.  Ornis  Scandinavica  23:33-42. 

Steele,  B.  B.  1993.  Selection  of  foraging  and  nesting 
sites  by  Black-throated  Blue  Warblers:  their  rela- 
tive influence  on  habitat  choice.  Condor  95:568- 
589. 

Steffes,  M.  W.  1999.  The  Black- throated  Blue  War- 
bler along  the  Superior  hiking  trail  in  northeastern 
Minnesota.  Loon  71:5-11. 

The  Nature  Conservancy.  2000.  Toward  a new  con- 
servation vision  for  the  Great  Lakes  region:  a sec- 
ond iteration.  The  Nature  Conservancy,  Great 
Lakes  Program,  Chicago,  Illinois. 

U.S.  Fish  and  Wildlife  Service.  2002.  Birds  of  con- 
servation concern  2002.  U.S.  Fish  and  Wildlife 
Service,  Division  of  Migratory  Bird  Management, 
Arlington,  Virginia. 

Van  Deelen,  T.  R.,  H.  Campa,  M.  Hamady,  and  J.  B. 
Haufler.  1998.  Migration  and  seasonal  range  dy- 
namics of  deer  using  adjacent  deeryards  in  north- 
ern Michigan.  Journal  of  Wildlife  Management 
62:205-13. 

Winnett-Murray,  K.  1991.  Veery  {Catharus  fusces- 
cens). Pages  350-351  in  The  atlas  of  breeding 
birds  of  Michigan  (R.  Brewer,  G.  A.  McPeek,  and 
R.  J.  Adams,  Jr.,  Eds.).  Michigan  State  University 
Press,  East  Lansing. 

Wolf,  A.  T„  R.  W.  Howe,  and  G.  J.  Davis.  1995. 
Detectability  of  forest  birds  from  stationary  points 
in  northern  Wisconsin.  Pages  19-23  in  Monitor- 
ing bird  populations  by  point  counts  (C.  J.  Ralph, 
J.  R.  Sauer,  and  S.  Droege,  Eds.).  General  Tech- 
nical Report  PSW-GTR-149,  USDA  Forest  Ser- 
vice, Pacific  Southwest  Research  Station,  Albany, 
California. 

Zar,  J.  H.  1999.  Biostatistical  analysis,  4th  ed.  Prentice 
Hall,  Upper  Saddle  River,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):47 1 — 477,  2006 


SOARING  AND  GLIDING  FLIGHT  OF  MIGRATING 
BROAD-WINGED  HAWKS:  BEHAVIOR  IN  THE  NEARCTIC  AND 

NEOTROPICS  COMPARED 

VINCENT  CAREAU,14  JEAN-FRAN^OIS  THERRIEN,15  PABLO  PORRAS,2 
DON  THOMAS,1 2 3 4 5 6  AND  KEITH  BILDSTEIN36 


ABSTRACT. — We  compared  migrating  behavior  of  Broad-winged  Hawks  ( Buteo  platypterus)  at  two  sites 
along  their  migration  corridor:  Hawk  Mountain  Sanctuary  in  eastern  Pennsylvania  and  the  Kekoldi  Indigenous 
Reserve  in  Limon,  Costa  Rica.  We  counted  the  number  of  times  focal  birds  intermittently  flapped  their  wings 
and  recorded  the  general  flight  type  (straight-line  soaring  and  gliding  on  flexed  wings  versus  circle-soaring  on 
fully  extended  wings).  We  used  a logistic  model  to  evaluate  which  conditions  were  good  for  soaring  by  calcu- 
lating the  probability  of  occurrence  or  absence  of  wing  flaps.  Considering  that  even  intermittent  flapping  is 
energetically  more  expensive  than  pure  soaring  and  gliding  flight,  we  restricted  a second  analysis  to  birds  that 
flapped  during  observations,  and  used  the  number  of  flaps  to  evaluate  factors  influencing  the  cost  of  migration. 
Both  the  occurrence  and  extent  of  flapping  were  greater  in  Pennsylvania  than  in  Costa  Rica,  and  during  periods 
of  straight-line  soaring  and  gliding  flight  compared  with  circle-soaring.  At  both  sites,  flapping  was  more  likely 
during  rainy  weather  and  early  and  late  in  the  day  compared  with  the  middle  of  the  day.  Birds  in  Costa  Rica 
flew  in  larger  flocks  than  those  in  Pennsylvania,  and  birds  flying  in  large  flocks  flapped  less  than  those  flying 
alone  or  in  smaller  flocks.  In  Pennsylvania,  but  not  in  Costa  Rica,  the  number  of  flaps  was  higher  when  skies 
were  overcast  than  when  skies  were  clear  or  partly  cloudy.  In  Costa  Rica,  but  not  in  Pennsylvania,  flapping 
decreased  as  temperature  increased.  Our  results  indicate  that  birds  migrating  in  large  flocks  do  so  more  efficiently 
than  those  flying  alone  and  in  smaller  flocks,  and  that  overall,  soaring  conditions  are  better  in  Costa  Rica  than 
in  Pennsylvania.  We  discuss  how  differences  in  instantaneous  migration  costs  at  the  two  sites  may  shift  the 
species’  migration  strategy  from  one  of  time  minimization  in  Pennsylvania  to  one  of  energy  minimization  in 
Costa  Rica.  Received  15  November  2005,  accepted  8 July  2006. 


Each  year,  more  than  one  million  Broad- 
winged Hawks  ( Buteo  platypterus ) make  a 
round-trip  migration  of  6,000-10,000  km 
along  the  Mesoamerican  Land  Corridor  when 
traveling  between  their  North  American 
breeding  grounds  and  wintering  areas  in  Cen- 
tral and  South  America  (Bildstein  and  Zalles 
2001).  Because  the  power  requirement  for 
continuous,  flapping  flight  has  an  allometric 


1 Dept,  de  biologie,  Univ.  de  Sherbrooke,  Sher- 
brooke, QC  J1K  2R1,  Canada. 

2 Asociacion  ANAI,  Costado  Norte  de  Cancha  de 
Futbol  del  Colegio  Monterrey,  Vargas  Araya,  San  Pe- 
dro, San  Jose,  Costa  Rica. 

3 Acopian  Center  for  Conservation  Learning,  Hawk 
Mountain  Sanctuary,  410  Summer  Valley  Road,  Or- 
wigsburg,  PA  17961,  USA. 

4 Current  address:  Dept,  des  sciences  biologiques, 
Univ.  du  Quebec  a Montreal,  141  Av.  President-Ken- 
nedy,  C.P.  8888  succursale  Centre-ville,  Montreal,  QC 
H3C  3P8,  Canada;  et  Centre  d’etudes  nordiques,  Univ. 
Laval,  QC  G1K  7P4,  Canada. 

5 Current  address:  Dept,  de  biologie  et  Centre 
d’etudes  nordiques,  Univ.  Laval,  QC  G1K  7P4,  Can- 
ada. 

6 Corresponding  author;  e-mail: 
bildstein@hawkmtn.org 


mass  exponent  of  1.17  (Pennycuick  1972), 
large-bodied  migrants  are  penalized  compared 
with  small-bodied  migrants  in  that  they  need 
a disproportionately  larger  fat  reserve  to  ac- 
complish a non-stop,  powered-flight  migration 
of  a given  distance.  As  such,  long-distance 
migration  represents  a potentially  acute  ener- 
getic challenge  for  large-bodied  migrants  such 
as  Broad-winged  Hawks  (265-560  g;  Good- 
rich et  al.  1996).  In  fact,  measures  of  fat  re- 
serves at  the  onset  of  migration  suggest  that 
Broad-winged  Hawks  do  not  carry  the  fuel 
supply  needed  to  sustain  powered  flight  be- 
tween their  breeding  and  wintering  grounds 
without  also  feeding  en  route  (Bildstein  1999). 

There  are  two  possible  solutions  to  this  en- 
ergetic challenge.  First,  large-bodied  migrants 
may  complete  their  migration  in  stages,  paus- 
ing periodically  to  feed  and  replenish  fat  re- 
serves en  route.  Second,  if  their  flight  me- 
chanics permit,  they  may  significantly  reduce 
the  energetic  costs  associated  with  powered 
flight  by  relying  instead  on  soaring  and  glid- 
ing flight.  Although  ducks,  geese,  and  many 
shorebirds  and  landbirds  exploit  the  first  strat- 
egy (Moore  2000),  Broad-winged  Hawks  do 


471 


472 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


not  feed  substantially  when  migrating,  partic- 
ularly in  the  tropics  (Bildstein  1999),  possibly 
because  their  sit-and-wait  foraging  strategy 
does  not  lend  itself  well  to  the  high  capture 
rates  needed  for  the  rapid  accumulation  of  fat 
reserves.  Instead,  they  rely  heavily  on  gliding 
and  soaring  flight  to  complete  their  long-dis- 
tance movements  (Smith  et  al.  1986).  Because 
basal  metabolic  rate  (BMR)  increases  with 
mass  by  an  allometric  exponent  of  approxi- 
mately 0.75,  soaring  and  gliding  flight  become 
increasingly  cost-efficient  as  mass  increases 
(Hedenstrom  1993).  Indeed,  it  has  been  esti- 
mated that  100  g of  fat  would  fuel  powered 
flight  for  only  about  5 days  in  Broad-winged 
Hawks,  but  it  would  sustain  soaring  flight  in 
the  species  for  an  estimated  20  days  (Smith  et 
al.  1986). 

Soaring  flight  is  based  on  the  conversion  of 
the  energy  in  atmospheric  air  currents  into  pri- 
marily potential  energy  (Pennycuick  1972).  In 
North  America,  soaring  Broad-winged  Hawks 
gain  altitude  while  circling  in  thermals  and  ri- 
ding deflection  updrafts  with  their  wings  and 
tails  fully  spread,  and  then  gliding  on  flexed 
wings  along  their  preferred  direction  of  travel 
as  they  convert  the  altitude  gained  into  dis- 
tance traveled  while  seeking  the  next  thermal 
or  updraft  along  their  migratory  route  (Kerlin- 
ger  1989).  In  Central  America,  where  the  spe- 
cies also  alternately  soars  and  glides  among 
small  thermals,  it  also  straight-line  soars  and 
glides  in  the  much  larger  tropical  thermals  and 
“thermal  streets”  ( sensu  Smith  1985)  found 
in  that  region. 

Because  the  distribution,  abundance,  and 
strength  of  thermals  and  updrafts  are  affected 
by  topography,  vegetation  cover,  vertical  tem- 
perature gradient  of  the  atmosphere,  and  in- 
tensity of  solar  radiation,  soaring  flight  im- 
poses constraints  on  the  spatial  and  temporal 
organization  of  migration  (Kerlinger  1989). 
Soaring  migrants  are  able  to  migrate  efficient- 
ly only  when  sufficient  solar  radiation  and  low 
cloud  cover  favor  the  production  of  thermals, 
thus  concentrating  individuals  in  specific  sea- 
sonal and  daily  time  windows.  Also,  in  the 
temperate  zone,  thermals  often  occur  in  small, 
localized  pockets,  which  sometimes  force 
soaring  birds  to  fly  close  to  each  other  when 
using  the  same  thermal.  This  has  led  some  to 
suggest  that  flocking  behavior  occurs  passive- 
ly among  soaring  migrants,  as  limited  spatial 


and  temporal  windows  of  soaring  opportunity 
act  to  group  the  birds  during  their  migrations 
(Smith  1985).  Alternatively,  others  have  spec- 
ulated that  soaring  migrants,  such  as  Broad- 
winged Hawks,  actively  form  groups  because 
doing  so  allows  them  to  gather  information 
(e.g.,  Danchin  et  al.  2004)  about  the  location 
and  strength  of  individual  thermals  passively 
provided  by  individuals  traveling  with  them 
(Kerlinger  1989). 

As  Broad-winged  Hawks  travel  south  in  au- 
tumn, it  is  likely  that  they  adjust  their  flight 
behavior  to  accommodate  changes  in  the 
abundance  and  strength  of  the  thermals  they 
encounter.  At  the  onset  of  migration  in  late 
summer  in  the  temperate  zone,  the  sun’s 
height  in  the  sky  and  overall  solar  intensity 
begin  to  decline  (Bildstein  1999);  the  stron- 
gest and  greatest  abundance  of  thermals  tends 
to  occur  episodically  during  the  several  days 
of  fair  weather  that  typically  follow  the  pas- 
sage of  cold  fronts  (Allen  et  al.  1996).  Farther 
south  in  the  tropics,  the  sun’s  height  in  the  sky 
and  solar  intensity  remain  relatively  more 
constant  during  the  migration  period  and  ther- 
mal strength  appears  to  vary  primarily  as  a 
function  of  local  cloud  cover  (Smith  1980). 

It  has  been  suggested  that  the  movements 
of  soaring  migrants  are  less  constrained  in  the 
tropics  than  in  the  temperate  zone  and  that 
their  flight  patterns  differ  in  the  two  regions 
(Bildstein  and  Saborio  2000).  For  example, 
Fuller  et  al.  (1998)  reported  that  the  migration 
speed  of  satellite-tracked  Swainson’s  Hawks 
( Buteo  swainsoni)  soaring  and  gliding  be- 
tween breeding  grounds  in  western  North 
America  and  wintering  areas  in  Argentina  was 
42%  greater  in  the  tropics  than  in  the  temper- 
ate zone.  Here,  we  compare  the  flight  behavior 
of  Broad-winged  Hawks  at  temperate  and 
tropical  sites  to  test  three  main  predictions:  (1) 
because  soaring  conditions  are  better  in  the 
tropics,  birds  would  begin  flying  earlier  in  the 
day  and  flap  less  there  than  in  the  temperate 
zone;  (2)  birds  within  a given  site  would  flap 
less  at  higher  temperatures  and  less  cloud  cov- 
er; and  (3)  birds  would  flap  less  when  trav- 
eling in  large  flocks  than  when  traveling  alone 
or  in  smaller  flocks. 

METHODS 

We  observed  migrating  Broad-winged 
Hawks  in  the  temperate  zone  at  Hawk  Moun- 


Careau  et  al.  • BROAD-WINGED  HAWK  MIGRATION 


473 


tain  Sanctuary  in  the  Central  Appalachian 
Mountains  of  eastern  Pennsylvania  (40°  58'  N, 
74°  59'  W;  464  m ASL)  on  10-28  September 
2002,  during  peak  passage  at  that  site.  Hawk 
Mountain  straddles  the  300-km-long  Kittatin- 
ny  Ridge,  which  acts  as  a leading  line  for  rap- 
tor migrants  in  the  region  (Bildstein  1999).  In 
the  tropics,  we  observed  migrating  Broad- 
winged Hawks  from  a 10-m  tower  at  the  Kek- 
oldi  Indigenous  Reserve,  southeast  of  Puerto 
Viejo  in  Talamanca,  Limon,  Costa  Rica  (9° 
38'  N,  82°  47'  W;  200  m ASL)  on  2-19  Oc- 
tober 2002,  during  peak  passage  at  that  site 
(Porras-Penaranda  et  al.  2004).  The  Caribbean 
Sea,  ~2  km  to  the  north,  and  the  Talamanca 
Mountains,  ~5  km  to  the  south,  funnel  birds 
through  the  region’s  coastal  lowlands,  making 
this  area  one  of  several  major  concentration 
points  along  the  Mesoamerican  Land  Corridor 
(Bildstein  and  Zalles  2001). 

We  used  7 X 35  binoculars  and  a 20-60 X 
zoom  telescope  to  watch  birds  at  each  site  be- 
tween sunrise  and  18:00  EST.  Individual  ob- 
servations were  made  on  a focal  individual 
during  a 30-sec  sample  period  beginning  as 
soon  as  the  bird  was  identified  as  a Broad- 
winged Hawk.  The  30-sec  length  represented 
a fair  trade-off  between  gaining  a representa- 
tive record  of  flight  behavior  and  losing  con- 
tact with  the  focal  bird  before  the  observation 
period  was  completed.  During  our  observa- 
tions, we  recorded  the  number  of  seconds  the 
focal  individual  spent  (1)  circle-soaring  in  an 
individual  thermal  and  (2)  straight-line  soar- 
ing and  gliding  between  thermals  and  along 
thermal  streets.  When  circle-soaring,  birds  as- 
cended thermals  on  fully  outstretched  wings 
with  their  tails  fanned.  When  straight-line 
soaring  and  gliding,  birds  flew  on  semi-flexed 
wings  with  their  wingtips  and  tails  partly  fold- 
ed. We  also  recorded  the  number  of  flaps  (i.e., 
individual  wing  beats)  and  used  it  as  a mea- 
sure of  powered  flight. 

We  determined  flock  size  by  counting  or 
estimating  the  number  of  birds  soaring  within 
the  same  thermal  or  soaring  and  gliding  in  the 
same  flight  line  as  the  focal  bird.  In  Pennsyl- 
vania, flocks  were  composed  of  only  Broad- 
winged Hawks.  In  Costa  Rica,  however. 
Broad-winged  Hawks  sometimes  commingled 
with  Swainson’s  Hawks  and  Turkey  Vultures 
(Cathartes  aura ) in  mixed-species  flocks.  We 
noted  temperature  and  cloud  cover  (clear  and 


partly  cloudy  skies  versus  complete  overcast) 
at  hourly  intervals.  We  also  noted  time  of  day 
as  time  after  sunrise  (06:45  EST  in  Pennsyl- 
vania and  05:25  CST  in  Costa  Rica)  and  then 
divided  the  day  into  three  periods  (early,  mid, 
and  late)  to  simplify  analyses.  At  both  sites, 
the  early  period  included  the  first  4 hr  after 
sunrise,  the  mid-period  the  next  4 hr,  and  the 
late  period  the  next  3 hr.  We  did  not  record 
flight  behavior  later  in  the  day. 

We  performed  all  analyses  using  the  JMP 
5.0.1  statistical  package  (SAS  Institute,  Inc. 
2002).  We  used  non-parametric  Mann-Whit- 
ney  U- tests  to  compare  mean  onset  of  activity 
and  flock  size  between  Pennsylvania  and  Cos- 
ta Rica.  To  allow  comparisons  between  soar- 
ing and  gliding  phases  of  flight,  we  restricted 
our  analyses  to  30-sec  sequences  in  which  the 
focal  bird  remained  in  one  flight  phase  (soar- 
ing or  gliding).  We  conducted  two  general 
analyses.  The  first  examined  which  conditions 
enabled  soaring  and  gliding  flight  without 
flaps.  The  second  examined  factors  that  influ- 
enced the  extent  of  flapping  when  it  did  occur. 

For  the  first  analysis,  we  divided  observa- 
tions into  those  during  which  the  bird  did  or 
did  not  flap.  We  ran  a stepwise  logistic  re- 
gression that  included  all  independent  vari- 
ables (site,  flight  phase,  flock  size,  tempera- 
ture, and  cloud  cover)  and  two-way  interac- 
tions. The  odds  ratio  (OR)  measures  how  the 
fitted  probability  is  multiplied  as  the  regressor 
changes  from  its  minimum  to  its  maximum  for 
continuous  data,  or  from  one  category  to  the 
other  for  nominal  data  (Hosmer  and  Leme- 
show  1989).  We  used  the  log-likelihood  ratio 
(L-R)  test  to  determine  P-values.  The  second 
analysis  was  restricted  to  birds  that  flapped  at 
least  once  while  we  were  observing  them.  For 
each  site,  we  conducted  an  ANCOVA  on  the 
number  (logl0-transformed)  of  flaps,  according 
to  the  flock  size,  flight  phase,  temperature,  and 
cloud  cover.  Data  are  presented  as  means  ± 
SE. 

RESULTS 

We  made  1,537  30-sec  observations  of 
Broad-winged  Hawks  during  1 3 days  in  Penn- 
sylvania and  2,103  observations  during  15 
days  in  Costa  Rica.  In  Costa  Rica,  flocks 
ranged  in  size  from  2 to  >1,000  individuals 
(mean  = 427  ±10;  median  = 140).  In  Penn- 
sylvania, flock  size  never  exceeded  350  indi- 


474 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol  118 . No.  4 . December  2006 


9 


Early  Mid-day  Late  Rainy  weather 

Time  of  day  and  weather 


FIG.  1.  Mean  numbers  of  wing  flaps  per  30-sec  observation  period  (±SE)  in  relation  to  time  of  day  and 
rain  condition  in  Pennsylvania.  USA.  10-28  September  2002.  and  in  Talamanca.  Costa  Rica.  2-19  October 
2002.  Numbers  above  the  error  bars  represent  sample  sizes. 


viduals  (mean  = 26  ± 1;  median  = 10).  Over- 
all, flock  size  was  significantly  greater  in  Cos- 
ta Rica  than  in  Pennsylvania  (U  = 1177.6,  P 
< 0.001).  The  first  migrant  of  the  day  was 
sighted  almost  one  hour  later  in  Pennsylvania 
than  in  Costa  Rica  (198  ± 53  min  after  sunrise 
versus  150  ± 11  min  after  sunrise,  U = 6.32, 
P = 0.022),  and  the  first  individuals  sighted 
each  day  were  more  likely  to  flap  in  Pennsyl- 
vania than  in  Costa  Rica  (35%  versus  16%, 
L-R  x2  = 162.7,  P < 0.001). 


TABLE  1.  Results  of  the  logistics  model  for  the 
occurrence  of  flapping  flight  among  Broad-winged 
Hawks  in  Pennsylvania.  USA.  10-28  September  2002. 
and  in  Talamanca.  Costa  Rica.  2-19  October  2002.  by 
temperature  (°C),  flock  size,  flight  type  (circle  or 
straight-line  soaring),  and  cloud  cover  (overcast  or 
not).  The  log-likelihood  (L-R)  x2  and  P- value  are 
shown.  Sample  size  is  2,153. 


Term 

df 

L-R  x2 

p 

Site 

1 

10.24 

<0.001 

Temperature 

1 

111.56 

<0.001 

Site  X temperature 

1 

55.13 

<0.001 

Flock  size 

1 

16.76 

<0.001 

Flight  type 

1 

77.63 

<0.001 

Cloud  cover 

1 

15.24 

<0.001 

At  both  sites,  birds  were  more  likely  to  flap 
early  and  late  in  the  day  than  at  mid-day 
(Pennsylvania:  L-R  x2=  67.1.  P < 0.001;  Cos- 
ta Rica:  L-R  x2  — 68.6.  P < 0.001;  Fig.  1). 
Flapping  was  greater  during  rainy  periods  at 
both  sites,  but  significantly  so  only  in  Costa 
Rica  (Pennsylvannia:  L-R  x2  = 3.84.  P = 
0.051:  Costa  Rica:  L-R  x2  = 78.6.  P < 0.001). 
To  account  for  these  effects,  we  excluded 
from  the  analyses  that  follow  any  observa- 
tions made  early  and  late  in  the  day  and  dur- 
ing rainy  weather. 

The  logistic  model  indicated  which  condi- 
tions favored  soaring  flight  (Table  1)  and  the 
ANCOVA  identified  which  factors  determined 
the  extent  of  powered  flight  when  it  occurred 
(Table  2).  Both  extent  and  probability  of  flap- 
ping were  greater  during  straight-line  soaring 
and  gliding  than  during  circle-soaring  (Fig.  2; 
OR  = 0.3).  The  overall  flapping  probability 
was  lower  when  birds  flew  in  larger  flocks 
than  in  smaller  flocks  or  alone  (OR  = 2.8). 
There  was  no  significant  difference  between 
flapping  rates  in  Pennsylvania  and  Costa  Rica 
when  birds  flew  in  flocks  of  up  to  50  birds 
(L-R  x2  = 3.75,  n = 1.038,  df  = 1,  P = 
0.053);  however,  when  birds  were  in  flocks 


Careau  et  al.  • BROAD-WINGED  HAWK  MIGRATION 


475 


TABLE  2.  Comparisons  of  factors  influencing  the  numbers  of  flaps  per  observation  of  Broad-winged  Hawks 
in  Pennsylvania,  USA,  10-28  September  2002,  and  in  Talamanca,  Costa  Rica,  2-19  October  2002.  The  AN- 
COVA  was  restricted  to  birds  that  flapped  at  least  once  during  the  observation. 


Pennsylvania  ( n = 208) 


Costa  Rica  ( n = 156) 


df 

/•"-ratio 

p 

/r-ratio 

p 

Flock 

1 

0.10 

0.76 

9.17 

0.003 

Flight  type 

1 

14.20 

<0.001 

18.63 

<0.001 

Temperature 

1 

1.42 

0.24 

8.34 

0.005 

Cloud  cover 

4 

3.29 

0.012 

1.70 

0.16 

that  ranged  in  size  from  5 1 to  350  birds,  flap- 
ping probability  was  significantly  lower  in 
Costa  Rica  than  it  was  in  Pennsylvania  (L-R 
X2  = 10.25,  n = 468,  df  = 1,  P = 0.001). 
More  than  94%  of  the  Broad-winged  Hawks 
seen  in  Costa  Rica  were  flying  in  flocks  of 
>50,  and  flapping  was  far  more  likely  in 
Pennsylvania  than  it  was  in  Costa  Rica  (OR 
= 1.87;  Fig.  2).  Moreover,  the  number  of  flaps 
decreased  with  flock  size  in  Costa  Rica,  but 
not  in  Pennsylvania  (Table  2). 

Overall,  the  probability  of  flapping  was 
greater  during  periods  of  complete  overcast 
than  it  was  when  cloud  cover  was  <75%  (OR 
= 1.7);  however,  cloud  cover  had  an  effect  on 
the  number  of  flaps  only  in  Pennsylvania  (Ta- 
ble 2).  Although  probability  of  flapping  de- 
creased as  temperature  increased  (minimum 
temperature  = 15°  C,  maximum  temperature 
= 31°  C,  OR  = 574.1);  the  relationship  was 
significantly  weaker  in  Pennsylvania  than  in 


449 


Circle  soaring  Straight-line  soaring  and  gliding 

Flight  type 


FIG.  2.  Mean  number  of  wing  flaps  per  30-sec  ob- 
servation period  (±  SE)  made  by  Broad-winged  Hawks 
in  circle  soaring  or  straight-line  soaring  and  gliding  flight 
in  Pennsylvania,  USA,  10-28  September  2002,  and  in 
Talamanca,  Costa  Rica,  2-19  October  2002.  Numbers 
above  the  error  bars  represent  sample  sizes. 


Costa  Rica  (site  X temperature  interaction 
term,  OR  = 0.03).  Accordingly,  temperature 
had  an  effect  on  the  number  of  flaps  in  Costa 
Rica  but  not  in  Pennsylvania  (Table  2). 

DISCUSSION 

Since  Huffaker  (1897)  first  provided  evi- 
dence of  the  existence  of  thermal  updrafts 
based  on  observations  of  soaring  birds,  many 
studies  have  shown  that  avian  flight  can  be 
used  to  gather  information  on  meteorological 
processes  (Shannon  et  al.  2002).  We  present 
our  data  as  a biological  method  for  measuring 
soaring  conditions  for  Broad-winged  Hawks 
traveling  between  the  temperate  zone  and  the 
tropics  during  southbound  migration  in  au- 
tumn, and  we  offer  a preliminary  indication 
of  how  differences  in  soaring  conditions  affect 
the  efficacy  of  migratory  flight  in  the  species. 

In  general,  our  observations  confirm  the 
flight  behavior  of  soaring  migrants  document- 
ed elsewhere  (Kerlinger  and  Gauthreaux 
1985,  Spaar  and  Bruderer  1997,  Spaar  et  al. 
1998).  For  example,  as  temperatures  and  solar 
radiation  increase  each  morning,  birds  rely 
less  on  flapping  flight  and  more  on  soaring 
and  gliding  flight,  presumably  to  reduce  the 
energetic  costs  of  travel  by  taking  advantage 
of  the  stronger  mid-day  thermals. 

The  negative  correlation  between  flapping 
rates  and  flock  size  suggests  that  Broad- 
winged Hawks  use  information  available  in 
flocks  to  increase  their  flight  efficiency  (Ker- 
linger 1989).  That  said,  although  smaller  flock 
sizes  and  higher  flapping  rates  in  Pennsylva- 
nia were  probably  due  at  least  in  part  to  this 
effect,  smaller  and  weaker  thermals  in  Penn- 
sylvania also  may  have  contributed  to  a great- 
er likelihood  of  flapping  at  the  site. 

We  suggest  that  migrating  Broad-winged 
Hawks  do  not  pursue  a pure  soaring  and  glid- 


476 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


ing  strategy  throughout  their  migration  be- 
cause they  are  constrained  from  doing  so  in 
two  ways.  First,  they  cannot  soar  when  ver- 
tical airspeeds  within  thermals  fail  to  reach  a 
critical  threshold  value,  and  second,  they  can- 
not glide  efficiently  when  inter-thermal  dis- 
tances exceed  their  maximum  gliding  range 
(Kerlinger  1989).  Our  data  indicate  that 
Broad-winged  Hawks  respond  to  these  con- 
straints by  using  powered  flight  preferentially 
during  straight-line  soaring  and  gliding  flight 
and  secondarily  when  circle-soaring.  This  tac- 
tic also  has  been  observed  in  migrating  Com- 
mon Cranes  ( Grus  grus\  Pennycuick  et  al. 
1979),  as  well  as  in  other  raptor  species.  By 
stretching  inter-thermal  glides  with  flapping 
flight,  birds  increase  the  distance  realized, 
thereby  extending  their  ability  to  reach  and 
use  the  next  thermal  (Pennycuick  1998).  Sec- 
ond, under  certain  circumstances,  soaring  and 
gliding  can  slow  travel  compared  with  flap- 
ping flight,  particularly  when  the  birds  are 
soaring  in  small  thermals.  Indeed,  migrants 
are  likely  to  pursue  a pure  soaring  strategy 
only  when  they  have  time  to  wait  for  the  prop- 
er conditions  and  are  able  to  move  slowly 
along  the  migration  corridor.  For  Broad- 
winged Hawks,  time  limitations  may  be  more 
important  in  Pennsylvania  than  in  Costa  Rica, 
because  solar  intensity  and  photoperiod  de- 
crease rapidly  during  September  in  Pennsyl- 
vania, thereby  forcing  birds  to  move  south  in 
a brief  window  of  time  (Bildstein  1999).  On 
the  other  hand,  solar  intensity  and  photoperiod 
remain  relatively  high  and  constant  year-round 
in  Costa  Rica,  resulting  in  a more  prolonged 
window  of  time  for  hawk  movements  (Porras- 
Penaranda  et  al.  2004). 

As  a result.  Broad-winged  Hawks  may  be 
more  likely  to  use  a time-minimization  strat- 
egy in  temperate  than  in  tropical  zones,  re- 
sulting in  a higher  flapping  rate  in  Pennsyl- 
vania. Assuming  an  energy  consumption  of 
approximately  4 X BMR  in  soaring  flight  and 
a climbing  rate  of  1 m/sec,  flight  theory  pre- 
dicts that  during  time-minimizing  migration, 
heavy  birds  (>132  g)  should  switch  from 
soaring  to  flapping  flight  (Hedenstrom  1993). 
For  energy-minimizing  migration,  the  switch 
from  soaring  to  flapping  flight  occurs  at  a low- 
er climbing  rate.  Thus,  as  the  rate  of  climbing 
decreases,  time-minimizing  migrants  should 
switch  from  soaring  to  flapping  flight  sooner 


than  energy-minimizing  migrants  (Heden- 
strom 1993).  These  temporal  and  energetic  as- 
pects may  explain  why  Broad-winged  Hawks 
are  more  likely  to  resort  to  flapping  in  Penn- 
sylvania than  in  Costa  Rica. 

Our  observations  indicate  that  Broad- 
winged Hawks  shift  from  a mixed  strategy  of 
soaring  and  gliding  supplemented  by  powered 
flight  to  a nearly  pure  strategy  of  soaring  and 
gliding  as  they  proceed  during  their  south- 
bound migrations,  suggesting  that  the  instan- 
taneous metabolic  cost  of  migration  declines 
from  north  to  south.  By  relying  more  on  pow- 
ered flight  in  the  north,  where  conditions  are 
less  favorable  for  soaring,  Broad-winged 
Hawks  may  trade  off  energy  against  time,  a 
phenomenon  also  observed  in  Levant  Spar- 
rowhawks  ( Accipiter  brevipes\  Spaar  et  al. 
1998).  This  would  allow  them  to  move  along 
the  corridor  at  a faster  rate  at  the  expense  of 
depleting  fat  reserves. 

Finally,  we  highlight  the  fact  that  we  did 
not  discriminate  adult  from  juvenile  Broad- 
winged Hawks,  and  that  we  observed  mi- 
grants at  only  two  sites.  Additional  observa- 
tions in  which  the  flight  behavior  of  adults  and 
juveniles  are  compared  and  in  which  other 
species  are  observed  at  other  temperate  and 
tropical  sites  are  likely  to  provide  important 
insights  into  the  extent  to  which  age  and  lat- 
itudinal geography  affects  the  flight  behavior 
of  migrating  birds  of  prey. 

ACKNOWLEDGMENTS 

This  study  was  conducted  as  part  of  VC  and  JFT’s 
undergraduate  research  project  at  the  Universite  de 
Sherbrooke  and  when  VC  and  JFT  were  Conservation 
Science  Interns  at  Hawk  Mountain  Sanctuary  (sup- 
ported by  foreign  studies  scholarships  from  the  Min- 
istere  de  1’ Education  du  Quebec).  We  thank  the  au- 
tumn 2002  interns  at  Hawk  Mountain  Sanctuary  and 
the  interns  and  Bribri  people  at  Kekoldi  Indigenous 
Reserve  for  their  assistance  and  hospitality.  Finally,  we 
are  grateful  to  R.  Spaar,  P.  Kerlinger,  C.  J.  Farmer,  and 
three  anonymous  referees  for  helping  us  improve  our 
manuscript.  This  is  Hawk  Mountain  Sanctuary  Contri- 
bution to  Conservation  Science,  number  135. 

LITERATURE  CITED 

Allen,  P.  E.,  L.  J.  Goodrich,  and  K.  L.  Bildstein. 
1996.  Within-  and  among-year  effects  of  cold 
fronts  on  migrating  raptors  at  Hawk  Mountain, 
Pennsylvania,  1934-1991.  Auk  113:329-338. 
Bildstein,  K.  L.  1999.  The  forced  migration  of  the 
Broad-winged  Hawk.  Pages  79-102  in  Gathering 


Careau  et  al.  • BROAD-WINGED  HAWK  MIGRATION 


All 


of  angels:  migrating  birds  and  their  ecology  (K.  P. 
Able,  Ed.).  Cornell  University  Press,  Ithaca,  New 
York. 

Bildstein,  K.  L.  and  M.  Saborio.  2000.  Spring  mi- 
gration counts  of  raptors  and  New  World  vultures 
in  Costa  Rica.  Ornitologia  Neotropical  11:197- 
205. 

Bildstein,  K.  L.  and  J.  I.  Zalles.  2001.  Raptor  mi- 
gration along  the  Mesoamerican  Land  Corridor. 
Pages  1 19-136  in  Hawkwatching  in  the  Americas 
(K.  Bildstein  and  D.  Klem,  Eds.).  Hawk  Migration 
Association  of  North  America,  North  Wales, 
Pennsylvania. 

Danchin,  E.,  L.  A.  Giraldeau,  T.  J.  Valone,  and  R. 
H.  Wagner.  2004.  Public  information:  from  nosy 
neighbors  to  cultural  evolution.  Science  305:487- 
491. 

Fuller,  M.  R.,  W.  S.  Seegar,  and  L.  S.  Schueck. 
1998.  Routes  and  travel  rates  of  migrating  Pere- 
grine Falcons  Falco  peregrinus  and  Swainson’s 
Hawks  Buteo  swainsoni  in  the  Western  Hemi- 
sphere. Journal  of  Avian  Biology  29:433-440. 

Goodrich,  L.  J.,  S.  C.  Crocoll,  and  S.  E.  Senner. 
1996.  Broad-winged  Hawk  ( Buteo  platypterus). 
The  Birds  of  North  America,  no.  218. 

Hedenstrom,  A.  1993.  Migration  by  soaring  or  flap- 
ping flight  in  birds — the  relative  importance  of  en- 
ergy-cost and  speed.  Philosophical  Transactions  of 
the  Royal  Society  of  London,  Series  B 342:353- 
361. 

Hosmer,  D.  W.  and  S.  Lemeshow.  1989.  Applied  lo- 
gistic regression.  J.  Wiley  and  Sons,  New  York. 

Huffaker,  E.  C.  1897.  On  soaring  flight.  Annual  re- 
port of  the  Board  of  Regents,  Smithsonian  Insti- 
tution, Washington,  D.C. 

Kerlenger,  P.  1989.  Flight  strategies  of  migrating 
hawks.  University  of  Chicago  Press,  Chicago,  Il- 
linois. 

Kerlinger,  P.  and  S.  A.  Gauthreaux.  1985.  Seasonal 
timing,  geographic  distribution,  and  flight  behav- 
ior of  Broad- Winged  Hawks  during  spring  migra- 
tion in  southern  Texas:  a radar  and  visual  study. 
Auk  102:735-743. 

Moore,  F.  R.  2000.  Stopover  ecology  of  Nearctic-Neo- 


tropical  landbird  migrants.  Studies  in  Avian  Bi- 
ology 20:1-133. 

Pennycuick,  C.  J.  1972.  Animal  flight.  Edward  Arnold 
Ltd.,  London,  England. 

Pennycuick,  C.  J.  1998.  Field  observations  of  thermals 
and  thermal  streets,  and  the  theory  of  cross-coun- 
try soaring  flight.  Journal  of  Avian  Biology  29: 
33-43. 

Pennycuick,  C.  J.,  T.  Alerstam,  and  B.  Larsson. 
1979.  Soaring  migration  of  the  Common  Crane 
Grus  grus  observed  by  radar  and  from  an  aircraft. 
Ornis  Scandinavica  10:241-251. 

PORRAS-PEN ARANDA,  P.,  L.  ROBICHAUD,  AND  F.  BRANCH. 
2004.  New  full-season  count  sites  for  raptor  mi- 
gration in  Talamanca,  Costa  Rica.  Ornitologia 
Neotropical  15:267-278. 

SAS  Institute,  Inc.  2002.  SAS  JMP,  ver.  5.0.1.  Cary, 
North  Carolina. 

Shannon,  H.  D.,  G.  S.  Young,  M.  A.  Yates,  M.  R. 
Fuller,  and  W.  S.  Seegar.  2002.  Measurements 
of  thermal  updraft  intensity  over  complex  terrain 
using  American  White  Pelicans  and  a simple 
boundary-layer  forecast  model.  Boundary-Layer 
Meteorology  104:167-199. 

Smith,  N.  G.  1980.  Hawk  and  vulture  migrations  in 
the  Neotropics.  Pages  51-66  in  Migrant  birds  in 
the  Neotropics  (A.  Keast  and  E.  S.  Morton,  Eds.). 
Smithsonian  Institution  Press,  Washington,  D.C. 

Smith,  N.  G.  1985.  Thermals,  cloud  streets,  trade 
winds,  and  tropical  storms:  how  migrating  raptors 
make  the  most  of  atmospheric  energy  in  Central 
America.  Pages  51-65  in  Proceedings  of  hawk 
migration  conference  IV  (M.  Harwood,  Ed.). 
Hawk  Migration  Association  of  North  America, 
Lynchburg,  Virginia. 

Smith,  N.  G.,  D.  L.  Goldstein,  and  G.  A.  Barthol- 
omew. 1986.  Is  long-distance  migration  possible 
for  soaring  hawks  using  only  stored  fat?  Auk  103: 
607-611. 

Spaar,  R.  and  B.  Bruderer.  1997.  Migration  by  flap- 
ping or  soaring:  flight  strategies  of  Marsh,  Mon- 
tagu’s and  Pallid  harriers  in  southern  Israel.  Con- 
dor 99:458-469. 

Spaar,  R.,  H.  Stark,  and  F.  Liechti.  1998.  Migratory 
flight  strategies  of  Levant  Sparrowhawks:  time  or 
energy  minimization?  Animal  Behaviour  56: 
1185-1197. 


The  Wilson  Journal  of  Ornithology  1 18(4):478— 484,  2006 


COLONIALITY,  MATE  RETENTION,  AND  NEST-SITE 
CHARACTERISTICS  IN  THE  SEMIPALMATED  SANDPIPER 

JOSEPH  R.  JEHL,  JR.1 


ABSTRACT. — Coloniality  is  unusual  among  Scolopacidae.  At  Churchill.  Manitoba,  however,  the  small,  rem- 
nant population  of  Semipalmated  Sandpipers  ( Calidris  pusilla)  is  highly  clumped,  with  nesting  density  approx- 
imating 3-4  pairs/ha,  and  should  be  considered  colonial.  The  species  exhibits  high  fidelity  to  territory,  mates, 
and  nest  sites — behaviors  that  promote  rapid  pair  formation  and  allow  experienced  birds  to  increase  their  repro- 
ductive success  by  nesting  earlier  than  pairs  forming  for  the  first  time.  The  value  of  experience  and  early  nesting 
was  evidenced  by  the  fact  that  six  of  seven  returning  young  were  produced  by  experienced  pairs  and  had  hatched 
on  the  first  day  of  their  respective  nesting  seasons.  Nests  were  placed  in  dry  locations  very  near  open  water. 
Those  adjacent  to  small  shrubs  had  slightly  greater  success,  and  young  produced  from  these  nests  had  much 
higher  rates  of  return  than  those  from  nests  placed  amid  sedges.  In  other  parts  of  their  breeding  range,  Semi- 
palmated Sandpipers  are  also  clumped  and  seem  likely  to  be  colonial.  If  so,  estimates  of  breeding  populations 
derived  from  indirect  methods,  such  as  habitat  assessment  from  aerial  photographs,  will  have  limited  applicability 
and  will  need  to  be  complemented  by  ground-truthing.  Received  3 October  2005,  accepted  2 May  2006. 


Spatial  distribution  in  breeding  birds  runs 
the  gamut  from  solitary  nesting  coupled  with 
strongly  developed  territorial  behavior  to 
highly  colonial,  with  the  defended  area  being 
limited  to  the  area  that  parents  can  protect 
without  leaving  their  nests.  Shorebirds  (Char- 
adrii)  exhibit  similar  variation.  Most  are  soli- 
tary nesters,  but  in  some  groups  (e.g.,  Dro- 
madidae,  Recurvirostridae,  Glareolinae)  co- 
loniality is  the  rule,  the  extreme  being  attained 
by  the  Banded  Stilt  ( Cladorhynchus  leucoce- 
phalus ),  in  which  densities  up  to  18  nests/m2 
have  been  reported  (Minton  et  al.  1995.  del 
Hoyo  et  al.  1996,  van  Gils  and  Wiersma 
1996).  Lacking  “objective  (or  even  widely  ac- 
cepted) criteria  as  to  how  clumped  nests  must 
be  to  constitute  a true  colony,”  ornithologists 
have  used  such  terms  as  “semicolonial,” 
“strongly  clumped,”  or  “loose  colony”  to  de- 
scribe situations  in  which  “rather  more  dis- 
persed nests  . . . are  . . . judged  to  be  in  a 
clump  relative  to  the  density  of  nests  in  the 
general  vicinity”  (Campbell  and  Lack  1985: 
95).  In  any  case,  the  essence  of  coloniality  is 
that  birds  of  a feather  are  disposed  to  nest  near 
each  other,  the  attraction  being  primarily  so- 
cial rather  than  to  a common  habitat. 

Among  Scolopacidae.  coloniality  of  any 
kind  is  rare,  and  in  the  calidridine  sandpipers 
(Calidridini)  “semi-coloniality”  has  been  re- 
ported or  suspected  only  in  the  Western  (Cal- 


1 Smithsonian  Ornithology,  U.S.  National  Museum 

of  Natural  History,  Washington  D.C.  20560.  USA: 
e-mail:  grebe5k@cs.com 


idris  mauri)  and  Broad-billed  {Limicola  fal- 
cinellus ) sandpipers  (Palmer  1967,  van  Gils 
and  Wiersma  1996).  To  this  small  list  may  be 
added  the  Semipalmated  Sandpiper  ( Calidris 
pusilla ),  a monogamous  and  highly  territorial 
species  that  breeds  in  the  Subarctic  and  lower 
latitudes  of  the  North  American  Arctic.  De- 
spite having  been  studied  in  only  a few  areas, 
its  breeding  biology  is  well-documented, 
mainly  through  comprehensive  studies  at  La 
Perouse  Bay,  Manitoba,  by  Gratto-Trevor 
(1992,  and  references  therein).  Although 
known  to  nest  at  relatively  high  densities,  the 
Semipalmated  Sandpiper  has  not  been  sus- 
pected of  nesting  colonially.  At  Churchill. 
Manitoba,  however,  that  appears  to  be  the 
case.  Here  I present  observations  on  Semipal- 
mated Sandpiper  spacing  and  nesting  behav- 
ior, along  with  information  on  nest-site  char- 
acteristics, philopatry,  and  other  aspects  of  the 
species'  breeding  biology  that  complement 
and  extend  Gratto-Trevor’s  findings. 

METHODS 

Observations  were  made  in  a potential  nest- 
ing area  of  7,000  ha  in  the  “immediate  Chur- 
chill Area”  (Jehl  and  Lin  2001,  map  in  Jehl 
2004:  58°  45 ' N.  94°  00'  W)  from  1993 
through  2004  as  part  of  a broader  study  on 
shorebird  biology  (Jehl  and  Lin  2001,  Jehl 
2004).  From  previous  studies  in  1964  through 
1967,  I was  familiar  with  the  status  of  shore- 
birds  in  the  Churchill  area  (Jehl  and  Smith 
1970).  When  I resumed  studies  in  1991.  I 


478 


Jehl  • SEMIPALMATED  SANDPIPER  BIOLOGY 


479 


failed  to  encounter  Semipalmated  Sandpipers 
until  1993,  when  I found  a few  pairs  nesting 
in  a small  meadow  (—25  ha)  25  km  east  of 
the  Churchill  townsite.  Then,  and  in  each  sub- 
sequent year,  I attempted  to  find  all  nests  and 
mark  all  individuals.  I trapped  adults  at  the 
nest  in  a simple  walk-in  trap  and  banded  them 
with  aluminum  bands  (or  stainless  steel,  when 
available)  and  individually  coded  colored 
plastic  bands.  I made  standard  measurements 
with  dial  calipers  (culmen  and  tarsus  to  0.1 
mm;  flattened  wing  to  1 mm)  and  weighed 
each  bird  on  a digital  scale  (to  0. 1 g).  Chicks 
were  banded  (but  not  color-marked)  before 
they  left  the  nest.  From  this  effort,  the  iden- 
tities of  most  adults  (88%  of  93  from  1993  to 
2001)  and  young  (73%  of  120  from  1993  to 
2000)  were  known,  which  allowed  their  sta- 
tus, mates,  distribution,  and  nesting  success  to 
be  followed  from  year  to  year.  I aged  adults 
on  the  basis  of  Gratto  and  Morrison’s  (1981) 
observation  that  most  first-year  birds  are  dis- 
tinguishable from  older  birds  by  having  up  to 
four  newly  replaced  outer  primaries.  Obser- 
vations in  2001  through  2004  focused  on  doc- 
umenting the  identities  of  returned  birds. 

In  most  calidridines,  males  are  typically 
smaller  (e.g.,  Jehl  and  Murray  1985),  but  there 
is  much  overlap.  To  determine  sex,  I also  used 
behavioral  information,  including  observa- 
tions that  males  defend  territories  much  more 
strongly,  sing  longer  and  more  complex  songs, 
and  are  bolder  around  the  nest.  For  birds  re- 
turning in  subsequent  years,  it  was  usually 
possible  to  use  behavior  to  test  earlier  deter- 
minations: in  only  2 of  25  cases  did  a tentative 
sexing  need  to  be  reconsidered. 

RESULTS 

Phenology  and  colony  designation. — Semi- 
palmated Sandpipers  migrate  through  the 
Churchill  region  between  the  last  days  of  May 
and  the  first  third  of  June.  Locally  nesting 
birds  move  immediately  to  breeding  areas, 
where  they  engage  in  prolonged  and  conspic- 
uous territorial  and  courtship  displays.  Dis- 
play flights  take  place  at  elevations  of  40-50 
m and  may  last  10  min  or  more.  Typically, 
these  displays  involve  several  birds,  which 
chase  back  and  forth  over,  and  well  beyond, 
the  nesting  area. 

From  1993  through  2004,  the  only  Semi- 
palmated Sandpipers  nesting  in  the  potential 


(7,000  ha)  nesting  area  occurred  in  the  25-ha 
meadow  described  above.  Bordered  by  two 
lakes  and  dotted  with  shallow  ponds  that  dried 
out  by  late  June,  the  area  was  relatively  wet 
and  contained  slightly  more  shrubby  vegeta- 
tion than  some  other  nearby  sites.  Because  ( 1 ) 
the  nesting  area  occupied  only  3-4  ha  of  this 
meadow,  (2)  nest  density  was  extremely  high 
(see  below),  (3)  similar  habitat  elsewhere  in 
the  Churchill  area  was  unused,  (4)  the  historic 
distribution  of  Semipalmated  Sandpipers  at 
Churchill  had  not  been  limited  to  this  type  of 
habitat,  and  (5)  nesting  areas  used  through  the 
1960s,  though  largely  unchanged,  were  no 
longer  used,  it  was  clear  that  the  birds  were 
attracted  to  each  other  and  not  to  any  specific 
habitat  or  topographic  conditions.  Conse- 
quently, their  nesting  behavior  could  be  de- 
scribed as  colonial.  Elsewhere  in  the  Churchill 
area,  I encountered  Semipalmated  Sandpipers 
only  twice  from  1993  through  2004:  one  un- 
mated male,  and  an  apparent  pair,  each  located 
>5  km  from  the  colony.  All  three  birds  dis- 
appeared after  a few  days. 

The  colony  contained  five  pairs  in  1993. 
Colony  size  had  increased  slightly  by  1995 
(11  nests;  Table  1)  and  (probably)  1996,  but 
runoff  in  1996  flooded  some  early  nests  and 
may  have  prevented  some  pairs  from  finding 
suitable  territories  or  renesting.  In  1997,  the 
number  of  adults  was  halved  and  I found  only 
two  nests.  Subsequently,  through  2001,  the 
colony  size  fluctuated  from  two  to  three  pairs, 
and  by  2003  (and  perhaps  2004)  there  was 
only  a single,  unpaired  male.  At  maximum 
size  in  1995  (Fig.  1),  the  colony  encompassed 
3.4  ha  (determined  by  a polygon  drawn 
around  the  outermost  nests;  this  area  included 
open-water  areas  where  nesting  was  impossi- 
ble), had  a maximum  linear  extent  of  416  m, 
and  a density  of  3.2  pairs/ha  (maximum  = 4.1 
in  1993).  Nests  were  tightly  packed,  the  near- 
est-neighbor distance  averaging  about  55  m 
(minimum  = 31  m). 

Mate  fidelity. — -As  in  some  other  calidridines 
(e.g.,  Least  Sandpiper,  Calidris  minutilla\  Stilt 
Sandpiper,  C.  himantopus ; Dunlin,  C.  alpina\ 
Jehl  1970;  JRJ  unpubl.  data),  Semipalmated 
Sandpipers  form  long-term  bonds  and  pairs 
tend  to  re-occupy  former  territories  as  long  as 
both  members  are  alive  (see  also  Gratto  et  al. 
1985).  In  16  cases  in  this  study,  both  partners 
returned,  pairs  reunited  13  times  in  the  follow- 


480 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  1.  Population  size  and  density  of  Semipalmated  Sandpipers  at  Churchill,  Manitoba,  1993-2001. 


Year 

Population  size3 

No.  nests 

Nesting  area 
(ha)c 

No.  pairs/hac 

Maximum  extent 
of  colony  (m) 

Distance  to  nearest  nest(s): 
[range]  and  median  (m) 

1993 

>10 

5 

1.2 

4.1 

126 

[54-181]  87.3 

1994 

16-19 

8 

2.8 

2.9 

268 

[52-63]  55.1 

1995 

22-24 

11 

3.4 

3.2 

416 

[31-101]  54.4 

1996 

21-22 

8b 

2.7 

2.9 

381 

[37-124]  88.4 

1997 

12 

2 

— 

— 

121 

121 

1998 

7 

3 

— 

— 

274 

[84-193] 

1999 

4 

2 

— 

— 

— 

55 

2000 

6-8 

3 

— 

— 

— 

90,  91 

2001 

>6 

3 

— 

— 

— 

— 

a Estimated  number  of  adults  in  colony  early  in  the  season. 
b Omits  one  renesting. 

c Could  not  be  calculated  from  two  points  or  when  nests  were  arranged  linearly. 


ing  season,  and  all  had  nested  successfully  in 
the  previous  year.  Three  pairs  divorced  (one 
previously  successful,  two  unsuccessful).  The 
successful  male  acquired  a new  mate  and  his 
old  mate  soon  disappeared.  Of  the  two  pre- 
viously unsuccessful  pairs,  the  nest  of  one  was 
flooded,  the  female  acquired  a new  mate,  and 
the  old  male  skipped  breeding;  both  birds  of 


the  other  pair  acquired  new  partners,  but  the 
males  retained  their  previous  nest  sites.  Of  the 
pairs  that  reunited,  two  remained  intact  for 
four  seasons,  three  for  three  seasons,  and  two 
for  two  seasons. 

Nineteen  pairs  failed  to  reunite.  The  reasons 
can  only  be  guessed,  as  banded  but  unidenti- 
fied birds  occasionally  showed  up  early  in  the 


FIG.  1.  Location  and  spacing  of  Semipalmated  Sandpiper  nests  (•)  at  Churchill,  Manitoba,  1995. 


Jehl  • SEMIPALMATED  SANDPIPER  BIOLOGY 


481 


TABLE  2.  Spacing  and  dispersal  of  Semipalmated  Sandpipers  at  Churchill  and  La  Perouse  Bay,  Manitoba. 


Churchill  1993-2001 

La  Perouse  Bay  1980-19843 

Variable 

n 

Spacing,  behavior 

n 

Spacing,  behavior 

Size  of  breeding  area;  habitat 
Population  size;  density 

3-4  ha;  a single  small 
meadow 

2-11  pairs;  3-4/ha 

2 km2;  on  delta  of  Mast 
River 

100  pairs;  1/ha 

Pairs  reuniting,  if  both  alive 

16 

13  (81%) 

79 

64  (81%) 

Reuse  of  old  nest  cup 

41 

8 (19.5%) 

305 

13  (4.3%) 

Rate  of  nest  reuse  if  both 
parents  returned 

13 

8 (61.5%) 

No  data 

Nest  shift:  reunited  pairs 

14 

Range  = 0-85  m;  mean  = 
25.4  ± 36  m;  mode  = 0 m 

168 

Range  = 0-575  m;  annual 
medians:  40-66  m 

Nest  shift:  female  mate  change 

8 

4-360  m,  mean  = 153  ± 
126  m;  median  = 115m 

33 

Range  = 23-825  m;  annual 
medians:  138-174  m 

a From  Gratto  et  al.  (1985). 


year  and  then  disappeared,  perhaps  without 
mating  or  perhaps  because  their  nest  was  lost 
before  I could  find  it.  In  several  cases,  the 
break-up  was  evidently  due  to  bad  timing  (one 
partner  returned  late)  or  the  unavailability  of 
a previous  nest  site  (see  below). 

Nest-site  selection  and  site  tenacity. — Just 
as  Semipalmated  Sandpipers  tend  to  retain 
mates  and  territories  from  year  to  year,  they 
also  retain  nest  sites,  as  long  as  the  previous 
nesting  attempt  was  successful,  the  mate  re- 
mains alive,  and  the  nest  is  in  suitable  con- 
dition and  does  not  contain  unhatched  eggs 
from  the  previous  season.  Of  13  cases  in 
which  both  mates  returned  and  reunited,  the 
distance  to  subsequent  nests  ranged  from  0 to 
85  m (mode  = 0 m;  Table  2).  One  pair  used 
the  same  nest  for  4 successive  years. 

Semipalmated  Sandpipers  selected  nest  lo- 
cations very  near  ponds  (mean  = 10.9  m ± 
8.8,  range  = 0.5-29.5  m,  n = 26),  but  placed 
their  nests  in  dry  situations  on  the  sides  or 
tops  of  small  hummocks  or  ridges.  Two  types 
of  nest  sites  were  used:  “shrub”  sites  were 
located  under,  or  adjacent  to,  small  bushes — 
in  this  case  sweetgale  {Myrica  gale ) or  dwarf 
birch  ( Betula  nana) — which  typically  allowed 
access  from  only  one  direction;  “sedge”  sites 
were  in  low,  damp  areas  and  nests  were  placed 
in  a clump  of  sedge  {Care x spp.).  At  41  doc- 
umented sites  (including  those  reused  by  the 
same  pair  in  subsequent  years),  30  were  in 
shrub  and  1 1 in  sedge.  Nesting  success  was 
slightly  (but  not  significantly)  greater  in  shrub 
sites  (83%  versus  72%),  which  are  better  con- 
cealed and  less  subject  to  flooding.  However, 


the  greater  desirability  of  shrub  sites  was  clear 
from  their  retention  rates.  Of  25  successful 
shrub  sites,  14  (56%)  were  reused,  13  by  a 
returning  pair  and  1 by  a male  with  a new 
mate.  Of  the  1 1 successful  shrub  sites  that 
were  not  reused,  the  nest  cup  or  habitat  had 
become  unusable  {n  = 3)  or  one  or  both  mates 
failed  to  return  ( n = 8).  In  sedge  sites,  8 of 
1 1 nestings  were  successful,  yet  none  was  re- 
occupied (1  site  was  used  several  years  later 
by  a pair  with  no  previous  breeding  experi- 
ence; the  nest  failed).  In  the  other  cases,  the 
habitat  had  changed  over  the  intervening  win- 
ter ( n = 3)  or  one  or  both  mates  failed  to 
return  ( n - 4). 

Among  individuals  that  moved  to  a new  lo- 
cation, males  {n  = 9)  tended  to  stay  near  their 
previous  nest  site  (median  distance  ==  40  m). 
Eight  paired  with  females  that  had  no  previous 
experience,  and  one  bred  successfully  in  the 
same  territory  for  4 successive  years,  each 
time  with  a new  mate  and  each  time  moving 
—50  m away  from  the  previous  site  before 
returning  to  the  original  nest  in  the  4th  year. 
Females  {n  = 8)  tended  to  move  farther  away 
from  previous  nest  sites  (median  = 115  m). 
Three  females  paired  with  experienced  males 
that  held  territories  near  the  center  of  the  col- 
ony; one  of  these  birds  failed  to  nest  one  year 
when  her  nest  was  flooded,  but  she  returned 
to  her  old  territory  (by  then  held  by  a different 
male)  and  nested  within  4 m of  the  original 
scrape.  The  other  five  females  bred  with  in- 
experienced males,  whose  nests  in  all  but  one 
case  were  on  the  periphery  of  the  colony.  One 
pair  in  its  2nd  year  moved  60  m,  then  80  m 


482 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


in  year  3,  and  80  m again  in  year  4.  When  the 
nest  was  flooded  in  year  4,  the  birds  moved 
85  m,  which  brought  them  to  within  4 m of 
their  original  nest. 

Of  120  local  chicks  banded,  7 returned  to 
breed.  At  least  six  of  these  were  produced  by 
pairs  in  which  at  least  one  parent  had  nested 
successfully  in  a previous  year;  five  (including 
two  from  the  same  clutch)  were  produced  by 
two  pairs.  All  returning  young  paired  with  in- 
experienced mates;  the  males  (n  = 5)  moved 
130-225  m (mean  = 197  m)  and  the  females 
(n  = 2)  moved  85  and  226  m from  their  natal 
sites.  When  the  colony  was  relatively  large, 
young  males,  with  one  exception,  were  only 
able  to  obtain  territories  at  the  colony  edge. 
One  bred  on  the  periphery  in  his  1st  year  and 
then  moved  to  a more  central  site  in  his  2nd 
year.  Another  male  obtained  a central  location 
at  first  breeding,  but  only  after  experienced 
neighbors  had  reduced  territory  defense  (cf. 
Jehl  1973)  and  started  incubating;  its  young 
hatched  a week  later  than  those  of  other  pairs. 

DISCUSSION 

Breeding  behavior. — The  aspects  of  mate 
and  territory  retention,  philopatry,  and  dis- 
persal treated  in  this  study  largely  conform  to 
those  reported  by  Gratto  et  al.  (1985)  at  La 
Perouse  Bay,  —30  km  to  the  east  (Table  2). 
At  Churchill,  nest  density  was  greater  than  it 
was  at  La  Perouse  Bay  (3-4  versus  1 pair/ha), 
returning  pairs  dispersed  much  less  (if  at  all) 
from  previous  nests,  and  reuse  of  the  nest  cup 
was  greater  (19.5%  versus  4.3%;  61.5%  [this 
study]  if  both  pair  members  returned).  These 
differences  were  probably  related  to  topogra- 
phy and  the  size  and  stability  of  the  respective 
nesting  areas.  Churchill  birds  were  restricted 
to  a small  meadow,  whereas  Semipalmated 
Sandpipers  at  La  Perouse  Bay  bred  on  a river 
delta  that  often  experienced  high  flows  during 
runoff,  resulting  in  greater  loss  of  old  nest 
cups.  At  Churchill,  young  males  tended  to 
breed  at  the  colony’s  edge  but  did  not  disperse 
as  far  from  their  natal  sites  as  they  did  at  La 
Perouse  Bay  (197  m versus  549  m,  respec- 
tively), probably  because  the  colony  was 
much  smaller. 

For  any  species,  the  timing  of  breeding  is 
critical  to  reproductive  success  (Lack  1968), 
and  it  is  widely  acknowledged  that  individuals 
nesting  earlier — nearly  always  experienced 


birds — typically  have  greater  success  than 
those  that  start  later  (e.g.,  Soikkeli  1967,  Jehl 
1970,  Gratto  et  al.  1983,  Black  1996,  Handel 
and  Gill  2000,  Ruthrauff  and  McCaffery 
2005).  Early  breeding  is  enhanced  by  high 
rates  of  territory,  mate,  and  nest-site  retention, 
which  allow  mates  to  begin  nesting  as  soon 
as  habitat  conditions  permit.  These  behaviors 
are  especially  important  where  breeding  sea- 
sons are  short,  so  it  is  not  surprising  that  they 
have  been  reported  in  a variety  of  shorebirds 
that  nest  in  the  Arctic,  including  Dunlin,  Least 
and  Stilt  sandpipers,  and  Black  Turnstone  ( Ar - 
enaria  melanocephala;  Soikkeli  1967;  Jehl 
1970,  1973;  Gratto  et  al.  1985;  Jonsson  1987; 
Handel  and  Gill  2000;  Sandercock  et  al.  2005; 
JRJ  unpubl.  data).  In  this  study  the  importance 
of  adult  experience  and  early  nesting  was  con- 
firmed by  the  observation  that  six  of  the  seven 
chicks  that  returned  to  nest  were  not  only  pro- 
duced by  experienced  parents  but  also  hatched 
on  the  1st  day  of  their  respective  hatching  pe- 
riods. The  one  exception  hatched  from  the 
penultimate  nest  of  its  season  and  was  pro- 
duced by  a pair  that  had  not  nested  together 
previously.  Although  the  female  had  no 
known  experience,  the  male  had  bred  suc- 
cessfully twice.  Whereas  the  experience  of 
both  parents  is  surely  relevant,  that  of  the 
male  is  paramount  because  in  this  species  and 
many  other  sandpipers,  he  takes  the  sole  or 
major  role  in  rearing  the  chicks  from  hatching 
to  fledging  (Jehl  1973,  Gratto-Trevor  1991; 
JRJ  unpubl.  data). 

Territory  function  and  spacing. — When  not 
incubating,  Semipalmated  Sandpipers  left 
their  territories  and  departed  the  colony  area. 
Some  moved  to  the  mudflats  of  Hudson  Bay, 
a minimum  distance  of  2-3  km,  whereas  when 
water  levels  were  low  inland,  several  might 
have  fed  together  on  mudflats  in  a lake  bor- 
dering the  colony.  Because  territory  in  this 
species  is  not  based  on  food  availability,  it 
appears  that  nest  spacing  is  determined  by  a 
balance  between  attraction  to  conspecifics  and 
the  need  to  maintain  sufficient  distance  be- 
tween neighbors  to  prevent  predators  from 
finding  nests. 

Density  and  population  estimates. — Semi- 
palmated Sandpipers  are  reported  to  nest  at 
greater  densities  than  other  sandpipers,  except 
perhaps  the  Western  Sandpiper.  On  the  North 
Slope  of  Alaska,  where  the  Semipalmated 


Jehl  • SEMIPALMATED  SANDPIPER  BIOLOGY 


483 


Sandpiper  is  abundant.  Cotter  and  Andres 
(2000)  reported  mean  densities  of  30  pairs/ 
km2;  farther  inland  they  noted  up  to  21.3 
nests/km2.  At  La  Perouse  Bay,  Manitoba, 
Gratto  et  al.  (1985)  estimated  territory  size  to 
be  1.0  ha,  including  defended  water  areas 
(maximum  density  was  2.3  pairs/ha,  based  on 
dry  land  areas).  At  Churchill,  density  was 
even  greater,  reaching  up  to  4 pairs/ha  (=  400 
pairs/km2,  inclusive  of  pond  areas).  While  all 
populations  of  Semipalmated  Sandpipers  do 
not  necessarily  have  the  same  nesting  habits 
(e.g.,  Gratto  and  Cooke  1987),  spacing  is  also 
clumped  in  the  three  breeding  localities  clos- 
est to  Churchill:  Gordon  Point  and  Fox  Island 
(Jehl  2004;  JRJ  unpubl.  data)  and  La  Perouse 
Bay  (C.  Gratto-Trevor  pers.  comm.).  This  and 
the  high  densities  reported  elsewhere  suggest 
that  the  species  is  probably  colonial  through- 
out its  range.  If  so,  estimates  of  breeding  pop- 
ulations derived  from  indirect  methods,  such 
as  habitat  assessment  from  satellite  photogra- 
phy or  vegetation  maps  (e.g.,  Gratto-Trevor 
1996),  will  have  limited  applicability.  Addi- 
tional documentation  of  the  kinds  of  breeding 
behavior  reported  in  this  paper,  complemented 
by  ground-truthing  of  nest  spacing  in  different 
geographic  regions,  will  be  useful. 

ACKNOWLEDGMENTS 

I was  assisted  in  the  field  by  J.  Klima,  W.  Lin,  J. 
Terp,  C.  MacDonald,  and  many  volunteers  associated 
with  the  Churchill  Northern  Studies  Centre.  S.  I.  Bond 
prepared  the  map  in  Figure  1 and  helped  compile  the 
data.  I am  grateful  to  C.  Gratto-Trevor  for  laying  the 
groundwork  for  this  study  and  to  her,  J.  Klima,  A. 
Henry,  and  R.  I.  G.  Morrison  for  insightful  comments 
on  this  manuscript. 

LITERATURE  CITED 

Black,  J.  M.  (Ed.).  1996.  Partnerships  in  birds:  the 
study  of  monogamy.  Oxford  University  Press, 
New  York. 

Campbell,  B.  and  E.  Lack  (Eds.).  1985.  A dictionary 
of  birds.  T.  & A.  D.  Poyser,  Calton,  United  King- 
dom. 

Cotter,  B.  A.  and  B.  A.  Andres.  2000.  Nest  density 
of  shorebirds  inland  from  the  Beaufort  Sea  Coast, 
Alaska.  Canadian  Field-Naturalist  114:287-291. 
del  Hoyo,  J.,  A.  Elliott,  and  J.  Sargatal  (Eds.). 
1996.  Handbook  of  the  birds  of  the  world,  vol.  3. 
Lynx  Edicions,  Barcelona,  Spain. 

Gratto,  C.  L.  and  F.  Cooke.  1987.  Geographic  vari- 
ation in  the  breeding  biology  of  the  Semipalmated 
Sandpiper.  Ornis  Scandinavica  18:223-235. 
Gratto,  C.  L.,  F.  Cooke,  and  R.  I.  G.  Morrison. 


1983.  Nesting  success  of  yearling  and  older  breed- 
ers in  the  Semipalmated  Sandpiper  Calidris  pus- 
ilia.  Canadian  Journal  of  Zoology  62:1889-1892. 

Gratto,  C.  L.  and  R.  I.  G.  Morrison.  1981.  Partial 
postjuvenal  molt  of  the  Semipalmated  Sandpiper 
( Calidris  pusilla).  Wader  Study  Group  Bulletin 
33:33-37. 

Gratto,  C.  L.,  R.  I.  G.  Morrison,  and  F.  Cooke. 
1985.  Philopatry,  site  tenacity,  and  mate  fidelity 
in  the  Semipalmated  Sandpiper.  Auk  102:16-24. 

Gratto-Trevor,  C.  L.  1991.  Parental  care  in  the 
Semipalmated  Sandpiper  Calidris  pusilla : brood 
desertion  by  females.  Ibis  133:394-399. 

Gratto-Trevor,  C.  L.  1992.  Semipalmated  Sandpiper 
( Calidris  pusilla).  The  Birds  of  North  America, 
no.  6. 

Gratto-Trevor,  C.  L.  1996.  Use  of  Landsat  TM  im- 
agery in  determining  important  shorebird  habitat 
in  the  Outer  Mackenzie  Delta,  Northwest  Terri- 
tories. Arctic  49:11-22. 

Handel,  C.  M.  and  R.  E.  Gill,  Jr.  2000.  Mate  fidelity 
and  breeding  site  tenacity  in  a monogamous  sand- 
piper, the  Black  Turnstone.  Animal  Behaviour  60: 
471-481. 

Jehl,  J.  R.,  Jr.  1970.  Sexual  selection  for  size  differ- 
ences in  two  species  of  sandpipers.  Evolution  24: 
311-319. 

Jehl,  J.  R.,  Jr.  1973.  Breeding  biology  and  systematic 
relationships  of  the  Stilt  Sandpiper.  Wilson  Bul- 
letin 85:115-147. 

Jehl,  J.  R.,  Jr.  2004.  Birdlife  of  the  Churchill  region. 
Trafford  Press,  Victoria,  British  Columbia. 

Jehl,  J.  R.,  Jr.,  and  W.  Lin.  2001.  Population  status 
of  shorebirds  nesting  at  Churchill,  Manitoba.  Ca- 
nadian Field-Naturalist  1 15:487-494. 

Jehl,  J.  R.,  Jr.,  and  B.  G.  Murray,  Jr.  1985.  The 
evolution  of  normal  and  reversed  sexual  size  di- 
morphism in  shorebirds  and  other  birds.  Current 
Ornithology  3:1-86. 

Jehl,  J.  R.,  Jr.,  and  B.  A.  Smith.  1970.  Birds  of  the 
Churchill  region,  Manitoba.  Special  Publication 
no.  1,  Manitoba  Museum  of  Man  and  Nature, 
Winnipeg,  Manitoba. 

Jonsson,  P.  E.  1987.  Sexual  size  dimorphism  and  as- 
sortative  mating  in  the  Dunlin  Calidris  alpina 
schinzii  in  southern  Sweden.  Ornis  Scandinavica 
18:257-264. 

Lack,  D.  1968.  Ecological  adaptations  for  breeding  in 
birds.  Methuen,  London,  United  Kingdom. 

Minton,  C.,  G.  Pearson,  and  J.  Lane.  1995.  History 
in  the  making:  Banded  Stilts  do  it  again.  Wing- 
span 5(2):  13-15. 

Palmer,  R.  1967.  Species  accounts.  Pages  43-267  in 
The  shorebirds  of  North  America  (G.  Stout,  Ed.). 
Viking  Press,  New  York. 

Ruthrauff,  D.  R.  and  B.  J.  McCaffery.  2005.  Sur- 
vival of  Western  Sandpipers  breeding  on  the  Yu- 
kon-Kuskokwim  Delta,  Alaska.  Condor  107:597- 
604. 

Sandercock,  B.  K.,  T.  Szekely,  and  A.  Kosztolanyi. 
2005.  The  effects  of  age  and  sex  on  the  apparent 


484 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


survival  of  Kentish  Plovers  breeding  in  southern 
Turkey.  Condor  107:583-596. 

Soikkeli,  M.  1967.  Breeding  cycle  and  population  dy- 
namics in  the  Dunlin  ( Calidris  alpina).  Annales 
Zoologici  Fennici  4:158-198. 


van  Gils,  J.  and  P.  Wiersma.  1996.  Family  Scolopa- 
cidae.  Pages  489-533  in  Handbook  of  the  birds 
of  the  world,  vol.  3 (J.  del  Hoyo,  A.  Elliott,  and 
J.  Sargatal,  Eds.).  Lynx  Edicions,  Barcelona, 
Spain. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):485-493,  2006 


EFFECTS  OF  HUMAN  RECREATION  ON  THE  INCUBATION 
BEHAVIOR  OF  AMERICAN  OYSTERCATCHERS 

CONOR  P.  McGOWAN1’3’4  AND  THEODORE  R.  SIMONS2 


ABSTRACT. — Human  recreational  disturbance  and  its  effects  on  wildlife  demographics  and  behavior  is  an 
increasingly  important  area  of  research.  We  monitored  the  nesting  success  of  American  Oystercatchers  ( Hae - 
matopus  palliatus)  in  coastal  North  Carolina  in  2002  and  2003.  We  also  used  video  monitoring  at  nests  to 
measure  the  response  of  incubating  birds  to  human  recreation.  We  counted  the  number  of  trips  per  hour  made 
by  adult  birds  to  and  from  the  nest,  and  we  calculated  the  percent  time  that  adults  spent  incubating.  We  asked 
whether  human  recreational  activities  (truck,  all-terrain  vehicle  [ATV],  and  pedestrian  traffic)  were  correlated 
with  parental  behavioral  patterns.  Eleven  a priori  models  of  nest  survival  and  behavioral  covariates  were  eval- 
uated using  Akaike’s  Information  Criterion  (AIC)  to  see  whether  incubation  behavior  influenced  nest  survival. 
Factors  associated  with  birds  leaving  their  nests  (n  = 548)  included  ATV  traffic  (25%),  truck  traffic  (17%), 
pedestrian  traffic  (4%),  aggression  with  neighboring  oystercatchers  or  paired  birds  exchanging  incubation  duties 
(26%),  airplane  traffic  (1%)  and  unknown  factors  (29%).  ATV  traffic  was  positively  associated  with  the  rate  of 
trips  to  and  away  from  the  nest  (p,  = 0.749,  P < 0.001)  and  negatively  correlated  with  percent  time  spent 
incubating  (3,  = -0.037,  P = 0.025).  Other  forms  of  human  recreation  apparently  had  little  effect  on  incubation 
behaviors.  Nest  survival  models  incorporating  the  frequency  of  trips  by  adults  to  and  from  the  nest,  and  the 
percentage  of  time  adults  spent  incubating,  were  somewhat  supported  in  the  AIC  analyses.  A low  frequency  of 
trips  to  and  from  the  nest  and,  counter  to  expectations,  low  percent  time  spent  incubating  were  associated  with 
higher  daily  nest  survival  rates.  These  data  suggest  that  changes  in  incubation  behavior  might  be  one  mechanism 
by  which  human  recreation  affects  the  reproductive  success  of  American  Oystercatchers.  Received  28  July  2005, 
accepted  24  April  2006. 


The  effect  of  human  recreational  activity  on 
wildlife  is  an  increasingly  important  area  of 
research  (Burger  1981,  Burger  and  Gochfeld 
1998,  Fitzpatrick  and  Bouchez  1998,  Whitta- 
ker and  Knight  1998,  Carney  and  Sydeman 
1999).  Human  disturbance  has  been  linked  to 
altered  foraging  behavior  (Burger  1981,  Bur- 
ger and  Gochfeld  1998,  Fitzpatrick  and 
Bouchez  1998,  Rodgers  and  Schwikert  2003, 
Stolen  2003)  and  diminished  reproductive 
success  of  many  waterbird  species  (Hunt 
1972,  Robert  and  Ralph  1975,  Tremblay  and 
Ellison  1979,  Safina  and  Burger  1983,  Rhulen 
et  al.  2003).  The  mechanisms  by  which  human 
disturbance  lowers  reproductive  success,  how- 
ever, are  poorly  understood. 

Current  data  indicate  that  American  Oys- 


1 North  Carolina  Coop.  Fish  and  Wildlife  Research 
Unit,  Dept,  of  Zoology,  North  Carolina  State  Univ., 
Campus  Box  7617,  Raleigh,  NC  27695,  USA. 

2 U.S.  Geological  Survey,  North  Carolina  Coop. 
Fish  and  Wildlife  Research  Unit,  Dept,  of  Zoology, 
North  Carolina  State  Univ.,  Campus  Box  7617,  Ra- 
leigh, NC  27695,  USA. 

3 Current  address:  Dept,  of  Fisheries  and  Wildlife, 
302  Anheuser  Busch  Natural  Resources  Bldg.,  Univ. 
of  Missouri,  Columbia,  MO  6521 1,  USA. 

4 Corresponding  author;  e-mail: 
cpm4h9  @ mizzou.edu 


tercatcher  ( Haematopus  palliatus ) populations 
in  the  Mid-Atlantic  states  are  declining  (Ma- 
whinney  and  Bennedict  1999,  Davis  et  al. 
2001).  The  U.S.  Shorebird  Conservation  Plan 
lists  the  American  Oystercatcher  as  a “Spe- 
cies of  High  Concern,”  due,  in  part,  to  human 
encroachment  on  breeding  habitat  (Brown  et 
al.  2001).  Evidence  that  humans  are  directly 
responsible  for  American  Oystercatcher  nest 
failure  is  limited  (Davis  et  al.  2001,  McGowan 
2004);  however,  human  recreation  is  often  as- 
sociated with  lower  oystercatcher  reproduc- 
tive success  (Hockey  1987,  Jeffery  1987, 
Novick  1996,  Davis  1999,  Leseberg  et  al. 
2000,  Verhulst  et  al.  2001,  McGowan  2004). 
Because  American  Oystercatcher  populations 
may  require  intensive  management  in  the  near 
future,  it  is  important  to  understand  the  rela- 
tionship between  human  recreation  and  oys- 
tercatcher nesting  success  (Brown  et  al.  2001, 
Davis  et  al.  2001). 

Skutch  (1949)  hypothesized  that  higher  lev- 
els of  parental  activity  during  the  nesting  pe- 
riod might  lead  to  greater  rates  of  predation 
because  more  activity  makes  nests  more  ob- 
vious to  predators.  Because  American  Oyster- 
catchers are  ground-nesting  shorebirds  that  are 
easily  flushed  from  their  nests  (Davis  1999), 


485 


486 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


we  similarly  hypothesized  that  human  recre- 
ation might  increase  the  activity  of  incubating 
oystercatchers,  thereby  leading  to  increased 
predation  rates.  Although  Skutch’s  hypothesis 
has  been  tested  extensively,  conclusions  are 
mixed  (Martin  1992,  Roper  and  Goldstein 
1987,  Martin  et  al.  2000,  Tewksbury  et  al. 
2002).  We  believe  that  nesting  American  Oys- 
tercatchers provide  a good  opportunity  to  test 
Skutch’s  hypothesis  because  their  nests  are 
relatively  easy  to  find  and  monitor,  and  they 
experience  high  rates  of  nest  predation  (Nol 
and  Humphrey  1994.  Davis  et  al.  2001,  Sa- 
bine et  al.  2005). 

In  this  study,  we  used  video  monitoring  to 
record  human  recreational  activity  and  the  be- 
havior of  incubating  oystercatchers  nesting  on 
the  Outer  Banks  of  North  Carolina.  We  asked 
whether  human  recreational  activity  altered 
the  behavior  of  nesting  birds,  and  whether  in- 
creased parental  activity  or  decreased  nest  at- 
tendance were  associated  with  higher  rates  of 
nest  failure. 

METHODS 

Study  areas. — We  monitored  nesting  suc- 
cess of  American  Oystercatchers  at  Cape 
Lookout  (76°  32'  W,  34°  36'  N)  and  Cape  Hat- 
teras  (75°  31'  W,  35°  24'  N)  national  seashores 
in  North  Carolina  during  2002  and  2003.  The 
seashores  comprise  >160  km  of  barrier  island 
habitat  that  supports  —90  breeding  pairs  of 
American  Oystercatchers.  All  work  at  Cape 
Lookout  National  Seashore  was  conducted  on 
North  Core  Banks  and  South  Core  Banks  (see 
Godfrey  and  Godfrey  1976  for  site  descrip- 
tion). Cape  Hatteras  National  Seashore  com- 
prises three  main  islands:  Bodie,  Hatteras.  and 
Ocracoke  Islands.  These  barrier  islands  are 
long,  narrow,  and  bordered  by  sandy  beaches 
on  the  ocean  side  and  salt  marshes  on  the 
sound  side.  American  Oystercatchers  nest  on 
the  ocean  side  beaches,  dunes,  and  adjacent 
sand  flats.  Raccoons  ( Procyon  lotor ) and  feral 
cats  ( Felis  catus ) are  common  on  all  islands 
except  Ocracoke,  which  has  no  raccoons.  The 
islands  are  open  to  the  public  and  most  beach- 
es are  open  to  vehicles.  Approximately 
650,000  people  visit  Cape  Lookout  each  year: 
the  visitation  rate  at  Cape  Hatteras  is  consid- 
erably higher  and  has  increased  steadily  from 
1.5  million  in  1986  to  2.2  million  in  2005 
(National  Park  Service  2005).  Park  visitors 


use  the  beaches  for  walking,  shell  collecting, 
swimming,  and  fishing,  and  they  drive  four- 
wheel  drive  passenger  vehicles  (ORVs)  and 
smaller,  all-terrain  vehicles  (ATVs)  on  the 
beach.  Vehicles  are  permitted  along  a network 
of  unpaved  roads  behind  the  primary  dunes 
and  anywhere  on  the  open  beach,  except  in 
designated  areas  that  are  closed  to  protect  veg- 
etation, nesting  sea  turtles,  and  shorebirds,  and 
to  prevent  erosion. 

Data  collection. — We  located  oystercatcher 
nests  ( n — 268)  and,  from  15  April  until  30 
July  in  2002  and  2003,  checked  their  status 
every  3-4  days  until  chicks  hatched  or  the 
nests  failed.  We  used  SONY  HI-8  video  cam- 
eras to  record  the  incubation  behavior  of  nest- 
ing adults  at  randomly  selected  nests  ( n = 72). 
We  videotaped  nests  on  Bodie  Island  and  Hat- 
teras Island  (Cape  Hatteras  National  Sea- 
shore), and  on  North  Core  Banks  and  South 
Core  Banks  (Cape  Lookout  National  Sea- 
shore). Nests  were  filmed  for  approximately 
4-hr  intervals  at  least  once  between  the  com- 
pletion of  egg  laying  and  hatching.  In  the  ab- 
sence of  human  recreational  activity,  we  as- 
sumed that  parental  behavior  would  be  natural 
and  homogenous  throughout  the  incubation 
period.  Evidence  indicates  that  both  American 
and  Black  ( Haematopus  bachmani)  oyster- 
catchers incubate  their  eggs  90-100%  of  the 
time  once  the  clutch  is  completed,  and  that  the 
amount  of  time  spent  incubating  does  not  vary 
during  the  incubation  period  (Nol  and  Hum- 
phrey 1994,  Andres  and  Falxa  1995).  Verbo- 
ven  et  al.  (2001)  showed  that  Eurasian  Oys- 
tercatchers incubated  85-90%  of  the  time  at 
undisturbed  nests,  and  that  the  percentage  of 
time  spent  incubating  was  constant  between 
the  end  of  the  laying  period  and  hatching. 
Studies  of  other  shorebird  species  indicate 
similar  incubation  patterns  (Norton  1972),  al- 
though Cartar  and  Montgomerie  (1987)  found 
that  nest  attendance  of  White-rumped  Sand- 
pipers ( Calidris  fuscicollis)  may  vary  daily, 
depending  on  weather  or  other  environmental 
factors. 

Novick  (1996)  reported  that  human  activity 
on  South  Core  Banks  at  Cape  Lookout  Na- 
tional Seashore  was  distributed  “fairly  even- 
ly” throughout  the  day  and  was  greater  on 
weekends  (Friday-Sunday)  than  on  weekdays. 
Novick  (1996)  also  reported  that  humans  con- 
centrated around  activity  centers,  such  as  the 


McGowan  and  Simons  • HUMAN  RECREATION  ALTERS  INCUBATION  BEHAVIOR 


487 


ferry  dock,  the  lighthouse,  and  the  ocean  in- 
lets at  the  north  and  south  ends  of  South  Core 
Banks.  Our  nests  were  filmed  between  07:00 
and  14:00  EST,  on  both  weekdays  and  week- 
ends, which  we  believe  provided  an  unbiased 
representation  of  human  disturbance  and  pa- 
rental activity  patterns  at  each  nest. 

Each  video  camera  was  housed  in  a weath- 
erproof plastic  container  attached  to  a metal 
stand,  and  placed  approximately  5 m from  the 
nest  to  avoid  disturbing  incubating  birds. 
Most  cameras  faced  the  ocean  and  recorded 
activity  both  in  the  vicinity  of  the  nest  and  on 
open  beach  beyond  the  nest.  Sometimes  cam- 
eras were  placed  at  nests  located  in  the  dunes 
or  other  locations  where  the  ocean-side  beach 
was  not  visible.  In  these  cases,  we  directed 
cameras  toward  the  most  likely  source  of  hu- 
man recreation  (e.g.,  the  dune  road  at  Cape 
Lookout).  The  area  sampled  by  the  video 
camera  was  different  for  each  nest  due  to  dif- 
ferences in  the  surrounding  landscape;  there- 
fore, detection  probabilities  for  human  activ- 
ities were  heterogeneous  among  nests.  We  re- 
viewed tapes  in  real  time  to  count  the  number 
of  trips  by  incubating  birds  to  and  from  the 
nest  per  hr,  and  the  percent  time  that  adults 
spent  incubating.  Herein,  the  term  “trip”  re- 
fers to  a bird  leaving  or  returning  to  its  nest. 
We  also  counted  the  number  of  ORVs,  ATVs, 
and/or  pedestrians  passing  each  nest  per  hr. 

Statistical  analyses. — We  used  the  Mayfield 
(1961,  1975)  method  to  estimate  daily  nest 
survival  rates  and  hatching  success  for  all 
nests  monitored.  We  applied  the  Mayfield  es- 
timate to  entire  clutches  and  did  not  consider 
individual  egg  survival.  Heterogeneity  in  sur- 
vival probabilities  during  the  incubation  stage 
was  not  considered,  and  the  midpoint  rule  was 
used  to  designate  the  time  of  failure  and  time 
of  hatching  for  nests  that  failed  or  hatched  be- 
tween visits.  We  considered  nests  successful 
if  at  least  one  egg  hatched,  and  failed  when 
all  eggs  were  lost.  Partial  nest  failure  was  not 
considered  in  this  study. 

Each  time  a bird  left  its  nest  we  estimated 
the  time  between  departure  and  the  time  at 
which  the  probable  causal  event  occurred. 
Possible  causal  factors  included:  ATV,  ORV, 
pedestrian,  and  airplane  traffic,  as  well  as  in- 
teractions between  territorial  pairs  and  ex- 
changes in  incubation  duties.  We  report  these 
data  as  the  percent  of  nest  departures  for 


which  one  of  the  above  causal  factors  fol- 
lowed. We  also  report  the  percent  of  observed 
human  recreational  activities  that  were  pre- 
ceded by  a bird  leaving  its  nest. 

We  used  linear  regression  models  (Neter  et 
al.  1996)  to  determine  whether  human  recre- 
ational factors  were  correlated  with  oyster- 
catcher  parental  activity.  Trips  per  hr  and  per- 
cent time  spent  incubating  were  modeled  as 
dependant  variables,  with  number  of  ORVs, 
ATVs,  and  pedestrians  passing  a nest  per  hr 
serving  as  the  independent  variables. 

For  camera-monitored  nests,  we  used  the 
logistic  exposure  method  to  estimate  daily 
nest  survival  (Shaffer  2004).  We  used  SAS 
(ver.  9.1;  SAS  Institute,  Inc.  2003)  to  generate 
survival  estimates  and  to  test  competing  mod- 
els of  nest  survival  with  parental  behaviors  as 
covariates  (Shaffer  and  Thompson  in  press). 
We  tested  1 1 a priori  models  (Table  1 ) that 
modeled  trip  rate  and  percent  time  incubating 
as  both  continuous  and  categorical  variables. 
We  used  two  methods  for  categorizing  the 
data:  one  purely  statistical  and  one  based  on 
behavioral  observations.  For  statistical  cate- 
gorical models,  we  split  the  data  for  number 
of  trips/hr  (Tripcat)  and  percent  time  incubat- 
ing (Inccat)  into  low  and  high  categories,  us- 
ing the  median  value  of  each  as  the  cut-off 
point  (Tripcatl:  <3.69  trips/hr  = low,  >3.69 
trips/hr  = high;  Inccat  1:  <85%  = low,  >85% 
= high).  For  the  second  method  (biological 
categorical  models),  we  used  the  average  val- 
ues from  seven  nests  that  had  no  evidence  of 
human  disturbance;  we  then  divided  the  data 
into  a new  set  of  low  and  high  categories.  In 
this  case,  undisturbed  nests  averaged  2.25 
trips  per  hr.  Therefore,  we  used  three  trips  per 
hr  as  a conservative  estimate  of  oystercatcher 
nest  site  activity  in  the  absence  of  human  dis- 
turbance (Tripcat2:  <3.0  trips/hr  = low,  >3.0 
trips/hr  = high).  Time  spent  incubating  by  un- 
disturbed birds  averaged  90%  of  the  obser- 
vation period;  thus,  we  used  90%  as  the  cut- 
off point  to  categorize  nests  as  low  or  high  in 
terms  of  percent  time  spent  incubating  (Inc- 
cat2:  <90%  = low,  >90%  = high).  We  mod- 
eled each  categorical  variable  separately  and 
in  a model  that  included  both  trip  rate  and 
percent  time  incubating  (Table  1).  One  model 
included  a year  effect,  and  we  tested  a null 
model  (null)  that  assumed  constant  survival 
over  the  season.  We  used  an  information  the- 


488 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  1.  Eleven  candidate  models  used  to  examine  the  relationship  between  daily  nest  survival  and  pa- 
rental incubation  behaviors  of  American  Oystercatchers  nesting  on  the  Outer  Banks  of  North  Carolina  in  2002 
and  2003. 


Candidate  model 


Model  covariates 


Global  Continuous  Year,  trips  to  and  from  the  nest  per  hr,  percent  incubation  time 

Year  Year 

Models  with  statistically  categorized  data  (splitting  low  and  high  data  at  the  median  value) 

Global  categorization  1 Year,  tripcatl,  inccatl3 

Tripcatl  + inccatl  Tripcatl,  inccatl 

Tripcat  1 Tripcat  1 

Inccatl  Inccatl 

Models  with  biologically  categorized  data  (splitting  data  at  the  average  value  for  undisturbed  nests) 

Global  categorization  2 Year,  tripcat2,  inccat2 

Tripcat2  + inccat2  Tripcat2,  inccat2 

Tripcat2  Tripcat2 

Inccat2  Inccat2 

Null  No  covariates,  assumes  constant  survival 

a Inccatl,  inccat2,  tripcatl,  and  tripcat2  are  categorical  variables  into  which  nests  were  categorized  as  low  or  high  in  terms  of  percent  time  adult  birds 
spent  incubating  (inccat)  or  the  number  of  trips  adults  made  to  and  from  the  nest/hr  (tripcat),  according  to  the  criteria  that  follow:  inccatl:  <85%  = low, 
>85%  = high;  inccat2:  <90%  = low,  >90%  = high;  tripcatl:  <3.69  trips/hr  = low,  >3.69  trips/hr  = high;  tripcat2:  ^3.0  trips/hr  = low,  >3.0  trips/hr 
= high. 


oretic  approach  to  rank  the  models  from  most 
to  least  supported,  based  on  Akaike’s  Infor- 
mation Criterion  (AIC) — using  AICc,  AAICc, 
and  Akaike  weights  (w,);  Burnham  and  An- 
derson 2002).  Means  are  reported  ±SE. 

RESULTS 

We  monitored  185  nests  at  Cape  Lookout 
and  83  nests  at  Cape  Hatteras.  The  overall 
Mayfield  estimate  of  daily  nest  survival  was 
0.92  ± 0.006  at  Cape  Lookout  and  0.94  ± 
0.007  at  Cape  Hatteras.  The  highest  daily  nest 
survival  rates  were  recorded  at  Cape  Hatteras 
in  2003  (0.96  ± 0.008),  and  the  lowest  were 
recorded  at  Cape  Lookout  in  2002  (0.90  ± 
0.007);  these  were  the  only  year  and  location 
comparisons  that  were  significantly  different 
(Z  = 4.83,  P < 0.001). 

We  filmed  72  nests  for  a total  of  320.18  hr 
and  a mean  of  4.45  ± 1.19  hr  per  nest.  Most 
nests  were  filmed  once  for  ~4  hr,  but  some 
were  filmed  twice  before  they  hatched  or 
failed.  We  excluded  one  nest  from  the  analysis 
where  it  appeared  that  the  bird’s  behavior  was 
affected  by  the  presence  of  the  video  camera. 
Of  the  72  nests  filmed,  chicks  successfully 
hatched  from  19  and  53  nests  failed.  Sixty  two 
percent  of  nest  failures  were  due  to  mamma- 
lian predation  (n  — 32),  28.5%  failed  for  un- 
known reasons  ( n = 15),  and  11%  were  lost 


to  weather,  human  destruction,  or  abandon- 
ment ( n = 6). 

Though  not  true  experimental  controls, 
there  were  seven  nests  at  which  we  observed 
no  human  disturbance  during  filming.  Birds  at 
those  nests  incubated  for  90%  ± 0.033  of  the 
filming  period  and  made  2.25  ± 0.60  trips/hr 
compared  to  82%  ± 0.017  incubation  and 
3.66  ± 0.17  trips/hr  at  all  other  nests.  The 
number  of  trips/hr  at  undisturbed  nests  was 
significantly  lower  ( t — 2.27 , P = 0.026)  than 
at  all  other  nests.  The  percent  of  time  spent 
incubating  at  undisturbed  nests  was  not  sig- 
nificantly greater  ( t = 1.34,  P = 0.19)  than  it 
was  at  disturbed  nests. 

We  recorded  539  instances  in  which  incu- 
bating birds  departed  their  nests.  Of  those  in- 
stances, ATVs  were  filmed  within  3 min  of 
nest  departure  on  136  occasions  (25%)  and 
ORVs  were  filmed  92  times  (17%)  within  3 
min  of  departure.  We  recorded  a total  of  284 
ATVs,  62%  ( n = 177)  of  which  passed  by  a 
nest  within  <3  min  of  a bird  departing  its 
nest.  We  observed  1,466  ORVs  pass  by  filmed 
nests,  but  only  1 1%  (n  = 168)  passed  by  with- 
in 3 min  of  a bird  leaving  its  nest.  Groups  or 
individual  pedestrians  were  filmed  19  times 
(4%)  within  10  min  of  nest  departures.  Of  all 
the  110  pedestrians  that  we  observed,  33%  ( n 
= 36)  passed  by  within  10  min  of  a bird  de- 


McGowan  and  Simons  • HUMAN  RECREATION  ALTERS  INCUBATION  BEHAVIOR 


489 


FIG.  1.  The  effect  of  all-terrain  vehicle  (ATV) 
beach  traffic  on  incubation  behavior  of  American  Oys- 
tercatchers  on  the  Outer  Banks  of  North  Carolina  dur- 
ing the  2002  and  2003  breeding  seasons:  (A)  relation- 
ship between  the  percent  of  time  spent  incubating  and 
the  average  number  of  AT  Vs  passing  per  hour  (P,  = 
-0.037,  P = 0.025),  and  (B)  relationship  between  the 
number  of  trips  to  and  from  the  nest  per  hr  and  the 
number  of  ATVs  passing  per  hr  (P,  = 0.749,  P < 
0.001). 


parting  its  nest.  Eight  percent  ( n = 44)  of  nest 
departures  were  associated  with  territorial  dis- 
putes and  18%  (n  = 108)  with  the  exchange 
in  incubation  duties.  Eight  departures  (1%) 
were  associated  with  low-flying  airplanes  that 
passed  within  3 min  of  nest  departure.  For  the 
remaining  29%  ( n = 154)  of  nest  departures, 
no  disturbances,  territorial  interactions,  or  in- 
cubation exchanges  took  place  following  de- 
parture. 

Regression  models  showed  that  there  was 
little  or  no  association  between  ORV  traffic 
and  the  rate  at  which  incubating  oystercatch- 
ers  made  trips  to  and  from  their  nests  ((3,  = 
0.018,  P = 0.064)  or  the  percent  time  they 
spent  incubating  (f3t  — 0.0006,  P = 0.57). 
Likewise,  pedestrian  traffic  was  not  associated 
with  a significant  reduction  in  the  percent  time 
incubating  ((3t  = -0.005,  P = 0.75)  or  birds 
making  more  trips  to  and  from  their  nests  per 
hr  ((3!  = -0.268,  P = 0.079).  Increased  ATV 
traffic,  however,  was  associated  with  a reduc- 
tion in  the  percent  time  spent  incubating  ((3! 
= —0.037,  P = 0.025)  and  an  increase  in  the 
rate  of  trips  to  and  from  the  nest  ((3!  = 0.749, 
P < 0.001;  Fig.  1). 

All  models  except  the  global  continuous 
model  received  some  level  of  support,  but  no 
model  had  overwhelming  support  (Table  2). 
The  tripcat2  model  (i.e.,  nests  divided  into 
low  and  high  categories  based  on  average  trip 
rate  for  nests  with  no  observed  human  distur- 


TABLE  2.  Candidate  models  examining  the  relationship  between  daily  nest  survival  and  parental  incubation 
behaviors  of  American  Oystercatchers  nesting  on  the  Outer  Banks  of  North  Carolina  in  2002  and  2003.  Models 
are  ranked  in  descending  order  of  support  based  on  Akaike’s  information  criteria  AICc,  AAICc,  and  Akaike 
weights  (w,). 

Model 

Log-likelihood 

No.  parameters 

AICc 

AAICc 

W/ 

Tripcat2a 

-159.62 

2 

323.27 

0.00 

0.28 

Null 

-161.08 

1 

324.16 

0.89 

0.18 

Tripcat2  -1-  inccat2a 

-159.62 

3 

325.29 

2.02 

0.10 

Inccatl 

-160.68 

2 

325.39 

2.11 

0.097 

Inccat2 

-160.77 

2 

325.56 

2.29 

0.089 

Tripcat  1 

-160.98 

2 

325.99 

2.72 

0.072 

Year 

-161.07 

2 

326.17 

2.90 

0.066 

Tripcatl  + inccatl 

-160.26 

3 

326.56 

3.29 

0.054 

Global  categorical2 

-159.56 

4 

327.18 

3.92 

0.040 

Global  categorical  1 

-160.24 

4 

328.54 

5.28 

0.020 

Global  continuous 

-261.36 

4 

530.79 

207.52 

0.000 

a Inccatl,  inccat2,  tripcatl,  and  tripcat2  are  categorical  variables  into  which  nests  were  categorized  as  low  or  high  in  terms  of  percent  time  adult  birds 
spent  incubating  (inccat)  or  the  number  of  trips  adults  made  to  and  from  the  nest/hr  (tripcat),  according  to  the  criteria  that  follow:  inccatl:  <85%  = low, 
>85%  = high;  inccat2:  <90%  = low,  >90%  = high;  tripcatl:  <3.69  trips/hr  = low,  >3.69  trips/hr  = high;  tripcat2:  <3.0  trips/hr  = low,  >3.0  trips/hr 
= high. 


490 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  3.  Daily  survival  estimates  and  hatching  probability  estimates  for  nests  in  two  categories  of  behav- 
ioral data  collected  from  American  Oystercatchers  nesting  on  the  Outer  Banks  of  North  Carolina  in  2002  and 
2003. 


Category 

No.  nests 

Daily  probability 
of  survival 

Lower  / upper 
confidence  intervals 

Hatching 

probability 

Median  cutoffs 

<3.69  trips/hr 

37 

0.958 

0.935  / 0.973 

0.314 

>3.69  trips/hr 

35 

0.948 

0.925  / 0.965 

0.240 

Incubation  <85% 

32 

0.961 

0.938  / 0.975 

0.338 

Incubation  >85% 

40 

0.945 

0.922  / 0.962 

0.218 

Zero-observed-disturbance  average  cutoffs 

<3.00  trips/hr 

26 

0.969 

0.946  / 0.982 

0.424 

>3.00  trips/hr 

46 

0.944 

0.924  / 0.960 

0.213 

Incubation  <90% 

50 

0.967 

0.945  / 0.980 

0.400 

Incubation  >90% 

22 

0.948 

0.926  / 0.964 

0.237 

bance  as  the  only  covariate)  had  the  highest 
rank  of  all  the  models  (AAICc  = 0.00,  w,  = 
0.28).  The  null  model  was  ranked  second 
(AAICc  = 0.89,  w,  = 0.18),  and  the  model 
incorporating  both  tripcat2  and  inccat2  was 
ranked  third  (AAICc  = 2.02,  w,  = 0.10).  All 
the  models  with  categorical  behavioral  vari- 
ables, the  year  model,  and  the  null  model  had 
a AAICc  of  <7  and  weights  between  0.02  and 
0.28  (Table  2).  Generally,  models  with  a 
AAICc  of  <7  cannot  be  ruled  out,  but  models 
with  weights  <0.70  cannot  be  exclusively  ac- 
cepted (Burnham  and  Anderson  2002). 

The  estimated  daily  survival  rate  for  nests 
with  <3.69  trips  to  and  from  the  nest  per  hr 
was  greater  than  the  daily  survival  rate  for 
nests  with  >3.69  trips  to  and  from  the  nest 
per  hr  (Table  3).  That  same  pattern  was  ob- 
served when  the  data  were  divided  into  cate- 
gories representing  nests  with  ^3  trips  per  hr 
and  >3  trips  per  hr.  Nests  in  which  the  parents 
incubated  for  <85%  of  the  observation  period 
had  higher  daily  survival  probabilities  than 
nests  in  which  incubation  percentages  were 
>85%.  The  same  pattern  was  observed  when 
we  categorized  the  data  by  nests  in  which 
adults  spent  <90%  and  >90%  time  incubat- 
ing. These  data  indicated  that  nests  in  which 
parents  made  more  trips  to  and  from  the  nest 
had  a lower  daily  survival  probability,  and 
that  nests  where  the  parents  spent  more  than 
85-90%  of  their  time  incubating  had  a lower 
chance  of  surviving  each  day. 

DISCUSSION 

Our  data  show  clear  associations  between 
human  recreation  and  incubation  behavior  of 


American  Oystercatchers.  ATV  traffic  was  as- 
sociated with  increased  rates  of  trips  to  and 
from  the  nest  and  reduced  time  incubating; 
other  forms  of  human  recreation  were  more 
weakly  associated  with  oystercatcher  nesting 
behaviors.  Sixty  two  percent  of  the  ATVs  that 
we  observed  passed  within  3 min  of  a bird 
departing  its  nest,  whereas  the  same  was  true 
for  only  1 1%  of  the  OR  Vs  that  we  observed. 
Birds  appear  to  have  habituated  to  the  pres- 
ence of  OR  Vs  (Whittaker  and  Knight  1998), 
but  they  view  ATVs  (and  to  a lesser  extent, 
pedestrians)  as  threats.  Peters  and  Otis  (2005) 
reported  that  wintering  American  Oyster- 
catchers habituated  to  boat  traffic  on  the  in- 
tercoastal waterway  in  South  Carolina.  Other 
studies  have  shown  that  birds  respond  differ- 
ently to  different  forms  of  human  recreational 
disturbance  (Burger  1981),  but  most  have  fo- 
cused only  on  changes  in  foraging  behavior 
(Burger  and  Gochfeld  1998.  Rodgers  and 
Schwikert  2003,  Stolen  2003).  Our  study  is 
one  of  the  few  to  investigate  how  human  rec- 
reational disturbance  affects  incubation  be- 
havior. ATVs  are  louder  and  move  faster  than 
ORVs  and  pedestrians,  which  might  explain 
why  the  birds  are  affected  more  by  ATV  traf- 
fic (Burger  1981,  Burger  and  Gochfeld  1998). 
ORVs  and  pedestrians  also  tend  to  stay  closer 
to  the  firm  sand  along  the  water’s  edge,  which 
means  they  generally  travel  farther  from  nest- 
ing birds. 

Although  the  probability  of  hatching  was 
low  in  all  nests,  regardless  of  parental  activity, 
we  did  find  evidence  that  human  recreational 
disturbance  may  reduce  the  nesting  success  of 


McGowan  and  Simons  • HUMAN  RECREATION  ALTERS  INCUBATION  BEHAVIOR 


491 


American  Oystercatchers  by  altering  incuba- 
tion behavior.  Analyses  based  on  AIC  model 
selection  indicated  that  the  rate  of  parental 
trips  to  and  from  the  nest  and  the  percent  time 
that  parents  spent  incubating  may  have  af- 
fected daily  nest  survival  rates.  Although  no 
model  received  overwhelming  support,  none 
of  the  categorical  behavioral  models  could  be 
ruled  out.  The  daily  survival  estimates  indi- 
cated that  nesting  adults  that  made  fewer  trips 
to  and  from  the  nest  had  greater  daily  nest 
survival  rates.  Conversely,  nests  where  the 
parents  incubated  for  less  time  had  higher  dai- 
ly survival  rates.  We  hypothesize  that  mam- 
malian nest  predators,  the  primary  nest  pred- 
ators in  this  system  (Davis  et  al.  2001),  are 
better  able  to  find  disturbed  nests  through 
smell  because  each  time  a parent  gets  up  and 
walks  away  from  a nest  it  leaves  a scent  trail 
that  raccoons  and  cats  may  follow.  Our  results 
differ  from  those  of  Verboven  et  al.  (2001), 
but  that  is  likely  because  the  primary  nest 
predators  in  that  system  were  avian  predators. 

ATV  traffic  is  not  the  only  factor  affecting 
oystercatcher  nesting  success  on  North  Caro- 
lina’s Outer  Banks.  Nest  predation  is  an  im- 
portant determinant  of  hatching  success  in  the 
Outer  Banks  (Davis  et  al.  2001,  McGowan  et 
al.  2005),  and  relationships  between  human 
recreation  and  nest  predators  are  poorly  un- 
derstood. Vehicular  traffic  also  may  affect  suc- 
cess during  the  chick-rearing  phase  of  repro- 
duction. In  the  2003  breeding  season,  we  con- 
firmed that  five  chicks  from  three  different 
nests  were  run  over  by  vehicles  on  the  beach- 
es of  South  Core  Banks  at  Cape  Lookout  Na- 
tional Seashore  and  Hatteras  Island  at  Cape 
Hatteras  National  Seashore  (McGowan  2004). 

The  negative  association  between  percent 
time  incubating  and  daily  nest  survival  seems 
counterintuitive.  Conway  and  Martin  (2000) 
showed  that  birds  balance  the  costs  of  egg  ex- 
posure with  those  of  high  parental  activity. 
Birds  with  high  levels  of  nest-predation  pres- 
sure minimize  nest-site  activity  by  taking  few- 
er, longer  trips  off  the  nest  (Conway  and  Mar- 
tin 2000).  This  behavior  helps  reduce  parental 
activity  around  the  nest,  but  it  also  reduces  the 
amount  of  incubation.  American  Oystercatch- 
er behavior  may  reflect  a similar  trade  off; 
their  eggs  can  tolerate  extensive  heating  and 
cooling  (Nol  and  Humphrey  1994).  In  our 
study,  several  clutches  exposed  for  approxi- 


mately 1 hr  at  mid  day  hatched  successfully. 
One  videotaped  nest  hatched  successfully, 
even  though  the  parents  incubated  for  only 
66.8%  of  the  4.07-hr  observation  period.  Egg 
hardiness  may  reflect  an  adaptation  that  en- 
ables parents  to  reduce  nest-site  activity.  Par- 
ents that  depart  their  nest  and  wait  until  mul- 
tiple disturbances  have  passed  before  return- 
ing may  have  greater  nesting  success  than  par- 
ents that  return  to  their  nests  quickly  and  flush 
repeatedly.  Future  analyses  should  assess  the 
effect  that  the  average  amount  of  time  birds 
spend  off  the  nest  has  on  nest  success. 

There  were  several  potential  sources  of 
measurement  error  in  our  study  that  might  ex- 
plain why  no  models  were  strongly  supported. 
Incubation  behavior  might  vary  as  birds  ha- 
bituate to  disturbance  (Whittacker  and  Knight 
1998).  Because  the  field  of  view  varied  at 
each  nest,  our  cameras  recorded  areas  of  dif- 
ferent size  for  each  nest,  and  we  were  unable 
to  control  for  these  differences  in  the  analyses. 
We  were  also  unable  to  measure  the  distance 
from  the  nests  to  the  disturbance  recorded  on 
our  video.  Several  studies  have  shown  that  the 
proximity  of  human  disturbance  has  a major 
effect  on  the  behavioral  responses  of  birds 
(Burger  and  Gochfeld  1998,  Rodgers  and 
Schwikert  2003).  It  is  likely  that  in  some  cas- 
es, recreational  activity  recorded  by  our  cam- 
eras did  not  elicit  a response  from  the  incu- 
bating bird  because  the  activity  was  too  far 
away.  Video  monitoring  is  an  extremely  use- 
ful tool  for  studying  avian  behavior;  however, 
future  studies  of  human  disturbance  using  vid- 
eo monitoring  should  entail  measuring  dis- 
tances to  sources  of  disturbance.  Recording 
nests  for  longer  periods  of  time  also  would 
alleviate  a great  deal  of  uncertainty.  Sabine  et 
al.  (2005)  were  very  successful  in  studying 
nest  success  of  oystercatchers  in  Georgia  by 
using  time-lapse  videography  throughout  the 
incubation  period. 

Our  simplified  approach  of  categorizing 
nests  into  low  or  high  levels  of  parental  activ- 
ity provided  a coarse-scale  observational  mea- 
sure of  behavioral  responses  to  recreation  and 
disturbance;  we  expected  this  to  reduce  ob- 
servation errors.  Other  researchers  that  have 
evaluated  the  effects  of  human  disturbance  on 
avian  behavior  used  experimental  designs 
with  defined  treatment  groups  (Robert  and 
Ralph  1975,  Tremblay  and  Ellison  1979,  Ver- 


492 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


hulst  et  al.  2001,  Stolen  2003).  We  studied  the 
effects  of  ambient  human  disturbance  caused 
by  park  staff  and  recreational  visitors  to  de- 
termine whether  it  was  linked  to  patterns  of 
nesting  success.  Future  studies  of  human  ac- 
tivity and  oystercatcher  nesting  success  that 
compare  the  behavior  of  birds  on  beaches 
closed  to  vehicle  and  pedestrian  traffic  with 
the  behavior  of  birds  exposed  to  different 
types  and  intensities  of  human  activity  are 
needed  to  improve  our  understanding  of  the 
patterns  suggested  by  this  study. 

ACKNOWLEDGMENTS 

We  are  grateful  to  J.  R.  Cordes  and  M.  Lyons  for 
their  tremendous  contributions  to  this  study.  We  thank 
the  National  Park  Service,  the  U.S.  Fish  and  Wildlife 
Service,  and  the  U.S.  Geological  Survey  for  support- 
ing this  research.  We  thank  M.  W.  Rikard  and  the  staff 
of  Cape  Lookout  National  Seashore,  and  S.  Harrison 
and  the  staff  at  Cape  Hatteras  National  Seashore,  for 
all  their  assistance.  We  thank  J.  A.  Collazo,  K.  H.  Pol- 
lock, and  F.  R.  Thompson,  III,  for  assistance  on  the 
design  and  analysis  of  this  study,  and  J.  Hostetler  and 
M.  S.  Pruett  for  their  assistance  with  data  collection 
and  management.  We  thank  C.  C.  McGowan  for  ed- 
iting this  manuscript.  We  are  grateful  to  three  anony- 
mous reviewers  for  their  comments  and  suggestions 
for  improving  this  manuscript. 

LITERATURE  CITED 

Andres,  B.  A.  and  G.  A.  Falxa.  1995.  Black  Oyster- 
catcher  ( Haematopus  bachmani).  The  Birds  of 
North  America,  no.  155. 

Brown,  S.,  C.  Hickey,  B.  Harrington,  and  R.  Gill 
(Eds.).  2001.  The  U.S.  shorebird  conservation 
plan,  2nd  ed.  Manomet  Center  for  Conservation 
Sciences,  Manomet,  Massachusetts. 

Burger,  J.  1981 . The  effect  of  human  activity  on  birds 
at  a coastal  bay.  Biological  Conservation  21:231- 
241. 

Burger,  J.  and  M.  Gochfeld.  1998.  Effects  of  eco- 
tourism  on  bird  behavior  at  Loxahatchee  National 
Wildlife  Refuge,  Florida.  Environmental  Conser- 
vation 25:13-21. 

Burnham,  K.  P.  and  D.  R.  Anderson.  2002.  Model 
selection  and  multimodel  inference:  a practical  in- 
formation-theoretic  approach.  Springer- Verlag, 
New  York. 

Carney,  K.  M.  and  W.  J.  Sydeman.  1999.  A review 
of  human  disturbance  effects  on  nesting  colonial 
waterbirds.  Waterbirds  22:68-79. 

Cartar,  R.  D.  and  R.  D.  Montgomerie.  1987.  Day- 
to-day  variation  in  nest  attentiveness  of  White- 
rumped  Sandpipers.  Condor  89:252-260. 
Conway,  C.  and  T.  Martin.  2000.  Evolution  of  pas- 
serine incubation  behavior:  influence  of  food,  tem- 


perature and  nest  depredation.  Evolution  54:670- 
685. 

Davis,  M.  B.  1999.  Reproductive  success,  status  and 
viability  of  American  Oystercatcher  ( Haematopus 
palliatus ).  M.Sc.  thesis,  North  Carolina  State  Uni- 
versity, Raleigh. 

Davis,  M.  B.,  T.  R.  Simons,  M.  J.  Groom,  J.  L.  Weav- 
er, and  J.  R.  Cordes.  2001.  The  breeding  status 
of  the  American  Oystercatcher  on  the  East  Coast 
of  North  America  and  breeding  success  in  North 
Carolina.  Waterbirds  24:195-202. 

Fitzpatrick,  S.  and  B.  Bouchez.  1998.  Effects  of  rec- 
reational disturbance  on  the  foraging  behaviours 
of  waders  on  a rocky  beach.  Bird  Study  45:157- 
171. 

Godfrey,  P.  G.  and  M.  M.  Godfrey.  1976.  Barrier 
island  ecology  of  Cape  Lookout  National  Sea- 
shore and  vicinity.  North  Carolina.  U.S.  Govern- 
ment Printing  Office,  Washington,  D.C. 

Hockey,  P.  A.  R.  1987.  The  influence  of  coastal  uti- 
lization by  man  on  the  presumed  extinction  of  the 
Canarian  Black  Oystercatcher.  Biological  Conser- 
vation 39:49-62. 

Hunt,  G.  L.  1972.  Influence  of  food  distribution  and 
human  disturbance  on  the  reproductive  success  of 
Herring  Gulls.  Ecology  53:1051-1061. 

Jeffery,  R.  G.  1987.  Influence  of  human  disturbance 
on  the  nesting  success  of  African  Black  Oyster- 
catchers.  South  African  Journal  of  Wildlife  Re- 
search 17:71-72. 

Leseberg,  A.,  P.  A.  R.  Hockey,  and  D.  Loewenthal. 
2000.  Human  disturbance  and  the  chick-rearing 
ability  of  African  Black  Oystercatcher:  a geo- 
graphic perspective.  Biological  Conservation  96: 
379-385. 

Martin,  T.  E.  1992.  Interaction  of  nest  predation  and 
food  limitation  in  reproductive  strategies.  Current 
Ornithology  9:163-197. 

Martin,  T.  E„  J.  Scott,  and  C.  Menge.  2000.  Nest 
predation  increases  with  parental  activity:  sepa- 
rating nest  site  and  parental  activity  effects.  Pro- 
ceedings of  the  Royal  Society  of  London,  Series 
B 267:2287-2293. 

Mawhinney,  K.  B.,  B.  Allen,  and  B.  Benedict.  1999. 
Status  of  the  American  Oystercatcher  (. Haemato- 
pus palliatus),  on  the  Atlantic  Coast.  Northeastern 
Naturalist  6:177-182. 

Mayfield,  H.  F.  1961.  Nesting  success  calculated  from 
exposure.  Wilson  Bulletin  73:255-261. 

Mayfield,  H.  F.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

McGowan,  C.  P.  2004.  Factors  affecting  nesting  suc- 
cess of  American  Oystercatchers  ( Haematopus 
palliatus ) in  North  Carolina.  M.Sc.  thesis,  North 
Carolina  State  University,  Raleigh. 

McGowan,  C.  P,  T.  R.  Simons,  W.  Golder,  and  J. 
Cordes.  2005.  A comparison  of  American  Oys- 
tercatcher reproductive  success  on  barrier  beach 
and  river  island  habitats  in  coastal  North  Carolina. 
Waterbirds  28:150-155. 

National  Park  Service.  2005.  Park  visitation  report. 


McGowan  and  Simons  • HUMAN  RECREATION  ALTERS  INCUBATION  BEHAVIOR 


493 


National  Park  Service,  Public  Use  Statistics  Of- 
fice, Denver,  Colorado. 

Neter,  J.,  M.  H.  Kunter,  C.  J.  Nachtsheim,  and  W. 
Wasserman.  1996.  Applied  linear  statistical  mod- 
els, 4th  ed.  WCB/McGraw-Hill,  New  York. 

Nol,  E.  and  R.  C.  Humphrey.  1994.  American  Oys- 
tercatcher  ( Haematopus  palliatus ).  The  Birds  of 
North  America,  no.  82. 

Norton,  D.  W.  1972.  Incubation  schedules  of  four  spe- 
cies of  Calidridine  sandpipers  at  Barrow,  Alaska. 
Condor  74:164-176. 

Novick,  J.  S.  1996.  An  analysis  of  human  recreational 
impacts  on  the  reproductive  success  of  American 
Oystercatchers  ( Haematopus  palliatus ):  Cape 
Lookout  National  Seashore,  North  Carolina. 
M.Sc.  thesis,  Duke  University,  Durham,  North 
Carolina. 

Peters,  K.  A.  and  D.  L.  Otis.  2005.  Using  the  risk- 
disturbance  hypothesis  to  assess  the  relative  ef- 
fects of  human  disturbance  and  predation  risk  on 
foraging  American  Oystercatchers.  Condor  107: 
716-725. 

Robert,  H.  C.  and  C.  J.  Ralph.  1975.  Effects  of  hu- 
man disturbance  on  the  breeding  success  of  gulls. 
Condor  77:495-499. 

Rodgers,  J.  A.  and  S.  T.  Schwikert.  2003.  Buffer 
zone  distances  to  protect  foraging  and  loafing  wa- 
terbirds  from  disturbance  by  airboats  in  Florida. 
Waterbirds  26:437-443. 

Roper,  J.  J.  and  R.  R.  Goldstein.  1997.  A test  of  the 
Skutch  hypothesis:  does  activity  at  nests  increase 
nest  predation  risk?  Journal  of  Avian  Biology  28: 
111-116. 

Ruhlen,  T.  D.,  S.  Abbot,  L.  E.  Stenzel,  and  G.  W. 
Page.  2003.  Evidence  that  human  disturbance  re- 
duces Snowy  Plover  chick  survival.  Journal  of 
Field  Ornithology  74:300-304. 

Sabine,  J.  B„  J.  M.  Meyers,  and  S.  H.  Schweitzer. 


2005.  A simple,  inexpensive  video  setup  for  the 
study  of  avian  nest  activity.  Journal  of  Field  Or- 
nithology 76:293-297. 

Safina,  C.  and  J.  Burger.  1983.  Effects  of  human 
disturbance  on  reproductive  success  in  the  Black 
Skimmer.  Condor  85:164-171. 

SAS  Institute,  Inc.  2003.  SAS,  ver.  9.1.  SAS  Institute, 
Cary,  North  Carolina. 

Shaffer,  T.  L.  2004.  A unified  approach  to  analyzing 
nest  success.  Auk  121:526-540. 

Shaffer,  T.  L.  and  F.  R.  Thompson,  III.  In  Press.  Mak- 
ing meaningful  estimates  of  nest  survival  with 
model-based  methods.  Studies  in  Avian  Biology. 

Skutch,  A.  1949.  Do  tropical  birds  rear  as  many 
young  as  they  can  nourish?  Ibis  91:431-455. 

Stolen,  E.  D.  2003.  The  effects  of  vehicle  passage  on 
foraging  behavior  of  wading  birds.  Waterbirds  26: 
429-436. 

Tewksbury,  J.  J.,  T.  E.  Martin,  S.  J.  Hejl,  M.  J. 
Kuehn,  and  J.  W.  Jenkins.  2002.  Parental  behav- 
ior of  a cowbird  host:  caught  between  the  costs  of 
egg-removal  and  nest  predation.  Proceedings  of 
the  Royal  Society  of  London,  Series  B 269:423- 
429. 

Tremblay,  J.  and  L.  N.  Ellison.  1979.  Effects  of  hu- 
man disturbance  on  breeding  of  Black-crowned 
Night  Herons.  Auk  96:364-369. 

Verboven,  N.,  B.  J.  Ens,  and  S.  Dechesne.  2001.  Ef- 
fect of  investigator  disturbance  on  nest  attendance 
and  egg  predation  in  Eurasian  Oystercatchers. 
Auk  118:503-508. 

Verhulst,  S.,  K.  Oosterbeek,  and  B.  J.  Ens.  2001. 
Experimental  evidence  for  effects  of  human  dis- 
turbance on  foraging  and  parental  care  in  Oyster- 
catchers. Biological  Conservation  101:375-380. 

Whittacker,  D.  and  R.  L.  Knight.  1998.  Understand- 
ing wildlife  responses  to  humans.  Wildlife  Society 
Bulletin  26:312-317. 


The  Wilson  Journal  of  Ornithology  1 18(4):494-501,  2006 


MOVEMENTS  OF  LONG-TAILED  DUCKS  WINTERING  ON  LAKE 
ONTARIO  TO  BREEDING  AREAS  IN  NUNAVUT,  CANADA 

MARK  L.  MALLORY,15  JASON  AKEAROK,1  NORM  R.  NORTH,2 
D.  VAUGHAN  WESELOH,3  AND  STEPHANE  LAIR4 5 


ABSTRACT. — We  used  implanted  satellite  transmitters  to  track  the  northbound  (spring)  and  southbound  (fall) 
migration  and  possible  breeding  locations  of  three  Long-tailed  Ducks  ( Clangula  hyemalis ) wintering  on  western 
Lake  Ontario  in  Ontario,  Canada.  The  birds  exhibited  short,  rapid  migration  movements  punctuated  by  extended 
periods  of  up  to  30  days  at  staging  areas.  For  much  of  the  nesting  period  (—10  June  to  10  July),  the  birds 
remained  inland  of  western  Hudson  Bay  in  Nunavut.  During  fall  migration,  they  circumnavigated  Hudson  Bay 
to  its  eastern  coast,  opposite  the  coast  they  had  followed  in  spring,  for  a mean  travel  distance  of  6,760  km. 
Identification  of  these  previously  unknown,  key  migration  sites  fills  some  important  information  gaps  on  Long- 
tailed Ducks  in  eastern  Canada,  and  it  augments  what  is  known  about  important  coastal  marine  habitats  in  the 
Arctic.  Received  28  June  2005,  accepted  24  March  2006. 


The  Long-tailed  Duck  ( Clangula  hyemalis ; 
formerly  Oldsquaw)  is  a medium-sized  sea 
duck  with  a circumpolar  distribution,  found 
across  North  America  (Robertson  and  Savard 
2002).  It  is  purportedly  the  most  numerous 
species  of  sea  duck,  although  population  es- 
timates are  unreliable  (Bellrose  1980,  Robert- 
son and  Savard  2002).  North  American  pop- 
ulations winter  principally  along  the  Pacific 
(45°  to  60°  N)  and  Atlantic  (35°  to  53°  N) 
coasts,  where  declines  in  abundance  have 
been  reported  (Robertson  and  Savard  2002); 
some  Long-tailed  Ducks  overwinter  on  the 
Great  Lakes.  Despite  the  species’  ubiquitous 
presence  along  coasts  and  on  large  lakes  in 
winter,  and  its  widespread  breeding  distribu- 
tion, we  know  little  of  the  biology  and  move- 
ments of  this  species  other  than  what  was  re- 
ported by  Alison  (1975)  and  Peterson  and  El- 
larson  (1979).  This  is  likely  attributable  to 
three  factors:  (1)  the  species  is  not  harvested 
heavily,  so  there  has  been  little  historical  pres- 
sure to  gather  information  about  it;  (2)  it 
breeds  in  low  densities  and  is  dispersed  across 


1 Canadian  Wildlife  Service,  Box  1714,  Iqaluit,  NU 
X0A  OHO,  Canada. 

2 Canadian  Wildlife  Service,  465  Gideon  Dr.,  PO. 
Box  490,  Lambeth  Station,  London,  ON  N6P  1R1, 
Canada. 

3 Canadian  Wildlife  Service,  4905  Dufferin  St., 
Downsview,  ON  M3H  5T4,  Canada. 

4 Faculte  de  medecin  veterinaire,  Univ.  de  Montreal, 
C.P.  5000,  St-Hyacinthe,  QC  J2S  7C6,  Canada. 

5 Corresponding  author;  e-mail: 
mark.mallory@ec.gc.ca 


remote  tundra  (e.g.,  Pattie  1990),  which 
makes  banding  studies  difficult  to  initiate;  and 
(3)  its  breeding  range  lies  outside  the  areas 
covered  by  annual  North  American  waterfowl 
surveys  (Cowardin  and  Blohm  1992).  How- 
ever, recent  concern  about  population  declines 
among  many  sea  duck  species  has  prompted 
scientific  investigation  of  the  Long-tailed 
Duck  (Sea  Duck  Joint  Venture  Management 
Board  2001). 

A significant  information  need  for  the 
Long-tailed  Duck  is  the  delineation  of  areas 
used  by  different  populations  and  the  bird’s 
movements  between  breeding,  molting,  and 
wintering  areas.  Prior  observations  during 
southbound  (fall)  migration  suggested  that 
Long-tailed  Ducks  in  Hudson  and  James  bays 
move  south,  probably  along  river  systems,  to 
the  Great  Lakes  (Bellrose  1980,  Leafloor  et  al. 
1996,  Robertson  and  Savard  2002).  More  re- 
cently, technological  advances  have  allowed 
biologists  to  track  birds  remotely,  thus  provid- 
ing new  insights  into  the  movements  and  ecol- 
ogy of  many  species  (e.g.,  Brodeur  et  al.  2002, 
Robert  et  al.  2002,  Petrie  and  Wilcox  2003). 
We  use  data  gathered  from  Long-tailed  Ducks 
implanted  with  satellite  transmitters  to  de- 
scribe their  movements  (1)  from  their  capture 
in  late  winter  on  the  Canadian  Great  Lakes  to 
breeding  areas  and  (2)  during  fall  migration 
from  the  eastern  Canadian  Arctic.  We  pre- 
dicted that  Long-tailed  Ducks  would  move 
north  from  the  Great  Lakes  to  James  Bay,  nest 
along  Hudson  Bay,  and  then  return  along  the 
same  route  in  fall  migration. 


494 


Mallory  et  al.  • LONG-TAILED  DUCK  MIGRATION 


495 


METHODS 

We  captured  Long-tailed  Ducks  on  27 
March  2003  and  30  March  2004  at  the  mouth 
of  the  Niagara  River  near  the  town  of  Niagara- 
On-The-Lake  (43°  15'  N,  79°  4'  W).  To  cap- 
ture the  birds,  we  used  mist  nets  suspended 
across  observed  feeding  areas,  similar  to  pro- 
cedures described  by  Brodeur  et  al.  (2002)  for 
capturing  Harlequin  Ducks  ( Histrionicus  his- 
trionicus ).  Captures  took  place  in  the  morning 
(—06:00  EDT)  when  light  was  still  low  and 
birds  probably  had  difficulty  seeing  the  mist 
net.  Captured  birds  were  placed  in  dark  con- 
tainers and  moved  to  a nearby,  temporary  sur- 
gical suite.  We  implanted  transmitters  into 
nine  ducks,  although  only  three  provided  us 
with  migration  data.  We  believe  that  the  trans- 
mitter antenna  on  one  bird  moved  or  was  im- 
paired, as  we  received  sporadic  transmissions 
without  location  information  for  2 months  af- 
ter surgery.  The  other  five  ducks  stopped 
transmitting  within  2 weeks  of  surgery,  prob- 
ably due  to  mortality. 

Satellite  transmitters  were  supplied  by  Mi- 
crowave Telemetry,  Inc.  (Columbia,  Mary- 
land; Model  PTT-100  Implantable),  and 
weighed  approximately  39  g.  As  such,  the  tar- 
get weight  for  birds  into  which  these  trans- 
mitters would  be  implanted  was  780  g (i.e., 
transmitters  were  5%  of  their  body  mass). 
However,  we  experienced  considerable  diffi- 
culty, both  in  capturing  birds  and  in  finding 
birds  of  this  size.  At  the  time  of  implantation, 
the  three  birds  that  we  tracked  weighed  779  g 
(male),  740  g (female),  and  700  g (male); 
thus,  the  transmitters  represented  5.0,  5.3  and 
5.6%  of  their  body  mass,  respectively.  Cap- 
tured birds  (2  males,  1 female)  were  held  in 
captivity  for  302  ± 80  (SD)  min,  which  in- 
cluded 71  ±5  min  of  anesthetization  and  33 
± 9 min  of  surgery.  Each  transmitter  was  sur- 
gically inserted  in  the  right  abdominal  air  sac 
of  the  anesthetized  duck,  and  each  had  a trans- 
cutaneous antenna  that  exited  cranially  to  the 
synsacrum.  Surgical  and  anesthetic  procedures 
followed  those  described  by  Fitzgerald  et  al. 
(2001).  Birds  were  released  at  the  capture  site 
after  the  effects  of  anesthesia  wore  off. 

Radio-marked  birds  were  tracked  using  the 
ARGOS  satellite  system.  Transmitters  were 
duty-cycled  on  a schedule  of  8 hr  on  followed 
by  72  hr  off  (for  24  cycles);  subsequently  (for 


the  remainder  of  their  battery  life,  approxi- 
mately 60  cycles),  their  schedule  shifted  to  8 
hr  on  followed  by  48  hr  off.  Because  our  sam- 
ple size  was  small,  we  used  data  with  ARGOS 
codes  0-3  (accuracy  <1,000  m);  however,  we 
also  included  some  Auxiliary  Processing  lo- 
cations (ARGOS  codes  A,  B,  C,  and  Z;  no 
estimate  of  accuracy;  ARGOS  1996),  despite 
the  reduced  confidence  in  their  accuracy.  To 
determine  whether  to  include  a given  location 
coded  as  A-C  or  Z,  we  compared  it  to  loca- 
tions documented  before  and  after  the  record 
in  question;  if  it  was  along  the  same  flight 
path  or  within  a few  km  of  areas  where  the 
birds  were  staging,  the  location  was  retained. 
Outlier  data  were  generally  obvious — well  off 
the  flight  path  and/or  indicating  distances  not 
achievable  from  the  high-accuracy  locations. 
On  days  when  we  received  only  data  with  low 
accuracy  codes,  data  were  excluded.  This  pro- 
ject was  carried  out  according  to  protocols  ap- 
proved by  the  Canadian  Council  on  Animal 
Care.  All  means  are  reported  ±SD. 

RESULTS 

Transmitter  performance. — We  received 
1,747  transmissions  from  the  three  implanted 
birds,  of  which  1,203  (69%)  provided  usable 
information  on  locations.  One  duck  provided 
67%  of  the  data,  but  this  was  attributed  to 
more  frequent  transmissions  per  day,  not  a 
longer  transmission  period.  The  three  trans- 
mitters provided  a mean  performance  of  582 
transmissions  and  401  locations  over  217  days 
and  6,760  km  of  travel. 

Bird  movements. — The  two  male  Long- 
tailed Ducks  spent  most  of  April  2003  on 
Lake  Ontario  near  the  capture  site;  on  27-28 
April,  they  moved  to  Georgian  Bay  on  Lake 
Huron  (45°  29' N,  80°  40' W),  where  they 
staged  for  the  next  23  and  30  days,  respec- 
tively (Figs.  1,  2A).  This  was  followed  by  a 
rapid  migration  to  northwestern  James  Bay 
(54°  N,  82°  W);  transmissions  were  3 days 
apart,  and  one  bird  had  arrived  at  this  site 
from  Lake  Huron  between  consecutive  trans- 
missions. The  males  stayed  in  northwestern 
James  Bay  for  approximately  2 weeks  (Fig. 
2B),  and  then  moved  to  western  Hudson  Bay 
(58°  3' N,  93°  14'  W and  63°53'N,  95°  31' 
W)  for  the  last  3 weeks  of  June  and  the  1st 
week  of  July  (Fig.  2C);  during  that  time,  they 
moved  only  very  short  distances  from  inland 


496 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


FIG.  1.  Movements  of  two  male  and  one  female  Long-tailed  Duck  captured  at  Niagara-On-The-Lake,  On- 
tario, Canada,  in  2003  (males:  squares  and  circles)  and  2004  (female:  triangles).  Lines  represent  tracked  or 
interpolated  flight  paths. 


Mallory  et  al.  • LONG-TAILED  DUCK  MIGRATION 


497 


FIG.  2.  Details  of  staging  and  apparent  breeding  locations  used  by  Long-tailed  Ducks  that  moved  from 
Ontario  to  Nunavut,  Canada,  in  2003  (males:  squares  and  circles)  and  2004  (female:  triangles). 


498 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


locations  and  we  assumed  that  each  was  at- 
tending a mate  at  a nest  site.  For  much  of  the 
summer  (12  July  to  18  September,  and  10  July 
to  31  August),  the  males  moved  to  a coastal 
location  near  Bibby  Island  (61°  56'  N,  93°  14' 
W).  One  male  departed  this  site  in  early  Sep- 
tember and  moved  farther  north  to  southern 
Melville  Peninsula  (66°  23'  N,  85°  46'  W), 
where  he  stayed  from  10  to  18  September.  The 
other  male  moved  to  southern  Southampton 
Island  (62°  53'  N,  83°  36'  W),  where  he  stayed 
from  24  to  28  September;  by  1 October,  this 
male  had  migrated  east  across  Hudson  Bay 
and  remained  near  the  Belcher  Islands  (Fig. 
2D;  56°  30'  N,  79°  30'  W)  and  eastern  Hudson 
Bay  until  transmissions  ceased  on  6 Novem- 
ber. The  second  male  remained  farther  north, 
but  by  20  October,  he  had  migrated  southward 
to  Coats  Island  (62°  30'  N,  82°  30'  W);  by  6 
November,  he  had  moved  farther  south  to 
eastern  James  Bay,  where  his  radiotransmitter 
failed  (also  on  6 November). 

In  2004,  the  female  exhibited  a movement 
pattern  similar  to  that  of  the  males  in  2003 
(Fig.  1).  The  duck  remained  in  western  Lake 
Ontario  until  27  April;  by  3 May,  she  had 
moved  to  Georgian  Bay,  Lake  Huron,  where 
she  remained  until  31  May.  By  2 June,  the 
bird  had  migrated  north  to  western  James  Bay, 
where  she  stayed  until  29  June.  Unlike  the 
males,  this  female  then  spent  from  30  June  to 
10  July  moving  northwest  across  Hudson  Bay, 
well  offshore,  before  heading  inland  in  Nu- 
navut south  of  Arviat  (61°  28'  N,  93°  48'  W). 
The  female  remained  inland  until  4 August, 
and  then  moved  slightly  north  and  offshore  to 
the  coast  around  Bibby  Island,  where  she  re- 
mained until  17  October.  By  20  October,  the 
bird  had  moved  north  to  the  southwestern 
coast  of  Southampton  Island  (63°  30'  N.  86° 
38'  W),  where  she  remained  until  at  least  31 
October,  at  which  time  her  radiotransmitter 
failed.  During  the  two  monitoring  periods 
(April  to  October.  2003  and  2004),  the  three 
radio-marked  birds  spent  12%  of  their  time  at 
Georgian  Bay,  7%  at  James  Bay,  and  30% 
near  Bibby  Island,  western  Hudson  Bay. 

Flight  speeds. — Flight  (ground)  speeds  of 
the  three  Long-tailed  Ducks  were  calculated 
for  several  days  when  their  transmissions  in- 
dicated continuous  movement  (i.e.,  locations 
traced  a linear  track).  The  birds  traveled  at 
50.2  ± 16.8  km/hr  (n  = 5 days).  On  22  Oc- 


tober 2003,  however,  one  male’s  transmitter 
recorded  a southward  movement  that  started 
at  08:35,  when  the  bird  was  positioned  at  59° 
6'  N,  84°  18'  W.  By  the  time  the  transmission 
period  ended  at  15:30,  the  bird  had  moved 
south  to  55°  54'  N,  78°48'W,  which  repre- 
sents a straight-line  distance  of  —600  km  in 
7 hr,  or  a flight  speed  of  86  km/hr.  Unfortu- 
nately. the  duty  cycle  on  the  transmitters  did 
not  allow  us  to  reliably  assess  whether  birds 
were  more  likely  to  move  at  day  or  night.  All 
of  the  movements  used  to  calculate  flight 
speeds  were  recorded  between  00:40  and  16:30. 

DISCUSSION 

The  data  gathered  in  this  study  provide  new 
insights  into  the  habitat  use  and  migration  pat- 
terns of  Long-tailed  Ducks  in  eastern  North 
America.  Radio-marked  Long-tailed  Ducks 
wintering  on  western  Lake  Ontario  moved 
northwest  to  breed  along  western  Hudson 
Bay,  and  then  appeared  to  circumnavigate 
Hudson  Bay  before  traveling  southward  along 
its  eastern  coast  during  fall  migration.  The  lat- 
ter finding  was  unexpected  and  counter  to  our 
predictions,  as  there  was  no  previous  evidence 
of  this  circuitous  movement  pattern.  Our  in- 
terpretation assumes  that  the  implantation  pro- 
cedure did  not  markedly  alter  the  birds’  travel 
routes  and  migration  patterns.  We  believe  this 
to  be  a reasonable  assumption  because  the 
findings  of  prior  studies  have  suggested  sim- 
ilar migratory  patterns  linking  these  regions 
(Bellrose  1980.  Leafloor  et  al.  1996).  The  in- 
formation provided  by  the  satellite  transmit- 
ters confirms  this  pattern,  and  we  identified 
some  key  staging  locations.  Despite  our  small 
sample  size,  the  similarity  of  movements  in 
both  years  and  by  both  sexes  attests  to  the 
importance  of  the  key  sites. 

Long-tailed  Duck  migration  northward 
from  the  Great  Lakes  takes  place  in  a series 
of  short,  rapid  movements,  separated  by  rel- 
atively long  stopovers  at  certain  major  coastal 
sites.  Northern  Georgian  Bay  in  Lake  Huron 
(Fig.  2A)  and  western  James  Bay,  particularly 
north  of  Akimiski  Island  (Fig.  2B),  appear  to 
be  critical  stopover  sites  for  this  species  dur- 
ing spring  migration,  as  birds  spent  nearly 
20%  of  their  time  between  April  and  October 
in  these  bays.  The  importance  of  James  Bay 
to  migrating  waterfowl  has  been  known  for 
some  time  and  led  to  creation  of  the  James 


Mallory  et  al.  • LONG-TAILED  DUCK  MIGRATION 


499 


Bay  Preserve  in  the  early  1900s  (reviewed  in 
Mallory  and  Fontaine  2004).  Our  data  provide 
further  evidence  of  the  importance  of  the 
northwestern  coast  of  James  Bay  to  certain  sea 
ducks  (Mallory  and  Fontaine  2004). 

Both  male  and  female  Long-tailed  Ducks 
wintering  on  western  Lake  Ontario  migrated 
north  and  apparently  bred  inland  along  west- 
ern Hudson  Bay.  We  believe  that  the  males 
attended  their  mates  for  a period  of  about  4 
weeks  before  moving  to  molting  sites  some 
time  between  10  and  12  July.  This  interpre- 
tation of  the  satellite  data  is  consistent  with 
Alison’s  (1975)  observations  that  males  left 
their  breeding  ponds  near  Churchill,  Manito- 
ba, on  about  10  July.  In  2004,  the  implanted 
female  staged  near  Akimiski  Island  much  lon- 
ger than  the  males  had  in  2003,  perhaps  due 
to  the  winter  conditions  that  persisted  rela- 
tively late  along  western  Hudson  Bay  that 
year.  When  the  female  finally  moved  to  the 
breeding  area,  she  stayed  well  offshore  and 
flew  over  sea-ice  (Environment  Canada  2005), 
counter  to  the  expected  pattern  of  following 
shorelines  (Johnson  1985).  The  female  was 
positioned  inland  at  potential  nesting  areas  for 
a period  of  25  days  beginning  around  10  July. 
If  she  nested,  her  nest  would  have  been  ini- 
tiated about  1 month  later  than  those  of  most 
Long-tailed  Ducks  nesting  in  that  region  (Al- 
ison 1975);  thus,  if  she  did  nest,  we  suspect 
that  her  nest  was  abandoned  or  depredated. 
Female  Long-tailed  Ducks  require  —33  days 
to  lay  and  hatch  an  average-sized  clutch  (7 
days  for  laying  plus  26  days  for  incubation), 
longer  than  the  amount  of  time  the  radio- 
marked  duck  spent  in  that  area.  It  is  also  pos- 
sible that  implantation  of  the  transmitter  into 
her  celomic  cavity  could  have  affected  ovi- 
position  and  normal  nesting  behavior,  or  it  is 
possible  that  she  had  not  yet  reached  breeding 
age  (which  also  could  have  explained  some  of 
her  erratic  movements). 

An  important  finding  of  our  study  was  the 
location  of  a molting  area  near  Bibby  Island, 
between  Arviat  and  Whale  Cove,  Nunavut, 
where  the  three  ducks  spent  30%  of  their  time 
during  the  study  period.  This  site  was  previ- 
ously unknown,  and  demonstrates  the  utility 
of  satellite  transmitters  for  revealing  impor- 
tant, but  remote  and  undiscovered,  sites  used 
by  some  migratory  bird  species  (e.g.,  Brodeur 
et  al.  2002).  The  male  Long-tailed  Ducks 


moved  to  the  area  around  Bibby  Island  after 
leaving  their  breeding  ponds,  whereas  the  fe- 
male arrived  somewhat  later;  both  the  male 
and  female  arrival  dates  were  similar  to  those 
reported  for  their  respective  sexes  at  molting 
sites  elsewhere  (Johnson  and  Richardson 
1982,  Johnson  1985).  All  three  birds  spent  up 
to  2 months  in  the  shallow  waters  around  the 
coast  near  Bibby  Island.  The  proportion  of  the 
overall  Long-tailed  Duck  population  that 
molts  at  this  site,  and  the  extent  to  which  this 
area  supports  molting  birds  of  other  waterfowl 
species,  should  be  investigated. 

There  was  considerably  more  variation  in 
the  pattern  of  fall  migration  among  the  three 
birds.  The  males  moved  east  from  molting 
sites,  then  south  along  eastern  Hudson  Bay. 
One  male  spent  a month  near  the  Belcher  Is- 
lands; the  other  male  followed  the  same  gen- 
eral pathway,  but  did  not  arrive  in  eastern 
Hudson  Bay  until  3 weeks  after  the  first  male. 
Given  that  many  Long-tailed  Ducks  overwin- 
ter in  polynyas  near  the  Belcher  Islands  and 
in  western  Hudson  Bay  (Robertson  and  Sa- 
vard  2002),  birds  in  our  study  may  not  have 
continued  southward.  The  female  appeared  to 
be  following  the  same  path  as  the  males,  but 
initiated  her  fall  migration  relatively  late  and 
had  only  moved  to  Southampton  Island  by  the 
time  her  radiotransmitter  failed  in  late  Octo- 
ber. During  fall  migration,  Leafloor  et  al. 
(1996)  collected  birds  in  northern  Ontario; 
given  that  the  birds  had  fat  stores  sufficient 
for  migration,  they  postulated  that  offshore 
sites  in  Hudson  and  James  bays  must  be  im- 
portant to  Long-tailed  Ducks  for  gathering  nu- 
trients. Our  data  support  their  hypothesis.  Giv- 
en the  varied  locations  where  our  radiomarked 
birds  spent  their  post-molting  period,  it  ap- 
pears that  there  may  be  many  locations  where 
the  birds  can  gather  food,  unlike  the  more  lim- 
ited number  of  locations  suggested  by  our 
spring  migration  data. 

The  transmitters  provided  performance  sim- 
ilar to  that  observed  for  swans  (Petrie  and  Wil- 
cox 2003),  with  almost  70%  of  the  data  being 
usable.  The  flight  speeds  we  calculated  were 
similar  to  values  reported  previously  for  Long- 
tailed Ducks  (up  to  90  km/hr;  Bergman  1974), 
but  a better  assessment  would  be  possible  with 
a duty  cycle  setting  on  the  transmitters  that 
would  provide  more  transmissions  during 
movement  periods.  The  ducks  in  our  study 


500 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


were  at  the  lower  body-size  limit  recommend- 
ed for  the  satellite  transmitters  available  to  us 
at  the  time  (e.g.,  Caccamise  and  Hedin  2003), 
and  some  of  our  birds  were  smaller  than  we 
would  have  preferred  (i.e.,  transmitter  weight 
>5%  of  body  mass).  The  newer,  smaller  trans- 
mitters (http://www.microwavetelemetry.com) 
available  today  should  allow  researchers  to  bet- 
ter track  smaller  birds. 

The  data  gathered  by  tracking  the  three 
Long-tailed  Ducks  in  our  study  has  provided 
valuable  new  information  on  the  species’ 
movements  and  habitat  use;  however,  the  util- 
ity of  these  data  are  not  restricted  to  this  spe- 
cies. For  example,  in  a study  of  Peregrine  Fal- 
cons ( Falco  peregrinus)  breeding  along  west- 
ern Hudson  Bay,  Johnstone  et  al.  (1996)  noted 
that  the  falcons  there  contained  higher  con- 
taminant loads  than  birds  elsewhere  in  the  Ca- 
nadian Arctic.  They  speculated  that  falcons 
were  accumulating  these  contaminants  from 
migratory  prey,  notably  Black  Guillemots 
( Cepphus  grylle\  marine  piscivores;  Mallory 
et  al.  2005)  and  Long-tailed  Ducks,  which 
presumably  have  been  accumulating  pollut- 
ants from  the  heavily  contaminated  Great 
Lakes.  Our  data  on  movements  of  Long-tailed 
Ducks  support  this  linkage  to  the  Great  Lakes 
and  raise  concerns  that  Long-tailed  Ducks 
may  transport  contaminants  to  Arctic  ecosys- 
tems. 

ACKNOWLEDGMENTS 

We  are  grateful  to  the  many  people  who  assisted 
with  various  aspects  of  this  project,  and  the  Niagara- 
On-The-Lake  Sailing  Club.  Two  anonymous  referees 
and  G.  J.  Robertson  provided  helpful,  critical  reviews 
of  the  manuscript.  Linancial  support  was  provided  by 
Environment  Canada  (CWS-OR  and  PNR),  and  the 
Sea  Duck  Joint  Venture.  This  research  was  conducted 
under  permits  CA  0118  and  2003PNR015. 

LITERATURE  CITED 

Alison,  R.  M.  1975.  Breeding  biology  and  behavior 
of  the  Oldsquaw  ( Clangula  hyemalis  L.).  Ornitho- 
logical Monographs  18:1-52. 

ARGOS.  1996.  User’s  manual.  Service  Argos,  Inc., 
Landover,  Maryland. 

Bellrose,  E C.  1980.  Ducks,  geese  and  swans  of 
North  America.  Stackpole  Books,  Harrisburg, 
Pennsylvania. 

Bergman,  G.  1974.  The  spring  migration  of  the  Long- 
tailed Duck  and  the  Common  Scoter  in  western 
Finland.  Ornis  Fennica  51:129—145. 

Brodeur,  S.,  J.-P.  L.  Savard,  M.  Robert,  P.  Laporte, 


P.  Lamothe,  R.  D.  Titman,  S.  Marchand,  S.  Gil- 
liland, and  G.  Fitzgerald.  2002.  Harlequin 
Duck  Histrionicus  histrionicus  population  struc- 
ture in  eastern  Nearctic.  Journal  of  Avian  Biology 
33:127-137. 

Caccamise,  D.  F.  and  R.  S.  Hedin.  2003.  An  aerody- 
namic basis  for  selecting  transmitter  loads  in 
birds.  Wilson  Bulletin  97:306-318. 

Cowardin,  L.  M.  and  R.  J.  Blohm.  1992.  Breeding 
population  inventories  and  measures  of  recruit- 
ment. Pages  423-445  in  Ecology  and  manage- 
ment of  breeding  waterfowl  (B.  D.  J.  Batt,  A.  D. 
Afton,  M.  G.  Anderson,  C.  D.  Ankney,  D.  H. 
Johnson,  J.  A.  Kadlec,  and  G.  L.  Krapu,  Eds.). 
University  of  Minnesota  Press,  Minneapolis. 

Environment  Canada.  2005.  Canadian  Ice  Service. 
http://ice-glaces.ec.gc.ca  (accessed  January  2005). 

Fitzgerald,  G.,  S.  Brodeur,  and  M.  Robert.  2001. 
Implantation  abdominale  d’emetteurs  sur 
l’Arlequin  plongeur  ( Histrionicus  histrionicus ) et 
le  Garrot  d’lslande  ( Bucephala  islandica).  Le  Me- 
decin  Veterinaire  du  Quebec,  Medecine  Zoolo- 
gique  31:39-43.  [in  French] 

Johnson,  S.  R.  1985.  Adaptations  of  the  Long-tailed 
Duck  ( Clangula  hyemalis  L.)  during  the  period  of 
molt  in  arctic  Alaska.  Proceedings  of  the  Inter- 
national Ornithological  Congress  18:530-540. 

Johnson,  S.  R.  and  W.  J.  Richardson.  1982.  Waterbird 
migration  near  the  Yukon  and  Alaskan  coast  of 
the  Beaufort  Sea:  II.  Moult  migration  of  seaducks 
in  summer.  Arctic  35:291-301. 

Johnstone,  R.  M.,  G.  S.  Court,  A.  C.  Fesser,  D.  M. 
Bradley,  L.  W.  Oliphant,  and  J.  D.  MacNeil. 
1996.  Long-term  trends  and  sources  of  organo- 
chlorine  contamination  in  Canadian  Tundra  Pere- 
grine Falcons,  Falco  peregrinus  tundrius.  Envi- 
ronmental Pollution  93:109-120. 

Leafloor,  J.  O.,  J.  E.  Thompson,  and  C.  D.  Ankney. 
1996.  Body  mass  and  carcass  composition  of  fall 
migrant  Oldsquaws.  Wilson  Bulletin  108:567- 
570. 

Mallory,  M.  L.,  B.  M.  Braune,  M.  Wayland,  and 
K.  G.  Drouillard.  2005.  Persistent  organic  pol- 
lutants in  marine  birds,  arctic  hare  and  ringed 
seals  near  Qikiqtarjuaq,  Nunavut,  Canada.  Marine 
Pollution  Bulletin  50:95-104. 

Mallory,  M.  L.  and  A.  J.  Fontaine.  2004.  Key  ma- 
rine habitat  sites  for  migratory  birds  in  Nunavut 
and  the  Northwest  Territories.  Canadian  Wildlife 
Service  Occasional  Paper,  no.  109,  Ottawa,  On- 
tario. 

Pattie,  D.  L.  1990.  A 16-year  record  of  summer  birds 
on  Truelove  Lowland,  Devon  Island,  Northwest 
Territories,  Canada.  Arctic  43:275-283. 

Peterson,  S.  R.  and  R.  S.  Ellarson.  1979.  Changes 
in  oldsquaw  carcass  weight.  Wilson  Bulletin  91: 
288-300. 

Petrie,  S.  A.  and  K.  L.  Wilcox.  2003.  Migration  chro- 
nology of  eastern  population  Tundra  Swans.  Ca- 
nadian Journal  of  Zoology  81:861-870. 

Robert,  M.,  R.  Benoit,  and  J.-P.  L.  Savard.  2002. 


Mallory  et  al.  • LONG-TAILED  DUCK  MIGRATION 


501 


Relationship  among  breeding,  molting  and  win- 
tering areas  of  male  Barrow’s  Goldeneye  ( Buce - 
phala  islandica)  in  eastern  North  America.  Auk 
119:679-684. 

Robertson,  G.  J.  and  J.-P.  L.  Savard.  2002.  Long- 
tailed Duck  ( Clangula  hyemalis).  The  Birds  of 
North  America,  no.  651. 


Sea  Duck  Joint  Venture  Management  Board. 
2001.  Sea  Duck  Joint  Venture  strategic  plan: 
2002-2006.  Sea  Duck  Joint  Venture  Continental 
Technical  Team.  Unpublished  Report.  U.S.  Fish 
and  Wildlife  Service,  Anchorage,  Alaska,  and 
Canadian  Wildlife  Service,  Sackville,  New 
Brunswick. 


The  Wilson  Journal  of  Ornithology  1 18(4):502-507,  2006 


FEMALE  TREE  SWALLOW  HOME-RANGE  MOVEMENTS  DURING 
THEIR  FERTILE  PERIOD  AS  REVEALED  BY  RADIO-TRACKING 

MARY  K.  STAPLETON1 2 3’23  AND  RALEIGH  J.  ROBERTSON1 


ABSTRACT. — Tree  Swallows  ( Tachycineta  bicolor ) show  one  of  the  highest  levels  of  extra-pair  mating 
among  bird  species,  yet  extra-pair  copulations  are  rarely  observed.  Despite  the  suggestion  that  extra-pair  cop- 
ulations could  be  taking  place  away  from  nest  sites,  very  little  is  known  about  movement  patterns  of  individual 
Tree  Swallows  during  the  pre-laying  and  laying  periods.  We  used  radio  telemetry  to  track  movement  patterns 
of  four  female  Tree  Swallows  at  dawn  and  dusk  during  the  pre-laying  and  laying  periods.  Our  tracking  results 
indicate  that  individual  females  differed  in  their  movement  patterns:  some  remained  close  to  their  nest  site  on 
multiple  nights  while  others  were  rarely  detected  near  their  nest  box  at  night.  Despite  differences  in  movement 
patterns,  all  four  females  that  we  tracked  produced  extra-pair  offspring  for  which  we  were  unable  to  identify 
extra-pair  sires,  even  after  sampling  the  majority  of  males  breeding  within  our  nest-box  grids.  Despite  the  small 
sample  size,  our  results  confirmed  extensive  Tree  Swallow  movement  away  from  nest-box  grids  during  the  pre- 
laying and  laying  periods.  This  highlights  the  need  for  future  studies  of  mating  behavior  away  from  the  nesting 
site,  particularly  for  species  that  forage  and/or  roost  in  communal  areas  during  their  fertile  period.  Received  25 
July  2005,  accepted  17  April  2006. 


While  genetic  evidence  of  extra-pair  fertil- 
izations among  birds  is  widespread,  less  is 
known  about  the  behaviors  that  lead  to  extra- 
pair copulations  (EPCs;  Westneat  and  Stewart 

2003) .  In  the  Tree  Swallow  {Tachycineta  bi- 
color),  within-pair  copulations  take  place  ex- 
tremely frequently  and  are  clearly  visible. 
EPCs,  however,  are  rarely  observed  (Venier 
and  Robertson  1991,  Lifjeld  et  al.  1993,  Ven- 
ier et  al.  1993),  despite  the  high  levels  of  ex- 
tra-pair paternity  (up  to  80%  of  all  females  in 
a population  produce  extra-pair  young;  Barber 
et  al.  1996).  Indeed,  extra-pair  copulations  can 
be  difficult  to  observe,  and  many  researchers 
have  used  radio  telemetry  for  following  both 
male  and  female  birds  during  their  extra-ter- 
ritorial forays  in  an  attempt  to  document  ex- 
tra-pair mating  behavior  (Smiseth  and 
Amundsen  1995,  Neudorf  et  al.  1997,  Pitcher 
and  Stutchbury  2000,  Mays  and  Ritchison 

2004) . 

Although  many  passerines  defend  all-pur- 
pose territories  for  foraging  and  nesting  (but 
see  Reyer  et  al.  1997),  Tree  Swallows  defend 
only  the  area  immediately  surrounding  their 
nest  site  (e.g.,  the  nest  box).  Often  they  leave 


1 Dept,  of  Biology,  Queen’s  Univ.,  4320  Bioscienc- 
es, Kingston,  ON  K7L  3N6,  Canada. 

2 Current  address:  107-F  N.  Rock  Glen  Rd.,  Balti- 
more, MD  21229,  USA. 

3 Corresponding  author;  e-mail: 
marykstapleton@gmail.com 


their  territory  for  long  periods  of  time,  pre- 
sumably to  forage  and  roost  (Hayes  and  Co- 
hen 1987,  Robertson  et  al.  1992;  MKS  pers. 
obs.).  During  these  off-territory  forays,  Tree 
Swallows  often  are  found  in  groups  compris- 
ing many  potential  copulation  partners  (Rob- 
ertson et  al.  1992,  Dunn  and  Whittingham 

2005).  Dunn  and  Whittingham  (2005)  found 
that,  on  subsequent  nights,  female  Tree  Swal- 
lows used  different  roost  sites  often  compris- 
ing hundreds  of  individuals.  Hayes  and  Cohen 
(1987),  however,  radio-tracked  several  breed- 
ing male  Tree  Swallows  at  dusk  and  reported 
that  they  “tended  to  return  to  the  same  grove 
night  after  night.” 

Examining  potential  intra-specific  variation 
in  behavioral  patterns  can  be  valuable  for  un- 
derstanding the  underlying  forces  that  shape  a 
species’  mating  system  (see  Westneat  and 
Stewart  2003).  In  this  study,  we  used  radio 
telemetry  to  track  female  movements  in  an 
Ontario  population  of  Tree  Swallows.  Specif- 
ically, we  recorded  first-  (dawn)  and  last- 
(dusk)  known  locations  of  individual  Tree 
Swallows  each  day  during  the  pre-laying  and 
laying  periods.  For  each  female,  we  deter- 
mined her  relative  roosting  location  (i.e.,  on 
or  off  the  nest-box  grid)  and  the  maximum 
distance  from  her  nest  box  she  was  detected 
each  day.  In  addition,  we  conducted  parentage 
analysis  on  the  offspring  of  all  four  focal  fe- 
males, evaluating  extra-pair  fertilizations  in 
light  of  their  movement  and  roosting  patterns. 


502 


Stapleton  and  Robertson  • RADIO-TRACKING  TREE  SWALLOWS 


503 


METHODS 

Our  study,  conducted  during  the  2002 
breeding  season  at  Queen’s  University  Biolog- 
ical Station  in  Chaffey’s  Locks,  Ontario,  Can- 
ada (44°  34'  N,  76°  20'  W),  focused  on  the 
area  surrounding  eight  grids  of  nest  boxes  (6— 
39  boxes  per  grid,  0.28-1.92  ha;  see  Kempen- 
aers  et  al.  [1998]  for  details  regarding  nest- 
box  arrangement).  During  the  early  part  of  the 
breeding  season.  Tree  Swallows  in  our  popu- 
lation generally  spent  the  morning  hours  de- 
fending their  nest  sites  as  well  as  building 
nests.  During  the  late  afternoon  and  evening 
hours,  however,  often  they  were  absent  from 
the  nesting  grid,  presumably  to  forage  in  areas 
with  higher  concentrations  of  insects.  Despite 
high  levels  of  extra-pair  paternity,  male  Tree 
Swallows  do  not  guard  their  mates  (Leffelaar 
and  Robertson  1984),  and  there  is  evidence 
that  females  are  able  to  select  and  reject  cop- 
ulation partners,  at  least  in  the  area  immedi- 
ately surrounding  the  nest  site  (Lifjeld  and 
Robertson  1992). 

Telemetry. — Four  female  Tree  Swallows 
were  radio-tracked  during  the  pre-laying  and 
laying  periods.  To  each  female,  we  attached 
an  LB-2  radio  transmitter  (0.52  g;  Holohil 
Systems  Ltd.,  Carp,  Ontario),  secured  with  a 
figure-eight  style  leg  harness  (Rappole  and 
Tipton  1991).  To  track  radio-tagged  birds,  we 
used  R-1000  receivers  (Communications  Spe- 
cialists, Inc.,  Orange,  California),  3-element 
and  5-element  hand-held  Yagi  antennae,  and 
an  omni-directional  antenna.  We  used  two 
methods  of  tracking:  opportunistic  and  sys- 
tematic. The  opportunistic  method  consisted 
of  constantly  monitoring  all  active  transmit- 
ters while  driving  along  roads,  as  well  as  hik- 
ing into  areas  inaccessible  to  vehicles  sur- 
rounding the  Tree  Swallow  grids.  The  other 
method  involved  systematically  surveying  a 
general  area  from  a pair  of  pre-established 
look-out  points  separated  by  —125-1,500  m: 
two  observers  (one  at  each  point)  equipped 
with  a receiver,  directional  antenna,  and  a 
handheld  two-way  communication  radio 
would  simultaneously  document  the  location 
of  a given  female.  We  were  able  to  detect  sig- 
nals up  to  —2,600  m away.  In  both  methods, 
all  frequencies  were  scanned  continuously, 
and,  when  a signal  was  detected,  observers 
would  simultaneously  record  the  compass 


bearing  of  the  signal.  When  the  precise  angle 
could  not  be  determined,  a range  of  angles 
that  encompassed  the  signal  was  recorded. 
Birds  were  tracked  opportunistically  through- 
out the  day  (04:00-22:00  EST),  as  well  as 
systematically  during  morning  (05:00-0:700) 
and  evening  (19:00-21:00)  hours.  At  the  end 
of  an  evening  tracking  session,  observers  vis- 
ited each  grid  to  confirm  the  presence  or  ab- 
sence of  focal  birds  in  their  nest  boxes.  The 
total  radio-tracking  effort  was  100  hr.  Radio 
transmitters  were  removed  from  birds  during 
the  incubation  period. 

Roosting  areas. — We  were  unable  to  visu- 
ally locate  any  roosting  sites  (except  when 
birds  roosted  in  their  nest  boxes)  because  con- 
sistent radio  signals  often  were  not  detectable 
during  nighttime  hours  (evidence  that  a bird 
had  settled  into  a roost  site).  We  also  attribut- 
ed the  lack  of  nighttime  signals  to  the  birds 
roosting  beyond  receiver  ranges  (i.e.,  >2.6 
km)  or  to  signals  being  blocked  by  terrain 
(i.e.,  birds  roosting  low  in  a valley).  When  a 
signal  was  detected  consistently  after  sun- 
down, it  was  usually  because  the  bird  was 
roosting  in  her  nest  box.  In  the  few  cases 
where  a bird  was  suspected  of  roosting  outside 
of  her  box  but  near  the  grid,  difficulties  with 
navigating  the  hilly  terrain  in  the  dark  pre- 
cluded visual  confirmation  of  the  roosting  site. 
Thus,  we  focused  our  efforts  on  estimating  the 
general  location  of  each  bird  through  trian- 
gulation early  in  the  morning  (05:00—07:00) 
and  at  dusk  (19:00-21:00).  We  used  the  first 
and  last  known  locations  of  individuals  as  an 
indication  of  general  roosting  area.  When  we 
were  unable  to  detect  a given  individual’s  sig- 
nal during  our  evening  observation  period,  we 
were  able  to  determine  only  that  the  bird  was 
not  in  the  box  (i.e.,  away  from  the  nest  grid). 

Mapping. — Compass  bearings  were  entered 
onto  a GIS-based  topographic  map  of  the  area, 
and  bird  locations,  as  determined  from  trian- 
gulation, were  plotted  using  AutoCAD  (Au- 
todesk 2000).  For  a given  individual  on  a giv- 
en day,  we  defined  “first-known  location”  as 
the  bird’s  location  when  detected  for  the  first 
time  prior  to  08:00;  “last-known  location” 
was  the  bird’s  location  when  detected  for  the 
last  time  after  21:00.  The  “farthest  location” 
was  the  greatest  distance  between  the  bird’s 
location  and  its  nest  box,  regardless  of  time 
of  day.  If  a signal  was  recorded  as  coming 


504 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


from  a range  of  compass  directions,  the  mean 
of  the  reported  directions  was  used  and  the 
location  of  the  bird  was  recorded  as  being  at 
the  intersection  of  the  two  vectors.  If  the  vec- 
tors of  the  means  did  not  cross,  then  the  range 
was  plotted  for  each  observer  and  we  recorded 
the  bird’s  location  as  being  at  least  as  far  as 
the  closest  point  where  the  two  ranges  over- 
lapped. If  the  two  vectors  did  not  overlap  but 
came  from  the  same  direction  (presumably 
due  to  a moving  bird),  the  bird’s  location  was 
plotted  as  being  at  least  as  far  as  the  observer 
look-out  points  when  these  points  were  be- 
tween the  nest  box  and  the  bird’s  presumed 
location.  Therefore,  our  reported  bird  loca- 
tions are  conservative  estimates,  reflecting  the 
closest  a bird  could  have  been  to  its  nest  box 
within  the  range  detected.  Distance  from  the 
focal  bird’s  nest  box  to  each  location  detected 
during  observation  periods  was  calculated 
with  AutoCAD  (Autodesk  2000)  as  the 
straight-line  distance  between  the  two  points. 

Movement. — For  each  female,  we  defined 
the  pre-laying  period  as  the  day  the  transmitter 
was  attached  until  the  day  before  the  first  egg 
was  laid  (i.e.,  day  “—X”  until  day  1”). 
The  laying  period  included  the  day  the  first 
egg  was  laid  (i.e.,  day  “0”)  and  continued  un- 
til the  day  the  penultimate  egg  was  laid  or  the 
transmitter  stopped  working,  whichever  was 
later  (i.e.,  day  0 until  day  “X”).  The  average 
maximum  distance  each  female  traveled  dur- 
ing each  period  (pre-laying  and  laying)  was 
calculated  by  summing  the  greatest  distance 
recorded  each  day  and  dividing  by  the  number 
of  days  on  which  the  bird’s  location  was  re- 
corded. The  number  of  days  on  which  we  had 
detected  a distance  varied  between  females 
due  to  differences  in  how  long  the  pre-laying 
period  lasted  and/or  failure  to  detect  a bird  on 
a particular  day. 

Paternity. — We  used  1 1 hypervariable  mi- 
crosatellite loci  (total  probability  of  exclusion 
= 0.999)  to  determine  parentage  of  eggs  and 
nestlings  produced  by  the  four  focal  females. 
To  assign  paternity  to  extra-pair  offspring,  we 
genotyped  all  males  caught  in  surrounding 
nest  boxes  (/?  = 78  males).  We  also  used  ge- 
notypic data  collected  from  males  for  a sep- 
arate study  in  1997,  2000.  2001.  and  2003  (n 
= 65).  because  some  of  those  males  may  have 
been  present,  but  not  caught  (e.g.,  breeding  in 


natural  cavities),  in  2002.  Genotyping  meth- 
ods are  described  in  detail  in  Stapleton  (2005). 

Statistical  analyses. — We  plotted  bird  loca- 
tions and  used  AutoCAD  to  calculate  distanc- 
es (Autodesk  2000).  Differences  in  distance 
from  nest  box  in  the  pre-laying  compared  with 
the  laying  period  were  calculated  with  JMPIN 
(SAS  Institute,  Inc.  2000)  using  a two-tailed 
matched-pair  f-test  at  the  0.05  significance 
level.  We  used  GERUD1.0  (Jones  2001)  to 
calculate  the  minimum  number  of  extra-pair 
sires  within  a given  brood,  based  on  the  max- 
imum number  of  unique  paternal  alleles  pre- 
sent in  all  offspring  of  the  brood.  Values  re- 
ported in  the  results  are  means  ± SE. 

RESULTS 

All  four  female  Tree  Swallows  were  tracked 
until  at  least  2 days  after  the  first  egg  was  laid 
(i.e.,  until  at  least  day  +2;  Table  1).  Due  to 
difficulties  in  locating  precise  roosting  sites, 
we  used  last-known  location  at  night  and  first- 
known  location  in  the  morning  as  a proxy  for 
roosting  location  (i.e.,  distance  and  direction 
from  nest  box).  Radio-tracking  effort,  calcu- 
lated for  each  individual,  varied  due  to  indi- 
vidual differences  in  first  egg  dates  (range  = 
49.7-79.3  hr.  1 1-18  days;  Table  1).  Dates  are 
reported  as  negative  and  positive  integers, 
with  0 representing  the  first  egg  day. 

Movement. — There  was  a tendency  for  fe- 
males to  be  detected  farther  from  the  nest  box 
in  the  pre-laying  period  (mean  661  ± 200  m) 
than  in  the  laying  period  (225  ± 200  m; 
matched-pair  t- test:  = -2.80.  P — 0.068). 

Two  females  (STA3  and  HUW2;  Table  1) 
tended  to  remain  in  or  near  their  nest  boxes, 
one  female  (SRBP1)  was  commonly  found  at 
intermediate  distances  from  her  nest  box,  and 
one  female  (NBF2)  routinely  roosted  >2500 
m from  her  nest  box. 

The  female  nesting  at  NBF2  was  detected 
the  farthest  from  her  nest  box.  Although  her 
nest  box  was  within  200  m of  three  other  Tree 
Swallow  grids,  she  was  frequently  located  in 
the  evenings  near  the  SRB  grid,  which  was 
approximately  2,300  m distant.  NBF2  did  not 
roost  on  her  grid  until  day  +3  (Table  1).  Prior 
to  that,  she  was  detected  >2,500  m from  her 
nest  box  on  the  evenings  of  day  —6  and  day 
-1.  SRBP1  female  was  detected  off  her  grid 
early  in  the  pre-laying  period  at  distances  of 
<883  m (day  -4),  but  then  she  stayed  close 


Stapleton  and  Robertson  • RADIO-TRACKING  TREE  SWALLOWS 


505 


TABLE  1.  Summary  information  for  four  Tree  Swallows  radio-tracked  in  May  2002  at  Queen’s  University 
Biological  Station,  Ontario,  Canada.  Day  (relative  days  tracked)  was  relative  to  the  first  egg  date  (day  0). 
Location  of  a female  during  the  pre-laying  and  laying  periods  was  designated  either  as  “on”  (<100  m from 
nest  box)  or  off  (>101  m from  nest  box)  a nest-box  grid. 


Female 

Agea 

First  egg 
date 

Relative  days 
tracked 

Pre-laying 

(on/off)b 

Laying 

(on/off)b 

Clutch 

size 

No. 

EPOc 

Min.  no. 
EP41  sires 

EPO  in 
brood  (%) 

Hr 

tracked 

NBF2 

SY 

23  May 

-6  to  +4 

0/5e 

1/4 

5 

1 

1 

20 

49.7 

SRBP1 

SY 

26  May 

-9  to  +2 

2/7 

3/0 

6 

1 

1 

17 

57.2 

STA3 

ATY 

22  May 

-5  to  +5 

2/2c 

5/0 

5f 

>1« 

1 

>33 

47.9 

HUW2 

ASY 

26  May 

-11  to  +2 

7/4 

2/0b 

5 

3 

1 

60 

75.3 

a SY  = second  year,  ASY  = after-second-year,  ATY  = after-third-year. 

b “On”  = the  number  of  times  a female  was  detected  si 00  m from  her  nest  box  (i.e.,  on  or  very  near  the  nest  grid)  during  each  last  nightly  check; 
“off”  = the  number  of  times  a female  was  either  detected  slOl  m from  her  nest  box  or  no  signal  was  obtained  from  the  nest  grid  during  each  last 
nightly  check. 

c Extra-pair  offspring. 

d Minimum  number  of  extra-pair  (EP)  sires  (calculated  in  GERUD1.0),  based  on  the  number  of  unique  paternal  alleles. 
e No  telemetery  information  recorded  for  females  NBF2  and  STA3  on  days  -5  and  -4,  respectively  during  the  pre-laying  period. 
fTwo  nestlings  were  not  genotyped  (one  nestling  disappeared  from  the  nest;  one  nestling  did  not  yield  DNA). 

8 Social  male  not  captured;  presence  of  extra-pair  young  is  based  on  number  of  unique  paternal  alleles. 
h No  telemetry  information  recorded  for  female  HUW2  on  day  + 1 in  laying  period. 


to  the  grid  for  the  remainder  of  the  tracking 
period.  From  the  evening  of  day  —2  until  the 
end  of  tracking  (day  +3),  she  was  never  de- 
tected >72  m from  her  nest  box  and  seemed 
to  be  roosting  near  the  grid  (Table  1).  STA3 
female  showed  very  little  movement  and  was 
not  detected  off  her  grid  between  day  - 1 and 
day  +4,  her  last  egg  day  (Table  1).  Her  max- 
imum detected  movement  was  1,646  m on  the 
morning  of  day  —2.  HUW2  female  showed 
the  least  amount  of  movement  and  was  never 
detected  off  her  grid  between  day  —6  and  day 
+2  (Table  1). 

Paternity. — All  four  focal  females  produced 
at  least  one  extra-pair  offspring  (Table  1).  For 
one  female,  we  were  unable  to  catch  her  social 
mate.  In  this  case,  we  used  number  of  paternal 
alleles  per  locus  in  the  offspring  to  estimate 
the  minimum  number  of  sires  represented  in 
the  brood  (i.e.,  greater  than  three  unique  al- 
leles at  a single  locus  in  offspring  indicates 
more  than  one  sire).  We  were  unable  to  assign 
any  extra-pair  mates  for  any  of  the  four  focal 
females,  despite  our  success  at  sampling  the 
majority  of  males  using  nest  boxes  in  this 
population. 

DISCUSSION 

Last-known  locations  at  night  combined 
with  first-known  locations  in  the  early  morn- 
ing indicated  that  individual  female  Tree 
Swallows  in  this  population  do  not  return  to 
the  same  roost  site  night  after  night.  In  addi- 
tion, individuals  varied  with  respect  to  how 


far  away  from  their  nest  sites  they  roosted. 
Although  two  females  (HUW2  and  STA3) 
were  rarely  detected  >50  m from  their  nest 
boxes,  one  female  (NBF2)  was  routinely  de- 
tected up  to  2 km  from  her  nest  box.  There 
was  a strong  tendency  for  females  to  remain 
closer  to  their  nest  boxes  in  the  laying  period 
than  in  the  pre-laying  period.  Overall,  our  re- 
sults indicate  that  movement  patterns  of  Tree 
Swallows  differ  both  within  and  among  indi- 
viduals. These  results  are  in  accordance  with 
those  of  a recent  study  on  a Wisconsin  pop- 
ulation of  Tree  Swallows  (Dunn  and  Whit- 
tingham  2005),  in  which  four  females  that 
were  tracked  to  their  roosting  sites  over  sev- 
eral evenings  prior  to  egg  laying  did  not  al- 
ways use  the  same  roost  on  subsequent  nights. 
Furthermore,  although  these  females  all  nest- 
ed within  0.5  km  of  each  other,  their  individ- 
ual roosting  sites  defined  an  area  of  at  least 
103  km2.  Together,  these  results  highlight  the 
importance  of  continued  studies  away  from 
the  area  immediately  surrounding  the  nest  site, 
particularly  for  passerines  such  as  Tree  Swal- 
lows that  spend  considerable  time  away  from 
their  territories  during  the  breeding  season. 

The  tendency  for  some  female  Tree  Swal- 
lows to  roost  away  from  their  nest  site  during 
their  fertile  period  has  implications  with  re- 
spect to  extra-pair  mating.  Unlike  many  other 
passerines,  most  extra-pair  sires  among  Tree 
Swallows  do  not  seem  to  be  neighboring 
males  (Dunn  et  al.  1994,  Kempenaers  et  al. 
1999,  Kempenaers  et  al.  2001).  In  our  study 


506 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


population,  we  were  able  to  identify  extra-pair 
sires  for  49%  of  extra-pair  young  (Stapleton 
2005),  a pattern  consistent  with  results  of  pre- 
vious studies  (Dunn  et  al.  1994,  Kempenaers 
et  al.  1999,  Kempenaers  et  al.  2001).  Dunn  et 
al.  (1994)  suggested  that  female  Tree  Swal- 
lows obtain  their  EPCs  at  roosting  sites.  Al- 
though we  were  unable  to  directly  observe 
birds  roosting  away  from  their  nest  boxes,  our 
data  did  allow  us  to  determine  whether  or  not 
a given  female  spent  the  night  at  her  nest  box. 
Initially,  we  had  predicted  that  the  extra-pair 
sires  for  a given  brood  would  be  neighboring 
males  if  the  female  tended  to  roost  on  or  very 
near  her  nest-box  grid;  however,  although  all 
females  in  this  study  produced  extra-pair 
young,  we  were  unable  to  identify  extra-pair 
sires  for  any  of  the  four  focal  females,  despite 
having  sampled  most  of  the  neighboring 
males.  Thus,  whether  or  not  a female  roosted 
away  from  her  nest  box  or  tended  to  remain 
nearby  did  not  affect  whether  she  produced 
extra-pair  offspring  sired  by  neighboring 
males  in  our  small  sample  of  radio-tagged 
birds. 

Our  study  provides  additional  evidence  that 
movements  of  female  Tree  Swallows  are  ex- 
tensive and  variable  during  their  fertile  period 
(see  Dunn  and  Whittingham  2005).  The  main 
difficulty  with  our  study  was  our  inability  to 
consistently  locate  the  Tree  Swallows  fitted 
with  transmitters.  Despite  these  difficulties, 
we  encourage  future  telemetry  studies  coupled 
with  parentage  analyses  on  Tree  Swallows, 
particularly  in  areas  with  flat  topography  and 
adequate  vehicular  access  to  aid  in  tracking 
these  birds  over  their  relatively  large  home 
ranges  during  the  breeding  season. 

ACKNOWLEDGMENTS 

We  would  like  to  thank  C.  A.  Dale  for  her  excellent 
assistance  in  the  field  and  W.  F.  Connor  for  assisting 
with  field  work  and  the  GPS/GIS  aspect  of  the  project. 
D.  J.  Mennill  provided  advice  and  guidance  on  harness 
construction.  We  are  grateful  to  J.  T.  Lifjeld,  D.  L. 
Neudorf,  L.  M.  Ratcliffe,  D.  W.  Winkler,  T.  E.  Steeves, 
and  two  anonymous  reviewers  for  comments  that  im- 
proved earlier  versions  of  this  manuscript.  All  methods 
in  this  study  were  approved  by  the  Queen’s  University 
Animal  Care  Committee  under  permit  # RobertsRJ- 
040.  Funding  was  provided  through  grants  from  the 
American  Ornithologists’  Union  (MKS),  the  Society 
for  Canadian  Ornithologists  (MKS),  Queen’s  Univer- 
sity (MKS),  and  an  NSERC  equipment  grant  and  op- 
erating grant  (RJR). 


LITERATURE  CITED 

Autodesk,  Inc.  2000.  AutoCAD,  ver.  4.5.  Autodesk, 
Inc.,  San  Rafael,  California. 

Barber,  C.  A.,  R.  J.  Robertson,  and  P.  T.  Boag.  1996. 
The  high  frequency  of  extra-pair  paternity  in  Tree 
Swallows  is  not  an  artifact  of  nestboxes.  Behav- 
ioral Ecology  and  Sociobiology  38:425-430. 

Dunn,  P.  O.,  R.  J.  Robertson,  D.  Michaud-Freeman, 
and  P.  T.  Boag.  1994.  Extra-pair  paternity  in  Tree 
Swallows:  why  do  females  mate  with  more  than 
one  male?  Behavioral  Ecology  and  Sociobiology 
35:273-281. 

Dunn,  P.  O.  and  L.  A.  Whittingham.  2005.  Radio- 
tracking of  female  Tree  Swallows  prior  to  egg- 
laying.  Journal  of  Field  Ornithology  76:259-263. 

Hayes,  S.  G.  and  R.  R.  Cohen.  1987.  Night-roosting 
behavior  of  radio-tagged  breeding  male  Tree 
Swallows  ( Tachycineta  bicolor).  Journal  of  the 
Colorado-Wyoming  Academy  of  Science  19:18. 

Jones,  A.  G.  2001.  GERUD1.0:  a computer  program 
for  the  reconstruction  of  parental  genotypes  from 
progeny  arrays  using  multilocus  DNA  data.  Mo- 
lecular Ecology  Notes  1:215-218. 

Kempenaers,  B.,  B.  Congdon,  P.  Boag,  and  R.  J.  Rob- 
ertson. 1999.  Extra-pair  paternity  and  egg  hatch- 
ability  in  Tree  Swallows:  evidence  for  the  genetic 
compatibility  hypothesis?  Behavioral  Ecology  10: 
304-311. 

Kempenaers,  B.,  S.  Everding,  C.  Bishop,  P.  Boag, 
and  R.  J.  Robertson.  2001.  Extra-pair  paternity 
and  the  reproductive  role  of  male  floaters  in  the 
Tree  Swallow  ( Tachycineta  bicolor ).  Behavioral 
Ecology  and  Sociobiology  49:251-259. 

Kempenaers,  B.,  R.  B.  Lanctot,  and  R.  J.  Robertson. 
1998.  Certainty  of  paternity  and  paternal  invest- 
ment in  Eastern  Bluebirds  and  Tree  Swallows.  An- 
imal Behaviour  55:845-860. 

Leffelaar,  D.  and  R.  J.  Robertson.  1984.  Do  male 
Tree  Swallows  guard  their  mates?  Behavioral 
Ecology  and  Sociobiology  16:73-80. 

Lifjeld,  J.  T.,  P.  O.  Dunn,  R.  J.  Robertson,  and  P.  T. 
Boag.  1993.  Extra-pair  paternity  in  monogamous 
Tree  Swallows.  Animal  Behaviour  45:213-229. 

Lifjeld,  J.  T.  and  R.  J.  Robertson.  1992.  Female  con- 
trol of  extra-pair  fertilization  in  Tree  Swallows. 
Behavioral  Ecology  and  Sociobiology  31:89-96. 

Mays,  H.  L.,  Jr.,  and  G.  Ritchison.  2004.  The  effect 
of  vegetation  density  on  male  mate  guarding  and 
extra-territorial  forays  in  the  Yellow-breasted  Chat 
(Icteria  virens).  Naturwissenschaften  91 : 195-198. 

Neudorf,  D.  L.,  B.  J.  M.  Stutchbury,  and  W.  H.  Pip- 
er. 1997.  Covert  extraterritorial  behavior  of  fe- 
male Hooded  Warblers.  Behavioral  Ecology  8: 
595-600. 

Pitcher,  T.  E.  and  B.  J.  M.  Stutchbury.  2000.  Extra- 
territorial forays  and  male  parental  care  in  Hooded 
Warblers.  Animal  Behaviour  59:1261-1269. 

Rappole,  J.  H.  and  A.  R.  Tipton.  1991.  New  harness 
design  for  attachment  of  radio  transmitters  to 


Stapleton  and  Robertson  • RADIO-TRACKING  TREE  SWALLOWS 


507 


small  passerines.  Journal  of  Field  Ornithology  62: 
335-337. 

Reyer,  H.-U.,  K.  Bollmann,  A.  R.  Schlapfer,  A. 
Schymainda,  and  G.  Klecack.  1997.  Ecological 
determinants  of  extrapair  fertilizations  and  egg 
dumping  in  Alpine  Water  Pipits  ( Anthus  spinolet- 
ta ).  Behavioral  Ecology  8:534-543. 

Robertson,  R.  J.,  B.  J.  Stutchbury,  and  R.  R.  Cohen. 
1992.  Tree  Swallow  ( Tachycineta  bicolor).  The 
Birds  of  North  America,  no.  1 1 . 

SAS  Institute,  Inc.  2000.  JMPIN,  ver.  4.0.3.  SAS  In- 
stitute, Inc,  Minneapolis,  Minnesota. 

Smiseth,  P.  T.  and  T.  Amundsen.  1995.  Female  Blue- 
throats  ( Luscinia  s.  svecica)  regularly  visit  terri- 
tories of  extrapair  males  before  egg  laying.  Auk 
112:1049-1053. 


Stapleton,  M.  K.  2005.  Extrapair  mating  in  Tree 
Swallows:  an  examination  of  the  genetic  compat- 
ibility hypothesis.  Ph.D.  dissertation.  Queen’s 
University,  Kingston,  Ontario,  Canada. 

Venier,  L.  A.,  P.  O.  Dunn,  J.  T.  Lifjeld,  and  R.  J. 
Robertson.  1993.  Behavioural  patterns  of  extra- 
pair copulation  in  Tree  Swallows.  Animal  Behav- 
iour 45:412-415. 

Venier,  L.  A.  and  R.  J.  Robertson.  1991.  Copulation 
behaviour  of  the  Tree  Swallow  ( Tachycineta  bi- 
color): paternity  assurance  in  the  presence  of 
sperm  competition.  Animal  Behaviour  42:939- 
948. 

Westneat,  D.  F.  and  I.  R.  K.  Stewart.  2003.  Extra- 
pair paternity  in  birds:  causes,  correlates,  and  con- 
flict. Annual  Review  of  Ecology  Evolution  and 
Systematics  34:365-396. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):508 — 5 1 2,  2006 


EFFECTS  OF  PRESCRIBED  FIRE  ON  CONDITIONS  INSIDE  A 
CUBAN  PARROT  (AMAZONA  LEUCOCEPHALA)  SURROGATE 
NESTING  CAVITY  ON  GREAT  ABACO,  BAHAMAS 

JOSEPH  J.  O’BRIEN,1-5  CAROLINE  STAHALA,2  GINA  P.  MORI,3 
MAC  A.  CALLAHAM,  JR.,1  AND  CHRIS  M.  BERGH4 


ABSTRACT. — Cuban  Parrots  ( Amazona  leucocephala ) on  the  island  of  Great  Abaco  in  the  Bahamas  forage 
and  nest  in  native  pine  forests.  The  population  is  unique  in  that  the  birds  nest  in  limestone  solution  holes  on 
the  forest  floor.  Bahamian  pine  forests  are  fire-dependent  with  a frequent  surface  fire  regime.  The  effects  of  fire 
on  the  parrots,  especially  while  nesting,  are  not  well  known.  We  measured  ambient  conditions  inside  a cavity 
characteristic  of  the  Cuban  Parrot’s  Abaconian  population  as  a prescribed  fire  passed  over  it.  Cavity  conditions 
were  relatively  benign;  although  temperatures  immediately  outside  the  cavity  rose  to  >800°  C,  inside  tempera- 
tures increased  only  5°  C at  30  cm  inside  the  entrance  and  0.4°  C at  the  cavity  floor  (cavity  depth  was  —120 
cm).  C02  levels  briefly  rose  to  2,092  ppm  as  the  flames  passed,  but  dropped  to  nearly  ambient  levels  approxi- 
mately 15  min  later.  Smoke  levels  also  were  elevated  only  briefly,  with  0.603  mg  of  total  suspended  particulates 
filtered  from  0.1  m3  of  air.  Smokey  conditions  lasted  approximately  20  min.  Received  23  September  2005, 
accepted  5 May  2006. 


In  the  Bahamas,  the  Cuban  Parrot  ( Ama- 
zona leucocephala)  currently  occurs  only  on 
the  islands  of  Great  Abaco  and  Great  Inagua. 
The  Bahamian  populations  of  Cuban  Parrots 
are  often  recognized  as  a subspecies  {Ama- 
zona leucocephala  bahamensis).  Regardless 
of  taxonomic  rank,  the  Great  Abaco  popula- 
tion is  distinct  because  the  parrots  nest  in  the 
ground,  exploiting  small  solution  holes  in  the 
exposed  limestone  bedrock  found  in  stands  of 
Caribbean  pine  {Pinus  caribaea  var.  baha- 
mensis)— a forest  type  known  locally  as 
“pineyards.”  This  ground-nesting  behavior  is 
unique,  as  all  other  populations  of  Cuban  Par- 
rots are  known  to  nest  in  tree  cavities.  Pine 
seeds  and  fruit  of  other  pineyard  plants  are 
important  food  sources  for  the  parrots  on 
Great  Abaco  during  the  breeding  season  (At- 
trill  1981,  Snyder  et  al.  1982).  Bahamian  pine- 
yard  ecosystems  are  fire-dependent:  frequent 
fires  suppress  competing  broad-leaved  vege- 
tation, remineralize  nutrients  bound  in  litter, 
and  prevent  fuel  buildups  that  increase  the  risk 


1 U.S.  Dept,  of  Agriculture  Forest  Service,  Southern 
Research  Station,  320  Green  St.,  Athens,  GA  30602, 
USA. 

2 U.S.  Fish  and  Wildlife  Service,  1601  Balboa  Ave., 
Panama  City,  FL  32405,  USA. 

3 Dept,  of  Natural  Sciences,  Univ.  of  Maryland, 
Eastern  Shore,  Princess,  MD  21853,  USA. 

4 The  Nature  Conservancy,  P.O.  Box  420237,  Sum- 
merland  Key.  FL  33042,  USA. 

5 Corresponding  author;  e-mail:  jjobrien@fs.fed. us 


of  greater  fire  intensity  when  accidental  fires 
occur.  In  the  absence  of  fire,  broad-leaved  for- 
est species  eventually  outcompete  and  replace 
the  overstory  pines.  In  analogous  pine  forests 
in  southern  Florida,  suppression  of  fire  result- 
ed in  forest  succession  to  broad-leaved  vege- 
tation in  as  few  as  25  years  (Robertson  1955, 
Loope  and  Dunevitz  1981).  Fires  have  been 
occurring  in  Great  Abaco  pineyards  every  3 
to  5 years  since  at  least  the  late  1700s  (H.  D. 
Grissino-Mayer  unpubl.  data).  Human  activi- 
ties are  currently  the  most  frequent  sources  of 
ignition,  although  lightning-ignited  fires  do 
occur  and  their  frequency  is  probably  under- 
estimated. 

Prescribed  fire  has  become  a popular  man- 
agement tool  in  many  protected  areas  contain- 
ing fire-dependent  vegetation.  Currently,  the 
extemporaneous  fire  management  practiced  by 
local  Abaconians  has  been  very  effective  in 
maintaining  the  pineyards.  Future  fire  man- 
agement in  the  Bahamas  will  likely  depend 
more  on  prescribed  fires  lit  by  trained  profes- 
sionals as  land-use  changes  complicate  fire- 
management  situations.  The  judicious  appli- 
cation of  prescribed  fire  as  a resource  man- 
agement tool  requires  knowledge  of  fire  im- 
pacts, both  direct  and  indirect,  on  ecosystem 
properties.  Although  the  relationship  between 
fire  and  pineyard  vegetation  is  relatively  clear, 
the  impact  of  fire  on  pineyard  wildlife,  espe- 
cially parrots,  is  not  as  well  known.  The 
ground-nesting  behavior  of  the  Abaconian 


508 


O'Brien  et  al.  • FIRE  EFFECTS  ON  SURROGATE  PARROT  NEST 


509 


FIG.  1.  Location  of  the  island  of  Great  Abaco  and  Abaco  National  Park  within  the  Commonwealth  of  the 
Bahamas. 


population  raises  several  important  questions 
regarding  the  ways  in  which  fires  might  affect 
nesting  parrots. 

Fire  can  impact  parrots  both  indirectly  and 
directly.  Indirect  effects  are  mediated  primar- 
ily through  vegetation  and  subsequent  impacts 
on  parrot  food  resources  and  nesting  cover. 
Direct  effects  would  likely  be  most  important 
during  the  nesting  season.  A passing  fire 
might  result  in  increased  temperatures,  smoke, 
and  C02  levels  inside  the  nesting  cavity  that 
could  stress  or  kill  parrot  nestlings  or  adults 
reluctant  to  abandon  the  nest.  Herein,  we  re- 
port the  ambient  conditions  inside  a limestone 
cavity  characteristic  of  Cuban  Parrot  nest  sites 
as  a prescribed  fire  passed  over  it.  Conditions 
are  reported  as  means  ± SD. 

METHODS 

The  study  site  bordered  Abaco  National 
Park  (ANP;  26°  2'  N,  77°  15'  W)  in  the  south- 
ern portion  of  the  island  of  Great  Abaco,  Ba- 
hamas (Fig.  1).  ANP  was  established  in  1994 


by  The  Bahamas  National  Trust  and  encom- 
passes 8,300  ha.  The  habitat  consists  of  pine- 
yard  vegetation  along  with  some  tropical  dry 
forest  known  locally  as  “coppice.”  A forest 
inventory  we  conducted  in  the  vicinity  of  the 
experimental  area  revealed  that  pine  trees  now 
occupying  the  park  are  growing  in  even-aged 
stands.  Mean  tree  height  was  16  m ± 0.6, 
mean  diameter  at  breast  height  was  18.6  cm 
± 1.81,  and  mean  density  was  364  ± 273 
trees/ha. 

On  3 1 October  2004,  a crew  led  by  person- 
nel of  The  Nature  Conservancy  lit  a pre- 
scribed fire  in  Abaco  National  Park  as  a train- 
ing exercise  for  Bahamian  fire  fighters  and  re- 
source managers.  The  crew  used  drip  torches 
to  ignite  the  fire  at  13:00  EST  under  moderate 
weather  conditions:  ~1  m/sec  wind  speed, 
56%  relative  humidity,  and  high  levels  of  fuel 
moisture  resulting  from  rainfall  the  previous 
evening.  The  area  burned  was  a —10  ha  block 
bounded  by  former  logging  roads  and  a high- 
way. Although  the  site’s  exact  fire  history  was 


510 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


unknown,  fuel  loads  were  typical  of  areas  that 
had  not  burned  for  about  3 yr.  The  study  plot 
was  embedded  in  an  area  of  high-density  par- 
rot nesting  activity  (Gnam  and  Burchsted 
1991,  Stahala  2005),  with  an  active  colony  <1 
km  distant.  The  fuel  loads  and  stand  structure 
in  both  the  study  area  and  the  nearby  colonies 
were  similar. 

Prior  to  ignition  in  the  area  to  be  burned, 
we  located  a solution  hole  characteristic  of 
those  used  by  parrots  as  nesting  cavities  (Sny- 
der et  al.  1982,  Gnam  1990).  This  cavity  en- 
trance was  —30  cm  in  diameter,  within  the 
diameter  range  previously  reported  for  parrot 
cavity  entrances,  and  was  approximately  120 
cm  deep,  also  within  the  range  reported  for 
parrot  cavities  (124.2  ± 55.4;  Gnam  1990). 
The  floor  was  dry  and  contained  a small  heap 
of  dried  grass — evidence  of  vertebrate  activity 
within  the  cavity.  In  order  to  measure  tem- 
peratures inside  the  cavity,  we  placed  two 
type-T  thermocouples  read  by  Hobo  datalog- 
gers (Hobo  Pro  Series,  Onset,  Inc.,  Bourne, 
Massachusetts)  on  the  cavity  floor,  and  sus- 
pended another  thermocouple  30  cm  inside 
the  entrance.  As  the  fire  passed  over  the  cav- 
ity, an  infrared  camera  (S60,  FLIR,  Inc.,  Wil- 
sonville,  Oregon)  was  used  to  measure  ground 
surface  temperatures  outside  the  cavity. 

Inside  the  cavity,  we  also  measured  C02 
concentration  and  total  suspended  particulate 
density  by  sampling  air  through  a 4-m-long, 
5-mm-diameter  copper  tube  with  the  end 
placed  10  cm  above  the  surface  of  the  cavity 
floor.  A particulate  matter  (PM)  2.5  filter  (col- 
lects particulate  matter  >2.5  pm)  was  at- 
tached to  the  tube  tip  inside  the  cavity.  At  its 
other  end,  the  tube  was  connected  to  an  air 
pump  set  at  a maximum  flow  rate  of  1 .5 
1/min.  We  measured  C02  levels  with  an  infra- 
red gas  analyzer  (EGM4,  PP  Systems,  Inc., 
Amesbury,  Massachusetts);  the  air  flow  rate 
was  measured  simultaneously  with  a mass 
flow  controller  (Top-Trak  822-OV1-PV1-V1, 
Sierra  Instruments,  Inc.,  Monterey,  Califor- 
nia). The  output  of  the  gas  analyzer  and  mass 
flow  controller  were  measured  every  second 
and  stored  as  1-min  averages  by  a datalogger 
(CR10X,  Campbell  Scientific,  Inc.,  Logan, 
Utah).  All  instruments  were  placed  in  a small 
plastic  enclosure.  To  prevent  fire  damage,  we 
raked  fuel  from  around  the  enclosure,  then 
covered  it  with  a U.S.  Department  of  Agri- 


culture Forest  Service  fire  shelter,  an  alumi- 
nized fiberglass  tent  designed  to  shield  an  en- 
trapped firefighter  from  radiant  energy. 

RESULTS 

Although  a variety  of  ignition  techniques 
were  employed  in  the  area,  a low-intensity 
backing  fire  arrived  at  the  cavity  area  at  ap- 
proximately 15:14.  The  low  fuel  loads  found 
in  the  area,  coupled  with  the  moderate  weather 
conditions,  created  short  flames  (—30  cm 
high)  and  a slow  rate  of  spread;  the  fireline 
crept  along  at  about  15  cm/min  as  the  fire 
passed  the  vicinity  of  the  cavity  entrance.  The 
residence  time  of  the  fire  within  1 m of  the 
cavity  entrance  was  —15  min.  The  maximum 
fire  temperature  recorded  outside  the  cavity 
entrance  was  803°  C.  We  observed  minor  tem- 
perature changes  inside  the  cavity  as  the  fire 
passed:  a 5°  C increase  occurred  30  cm  inside 
the  entrance,  and  a 0.4°  C increase  occurred  at 
the  cavity  floor  (Fig.  2A). 

A total  of  0.903  mg  of  suspended  particu- 
lates was  captured  on  the  PM  2.5  air  filter  af- 
ter 0. 1 m3  of  air  had  been  filtered.  Changes  in 
air  flow  through  the  filter  indicated  that  smoke 
accumulation  was  constant  for  a brief  period, 
causing  a steep,  linear  decrease  in  air  flow,  but 
then  smoke  concentration  declined  toward  an 
asymptote  (Fig.  2B).  There  was  little  lingering 
smoke  production,  as  almost  no  smoldering 
occurred  following  passage  of  the  flaming 
front. 

C02  levels  in  the  cavity  rose  sharply  when 
the  fire  approached  the  entrance  and  then 
dropped  sharply  as  the  fire  moved  past  (Fig. 
2B).  The  maximum  concentration  recorded 
was  2,092  ppm.  Concentrations  of  C02 
>2,000  ppm  occurred  for  5 min,  and  concen- 
trations >1,000  ppm  occurred  for  19  min. 

DISCUSSION 

We  observed  relatively  benign  conditions 
inside  the  cavity  as  the  fire  passed.  The  mag- 
nitude of  temperature  change  caused  by  the 
fire  was  similar  to  that  observed  during  a typ- 
ical diurnal  cycle  in  the  absence  of  fire  (GPM 
unpubl.  data).  Inside  the  cavity,  smoke  levels 
were  low,  and  C02  levels  rose  moderately,  but 
declined  quickly  as  the  fire  passed.  The  C02 
concentrations  we  observed  probably  would 
not  have  had  much  effect  on  parrots:  although 
data  on  C02  effects  on  birds  were  not  avail- 


O’Brien  et  al.  • FIRE  EFFECTS  ON  SURROGATE  PARROT  NEST 


51  1 


Time  (HH:MM) 

FIG.  2.  Ambient  conditions  inside  a Cuban  Parrot  surrogate  nest  cavity  in  Abaco  National  Park,  Great  Abaco, 
Bahamas.  (A)  Temperatures  30  cm  inside  the  cavity  entrance  and  on  the  cavity  floor  100  cm  from  the  entrance. 
(B)  C02  concentration  and  air  flow  rate  through  a particulate  filter  as  a fire  passed  by  the  cavity;  the  flaming 
front  approached  the  cavity  entrance  at  15:14  EST  and  passed  at  approximately  15:35. 


able,  the  maximum  permissible  exposure  for 
humans,  as  determined  by  the  Occupational 
Safety  and  Health  Administration  (1997),  is 
an  8-hr  time- weighted  average  of  5,000  ppm 


with  a short-term  (<30  min)  exposure  limit  of 

30.000  ppm.  Concentrations  lower  than 

15.000  ppm  have  no  detectable  effect  on  peo- 
ple. 


512 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Although  the  tolerance  of  Cuban  Parrots  to 
C02  and  smoke  is  unknown,  they  are  capable 
of  surviving  fires  while  nesting.  In  2003,  a 
wildfire  passed  over  20  occupied  nests  and  did 
not  result  in  decreased  fledging  success  (Sta- 
hala  2005).  Another  wildfire  that  occurred  in 
ANP  in  2005  resulted  in  a similar  lack  of  mor- 
tality (GPM  pers.  obs.).  Our  measurements 
also  provide  direct  evidence  that  fire-induced 
elevations  in  temperature  and  C02  concentra- 
tion would  cause  minimal  stress.  Although  we 
sampled  only  a single  cavity  (thus  limiting  our 
sphere  of  inference),  our  results  are  likely  rep- 
resentative, given  the  low  fuel  loads  that  are 
typically  found  in  nesting  colonies. 

Burning  while  the  birds  are  actively  nesting 
might  have  a relatively  minor  impact  on  con- 
ditions inside  the  cavity.  Nonetheless,  the 
threatened  status  and  restricted  range  of  the 
ground-nesting  population,  as  well  as  the  am- 
ple opportunity  to  set  fires  outside  the  breed- 
ing season,  indicates  that  setting  prescribed 
fires  when  cavities  are  occupied  needs  to  be 
considered  carefully.  The  timing  of  a fire  ap- 
pears to  be  important,  as  parrot  pairs  seem  to 
choose  new  nesting  sites  in  recently  burned 
areas.  Although  it  appears  that  reduced  cover 
due  to  fire  has  no  significant  effect  on  preda- 
tion rates  of  nesting  parrots  (Stahala  2005), 
unbumed  patches  near  nests  might  attract 
predators  in  otherwise  burned  areas.  If  this 
were  true,  creating  firebreaks  around  colonies 
to  protect  parrots  from  fire  might  lead  to  in- 
creased parrot  mortality  and  would  not  be  jus- 
tifiable. While  the  direct  effects  of  fire  on  con- 
ditions inside  a nest  cavity  of  Abaco’s  Cuban 
Parrots  appear  negligible,  indirect  effects  of 
frequent  fires  are  of  paramount  importance, 
mainly  because  they  reduce  fuel  loads  and  fire 


intensities  and  are  critical  for  maintaining 
pineyard  ecosystems. 

ACKNOWLEDGMENTS 

We  would  like  to  thank  the  Friends  of  the  Environ- 
ment (Bahamas)  for  providing  invaluable  logistical 
support,  the  Bahamas  Department  of  Agriculture,  es- 
pecially D.  Knowles,  the  Bahamas  National  Trust,  and 
the  U.S.  Department  of  Agriculture  Forest  Service  In- 
ternational Program  for  facilitating  our  research.  Three 
anonymous  referees  provided  valuable  comments  that 
improved  this  manuscript.  K.  Smith  provided  logistical 
support  and  assisted  with  manuscript  preparation. 

LITERATURE  CITED 

Attrill,  R.  1981.  The  status  and  conservation  of  the 
Bahamas  Amazon  ( Amazonas  leucocephala  ba- 
hamensis).  Pages  81-87  in  Conservation  of  New 
World  parrots  (R.  F.  Pasquir,  Ed.).  ICBP  Technical 
Publication  no.l.  Smithsonian  Institution  Press, 
Washington,  D.C. 

Gnam,  R.  S.  1990.  Conservation  of  the  Bahama  Parrot. 
American  Birds  44:32-36. 

Gnam,  R.  and  A.  Burchsted.  1991.  Population  esti- 
mates for  the  Bahama  Parrot  on  Abaco  Island,  Ba- 
hamas. Journal  of  Field  Ornithology  62:139-146. 
Loope,  L.  L.  and  V.  I.  Dunevitz.  1981.  Impact  of  fire 
exclusion  and  invasion  of  Schinus  terebinthifolius 
on  limestone  rockland  pine  forest  of  southeastern 
Florida.  Report  T-645.  Everglades  National  Park, 
South  Florida  Research  Center,  Homestead,  Flor- 
ida. 

Occupational  Safety  and  Health  Administration. 
1997.  Limits  for  air  contaminants.  Occupational 
Health  and  Safety  Administration,  Code  of  federal 
regulations  29  C.F.R.  1910.1000,  Table  Z-l.  U.S. 
Government  Printing  Office,  Washington,  D.C. 
Robertson,  W.  B.,  Jr.  1955.  An  analysis  of  the  breed- 
ing bird  populations  of  tropical  Florida  in  relation 
to  the  vegetation.  Ph.D.  dissertation.  University  of 
Illinois,  Urbana. 

Snyder,  N.  F.  R..  W.  B.  King,  and  C.  B.  Kepler.  1982. 
Biology  and  conservation  of  the  Bahama  Parrot. 
Living  Bird  19:91-114. 

Stahala,  C.  2005.  Demography  and  conservation  of 
the  Bahama  Parrot  on  Great  Abaco  Island.  M.Sc. 
thesis.  North  Carolina  State  University,  Raleigh. 


The  Wilson  Journal  of  Ornithology  1 18(4):5 13-526,  2006 


UTILITY  OF  OPEN  POPULATION  MODELS:  LIMITATIONS  POSED 
BY  PARAMETER  ESTIMABILITY  IN  THE  STUDY  OF 
MIGRATORY  STOPOVER 

SARA  R.  MORRIS,1 3 AMANDA  M.  LARRACUENTE,1  KRISTEN  M.  COVINO,1 
MELISSA  S.  MUSTILLO,1  KATHRYN  E.  MATTERN,1  DAVID  A.  LIEBNER,1  AND 

H.  DAVID  SHEETS2 3 


ABSTRACT. — Open  population  models  using  capture-mark-recapture  (CMR)  data  have  a wide  range  of  uses 
in  ecological  and  evolutionary  contexts,  including  modeling  of  stopover  duration  by  migratory  passerines.  In 
using  CMR  approaches  in  novel  contexts  there  is  a need  to  determine  the  conditions  under  which  open  population 
models  may  be  employed  effectively.  Our  goal  was  to  determine  whether  there  was  a simple  a priori  mechanism 
of  determining  the  conditions  under  which  CMR  models  could  be  used  effectively  in  the  study  of  avian  stopover 
ecology.  Using  banding  data  (n  = 188  capture  histories),  we  examined  the  challenges  of  using  CMR-based 
models  due  to  parameter  inestimability,  adequacy  of  descriptive  power  (Goodness-of-Fit,  GOF),  and  parameter 
uncertainty.  These  issues  become  more  apparent  in  studies  with  limited  observations  in  a capture  history,  as  is 
often  the  case  in  studies  of  avian  stopover  duration.  Limited  sample  size  and  sampling  intensity  require  an 
approach  to  reducing  the  number  of  fitted  parameters  in  the  model.  Parameter  estimability  posed  the  greatest 
restriction  on  the  utility  of  open  population  models,  with  high  parameter  uncertainty  posing  a lesser  challenge. 
Results  from  our  study  also  indicate  the  need  for  >10  observations  per  estimated  parameter  (approximately  3 
birds  captured  or  recaptured  per  day)  to  provide  a reasonable  chance  of  successfully  estimating  all  model 
parameters.  Received  13  July  2005,  accepted  20  May  2006. 


Migratory  birds  frequently  use  stopovers  to 
complete  migration  successfully  between  their 
breeding  and  wintering  grounds.  Stopover 
sites  provide  refuge  from  predators,  protection 
against  inclement  weather,  and  food  resources 
to  allow  fat  deposition  to  fuel  migratory  flight. 
It  is  thought  that  many  migrating  passerines 
cannot  store  enough  fat  to  complete  their  mi- 
gration in  a single  transit,  but  must  refuel  by 
foraging  at  stopover  sites  along  their  routes 
(Dunn  2001,  Schwilch  and  Jenni  2001).  Pro- 
viding evidence  for  the  use  of  stopover  sites 
for  refueling,  Moore  and  Abom  (2000)  doc- 
umented increased  activity  patterns  and  dif- 
ferential habitat  use  by  lean  versus  fat  mi- 
grants. Lean  migrants  needing  to  refuel  may 
stay  longer  at  stopover  sites  than  fat  migrants 
(Moore  and  Kerlinger  1987,  Yong  and  Moore 
1997),  and  the  rate  of  mass  gain  also  may  af- 
fect stopover  duration.  The  length  of  time  that 
migrants  stay  at  stopover  sites  will  affect  the 
total  duration  of  migration  and  may  affect  the 
ability  of  birds  to  obtain  quality  territories. 


1 Dept,  of  Biology,  Canisius  College,  2001  Main  St., 
Buffalo,  NY  14208,  USA. 

2 Dept,  of  Physics,  Canisius  College,  2001  Main  St., 
Buffalo,  NY  14208,  USA. 

3 Corresponding  author;  e-mail: 
morriss@canisius.edu 


Species-specific  stopover  patterns  may  reflect 
both  intrinsic  characteristics  and  ecological 
factors  associated  with  individual  stopover 
sites  (Kaiser  1999).  Schaub  et  al.  (2001)  argue 
for  accurate  estimates  of  stopover  duration  to 
test  models  of  optimal  migration  strategy,  spe- 
cifically the  trade-off  between  time  spent  in 
flight  or  at  stopovers. 

Although  the  importance  of  en  route  mi- 
gratory stopover  sites  is  well  recognized 
(Moore  2000,  Petit  2000,  Sillett  and  Holmes 
2002,  Heglund  and  Skagen  2005),  all  sites  are 
not  equal.  Mehlman  et  al.  (2005)  recommend 
that  important  stopover  sites  be  identified 
based  on  the  relative  migrant  abundance,  the 
availability  of  resources  that  allow  birds  to  re- 
plenish fat  reserves,  and  the  location  of  the 
site  relative  to  other  sites  and  ecological  bar- 
riers. However,  specific  criteria  for  assessing, 
and  statistical  approaches  for  comparing,  sites 
have  not  been  established.  Furthermore,  there 
is  a recognized  need  for  research  on  how  sites 
differ  by  season,  species,  and  species  demog- 
raphy (Mehlman  et  al.  2005,  Partners  in  Flight 
Research  Working  Group  2002). 

Since  the  mid-1980s,  numerous  researchers 
have  described  the  basics  of  the  stopover  ecol- 
ogy of  migratory  landbirds  at  individual  sites 
along  the  northern  coast  of  the  Gulf  of  Mexico 


513 


514 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  4,  December  2006 


(Moore  and  Kerlinger  1987,  Moore  et  al. 
1990,  Kuenzi  et  al.  1991),  the  New  England 
coast  (Morris  et  al.  1994,  1996;  Parrish  2000), 
the  Great  Lakes  coasts  (Jones  et  al.  2002, 
Bonter  2003),  and  in  western  states  (Winker 
et  al.  1992,  Finch  and  Yong  2000).  Most  of 
these  studies  provide  simple  analyses  of  stop- 
over duration  based  on  recapturing  banded 
birds.  Calculating  the  amount  of  time  lapsing 
between  the  first  capture  and  the  last  recapture 
(Cherry  1982)  has  been  the  traditional  method 
of  estimating  stopover  duration  at  a given  site; 
however,  including  only  recaptured  birds  pro- 
vides conservative  estimates  of  stopover  du- 
ration because  birds  not  recaptured  have  not 
necessarily  left  the  field  site.  If  only  recap- 
tured birds  are  used  in  analyses  (regularly 
<5%  of  all  banded  migrants  are  recaptured), 
this  simple  approach  might  provide  a biased 
view  of  site  use  because  >95%  of  migrants 
are  excluded  from  analyses. 

The  limitations  of  the  minimum  stopover 
approach  have  resulted  in  the  suggestion  that 
open  population  models  based  on  capture- 
mark-recapture  (CMR)  data  be  used  to  esti- 
mate stopover  duration  (Lavee  et  al.  1991, 
Holmgren  et  al.  1993,  Kaiser  1995,  Schaub  et 
al.  2001).  The  Pradel  (1996)  extension  of  the 
Cormack-Jolly-Seber  (CJS)  models  allows  for 
a range  of  models  of  the  probabilities  of  ani- 
mal capture,  arrival,  and  departure  within  each 
interval  of  a given  study  period.  A number  of 
useful  statistics  may  be  derived  from  the  sto- 
chastic models,  including  mean  time  animals 
are  present  in  the  study  area,  mean  capture 
probability,  and  temporal  patterns  of  arrival, 
departure,  and  population  size.  These  models 
also  could  allow  meaningful  comparisons  of 
several  stopover  characteristics  among  sites. 

Although  the  assumptions  used  in  deriving 
open  population  models  are  widely  known 
(e.g.,  Pollock  et  al.  1990,  Cooch  and  White 
2005),  the  conditions  under  which  these  mod- 
els can  be  used  are  rarely  discussed.  Charac- 
teristics of  the  data  (i.e.,  capture/recapture  his- 
tories)— especially  sample  size,  number  of 
temporal  sampling  intervals  available,  recap- 
ture/resighting/recovery rate,  etc. — may  great- 
ly impact  the  potential  usefulness  of  these 
models.  To  use  a given  open  population  mod- 
el, first  all  the  model  parameters  must  be  es- 
timated. Typically,  parameter  estimates  are 
obtained  using  numerical  maximum  likeli- 


hood methods;  characteristics  of  the  capture 
history  and  the  model’s  mathematical  struc- 
ture will  determine  the  number  of  parameters 
that  can  be  reliably  estimated.  Parameters  that 
are  inestimable  due  to  limitations  of  a given 
capture  history  are  extrinsically  non-identifi- 
able  (McCullagh  and  Nelder  1989,  Viallefont 
et  al.  1998).  Capture  histories  that  involve 
long  periods  of  time,  particularly  those  with 
relatively  few  captures  and/or  recaptures,  of- 
ten prevent  successful  estimation  of  all  param- 
eter values;  the  resulting  extrinsic  non-identi- 
fiability  of  parameters  either  precludes  the  use 
of  open  population  models  or  requires  reduc- 
ing the  number  of  parameters. 

One  approach  to  reducing  the  number  of 
parameters  that  must  be  fitted  for  a given 
model  is  to  pool  observations  over  several 
consecutive  observation  periods  (e.g.,  Schaub 
and  Jenni  2001,  Schaub  et  al.  2001).  However, 
pooling  may  bias  the  parameter  estimates  and 
preclude  comparing  models  with  different 
pooling  intervals  (Hargrove  and  Borland 
1994,  Morris  et  al.  2005b).  The  difficulty  as- 
sociated with  the  need  to  establish  this  basic 
temporal  interval  has  been  recognized  in  the 
paleontological  literature  (Connolly  and  Mil- 
ler 2001,  Xu  et  al.  2005),  where  it  has  been 
addressed  by  determining  whether  or  not  anal- 
ysis results  remain  consistent  as  the  pooling 
interval  is  changed.  Additional  detailed  dis- 
cussion of  pooling  and  its  effects  appears  to 
be  lacking  in  both  the  statistical  and  ecologi- 
cal literature.  An  alternative  to  pooling  is  to 
use  multiple-day  constancy  (MDC;  Fig.  1), 
which  holds  parameter  values  fixed  over  a 
given  “constancy”  interval,  thus  reducing  the 
number  of  parameters  while  retaining  all  in- 
formation in  the  capture  history  (Morris  et  al. 
2005a).  Regardless  of  the  method  used  to  re- 
duce the  number  of  parameters,  decreasing  the 
number  of  parameters  in  a model  will  increase 
the  likelihood  that  all  parameters  can  be  suc- 
cessfully estimated,  by  reducing  the  incidence 
of  extrinsic  non-identifiability. 

When  using  open  population  models,  good- 
ness-of-fit  (GOF)  tests  must  be  applied  to  de- 
termine whether  the  models  have  adequate  de- 
scriptive power  prior  to  biological  applica- 
tions. Two  distinct  approaches  (analytical  tests 
based  on  contingency  tables  and  numerical 
tests  based  on  comparing  observed  model 
misfit  or  deviance  to  estimates  of  misfit  de- 


Morris  et  al.  • UTILITY  OF  OPEN  POPULATION  MODELS 


515 


A 


Day 

Day 

Day 

Day 

Day 

Day 

Day 

Day 

Day 

1 

2 

3 

4 

5 

6 

7 

8 

9 

Pi 

P2 

Pi 

P4 

Ps 

Ps 

Pv 

Ps 

Ps 

| 0;  I I >2  I I foi  I I 04  I I §5  I I <)>6  1 1 4*7  1 1 08  | | 4*9  • • • 

~T1  | Yi  | | Y>  | | YS  | m | Y7  | 1 YS  | | 79  1 ■ ■ . 


B 


Interval  1 
(Days  1-3) 

Interval  2 
(Days  4-6) 

Interval  3 
(Days  7-9) 

Pi 

Pi 

Pi 

<t>2 

Yi 

Yi 

C 


FIG.  1 . Open  population  models  may  be  used  to  estimate  stopover  duration  by  migratory  birds  by  estimating 
daily  rates  of  capture,  arrival,  and  departure.  Large  numbers  of  parameters  are  required  to  work  with  (A)  raw 
data,  while  both  (B)  pooled  data  (3-day  pooling  interval)  and  (C)  multiple-day  constancy  (MDC,  3-day  MDC 
interval)  provide  a reduction  in  the  number  of  parameters  in  the  open  population  models  fitted  to  bird  banding 
data.  Since  limited  sample  sizes  make  parameter  estimation  difficult,  some  reduction  in  the  number  of  parameters 
may  allow  use  of  these  models  with  smaller  data  sets.  Both  pooling  and  MDC  approaches  reduce  the  number 
of  fitted  parameters:  p = probability  of  capture;  4>  = probability  that  a bird  captured  on  one  day  remained  until 
the  following  day  (i.e.,  survival);  and  y = probability  that  a bird  captured  on  one  day  was  there  the  day  before 
(i.e.,  seniority).  Pooling,  however,  loses  information  from  multiple  captures  in  the  same  interval,  whereas  MDC 
retains  information  on  all  captures.  Figure  adapted  from  Morris  et  al.  (2005a). 


516 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


rived  from  simulations)  have  been  used  to  de- 
termine whether  open  population  models  fit 
the  data.  Once  the  most  complex  model  passes 
the  GOF  test,  selection  of  the  most  appropri- 
ate model  (of  those  nested  within  this  most 
complex  model)  for  the  data  using  Akaike’s 
Information  Criterion  (AIC)  can  occur.  Even 
when  models  can  be  chosen  and  fit,  the  vari- 
ances of  parameter  estimates  obtained  from 
open  population  models  may  be  too  large  for 
the  estimates  to  be  useful.  The  coefficient  of 
variation  (CV;  the  standard  deviation  of  the 
estimate/the  value  of  the  estimate  X 100)  may 
be  used  to  assess  the  potential  utility  of  stop- 
over estimates.  A low  CV  is  necessary  for  ef- 
fective comparison  of  statistical  measures 
among  species,  locations,  and/or  time  periods. 
However,  little  attention  has  been  paid  to  the 
dependence  of  the  CV  on  the  characteristics 
of  the  capture  history. 

In  this  study,  we  examined  capture  histories 
from  migration  banding  data  to  determine  the 
utility  of  open  population  models  for  estimating 
avian  stopover  duration.  We  used  a large  num- 
ber of  field  capture  histories  ( n = 188)  from 
migration  banding  datasets  rather  than  relying 
on  computer  simulations.  Whereas  computer 
simulations  would  provide  greater  control  over 
parameters,  we  wanted  to  be  sure  to  cover  a 
wide  range  of  natural  conditions  represented  by 
empirical  data.  Specifically,  we  were  interested 
in  determining  how  data  characteristics  affect 
parameter  estimability  (through  extrinsic  non- 
identifiability),  the  ability  of  models  to  pass 
GOF  tests,  and  the  CV  of  stopover  duration  es- 
timates. Estimating  the  range  of  sample  sizes 
and  recapture  rates  to  which  open  population 
models  can  be  fitted  may  help  us  determine 
whether  these  approaches  are  appropriate  for  a 
particular  capture  history.  To  that  end,  our  re- 
sults indicate  the  conditions  under  which  open 
population  models  can  be  used  effectively  with 
banding  data. 

METHODS 

Data  collection. — Migrating  birds  were 
captured  in  mist  nets  at  Appledore  Island, 
Maine  (1996-2002);  Star  Island,  New  Hamp- 
shire (1999  and  2000);  and  Hamlin  Beach 
State  Park,  near  Rochester,  New  York  (1999 
and  2000).  Mist  nets  were  operated  daily  dur- 
ing the  spring  and  fall  migration  seasons  ex- 
cept during  inclement  weather.  All  birds  cap- 


tured or  recaptured  were  transported  to  a cen- 
tral location  for  banding  and  data  collection. 

For  species  with  a sample  size  >50  indi- 
viduals in  a single  season,  we  created  a cap- 
ture history  that  indicated  whether  any  one  in- 
dividual was  captured  on  a given  day.  Using 
this  capture  history,  we  calculated  minimum 
stopover  by  subtracting  the  date  of  first  cap- 
ture from  the  date  of  final  capture,  following 
Cherry  (1982).  Additionally,  we  calculated  a 
variety  of  descriptive  statistics  that  were  used 
for  discriminant  function  analyses  (see  be- 
low). 

Capture-mark-recapture. — The  first  step  in 
the  analysis  was  to  determine  the  most  com- 
plex model  for  which  all  parameters  could  be 
estimated.  Numerical  maximum  likelihood 
methods  were  used  to  fit  Pradel’s  (1996)  ex- 
tension of  the  CJS  open  population  models  to 
each  capture  history.  Pradel’s  model  requires 
estimation  of  sighting  (p  = probability  of  cap- 
ture), seniority  (y  = probability  that  the  bird 
was  present  at  a stopover  site  during  the  pre- 
vious day),  and  survival  (4>  = probability  of 
remaining  at  a stopover  site  until  the  next 
day).  We  considered  time-dependent  open 
population  models  with  MDC  intervals  (Mor- 
ris et  al.  2005a)  ranging  from  1 to  7 days.  In 
the  MDC  approach  to  time-varying  parame- 
ters, the  parameters  are  fixed  over  the  MDC 
interval.  However,  all  captures  and  recaptures 
within  and  between  MDC  intervals  have  an 
influence  on  the  likelihood  function  and, 
hence,  the  parameter  estimates.  Each  of  these 
time-dependent  models  (in  which  sighting, 
survival,  and  seniority  probabilities  were  all 
free  to  vary  from  one  constancy  interval  to 
the  next)  was  fitted  to  the  capture  history,  and 
the  number  of  extrinsically  non-identifiable 
parameters  was  identified  using  an  estimate  of 
the  rank  of  the  Hessian  matrix  (Viallefont  et 
al.  1998).  Rank  deficiency  in  the  Hessian  ma- 
trix was  estimated  by  using  finite-difference 
methods,  and  then  tested  using  the  singular 
value  decomposition  method  (Viallefont  et  al. 
1998).  Rank  deficiency  was  taken  as  indicat- 
ing extrinsic  parameter  non-identifiability  in  a 
model.  While  some  parameters  in  Pradel’s  ex- 
tension of  the  CJS  model  are  non-identifiable 
due  the  model’s  structure  (i.e.,  intrinsic  ines- 
timability),  this  form  of  inestimability  is  part 
of  the  model,  and  does  not  negatively  impact 
its  further  use.  We  are  concerned  here  with 


Morris  et  al.  • UTILITY  OF  OPEN  POPULATION  MODELS 


517 


extrinsically  inestimable  parameters  in  band- 
ing data.  Inestimability  makes  it  difficult  to 
use  either  the  Schaub  et  al.  (2001)  formulation 
of  the  stopover  duration  or  the  more  recent 
estimate  put  forward  by  Efford  (2005).  Al- 
though Efford’s  approach  appears  simpler 
than  that  of  Schaub  et  al.  (2001),  it  still  re- 
quires an  estimate  of  the  distribution  of  arrival 
times,  thus  necessitating  the  estimation  of  the 
same  number  of  parameters  (See  Efford’s 
equation  5 and  discussion).  To  be  useful  in 
estimating  stopover  duration  (Schaub  and  Jen- 
ni  2001,  Schaub  et  al.  2001),  all  intrinsically 
estimable  parameters  in  a model  had  to  be 
completely  identifiable,  so  those  capture  his- 
tories with  non-identifiable  parameters  due  to 
the  structure  of  the  data  in  all  MDC  intervals 
tested  were  judged  unusable  for  further  analysis. 

We  used  software  written  by  HDS  and  DAL 
using  MATLAB  (The  Math  Works,  Inc.  1992) 
to  implement  Pradel’s  population  growth  rate 
(PGR)  method  (Pradel  1996).  We  compared 
the  performance  of  our  software  to  that  of 
MARK  (White  and  Burnham  1999,  Cooch 
and  White  2005)  and  SURGE  (Lebreton  et  al. 
1992,  Pradel  and  Lebreton  1993,  Cooch  et  al. 
1997);  it  produced  identical  results  for  a num- 
ber of  capture  histories,  both  from  our  data 
and  from  example  files  distributed  with 
MARK.  When  using  very  sparse  data,  our 
software  and  SURGE  had  similar  convergence 
properties,  with  results  depending  less  on 
sample  size  than  they  did  in  MARK,  which 
may  be  attributable  to  differences  in  the  par- 
ticular link  function  (the  default  choice)  we 
used  in  MARK  (Cooch  and  White  2005);  this 
particular  difference  in  performance  was  not 
investigated  in  depth. 

Since  capture  histories  included  a range  of 
sample  sizes  and  durations,  comparing  capture 
histories  required  a time-invariant  measure  of 
sampling  intensity.  We  used  the  number  of  ob- 
servations (sum  of  all  capture  and  recapture 
events)  per  estimated  parameter  in  a 7-day, 
time-dependent  MDC  model  as  the  measure 
of  observations  per  parameter.  The  7-day 
MDC  model  had  the  lowest  number  of  param- 
eters of  any  model  used  in  the  estimability 
determination  procedure  discussed  above.  We 
divided  the  capture  histories  into  three  cate- 
gories, based  on  the  number  of  observations 
(#)  per  estimated  parameter:  (1)2<#<5, 
(2)  5 < # < 10,  and  (3)  # > 10.  Our  highest 


category  (>10  observations  per  parameter) 
roughly  corresponds  with  three  birds  of  that 
species  captured  or  recaptured  per  day.  This 
categorization  allowed  us  to  examine  the  de- 
pendence of  estimability  on  the  ratio  of  ob- 
servations to  parameters,  and  does  not  require 
that  the  sampling  intervals  used  in  a study  be 
in  units  of  days. 

Capture  histories  were  tested  for  GOF  by 
assessing  the  ability  of  time-dependent  (i.e., 
the  most  complex)  models  to  fit  the  data.  Both 
analytical  tests  (based  on  contingency  tables) 
and  numerical  tests  (based  on  parametric 
bootstrap  procedures)  have  been  used  in  con- 
junction with  CMR  models.  The  first  approach 
is  to  use  contingency  tables  to  test  whether 
assumptions  of  the  open  population  models 
are  violated.  Specifically,  contingency  tables 
are  used  to  test  the  assumptions  that  each 
marked  animal  in  the  population  at  time  t has 
(1)  the  same  probability  of  recapture,  and  (2) 
the  same  probability  of  survival  (Pollock  et 
al.  1990).  Several  variations  on  these  tests 
have  been  incorporated  into  the  programs  RE- 
LEASE (Lebreton  et  al.  1992,  Burnham  et  al. 
1987),  MARK  (White  and  Burnham  1999), 
and  U-CARE  (Choquet  et  al.  2005).  The  con- 
tingency tables  can  be  pooled  to  produce  an 
overall  chi-square  statistic  for  the  capture  his- 
tory as  a whole,  as  well  as  testing  specific 
hypotheses  about  violations  of  model  assump- 
tions. When  faced  with  sparse  data,  the  con- 
tingency tables  may  be  pooled  to  improve 
their  performance,  particularly  when  the  num- 
ber of  expected  outcomes  in  one  or  more  cat- 
egories of  the  contingency  table  is  very  low. 
Pooling  contingency  tables,  however,  does  not 
always  result  in  tables  with  enough  entries  in 
each  cell  to  be  useful.  All  of  our  capture  his- 
tories that  had  estimable  models  for  MDC  in- 
tervals of  <7  days  were  submitted  to  GOF 
testing  using  the  contingency  table  methods  in 
U-CARE  (Choquet  et  al.  2005). 

The  second  alternative  is  to  use  numerical 
simulations  to  determine  whether  the  ob- 
served model  deviance  is  consistent  with  the 
deviance  distribution  obtained  by  using  the 
model  in  a parametric  bootstrap  procedure 
(also  called  a Monte  Carlo  simulation).  The 
model  deviance  is  the  difference  between  the 
observed  log-likelihood  and  the  log-likelihood 
for  a “saturated”  model,  and  it  serves  as  a 
model’s  measure  of  fit.  In  such  a procedure 


518 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


(as  implemented  in  MARK  and  our  software), 
the  model  is  used  to  generate  a series  of  sim- 
ulated capture  histories  of  the  same  size  as  the 
original  capture  history.  The  model  is  fit  to 
each  of  the  simulated  capture  histories  in  turn, 
and  a confidence  interval  for  the  deviances 
observed  over  the  simulated  data  is  obtained. 
If  the  observed  deviance  is  high  (above  the 
95%  upper  bound  of  the  simulation  devianc- 
es), then  it  may  be  possible  to  continue  the 
analysis  by  computing  an  estimated  variance 
inflation  factor  (c)  and  using  this  to  adjust  the 
statistics  of  model  choice  (White  2002,  Cooch 
and  White  2005).  Data  sparseness  also  affects 
this  parametric  bootstrap  approach  to  GOF 
testing  because  the  model  must  be  fit  to  the 
simulation  data  during  the  estimation  of  the 
range  of  deviances.  Each  capture  history  was 
tested  for  GOF  at  the  lowest  MDC  interval 
for  which  the  model  parameters  were  identi- 
fiable, using  software  written  by  HDS  and 
DAL.  Parameter  identifiability  was  monitored 
during  the  GOF  testing  procedure,  as  it  also 
poses  a problem  when  conducting  Monte  Car- 
lo simulations.  Similarly,  capture  histories  ex- 
hibiting evidence  of  a lack-of-fit  (i.e.,  those 
with  deviances  outside  the  95%  confidence  in- 
tervals from  the  simulations)  over  all  seven 
intervals  were  not  subjected  to  further  analy- 
sis. We  did  not  make  use  of  the  c estimation 
procedure  (White  2002,  Cooch  and  White 
2005),  as  it  turned  out  that  only  two  capture 
histories  fell  into  this  category  of  results. 

After  a time-dependent  model  was  shown 
to  exhibit  GOF,  we  compared  competing  mod- 
els to  determine  which  model  was  optimal  for 
producing  stopover  estimates.  Model  selection 
compared  all  prospective  models  over  several 
MDC  intervals  for  each  capture  history,  be- 
ginning with  the  smallest  MDC  interval  that 
passed  GOF.  We  excluded  prospective  models 
that  had  both  constant  seniority  and  survival 
because  they  predict  a population  size  that  is 
constant  or  monotonically  increasing  or  de- 
creasing. Based  on  field  observations,  we 
know  that  during  the  migration  period  the 
population  present  at  a stopover  site  increases 
to  a maximum  value  and  then  declines  to  zero, 
making  any  model  predicting  constant  popu- 
lation size  or  a monotonic  pattern  of  change 
in  population  size  biologically  unreasonable 
(see  Burnham  and  Anderson  1998  for  a dis- 
cussion of  the  exclusion  of  biologically  un- 


reasonable models).  The  lowest  AICc  value  in- 
dicated the  most  appropriate  model  for  a given 
capture  history,  thus  determining  the  appro- 
priate MDC  interval  and  whether  each  param- 
eter was  constant  or  time-dependent.  In  addi- 
tion to  determining  which  model  was  the  most 
appropriate,  the  AICc  score  was  used  to  assign 
a relative  AICc  weight  (w)  to  each  model, 
which  reflects  the  relative  probability  that 
each  model  is  correct.  If  the  AICc  weight  of 
the  chosen  model  was  <0.95,  we  also  includ- 
ed additional  models  with  relatively  high  AICc 
weights.  Thus,  the  number  of  models  included 
was  determined  by  a cumulative  AICc  weight 
of  0.95,  so  that  all  models  with  a reasonable 
chance  of  being  correct  were  considered.  We 
used  a bootstrapping  procedure  to  determine 
the  total  stopover  duration  estimate  and  the 
standard  deviation  of  this  estimate  (following 
Schaub  et  al.  2001). 

Schaub  et  al.  (2001)  present  a derivation  of 
the  expected  total  stopover  duration  calculated 
as  a daily  value;  we  report  the  average  total 
stopover  duration  over  the  migration  season. 
In  our  method,  the  daily  stopover  is  weighted 
by  the  estimated  probability  of  arrival  times, 
using  the  estimated  population  growth  rate  as 
presented  by  Pradel  (1996).  Efford  (2005)  ar- 
gues that  the  total  stopover  duration  (Schaub 
et  al.  2001)  produces  an  overestimate  of  the 
actual  duration.  Efford  (2005)  advocates  using 
a weighted  average  of  Schaub  et  al.’s  “stop- 
over-after” estimate  using  a weighting  derived 
from  Schwarz  and  Arnason’s  (1996)  estimates 
of  the  distribution  of  arrival  times  (Equation 
5 in  Efford  2005).  We  also  present  the  stop- 
over-after statistic,  again  weighted  using  the 
estimated  population  growth  rate  as  derived 
from  Pradel  (1996).  Conceptually,  this  ap- 
proach is  the  same  as  that  presented  by  Efford, 
although  the  computations  may  differ  slightly, 
as  the  Pradel  (1996)  parameterization  of  the 
problem  differs  from  that  used  by  Schwarz 
and  Amason  (1996). 

In  addition  to  having  adequate  descriptive 
power  and  being  estimable,  the  chosen  model 
must  yield  a useful  statistic  for  comparisons. 
The  coefficient  of  variation  (CV)  was  used  to 
determine  usefulness  of  the  total  stopover  sta- 
tistic estimated  for  each  species  in  each  sea- 
son. CV  was  calculated  by  dividing  the  stan- 
dard error  of  the  total  stopover  estimate  by  its 
mean  and  multiplying  by  100.  In  this  study. 


Morris  et  al.  • UTILITY  OF  OPEN  POPULATION  MODELS 


519 


TABLE  1.  Summary  of  the  utility  of  open  opulation  models  in  three  categories  representing  the  number  of 
observations  (#)  per  estimated  parameter  for  a given  capture  history  from  avian  banding  data.  To  be  applicable, 
models  had  to  have  estimable  parameters  and  pass  goodness-of-fit  (GOF)  testing.  As  the  number  of  observations 
per  parameter  increased,  the  number  of  capture  histories  that  could  be  analyzed  using  open  population  models 
also  increased.  Parameter  inestimability  in  both  model  fitting  and  GOF  testing  poses  the  greatest  impediment  to 
the  use  of  open  population  models  at  these  sample  sizes.  Bird  banding  data  were  collected  during  spring  and 
fall  migration  on  Appledore  Island,  Maine  (1996-2002);  Star  Island,  New  Hampshire  (1999-2000);  and  Hamlin 
Beach  State  Park,  New  York  (1999-2000).  The  banding  data  were  used  to  create  capture  histories,  which  indicate 
whether  and  individual  bird  was  captured  on  a particular  day;  a separate  capture  history  was  created  for  each 
bird  species  for  which  there  were  >50  captures  at  a single  location  during  a specific  season. 


No.  observations  per  estimated  parameter 


Capture  histories  that: 

2 < # < 5 
(n  = 42) 

5 < # < 10 
(«  = 81) 

<10  (n  = 65) 

Had  inestimable  parameters 

24  (57%) 

29  (36%) 

16  (25%) 

Were  inestimable  in  simulation  GOF 

15  (36%) 

30  (37%) 

6 (9%) 

Failed  simulation  GOF 

0 (0%) 

0 (0%) 

2 (3%) 

Failed  U-CARE  “transients"  test 

0 (0%) 

1 (1%) 

4 (6%) 

Had  an  applicable  model 

3 (7%) 

21  (26%) 

37  (57%) 

Had  a CV  <50%  in  total  stopover  duration3 

1 (2%) 

7 (9%) 

15  (23%) 

Had  a CV  >50%  in  total  stopover  duration3 

2 (5%) 

14  (17%) 

22  (34%) 

Had  a CV  <50%  in  stopover-afterb 

1 (2%) 

9 (11%) 

18  (28%) 

Had  a CV  >50%  in  stopover-afterb 

2 (5%) 

12  (15%) 

19  (29%) 

3 Total  stopover  estimates  are  based  on  open  population  models  and  estimates  from  stopover  duration  analysis  (SODA)  described  in  Schaub  et  al.  (2001); 
CV  (coefficient  of  variation)  = (SE/mean)  X 100. 

b Stopover-after  estimates  are  based  on  open  population  models  and  estimates  using  equation  5 from  Efford  (2005). 


only  CV  values  <50%  were  considered  useful 
because  comparing  different  stopover  esti- 
mates is  impossible  when  CV  values  are  sub- 
stantially >50%.  CV  values  could,  of  course, 
be  determined  for  any  estimated  parameters  in 
the  model;  we  focus  here  on  the  derived  sta- 
tistic (stopover  duration)  relevant  to  the  study 
of  migration  ecology. 

Discriminant  function  analyses. — We  used 
discriminant  function  analyses  to  examine 
which  conditions  led  to  estimability  of  param- 
eters in  the  original  capture  history  and  during 
GOF  testing.  We  used  a range  of  simple  sta- 
tistics that  could  be  calculated  without  em- 
ploying the  complex  CMR  models.  The  vari- 
ables included  in  these  analyses  were  the 
number  of  individuals  captured,  number  of 
days  sampled,  percent  of  individuals  recap- 
tured at  least  once,  total  number  of  captures 
and  recaptures,  total  number  of  recaptures, 
number  of  captures  per  day,  median  captures 
per  day,  recaptures  per  day,  number  of  days 
with  no  captures  or  recaptures,  minimum  stop- 
over estimate,  standard  deviation  of  the  min- 
imum stopover  estimate,  standard  deviation  in 
the  number  of  captures  per  day,  and  several 
measures  of  capture  consistency,  which  we 
term  “completeness.”  Completeness  is  the 


percentage  of  days  on  which  there  was  >1 
capture  event,  while  “completeness  two”  re- 
fers to  the  percentage  of  days  with  >2  capture 
events.  “Recapture  completeness”  and  “re- 
capture completeness  two”  refer  to  the  per- 
centage of  days  with  >1  or  >2  recaptures, 
respectively.  Backwards  stepwise  discriminant 
analyses  were  performed  in  SYSTAT  10.2 
(SYSTAT  Software,  Inc.  2002). 

RESULTS 

We  examined  the  parameter  estimability  of 
1-  to  7-day  MDC  models  applied  to  188  cap- 
ture histories  representing  34  different  species 
(97  capture  histories  from  fall  and  91  from 
spring  migration).  Of  these,  we  were  able  to 
obtain  estimable  parameters  of  a completely 
time-dependent  MDC  model  for  119  capture 
histories.  The  MDC  interval  at  which  models 
could  be  estimated  varied  among  capture  his- 
tories. The  shortest  interval  that  could  be  used 
ranged  from  3 to  7 days  (3 -day  n = 15,  4-day 
n = 22,  5-day  n = 40,  6-day  n — 21,  7-day 
n = 21).  Parameter  estimability  was  strongly 
dependent  on  the  number  of  observations  per 
parameter  (Table  1).  Estimability  also  played 
a large  role  in  the  outcome  of  GOF  testing. 
Relatively  few  capture  histories  failed  GOF 


520 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 . No.  4 . December  2006 


testing  in  an  absolute  sense.  Five  capture  his- 
tories showed  evidence  of  differences  in  cap- 
ture probabilities  of  previously  recaptured  in- 
dividuals relative  to  new  captures  (the  tran- 
sience test)  in  U-CARE.  and  two  had  excess 
deviance  in  the  parametric  bootstrap  test  (sim- 
ulation GOF).  The  remaining  capture  histories 
that  “failed"  GOF  did  so  because  of  param- 
eter inestimability  in  the  bootstrap  procedure. 
In  these  instances,  the  models  could  not  be  fit 
reliably  to  the  simulated  data  (i.e..  there  were 
problems  with  estimability  in  >10%  of  the 
simulated  capture  histories).  The  ability  of 
models  to  satisfy  the  GOF  criteria  was  sub- 
stantially greater  for  capture  histories  in  our 
highest  category  (>10  observations  per  pa- 
rameter) than  in  those  in  the  other  two  cate- 
gories (2  < # < 5 and  5 < # < 10  observa- 
tions per  parameter;  Table  1).  Data  sparseness 
also  affected  the  contingency  tests  imple- 
mented in  U-CARE;  42%  ( n — 119)  of  the 
capture  histories  with  estimable  parameters 
produced  useful  contingency  tables,  although 
the  percentage  varied  among  our  three  cate- 
gories (2  < # < 5:  0%,  n = 18;  5 < # ^ 10; 
38%,  n = 52;  >10:  61%,  n = 49). 

A discriminant  function  analysis  of  all  cap- 
ture histories  with  >10  observations  per  pa- 
rameter produced  a moderately  effective,  sta- 
tistically significant  discriminant  function  de- 
scribing parameter  estimability  (Wilks'  A.  = 
0.53,  F559  = 10.41.  P < 0.001)  with  positive 
loadings  on  duration,  recapture  completeness, 
and  median  captures  per  day.  There  were  neg- 
ative loadings  on  number  of  recaptured  birds 
and  minimum  stopover.  To  extract  biological 
information  from  discriminant  function  load- 
ings. we  examined  a range  of  bivariate  plots 
depicting  the  various  loadings.  The  plots 
yielded  only  one  clear  biological  interpreta- 
tion: capture  histories  with  high  minimum 
stopover  duration  often  had  inestimable  pa- 
rameters (Fig.  2).  Parameter  estimability  dur- 
ing GOF  testing  limited  the  number  of  capture 
histories  that  could  be  analyzed;  however,  a 
discriminant  function  analysis  to  predict  pa- 
rameter estimability  during  GOF  testing  of  the 
49  capture  histories  that  were  estimable  and 
had  >10  observations  per  parameter  was  not 
significant  (Wilks'  A = 0.83,  F4  44  = 2.20,  P 
= 0.085). 

Optimal  models  for  the  capture  histories  that 
passed  GOF  testing  varied  in  the  incorporation 


of  time-dependent  parameters  and  in  the  MDC 
interval  used  in  the  models.  When  the  AICc 
was  used  to  compare  the  estimable  candidate 
models,  regardless  of  the  number  of  observa- 
tions per  parameter.  88  viable  models  were 
identified  for  the  61  capture  histories.  The  total 
number  of  models  exceeded  the  number  of 
capture  histories,  as  multiple  models  were  con- 
sidered for  some  capture  histories.  For  46  of 
the  61  capture  histories,  a single  model  had  an 
overwhelming  AICt.  weight  (>0.95).  indicating 
that  a unique  model  was  identified.  Two  alter- 
native models  were  identified  for  seven  capture 
histories,  three  alternative  models  were  identi- 
fied for  six  capture  histories,  and  four  and  six 
models  were  identified  for  one  capture  history 
each.  Parameters  that  were  time-dependent  also 
varied  among  the  chosen  models.  All  three  pa- 
rameters were  time-dependent  in  14  capture 
histories,  two  parameters  were  time-dependent 
in  38  capture  histories  (p  and  6:  15;  p and  y: 
17;  d>  and  y:  6),  and  a single  parameter  was 
time-dependent  in  36  capture  histories  (p : 0;  <b: 
13;  y:  23).  The  MDC  time  interval  chosen  for 
all  61  capture  histories  varied  from  3 to  7 days 
(3-day  n = 5;  4-day  n = 2;  5-day  n = 18;  6- 
day  n — 19;  7-day  n = 44).  Although  52%  of 
our  original  capture  histories  were  collected 
during  the  fall.  75%  of  the  capture  histories 
with  applicable  models  were  collected  during 
the  fall. 

Estimated  total  stopover  duration  values 
ranged  from  0.76  to  17.08  days  (Table  2).  and 
the  CV  values  were  highly  variable  (ranging 
from  13%  to  274%).  Of  the  61  capture  his- 
tories that  were  useable  after  GOF  testing.  23 
had  a total  stopover  CV  of  <50%  (Table  1). 
Stopover-after  estimates  ranged  from  0.38  to 
10.13  days,  which  were  shorter  than  the  esti- 
mates of  total  stopover.  Despite  the  difference 
in  stopover  duration  estimates  obtained  by  es- 
timating total  stopover  and  stopover-after, 
stopover-after  had  a slightly  wider  range  of 
CV  values  than  total  stopover.  CV  values  for 
stopover-after  ranged  from  13%  to  365%. 
Most  of  the  estimates  involving  CV  values  of 
<50%  were  capture  histories  from  the  fall  mi- 
gration season  (18  of  the  23  estimates  for  total 
stopover  and  24  of  28  estimates  for  stopover- 
after),  approximately  mirroring  the  distribu- 
tion of  spring  and  fall  capture  histories  (75% 
of  estimable  capture  histories  were  collected 
during  the  fall).  These  useful  estimates  were 


Morris  et  al.  • UTILITY  OF  OPEN  POPULATION  MODELS 


521 


10 


<D  ^ 
> C/3 

s >» 

Q.  03 
O -o 
c n 'w' 

E 

13 

E 

c 


2 - 


X 


x*x 


X 

X* 


• 

m VV*-  • 


200 


400 


x Inestimable 
• Estimable 


x 


600  800 


Sample  size 


FIG.  2.  The  relationship  between  parameter  estim- 
ability,  minimum  stopover  duration  (days),  and  sample 
size.  Among  capture  histories  of  landbird  species  at 
migratory  stopover  sites  that  had  10  or  more  (by  spe- 
cies) capture  events  per  estimated  parameter,  those 
with  high  minimum  stopover  duration  often  had  ines- 
timable parameters. 


obtained  for  a variety  of  species  including  two 
vireos,  Red-breasted  Nuthatch  ( Sitta  canaden- 
sis),  two  kinglets,  two  thrushes.  Gray  Catbird 
(Dumetella  carolinensis ),  many  warbler  spe- 
cies, and  White-throated  Sparrow  {Zonotri- 
chia  leucophrys ; Table  2). 

DISCUSSION 

Our  study  provided  some  insights  about  the 
conditions  under  which  CMR  models  can  be 
effectively  used  to  estimate  migratory-stop- 
over duration.  Dividing  the  data  into  three  cat- 
egories based  on  the  number  of  observations 
per  parameter  revealed  the  importance  of  the 
observation: parameter  ratio  in  predicting  the 
utility  of  CMR  models.  Models  with  >10  ob- 
servations per  parameter  were  estimable  and 
—62%  satisfied  GOF  testing;  most  “failures” 
to  satisfy  GOF  were  due  to  the  difficulty  of 
estimating  parameters  during  the  GOF  proce- 
dure when  using  simulations.  If  our  banding 
data  are  representative,  then  the  presence  of 
>10  observations  per  parameter  (roughly 
three  birds  captured  or  recaptured  per  day) 
may  connote  a reasonable  probability  that 
CMR  models  will  be  useful  for  characterizing 
a given  capture  history. 

Although  we  present  analyses  based  on  to- 
tal number  of  observations  (summed  capture 
and  recapture  events)  per  parameter,  we  also 
conducted  similar  analyses  using  number  of 
individual  birds  banded  per  parameter,  yield- 


ing similar  results.  The  capture  histories  were 
also  divided  into  different  categories  based 
only  on  total  sample  size  (50  < n < 100,  100 
< « < 150,  and  n > 150).  The  division  by 
sample  size  alone  was  not  effective,  because 
sample  size  is  a product  of  both  sampling  du- 
ration and  sampling  intensity. 

Extrinsic  parameter  inestimability  proved  to 
be  the  largest  impediment  to  using  open  pop- 
ulation models  in  our  study,  affecting  both  the 
initial  model  fitting  and  GOF  testing.  The  dis- 
criminant function  analysis  revealed  that  a 
long  minimum  stopover  (>4  days)  was  a good 
indicator  that  the  parameters  would  not  be  es- 
timable. Because  most  birds  that  are  recap- 
tured at  stopover  sites  have  minimum  stop- 
overs of  only  a few  days,  long  minimum  stop- 
over statistics  likely  represent  multiple  birds 
with  unusually  long  stopovers.  Such  a scenar- 
io would  yield  a large  stopover  estimate  CV 
and  indicate  large  biological  differences 
among  migrants  at  a given  stopover  site.  Ex- 
amining the  16  capture  histories  with  >10  ob- 
servations per  parameter  but  with  inestimable 
parameters  revealed  that  3 histories  had  no  re- 
captures at  all  and  2 histories  had  only  2 re- 
captures. Ten  of  the  capture  histories  were 
from  three  Nearctic-Nearctic  migratory  spe- 
cies: five  White-throated  Sparrows  ( Zonotri - 
chia  albicollis ),  four  Yellow-rumped  Warblers 
( Dendroica  coronata ),  and  one  Ruby-crowned 
Kinglet  ( Regulus  calendula).  Three  of  the  oth- 
er capture  histories  represented  local  breeding 
species.  All  of  these  factors  led  us  to  believe 
that  the  inestimability  in  these  cases  might 
have  been  related  to  heterogeneous  migration 
behavior  (either  among  individuals  or  subpop- 
ulations). 

Unlike  what  we  found  for  parameter  estim- 
ability,  there  was  no  clear  single  factor  ex- 
plaining parameter  inestimability  in  GOF  test- 
ing. The  discriminant  function  had  low  pre- 
dictive power,  with  only  a 67%  chance  of  cor- 
rectly predicting  the  outcome  of  the  GOF  test, 
again  indicating  the  lack  of  strong  factors  in- 
fluencing estimability  in  GOF.  Biological  fac- 
tors related  to  heterogeneity  of  the  captured 
specimens  (Pollock  et  al.  1990,  Cooch  and 
White  2005)  can  easily  lead  to  failures  of 
GOF  testing.  Additionally,  there  may  be  sta- 
tistical reasons  for  some  of  the  observed  fail- 
ures in  GOF  testing.  The  GOF  test  is  based 
on  a Monte-Carlo  simulation  test  run  at  a 95% 


522 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


(/) 

3 •- 

T3  _ 

« -8 
a c 
z £ 


8 § 

'5  > 
£ > 

- >> 
■g  % 


3 X 

.2 

00 

% 

© 

X 

© 

in 

m 

X 

in 

o 

r-; 

00 

CM 

CM 

NO 

X 

Tt 

O' 

_ 

»n 

m 

in 

CM 

r^ 

Tf 

o 

o 

Z GO 

03 

00 

00 

CM 

00 

fCl 

Tt 

■'t 

o 

V- 

00 

r| 

NO 

m 

00 

r- 

o 

in 

fCJ 

cn 

X 

© 

in 

in 

cn 

x C 

_W) 

^ 0 

o 

© 

© 

NO 

1-H 

© 

© 

o 

d 

o 

cn 

o 

z 

© 

o 

© 

© 

o 

CM 

o 

CM 

© 

cd 

cd 

o 

x 

d 

o 

2 jo 

’5 

S | >, 

> .5  x 

+1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+! 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+l 

+1 

+1 

<D  > 

X X 

— 

cn 

CM 

m 

o 

X 

ON 

X 

CO 

CM 

CM 

m 

CM 

-f 

r- 

X 

X 

O' 

NO 

CM 

X 

m 

m 

r- 

»n 

X 

2 

<n 

<N 

90 

NO 

r- 

FH 

cn 

in 

X 

r" 

— 

CM 

NO 

-r 

CM 

CM 

ON 

CM 

ON 

X 

q 

ON 

cn 

r- 

m 

m 

■5^ 

X) 

— 

— 1 

CM 

cd 

cd 

CM* 

CM 

o 

o 

o 

o 

cd 

© 

© 

o 

CM* 

© 

NO 

— 

CM 

On 

r-' 

o 

CM 

o 

o 

C ^ 
3 2 


s.  ZS 

a>  g 

1 e 

2 s* 

3 > 
^ C 
T3  2 
<D  Z 

' 03 


C 

3 

GD 


& o 
W)g 

.2  <M 

3 On 
T3  os 
T5  On 


*-  3 -r  7? 


O 

3 Z* 


.2  o 

3 S 

a3  & 
> .2 
O o 
Q.  <u 
o a 
% * 
<u  .2 

03  £ 

I'S 

8i 

O 03 

2 S 

3 <u 
TJ  > 
GXj  4J 
.£  •£> 
"1  •£ 


4-  ° 

0 C/5 
4-  C 
CD  <u 

11 

1 « 

O 

2 U 


<u  — 

CD  3 

£ £ 

>>  r n 


2 £ 

I z 

3 4? 


3 


■c  o 

3 »0 

.2  A 

Z ^ 
.2  c 
o «f  •§ 


•a  8 
.2  « 
CO  00 

« .s 

1 6 

2 03 

S = 

<D  T3 

1 s 

2 © 

3 S 

T3  O 

t-  04 

<u  ' 

O ON 

a on 
o C 
00  <u 

<U  V- 
•£  -C 

C/5 

>>  o. 
-o  c 


"O  03 
D \S 
— C 

a,  3 

2*  tc 

c3  X) 

c/5  3 

*aj  00 
■a  -a 
o <u 

£ '5 

3 

C > 
O „ 


a £ 
^ « 

a > 
o o 
4-  a. 
° 2 
% ^ 
3 


3 .2 

l £ 

0)  > - 

*2  c/5 


'a  £ 

& « 
2 Z 


3 4-h 

— O 

0 C/5 

Q c 

S .a 

O 

GO  cc 

1 8 

>s  ° 

10  Z 

•3  W> 


~ 
e 

<D  3 

0C  _r 


N N 1) 

W c 
*8 


° •§ 
<D  33 
x 

3 x 
<U 

0)  ti 

« 1 
•O  ‘3 
U x 
.2  u 

ll 


§ c 

£•3 

>s  '— 1 

Z u 
T3  2 
« CO 
03  *r 

2 CM 

•a  o 

x O 

<U  Csl 

d sc 

O O' 
Z ON 

3 — 


<u  3 

2 ’g 

M e3 
^ 2 
S 2 

fc  o 

u "S 

1 a 

■g  < 

2 c 

x O 


X 

CM 

CM 

NO 

CM 

X 

X 

— 

ON 

CO 

ON 

X 

© 

NO 

NO 

X 

CM 

in 

CO 

CO 

r- 

m 

m 

_ 

in 

CO 

o 

CM 

X 

X 

nC 

o 

o 

© 

in 

o 

X 

o 

’t 

■ct 

O' 

ON 

■"t 

X 

CO 

q 

CM 

m; 

X 

X 

— 

r- 

NO 

CM 

©’ 

© 

© 

r-’ 

cd 

— 

— 

o 

CO 

z 

cd 

© 

z 

z 

© 

z 

NO 

z 

© 

’t* 

CM 

z 

cd 

Z 

o 

cd 

o 

o 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

cn 

CM 

r- 

t"- 

© 

r- 

CO 

NO 

CM 

© 

'-r 

co 

CM 

^t 

CM 

CM 

NO 

•X 

-t 

© 

X 

CM 

m 

■"Cf 

NO 

M; 

h; 

■ct 

co 

CO 

co 

Q) 

in 

«n 

CM 

co 

X 

© 

CM 

00 

CM 

o 

— • 

o 

oo 

NO 

CO 

CM 

© 

O) 

cd 

c-H 

cd 

mi 

00 

rr 

z 

z 

z 

z 

CM 

Z 

© 

CM 

z 

CM 

CM 

CM 

z 

cd 

cd 

in 

cd 

z 

'Tr 

z 

o 

CM  — 

— 


O o > o - vooor^cNoooNoo^tocONoo  — ONi^tnC'ONON-'nM- 1 

^t^t’^-^tCMCMCMCOTl-Tt  — CM  (N  cnCMCMCMCOCM  — (N  CO  ^ M 


rj-  — t^-0000  — Tfr  f" 
ITi  O O'  X IC|  r.  IT;  h 
— CM  — CM  (M  cn  CM 


o—i  oooNOmoNoooovototoot-NONOoo 
Ovcomr^ONOON  — — xcor^ootor'in 
— cm  m-  in  cm  — — — — — 


On 

NO 

NO 

t> 

X 

ON 

o 

— 

NO 

X 

ON 

o 

On 

o 

ON 

o 

o 

o 

On 

On 

o 

ON 

o 

o 

X 

CM 

O 

X 

O 

o 

o 

O 

ON 

a 

c 

o 

o 

O' 

o 

o 

o 

O 

o 

o 

O 

o 

O 

o 

c 

c 

ON 

o 

ON 

On 

ON 

On 

ON 

ON 

O' 

o 

o 

On 

ON 

ON 

o 

O' 

o 

o 

On 

o 

o 

o 

On 

ON 

o 

On 

o 

ON 

o 

o 

ON 

— 

— • 

•“l 

CM 

<N 

— ' 

— 

CM 

CM 

— 

(N 

(N 

(N 

<N 

(N 

CM 

(N 

CM 

CM 

CM 

CM 

GO  GO 


GO  GO  GO  GO 

c c c c 


cxo-Ig  os  13  23  13  23  o.o.o.o.c3,3232323 

C/J&OU.U.U-PUU.UH&OC^cn&OUUtt.U-PUtt. 


GO  GO  GO 

C 3 3 

232323  00.2323232323232323  o 
U-U-Lx.cncnu_ti.P-ii.Li.u_tJ-U-cn' 


I<<<<<<<<<<<W<<<IHOIIIIIIII<<W< 


o s 


. £ 5 

2 "c 


t-  o 

«u  <u 

X .3 
2 > 
5 T3 
Z'-S 

U,  3 


Go 

oc 

”60 


q S 3 
U » a 


U s 

W 00 


— O ' 
— 3 ~ 

2 Z <u 

H PQ  C2 


I ^ 


v c 

c2  a 


vC  ^ 

O cn 


n a 


■d  ^ 

^ O 

£ CQ 


c3  ^ 

2 <u 

o u 


Morris  et  al.  • UTILITY  OF  OPEN  POPULATION  MODELS 


523 


sd  d 

u.  03 

¥ E 


o.  c -a 

O'S  w 


in 

3 

© 

3 

CN 

r- 

m 

X 

co 

rr 

■St 

, 

X 

l"; 

X 

C0 

-St 

© 

CN 

fO 

3 

as 

ON 

ON 

3 

m 

CN 

X 

On 

o 

O 

ON 

co 

t-; 

r\ 

>n 

X 

X 

co 

in 

X 

CN 

-r 

'n 

O 

ON 

rr, 

CO 

rr, 

Tt 

>n 

NO 

co 

CN 

o 

in 

cn 

— i 

cn 

d 

© 

— 

©* 

d 

© 

© 

© 

CO 

©* 

© 

©’ 

© 

oi 

cN 

© 

© 

© 

o 

— 

o 

+ 1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

CN 

CN 

so 

r- 

't 

On 

ro 

r- 

© 

NO 

»n 

CO 

_ 

X 

Tt 

<N 

ON 

so 

't 

m 

SO 

CO 

CO 

(N 

m 

3 

NO 

O 

O 

CN 

x 

l"; 

On 

co 

r- 

't 

o 

r" 

On 

O 

c- 

Tf 

ON 

in 

©_ 

ro 

ON 

in 

co 

rr, 

© 

o; 

"t 

1-H 

1-H 

co 

d 

© 

— 

** 

— 

CN 

in 

-- 

CO 

-- 

o 

ri 

Tt 

’t 

CO 

r- 

X 

X 

NO 

NO 

CN 

X 

3 

On 

•t 

ON 

co 

in 

ro 

On 

co 

co 

3 

so 

CO 

X 

© 

in 

ON 

co 

m 

3 

<N 

in 

NO 

X 

On 

in 

X 

in 

in 

>n 

"t 

X 

co 

NO 

NO 

CN 

Tt 

CO 

Tf 

o 

© 

h 

X 

CN 

© 

«n 

co 

CN 

CN 

© 

CN 

— 

co 

CN 

CN 

r* 

co 

© 

iri 

CO 

** 

co 

>n 

© 

co 

©* 

■st 

CN 

o 

+ 1 

+1 

+1 

+1 

+ 1 

+ 1 

+ 1 

+1 

+ 1 

+1 

+ 1 

+ 1 

+1 

+1 

+1 

+1 

+ 1 

+1 

+1 

+ 1 

+1 

+1 

+1 

+ 1 

+1 

+ 1 

+1 

+1 

+1 

+1 

X 

CN 

co 

NO 

NO 

X 

3 

© 

ON 

CO 

CN 

CO 

r- 

On 

On 

so 

O' 

SO 

»n 

r- 

"t 

O 

so 

ON 

© 

r- 

-r 

>n 

m 

CO 

in 

X 

CN 

X 

m 

t"; 

O 

Tt 

X 

l"; 

>n 

NO 

in 

in 

in 

NO 

CN 

CN 

Tf 

X 

X 

CN 

NC 

— 

© 

On 

CN 

co 

NO 

t-H 

CN 

— 

co 

fO 

CN 

co 

't 

co 

co 

■st 

CO 

co 

NO 

mi 

t"’ 

X* 

co 

co 

CN 

co 

©* 

O 

-0-  -0-  -0-  -©-  -©-  -0-  ■©■  -0-  -0-  -0-  -0-  -©-  *0-  -0-  -0-  ■©■  -0-  -0-  ■©■  -0-  -©-  -0-  ■©■  -©-  -©-  -0-  ■©■  ■©■  -0-  ■©- 

^t-~  ^ r-.  ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ sc  ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ 


o © in  't  c — 'Oninmincnooin- 

m^(N(Nmc<)TtTtTt^Tt(NMCNimN 


r-  — in  oo  vC  tr,  on  xo 

m^CioinhirihtN 

(N  in  M n - — — — 


00  — CN  © O 

t"-  't  't  © — 

— — CN 


C©cOt^CNOOOt^©© 
^oinoMOOinO'tmts 
CN  — — — — -(NT,  — 


X 

© 

H 

o 

o 

© 

X 

X 

© 

r- 

© 

c 

CN 

r- 

X 

o 

r- 

© 

CN 

o o 

CN 

O 

© 

ON 

Os 

© 

o 

o 

© 

© 

© 

© 

© 

© 

o 

o 

C 

© 

© 

o 

© 

© 

© 

o 

o 

o o 

o 

© 

© 

© 

© 

o 

o 

o 

o 

© 

© 

© 

© 

© 

© 

o 

o 

c 

© 

© 

o 

© 

© 

© 

o 

3 

o 

3 

© 

1 

— 

— 

<N 

(N 

CN 

CN 

*— 1 

— 

— 

— 

CN 

(N 

CN 

— 

CN 

— 

N— 1 

CN 

CN 

CN 

CN  CN 

CN 

CN 

O-D-D-C-Cj  Q.  O.  c3  cs 

OOCncncntUXcntL-U, 


c3tdcascdBcsOiCi.ci«aca!3t3 

LUtUtULUtUtJUtL-cncnoOCOLUtJUtU 


a a a a s a 
n.  (ju  cn  oo  pl,  on 


<<<<IIW<<<<<<<<<<<<<<<<<WII<U 


II 

s -5 

^ s 
.0  •* 

I I 

ju  S 

n ^ 

•£  w 

> c 

<u  S 

« c/3 

•j=  ^ 

> 3J 

£ OC 


s 

<u 

o 

5 

-Cl 

3 


13.  K 

3 


to 


«*  o 


QQ  < 


.2  <u 
Ss 

to  •> 


X)  « 

c £ 
<U  t. 

> c 
O Z 


£ 

o 

— 

— 

53 

Ou 

CO 


70  X 

£ £ 


a A = Appledore  Island,  Maine;  H = Hamlin  Beach  State  Park,  New  York;  S = Star  Island,  New  Hampshire. 

b Mean  ± SE  of  the  total  stopover  estimate  (following  Schaub  et  al.  2001).  Estimates  in  boldface  had  a CV  of  <50%.  CV  values  were  calculated  as  (SE/mean)  X 100. 
c Mean  ± SE  of  the  stopover-after  estimate  (using  equation  5 in  Efford  2005).  Estimates  in  boldface  had  a CV  of  <50%.  CV  values  were  calculated  as  (SE/mean)  X 100. 


524 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


confidence  level.  It  is  worth  noting  that  this 
simulation  test  has  a Type  I error  rate  of  5% 
(i.e.,  5%  chance  of  passing  the  GOF  test  when 
the  model  does  not  have  adequate  descriptive 
power);  however,  the  expected  Type  II  error 
rate  (the  chance  that  the  model  has  failed  GOF 
when,  in  fact,  it  has  adequate  descriptive  pow- 
er) is  not  known,  so  we  cannot  even  say  with 
certainty  that  the  rate  of  GOF  failure  is  greater 
than  expected  by  chance.  The  contingency  ta- 
ble GOF  tests  implemented  in  U-CARE  also 
were  severely  limited  by  the  sparseness  of  the 
data  (only  42%  of  estimable  capture  histories 
could  be  tested  using  U-CARE). 

For  all  capture  histories  used  in  this  study, 
it  was  necessary  to  reduce  the  number  of  pa- 
rameters in  the  fitted  model  from  the  number 
present  in  a fully  time-dependent  model  to  es- 
timate all  parameters  successfully.  Our  results 
indicated  that  MDC  intervals  from  3 to  7 days 
were  necessary  to  reduce  the  parameter  count 
in  the  models  sufficiently  to  estimate  all  pa- 
rameters. Parameter  reduction  was  necessary 
even  for  relatively  large  sample  sizes  (up  to 
595  specimens  captured  over  38  days).  The 
only  current  alternative  to  the  MDC  method 
of  reducing  the  number  of  parameters  is  pool- 
ing the  data — with  its  attendant  problems  of 
possible  parameter  bias  (Hargrove  and  Bor- 
land 1994,  Morris  et  al.  2005b).  If  pooling  is 
desirable  in  a given  study,  the  MDC  interval 
approach  outlined  here  could  be  adapted  to 
determine  the  minimum  pooling  interval  nec- 
essary, based  on  parameter  estimability.  Re- 
gardless of  the  method,  successful  use  of 
CMR  models  on  banding  data  will  often  re- 
quire some  form  of  parameter  reduction. 

In  our  current  work,  the  CV  of  total  stop- 
over duration  measures  the  relative  uncertain- 
ty in  the  derived  parameter  of  interest.  The 
CV  includes  both  biological  variability  and 
variability  due  to  parameter  estimation  uncer- 
tainty. Given  our  current  available  data,  it  is 
somewhat  difficult  to  determine  the  extent  of 
the  biological  contribution  versus  the  sam- 
pling-related contribution.  Again,  long  mini- 
mum stopover  duration  might  indicate  hetero- 
geneity in  the  population.  However,  corre- 
sponding increases  in  (1)  the  fraction  of  cap- 
ture histories  with  a CV  of  <50%  and  (2)  the 
number  of  observations  per  parameter  (Table 
1)  indicate  some  variation  due  to  sample  size. 
Overall,  more  estimates  of  stopover  duration 


had  a CV  of  <50%  when  using  the  stopover- 
after  statistic  (28  capture  histories)  than  when 
using  the  total  stopover  statistic  (23  capture 
histories).  Thus,  in  addition  to  the  theoretical 
points  raised  by  Efford  (2005),  the  statistic 
based  on  his  equation  5 resulted  in  more  use- 
able  estimates  of  stopover  duration  based  on 
banding  data. 

Most  of  the  capture  histories  that  were  es- 
timable and  had  applicable  models  in  this 
study  were  collected  during  fall  migration  (Ta- 
ble 2).  Previous  work  on  Appledore  Island  re- 
sulted in  higher  rates  of  recapture  and  docu- 
mented longer  minimum  stopover  durations 
during  fall  migration  than  in  spring  migration 
(Morris  et  al.  1994,  Morris  and  Glasgow 
2001);  this  may  have  helped  increase  the  num- 
ber of  observations  per  parameter  available  in 
our  study,  which,  in  turn,  may  have  resulted 
in  higher  estimability.  We  did  not  see  a spe- 
cific pattern  related  to  avian  biology  that  ex- 
plained the  pattern  of  capture  histories  with 
low  CV  values.  Although  most  of  the  capture 
histories  with  low  CV  values  were  obtained 
during  fall  banding,  this  proportion  was  sim- 
ilar to  the  proportion  of  fall  capture  histories 
that  had  applicable  models.  The  capture  his- 
tories with  low  CV  values  represented  a wide 
range  of  species  (Table  2).  Species  that  had 
low  CV  values  over  multiple  seasons  included 
those  captured  in  high  numbers,  such  as  Red- 
eyed Vireo  ( Vireo  olivaceus ),  American  Red- 
start ( Setophaga  ruticilla),  and  Northern  Wa- 
terthrush  ( Seiurus  noveboracensis). 

Our  results  document  the  difficulty  associ- 
ated with  parameter  estimability  when  using 
passerine  banding  data  for  capture-mark-re- 
capture models  of  stopover  duration.  We  are 
not  implying  that  these  methods  cannot  or 
should  not  be  used  on  this  type  of  data,  but 
rather  they  should  be  used  cautiously,  partic- 
ularly when  sample  sizes  are  small.  Efford 
(2005)  suggests  using  a constant  c}>  model  for 
populations  with  no  consistent  trend  in  cf), 
which  would  reduce  problems  with  estimabil- 
ity. Researchers  planning  to  use  these  methods 
in  migration  banding  studies  should  attempt 
to  maximize  the  number  of  captures  and  re- 
captures during  sampling  periods  to  increase 
the  likelihood  of  parameter  estimability. 

ACKNOWLEDGMENTS 

This  research  was  funded,  in  part,  by  Canisius  Col- 
lege faculty  research  funding  to  SRM  and  HDS,  and 


Morris  et  al.  • UTILITY  OF  OPEN  POPULATION  MODELS 


525 


HHMI  Research  Assistantships  to  AML,  DAL,  and 
MSM.  We  are  very  grateful  to  the  many  people  who 
assisted  at  the  Appledore  Island  Migration  Banding 
Station.  R.  W.  Suomala  (Star  Island)  and  D.  Bonter 
(Braddock  Bay  Bird  Observatory)  generously  provid- 
ed their  banding  data  for  use  in  our  analyses.  We  also 
gratefully  acknowledge  the  assistance  of  R.  J.  Morris 
and  two  anonymous  referees,  who  provided  valuable 
comments  on  this  paper.  This  paper  is  contribution  12 
of  the  Appledore  Island  Migration  Banding  Station  and 
contribution  127  of  the  Shoals  Marine  Laboratory. 

LITERATURE  CITED 

Bonter,  D.  2003.  Migration  and  ecology  of  landbirds 
during  migration  in  the  Great  Lakes  Basin.  Ph.D. 
dissertation,  University  of  Vermont,  Burlington. 
Burnham,  K.  P.,  D.  R.  Anderson,  G.  C.  White,  C. 
Brownie,  and  K.  H.  Pollock.  1987.  Design  and 
analysis  methods  for  fish  survival  experiments 
based  on  release-capture.  American  Fisheries  So- 
ciety, Monograph  5. 

Burnham,  K.  P.  and  D.  R.  Anderson.  1998.  Model 
selection  and  inference:  a practical  information- 
theoretic  approach.  Springer,  New  York. 

Cherry,  J.  D.  1982.  Fat  deposition  and  length  of  stop- 
over of  migrant  White-crowned  Sparrows.  Auk 
99:725-732. 

Choquet,  R.,  A.  M.  Reboulet,  J.-D.  Lebreton,  O.  Gi- 
menez,  AND  R.  Pradel.  2005.  U-CARE  2.2  user’s 
manual.  Centre  d’Ecologie  Fonctionelle  et  Evo- 
lutive, Montpellier,  France,  www.cefe.cnrs.fr/ 
biom/PDF/Choquet-USER%20MANUAL%20U- 
CARE%202.2.pdf  (accessed  10  February  2006). 
Connolly,  S.  R.  and  A.  I.  Miller.  2001.  Global  Or- 
dovician faunal  transitions  in  the  marine  benthos: 
proximate  causes.  Paleobiology  27:779-795. 
Cooch,  E.  G.,  R.  Pradel,  and  N.  Nur.  1997.  A prac- 
tical guide  to  mark-recapture  analysis  using 
SURGE  2nd  ed.  Centre  d’Ecologie  Fonctionelle 
et  Evolutive  - CNRS,  Montpellier,  France. 

Cooch,  E.  and  G.  White.  2005.  Program  MARK:  a gen- 
tle introduction  4th  ed.  www.phidot.org/software/ 
mark/docs/book/mark_book.pdf  (accessed  6 June 
2005). 

Dunn,  E.  H.  2001.  Mass  change  during  migration  stop- 
over: a comparison  of  species  groups  and  sites. 
Journal  of  Field  Ornithology  72:419-432. 

Efford,  M.  G.  2005.  Migrating  birds  stop  over  longer 
than  usually  thought:  comment.  Ecology  86:3415- 
3418. 

Finch,  D.  M.  and  W.  Yong.  2000.  Landbird  migration 
in  riparian  habitats  of  the  Middle  Rio  Grande:  a 
case  study.  Studies  in  Avian  Biology  20:88-98. 
Hargrove,  J.  W.  and  C.  H.  Borland.  1994.  Pooled 
population  parameter  estimates  from  mark-recap- 
ture data.  Biometrics  50:1129-1 141. 

Heglund,  P.  J.  and  S.  K.  Skagen.  2005.  Ecology  and 
physiology  of  en  route  Nearctic-Neotropical  mi- 
gratory birds:  a call  for  collaboration.  Condor  107: 
193-196. 


Holmgren,  N.,  H.  Ellegren,  and  J.  Pettersson. 
1993.  Stopover  length,  body  mass,  and  fuel  de- 
position rate  in  autumn  migrating  adult  Dunlins 
Calidris  alpine : evaluating  the  effects  of  molting 
status  and  age.  Ardea  81:9-20. 

Jones,  J.,  C.  M.  Francis,  M.  Drew,  S.  Fuller,  and 
M.  N.  S.  Ng.  2002.  Age-related  differences  in 
body  mass  and  rates  of  mass  gain  of  passerines 
during  autumn  migratory  stopover.  Condor  104: 
49-53. 

Kaiser,  A.  1995.  Estimating  turnover,  movements  and 
capture  parameters  of  resting  passerines  in  stan- 
dardized capture-recapture  studies.  Journal  of  Ap- 
plied Statistics  22:1039-1047. 

Kaiser,  A.  1999.  Stopover  strategies  in  birds:  a review 
of  methods  for  estimating  stopover  length.  Bird 
Study  46  (suppl.):S299-308. 

Kuenzi,  A.  J.,  F.  R.  Moore,  and  T.  R.  Simons.  1991. 
Stopover  of  Neotropical  landbird  migrants  on  East 
Ship  Island  following  trans-Gulf  migration.  Con- 
dor 93:869-883. 

Lavee,  D.,  U.  Safriel,  and  I.  Meilijson.  1991.  For 
how  long  do  trans-Saharan  migrants  stop  over  at 
an  oasis?  Ornis  Scandinavica  22:33-44. 

Lebreton,  J.-D.,  K.  P.  Burnham,  J.  Clobert,  and  D. 
R.  Anderson.  1992.  Modeling  survival  and  test- 
ing biological  hypotheses  using  marked  animals: 
a unified  approach  with  case  studies.  Ecological 
Monographs  62:67-118. 

McCullagh,  P.  and  J.  A.  Nelder.  1989.  Generalized 
linear  models.  Chapman  and  Hall,  London,  United 
Kingdom. 

Mehlman,  D.  W.,  S.  E.  Mabey,  D.  N.  Ewert,  C.  Dun- 
can, B.  Abel,  D.  Cimprich,  R.  D.  Sutter,  and  M. 
Woodrey.  2005.  Conserving  stopover  sites  for 
forest-dwelling  migratory  landbirds.  Auk  122: 
1281-1290. 

Moore,  F.  R.  (Ed.).  2000.  Stopover  ecology  of  Nearc- 
tic-Neotropical migrants:  habitat  relations  and 
conservation  implications.  Studies  in  Avian  Biol- 
ogy, no.  20. 

Moore,  F.  R.  and  D.  A.  Aborn.  2000.  Mechanisms  of 
en  route  habitat  selection:  how  do  migrants  make 
habitat  decision  during  stopover?  Studies  in  Avian 
Biology  20:34-42. 

Moore,  F.  R.  and  P.  Kerlinger.  1987.  Stopover  and 
fat  deposition  by  North  American  wood-warblers 
(Parulinae)  following  spring  migration  over  the 
Gulf  of  Mexico.  Oecologia  74:47-54. 

Moore,  F.  R.,  P.  Kerlinger,  and  T.  R.  Simons.  1990. 
Stopover  on  a Gulf  coast  barrier  island  by  spring 
trans-Gulf  migrants.  Wilson  Bulletin  102:487- 
500. 

Morris,  S.  R.  and  J.  Glasgow.  2001.  Comparison  of 
spring  and  fall  migration  of  American  Redstarts 
on  Appledore  Island,  Maine.  Wilson  Bulletin  1 13: 
202-210. 

Morris,  S.  R.,  D.  W.  Holmes,  and  M.  E.  Richmond. 
1996.  A ten-year  study  of  the  stopover  patterns  of 
migratory  passerines  during  fall  migration  on  Ap- 
pledore Island,  Maine.  Condor  98:395-409. 


526 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Morris,  S.  R.,  D.  A.  Liebner,  A.  M.  Larracuente,  E. 
M.  Escamilla,  and  H.  D.  Sheets.  2005a.  Multi- 
ple-day constancy  as  an  alternative  to  pooling  for 
estimating  mark-recapture  stopover  length  in  Ne- 
arctic-Neotropical  migrant  landbirds.  Auk  122: 
319-328. 

Morris,  S.  R.,  M.  E.  Richmond,  and  D.  W.  Holmes. 
1994.  Patterns  of  stopover  by  warblers  during 
spring  and  fall  migration  on  Appledore  Island, 
Maine.  Wilson  Bulletin  105:703-718. 

Morris,  S.  R.,  E.  M.  Turner,  D.  A.  Liebner,  A.  M. 
Larracuente,  and  H.  D.  Sheets.  2005b.  Prob- 
lems associated  with  pooling  mark-recapture  data 
prior  to  estimating  stopover  length  for  migratory 
passerines.  Pages  673-679  in  Bird  conservation 
implementation  and  integration  in  the  Americas: 
proceedings  of  the  third  international  Partners  in 
Flight  conference  2002  (C.  J.  Ralph  and  T.  D. 
Rich,  Eds.).  General  Technical  Report  PSW-GTR- 
191,  USDA  Forest  Service,  Pacific  Southwest  Re- 
search Station,  Albany,  California. 

Parrish,  J.  D.  2000.  Behavioral,  energetic,  and  conser- 
vation implications  of  foraging  plasticity  during  mi- 
gration. Studies  in  Avian  Biology  20:53-70. 

Partners  in  Flight  Research  Working  Group.  2002. 
Priority  research  needs  for  the  conservation  of 
Neotropical  migrant  landbirds.  Journal  of  Field 
Ornithology  73:329-339. 

Petit,  D.  R.  2000.  Habitat  use  by  landbirds  along  Ne- 
arctic-Neotropical  migration  routes:  implications 
for  conservation  of  stopover  habitats.  Studies  in 
Avian  Biology  20:15-33. 

Pollock,  K.  H.,  J.  D.  Nichols,  C.  Brownie,  and  J.  E. 
Hines.  1990.  Statistical  inference  for  capture-re- 
capture  experiments.  Wildlife  Monographs  107:1- 
97. 

Pradel,  R.  1996.  Utilization  of  capture-mark-recapture 
for  the  study  of  recruitment  and  population  growth 
rate.  Biometrics  52:703-709. 

Pradel,  R.  and  J.-D.  Lebreton.  1993.  User’s  manual 
for  program  SURGE,  ver.  4.2.  Centre  D’Ecologie 
Fonctionnelle  et  Evolutive,  C.N.R.S.,  Montpellier- 
Cedex,  France. 

Schaub,  M.  and  L.  Jenni.  2001.  Stopover  durations  of 


three  warbler  species  along  their  autumn  migra- 
tion route.  Oecologia  128:217-227. 

Schaub,  M.,  R.  Pradel,  L.  Jenni,  and  J.-D.  Lebreton. 
2001.  Migrating  birds  stopover  longer  than  usu- 
ally thought:  an  improved  capture  recapture  anal- 
ysis. Ecology  82:852-859. 

Schwarz,  C.  J.  and  A.  N.  Arnason.  1996.  A general 
method  for  the  analysis  of  capture-recapture  ex- 
periments in  open  populations.  Biometrics  52: 
860-873. 

Schwilch,  R.  and  L.  Jenni.  2001.  Low  initial  refueling 
rate  at  stopover  sites:  a methodological  effect? 
Auk  118:698-708. 

Sillett,  T.  S.  AND  R.  T.  Holmes.  2002.  Variation  in 
survivorship  of  a migratory  songbird  throughout 
its  annual  cycle.  Journal  of  Animal  Ecology  71: 
296-308. 

SYSTAT  Software  Inc.  2002.  SYSTAT,  ver.  10.2. 
SYSTAT  Software,  Inc.,  Point  Richmond,  Cali- 
fornia. 

The  Math  Works,  Inc.  1992.  MATLAB.  The  Math 
Works,  Inc.,  Natick,  Massachusetts. 

V IALLEFONT,  A.,  J.-D.  LEBRETON,  A.-M.  REBOULET, 
and  G.  Gory.  1998.  Parameter  identifiability  and 
model  selection  in  capture-recapture  models:  a nu- 
merical approach.  Biometrical  Journal  40:313- 
325. 

White,  G.  C.  2002.  Discussion  comments  on:  the  use 
of  auxiliary  variables  in  capture-recapture  mod- 
eling. An  overview.  Journal  of  Applied  Statistics 
29:103-106. 

White,  G.  C.  and  K.  P.  Burnham.  1999.  Program 
MARK:  survival  estimation  from  populations  of 
marked  animals.  Bird  Study  46  (suppl.):  120-138. 

Winker,  K.,  D.  W.  Warner,  and  A.  R.  Weisbrod. 
1992.  Daily  mass  gains  among  woodland  migrants 
at  an  inland  stopover  site.  Auk  109:853-862. 

Xu,  C.,  M.  J.  Melchin,  H.  D.  Sheets,  C.  E.  Mitchell, 
and  F.  Jun-Xuan.  2005.  Patterns  and  processes  of 
latest  Ordovician  graptolite  extinction  and  recov- 
ery based  on  the  data  from  South  China.  Journal 
of  Paleontology.  79:842-861. 

Yong,  W.  and  F.  R.  Moore.  1997.  Spring  stopover  of 
intercontinental  migratory  thrushes  along  the 
northern  coast  of  the  Gulf  of  Mexico.  Auk  114: 
263-278. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):527— 53 1 , 2006 


MAXIMUM  DIVING  DEPTH  IN  FLEDGING  BLUE-FOOTED 
BOOBIES:  SKILL  DEVELOPMENT  AND  TRANSITION 
TO  INDEPENDENCE 

JOSE  ALFREDO  CASTILLO-GUERRERO1  AND  ERIC  MELLINK1 2 


ABSTRACT. — We  evaluated  maximum  diving  depth  and  time  spent  at  the  nest  of  fledging  Blue-footed  Boo- 
bies ( Sula  nebouxii)  at  Isla  El  Rancho,  Sinaloa,  in  the  Gulf  of  California,  Mexico.  Within  three  consecutive  10- 
day  post-fledging  intervals,  maximum  diving  depth  was  highly  variable,  but  was  not  affected  by  sex,  weight,  or 
body  condition.  During  the  first  days  of  post-fledging  flight,  maximum  diving  depth  increased  rapidly.  By  the 
second  week  after  first  flight,  the  plunge-dives  of  juveniles  were  almost  as  deep  as  those  of  adults.  Parental  care 
and  attachment  to  the  nest  lasted  several  additional  weeks  (up  to  40  days  after  first  flight).  Although  their  diving 
capacity  rapidly  reached  a level  similar  to  that  of  the  adults,  it  appeared  that  juvenile  boobies  took  much  longer 
in  acquiring  other  foraging  skills.  Received  1 August  2005,  accepted  5 July  2006. 


The  speed  with  which  juvenile  birds  ac- 
quire foraging  abilities  has  important  impli- 
cations for  the  evolution  of  life  histories 
(Wheelwright  and  Templeton  2003).  It  has 
been  hypothesized  that  parental  care  continues 
until  young  birds  acquire  mobility  and  forag- 
ing skills  adequate  for  survival.  Additional  pa- 
rental care  improves  the  survival  of  the  off- 
spring, but  decreases  long-term  survival  of  the 
parents  (Burger  1980). 

Juvenile  birds  face  major  challenges  in 
learning  how  to  identify  foraging  areas  and 
developing  foraging  techniques  as  the  period 
of  parental  care  ends  (Burger  1980,  Wheel- 
wright and  Templeton  2003).  The  study  of 
newly  volant  birds  can  help  elucidate  the  pro- 
cess of  such  learning.  However,  this  is  com- 
plicated in  the  wild,  as  fledglings  can  move 
freely  through  the  colony  site.  Most  of  the  few 
studies  on  the  subject  have  focused  on  pas- 
serines, which  have  a rather  short  transition  to 
independence  (Moreno  1984,  Wheelwright 
and  Templeton  2003).  In  seabirds,  the  devel- 
opment of  foraging  skills  and  its  relationship 
to  parental  care  are  not  well  known  (Yoda  et 
al.  2004).  We  are  aware  of  only  one  such  sea- 
bird study  (Brown  Booby,  Sula  leucogaster), 
although  the  birds  were  raised  by  humans 
(Yoda  et  al.  2004),  which  could  have  inter- 
fered with  social  learning  processes.  Even  less 
is  known  about  possible  intersexual  differenc- 


1  Centro  de  Investigation  Cientffica  y de  Educacion 
Superior  de  Ensenada,  A.R  2732,  Ensenada,  Baja  Cal- 
ifornia, Mexico. 

2 Corresponding  author;  e-mail: 
emellink@cicese.mx 


es  in  the  acquisition  of  foraging  skills  (Wheel- 
wright et  al.  2003). 

The  Blue-footed  Booby  ( S . nebouxii ) is  a 
sexually  dimorphic  seabird:  females  are  larger 
than  males  at  fledging  (Drummond  et  al. 
1991).  Parental  care  continues  for  a 6-week, 
post-fledging  period  (Nelson  1978).  During 
this  period,  young  birds  fly  out  to  sea  but  re- 
turn to  their  nests,  where  they  continue  re- 
ceiving food  from  the  parents.  In  this  study, 
we  determined  the  maximum  diving  depths 
(MDD)  of  wild  fledgling  Blue-footed  Boobies 
to  (1)  examine  the  ontogeny  of  MDD  and 
compare  it  with  the  diving  depths  achieved  by 
adults,  and  (2)  examine  the  relationship  be- 
tween the  development  of  diving  skills  and 
sexually  related  size  dimorphism. 

METHODS 

Field  work  was  conducted  at  Isla  El  Rancho 
(25°  10'  N,  108°  23' W),  a sandy,  120-ha  is- 
land in  the  south-central  Gulf  of  California, 
Mexico,  at  the  mouth  of  Bahia  de  Santa  Ma- 
ria-La  Reforma — a large  coastal  lagoon.  The 
colony  studied  was  located  on  the  northeast- 
ern part  of  the  island  among  4-m-high  sand 
dunes.  About  500  pairs  of  Blue-footed  Boo- 
bies nested  in  an  area  of  <1  ha,  with  a max- 
imum density  of  0.6  nests/m2. 

Between  January  and  May  2004,  we  visited 
the  island  12  times  for  periods  of  5 days  and 
monitored  100  nests  and  108  chicks  that  we 
had  marked  with  unique  combinations  of  col- 
or bands.  During  each  visit,  we  checked  the 
nests  daily,  and  weighed  and  measured  (cul- 
men,  ulna,  and  tarsus)  all  banded  chicks  every 
other  day.  Sex  was  determined  from  the  length 


527 


528 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


of  ulna  at  fledging  (males  = 191-207  mm; 
females  = 213-233  mm;  Drummond  et  al. 
1991).  Fledging  (age  at  first  flight)  was  in- 
ferred when  a bird  with  complete  juvenile 
plumage  left  its  nest  site  and  returned  with 
clean  feet  (feet  were  covered  with  excrement 
before  the  first  trip  to  sea).  For  most  birds,  we 
could  estimate  the  exact  age  at  fledging  (es- 
timates were  ± 2 days  in  some  cases). 

From  20  April  to  26  May,  we  estimated 
MDD  by  attaching  a capillary  tube  (Tygon,  8 
mm  internal  diameter;  Burger  and  Wilson 
1988)  to  the  lower  side  of  a booby’s  central 
rectrix.  Tubes  were  recovered  one  day  after 
application.  A total  of  99  capillary  tubes  pro- 
duced usable  data:  67  from  fledglings  (48  in- 
dividuals, 15  of  which  provided  data  for  more 
than  one  date),  17  from  adult  males,  and  15 
from  adult  females.  In  addition,  we  estimated 
the  amount  of  time  that  young  spent  at  their 
nests  by  monitoring  38  nests  hourly  during 
14-hr  diurnal  periods. 

We  tested  the  data  for  normality  and  homo- 
cedasticity  with  Kolmogorov-Smirnov  and 
Levene’s  tests,  respectively,  for  every  group 
to  be  compared.  We  used  parametric  proce- 
dures when  both  requirements  were  met.  We 
grouped  the  MDD  data  for  post-fledging  ju- 
veniles into  5-day  age  intervals.  We  then  con- 
ducted a Mann-Whitney  U-test  to  compare  the 
MDD  attained  by  male  and  female  fledglings 
for  each  5-day  period. 

We  used  a mixed-model  ANOVA-ANCO- 
VA  for  comparing  5-day  periods  (normality: 
D = 0.17-0.39;  homocedasticity:  F439  = 0.98, 
P = 0.42)  to  evaluate  the  possible  effects  of 
ontogeny  on  MDD.  The  number  of  days  since 
first  fledging  was  included  as  a covariate,  with 
the  5-day  periods  as  the  fixed  factor.  Multiple 
flights  of  the  same  bird  in  10-day  intervals  (1- 
10,  1 1-20,  and  >21  days  after  fledging)  were 
compared  with  /-tests  for  dependent  samples. 

We  found  no  significant  differences  be- 
tween adult  male  and  female  MDD  (3.4  ± 2.1 
m and  3.6  ± 1.6  m,  nx  = 15  and  n2  = 17, 
respectively;  / = 0.33,  P = 0.73;  normality: 
D = 0.22  and  0.23,  respectively;  homocedas- 
ticity: F131  = 1.7,  P = 0.26).  Therefore,  we 
pooled  the  MDD  of  both  sexes  to  compare 
adult  MDD  with  that  of  juveniles  that  had 
fledged  at  least  15  days  previously.  We  tested 
for  age-related  differences  in  MDD  using  a 
/-test  (normality:  D = 0.19  and  0.13,  respec- 


tively; homocedasticity:  F149  = 3.96,  P = 
0.55). 

We  used  linear  regressions  to  assess  wheth- 
er MDD  might  be  a function  of  weight  or 
body  condition.  Residuals  from  the  regression 
of  weight  on  culmen  length  were  used  as  a 
body  condition  index.  Using  residuals  of  a re- 
gression between  weight  and  body  measure- 
ments as  an  index  of  condition  is  adequate 
when  measurement  errors  and  variations  in 
body  size  are  low  (Schulte-Hostedde  et  al. 
2005);  the  major  assumption  to  be  met  is  that 
the  relationship  between  variables  is  linear, 
which  was  the  case  in  our  study  (r2  = 0.73, 
P < 0.001).  To  explore  the  relationship  be- 
tween days  since  first  flight  and  time  spent  at 
the  nest,  we  used  a mixed  ANOVA-ANCOVA 
model,  with  gender  serving  as  the  fixed  factor 
and  days  since  fledging  included  as  a covari- 
ate. All  statistical  tests  were  considered  sig- 
nificant at  a = 0.05,  and  reported  values  are 
means  ± SD. 

RESULTS 

Female  Blue-footed  Booby  chicks  reached 
their  maximum  pre-fledging  weight  (2,071  ± 
125.2  g)  between  60  and  75  days  of  age,  while 
males  reached  it  (1,628  ± 117.5  g)  between 
60  and  70  days  of  age.  Females  were  signifi- 
cantly heavier  than  males  (/4954  = 18.43,  P < 
0.001).  After  reaching  their  maximum  weight, 
female  chicks  lost  8.5%  of  their  weight  and 
weighed  1,830  g ± 72.2  at  first  flight,  whereas 
males  lost  7%  and  weighed  1,470  g ± 63.5. 
Males  began  to  fly  earlier  than  females  (83.4 
± 2.64  and  87.9  ± 3.8  days  of  age,  nx  = 23, 
n2  = 19,  respectively;  U = 67,  P < 0.001). 
MDD  within  any  given  period  was  highly  var- 
iable (Fig.  1),  and  there  were  no  statistical  dif- 
ferences between  male  and  female  fledglings 
(1-5  days  after  first  flight:  nx  = 9,  n2  = 6,  U 
= 19,  P = 0.34;  6-10  days:  nx  = 6,  n2  = 9, 
U = 16,  P = 0.38;  11-15  days:  nx  = l,n2  = 
9,  U = 25.5,  P = 0.52;  16-20  days:  nx  = 4, 
n2  = 6,  U = 11,  P = 0.83). 

We  did  not  detect  an  effect  of  date  on 
MDD,  per  se  (F1>38  = 3.31,  P = 0.10),  but 
despite  great  within-interval  variability,  MDD 
increased  with  time  since  first  flight  through- 
out the  first  15  days  of  flight  (0-5  days  = 1.68 
± 0.66  m,  6-10  days  = 2.69  ± 0.81  m,  11- 
15  days  = 3.02  ± 0.53  m;  F439  = 3.64,  P = 
0.012;  Fig.  1).  By  16-20  days  (3.11  ± 0.76 


Castillo-Guerrero  and  Mellink  • DIVING  DEPTHS  OF  FLEDGING  BOOBIES 


529 


4.5 

35 

4.0 


7 


1.5 


1.0  * * * * — 

1-5  6-10  11-15  16-20  21-40  Adults 

Days  since  fledging 


FIG.  1.  Maximum  diving  depth  of  Blue-footed  Boobies  increased  rapidly  during  the  first  15  days  after  their 
first  flight  at  Isla  El  Rancho,  Sinaloa,  Mexico,  2004.  Fledglings  then  dived  almost  as  deep  as  adults.  Means  ± 
SE  (white  zone)  and  95%  confidence  intervals  (whiskers)  are  shown.  Sample  size  is  indicated  above  whiskers. 


m)  and  21-40  days  (3.18  ± 0.55  m;  Fig.  1) 
since  flight,  MDD  stabilized.  The  15  juveniles 
for  which  we  had  > 1 MDD  value  (there  were 
2 values  for  9 birds  and  >2  for  6 birds)  ex- 
hibited a similar  tendency:  during  the  first  10 
days  after  fledging,  dives  were  shallower  than 
they  were  during  the  11-20  day  interval  (1- 
10  days  = 2.12  ± 0.70  m and  11-20  days  = 
3.03  ± 0.90  m,  tn  = -2.44,  P = 0.032).  Birds 
for  which  we  had  >2  records  made  shallower 
dives  during  the  first  10  days  after  fledging 
than  they  did  >21  days  post-fledging  (1-10 
days  = 1.94  ± 0.35  m and  >21  days  = 3.42 
± 0.69  m,  t4  = 5.05,  P = 0.007);  there  were 

no  significant  differences  between  the  two  lat- 
er periods  (11-20  days  = 2.64  ± 0.71  m and 
>21  days  = 2.90  ± 0.96  m,  t7  = -0.51,  P = 
0.62). 

MDD  of  juveniles  that  had  flown  for  at 
least  15  days  did  not  differ  from  that  of  adult 
birds  (2.99  ± 0.75  m and  3.51  ± 1.88  m,  re- 
spectively, t5A  = -1.27,  P = 0.26).  Weight 
was  not  correlated  with  diving  depth  within 
sex  (males:  P = 0.71;  females:  P = 0.90).  The 
regression  between  MDD  and  body  condition 
also  was  not  significant  ( P = 0.23). 

Juvenile  birds  progressively  reduced  their 


time  at  the  nest  after  their  first  flight  (r2  = 
0.33,  P < 0.001),  with  no  differences  between 
males  and  females  (F129  = 0.11,  P = 0.73). 
After  25  days  of  flight,  some  individuals  left 
the  nest  for  at  least  the  entire  daylight  period. 
Other  young  birds  remained  at  their  nests  for 
>40  days  (Fig.  2). 

DISCUSSION 

Blue-footed  Booby  parents  reduce  their 
provisioning  to  offspring  just  before  the  nest- 
lings take  their  first  flights  ( sensu  Nelson 
1978;  JAC-G  unpubl.  data).  This  reduction 
may  stimulate  fledging  and  encourage  the 
fledglings  to  develop  foraging  skills  away 
from  their  nest.  Juveniles  make  their  first 
plunge  dives  on  their  first  day  of  flight  (every 
recovered  capillary  tube  showed  evidence  of 
immersion,  including  four  that  were  attached 
to  birds  just  prior  to  their  first  flight). 

Clearly,  15  days  of  learning  were  enough 
for  juveniles  to  dive  almost  as  deep  as  adults. 
Based  on  our  observations,  the  fledglings 
made  their  first  plunges  at  low  angles  and 
from  low  heights.  As  the  days  passed,  the 
birds  increased  the  plunge  height  and  dives 
became  more  vertical.  During  the  first  days 


530 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


70  1 


~ 60 

</) 

CD 

~ 50 

2 ^ 
c Q 
<D  - 40  - 

CL  £ 

C/5  Cl 

0 ro  30 

c E 20 

0 

o 


o Males  • Females 


0 


10 

0 


oo 


5 10  15  20  25  30  35  40 

Days  since  fledging 


FIG.  2.  The  percent  of  diurnal  time  fledgling  Blue-footed  Boobies  spent  at  the  nest  decreased  with  time 
since  their  first  flight  (r2  = 0.33,  P < 0.001)  at  Isla  El  Rancho,  Sinaloa,  Mexico,  2004.  This  relationship  is 
described  by  the  equation  tn  = 0.5147  - 0.107  X days  since  fledging,  where  tn  = percentage  of  diurnal  time  at 
nest. 


after  initiation  of  flight,  fledglings  also  tended 
to  fly  in  groups  around  the  island,  suggesting 
that  social  interactions  might  facilitate  their 
development  of  diving  and,  perhaps,  foraging 
skills. 

For  several  weeks  after  their  first  flight, 
fledglings  continued  begging  for  food  from 
their  parents.  Juveniles  of  other  species  usu- 
ally cease  begging  when  foraging  for  them- 
selves becomes  more  profitable  (Moreno 
1984,  Heinsohn  1991,  Wheelwright  and  Tem- 
pleton 2003);  thus,  the  young  birds  in  our 
study  apparently  required  several  additional 
weeks  to  become  adequate  foragers.  Similar 
to  other  sulids  (Burger  1980,  Yoda  et  al. 
2004),  the  Blue-footed  Boobies  at  El  Rancho 
exhibited  gradual  separation  from  their  par- 
ents. Based  on  our  observations,  we  hypoth- 
esize that  there  are  two  periods  in  the  devel- 
opment of  foraging  skills:  (1)  an  initial  rapid 
improvement  in  the  depth  attained  during 
plunge-dives,  followed  by  (2)  improvement  in 
other  behaviors,  such  as  locating  and  captur- 
ing prey.  Presumably,  once  birds  begin  catch- 
ing fish,  begging  frequency  and  presence  at 
the  nest  decrease.  Some  juveniles  apparently 
achieved  this  level  of  independence  at  25  days 


after  fledging,  while  others  required  >40  days 
to  do  so. 

It  is  unlikely  that  temporal  changes  in  the 
depth  at  which  prey  were  found  affected  our 
recorded  MDD  in  fledglings.  Our  data  did  not 
exhibit  any  effects  of  date,  and  fledglings  did 
not  exhibit  much  synchrony  in  dates  of  first 
flight  that  could  confound  our  data.  Some 
fledglings  were  already  independent  by  the 
time  others  began  to  fly  and,  in  some  cases, 
>2  months  had  passed  between  early-  and 
late-fledging  birds.  We  did  not  find  evidence 
of  temporal  patterns  in  adult  MDDs. 

Despite  the  Blue-footed  Booby’s  distinct 
sexual  dimorphism  in  size  and  gender-influ- 
enced differences  in  growth  and  date  of  first 
flight  (Torres  and  Drummond  1999;  this 
study),  we  found  no  gender  differences  in 
MDD.  Given  the  limitations  of  capillary 
tubes,  however,  further  study  of  the  relation- 
ship between  sexual  dimorphism  in  size  and 
booby  diving  performance  is  warranted.  It 
seems  that  fledging  Blue-footed  Boobies  de- 
velop plunge-diving  skills  and  attain  MDDs 
similar  to  those  of  adults  relatively  quickly. 
However,  this  does  not  imply  that  juvenile 
feeding  success  and/or  foraging  performance 


Castillo-Guerrero  and  Mellink  • DIVING  DEPTHS  OF  FLEDGING  BOOBIES 


531 


is  equivalent  to  that  of  adults.  Their  nest  at- 
tendance and  insistent  begging  for  long  peri- 
ods indicate  that  foraging  for  themselves, 
along  with  developing  prey-finding  and  prey- 
capturing skills,  delays  the  full  independence 
of  young  Blue-footed  Boobies. 

ACKNOWLEDGMENTS 

We  are  grateful  to  CONACYT  and  SEMARNAT  for 
funding.  A.  Aguilar,  and  the  M.  A.  Gonzalez-Bernal 
family  provided  logistical  support.  M.  Prado  and  E.  A. 
Penaloza  assisted  during  field  work.  J.  Awkerman  pro- 
vided editorial  advice.  S.  Herzka,  E.  A.  Schreiber,  D. 
J.  Anderson,  and  an  anonymous  reviewer  greatly  im- 
proved this  manuscript;  we  are  grateful  to  all  of  them. 

LITERATURE  CITED 

Burger,  J.  1980.  The  transition  to  independence  and 
postfledging  parental  care  in  seabirds.  Pages  367 - 
447  in  Behavior  of  marine  animals,  vol.  4 (J.  Bur- 
ger, B.  L.  Olla,  and  H.  E.  Winn,  Eds.).  Plenum, 
New  York. 

Burger,  A.  E.  and  R.  P.  Wilson.  1988.  Capillary-tube 
depth  gauges  for  diving  animals:  an  assessment  of 
their  accuracy  and  applicability.  Journal  of  Field 
Ornithology  59:345-354. 

Drummond,  H.,  J.  L.  Osorno,  R.  Torres,  C.  Garcia- 
Cha velas,  and  L.  H.  Merchant.  1991.  Sexual 
size  dimorphism  and  sibling  competition:  impli- 


cations for  avian  sex  ratios.  American  Naturalist 
138:623-641. 

Heinsohn,  R.  G.  1991.  Slow  learning  of  foraging  skills 
and  extended  parental  care  in  cooperatively  breed- 
ing White-winged  Choughs.  American  Naturalist 
137:864-881. 

Moreno,  J.  1984.  Parental  care  of  fledged  young,  di- 
vision of  labor,  and  development  of  foraging  tech- 
niques in  the  Northern  Wheatear  ( Oenanthe  oen- 
anthe ).  Auk  101:741-752. 

Nelson,  J.  B.  1978.  The  Sulidae:  gannets  and  boobies. 
Oxford  University,  Oxford,  United  Kingdom. 

Schulte-Hostedde,  A.,  B.  Zinner,  J.  S.  Millar,  and 
G.  J.  Hickling.  2005.  Restitution  of  mass-size  re- 
siduals: validating  body  condition  indices.  Ecol- 
ogy 86: 155-163. 

Torres,  R.  and  H.  Drummond.  1999.  Does  large  size 
make  daughters  of  the  Blue-footed  Booby  more 
expensive  than  sons?  Journal  of  Animal  Ecology 
68:1133-1141. 

Wheelwright,  N.  T.  and  J.  Templeton.  2003.  Devel- 
opment of  foraging  skills  and  the  transition  to  in- 
dependence in  juvenile  Savannah  Sparrows.  Con- 
dor 105:279-287. 

Wheelwright,  N.  T.,  K.  A.  Tice,  and  C.  R.  Freeman- 
Gallant.  2003.  Postfledging  parental  care  in  Sa- 
vannah Sparrows:  sex,  size  and  survival.  Animal 
Behaviour  65:435-443. 

Yoda,  K.,  H.  Cono,  and  Y.  Naito.  2004.  Development 
of  flight  performance  in  the  Brown  Booby.  Pro- 
ceedings of  the  Royal  Society  of  London,  Series 
B (suppl.)  27LS240-S242. 


The  Wilson  Journal  of  Ornithology  1 18(4):532— 536,  2006 


VEGETATIVE  AND  THERMAL  ASPECTS  OF  ROOST-SITE 
SELECTION  IN  URBAN  YELLOW-BILLED  MAGPIES 

SCOTT  P.  CROSBIE,1’34  DOUGLAS  A.  BELL,1  AND  GINGER  M.  BOLEN2 3 4 


ABSTRACT. — We  examined  vegetative  and  thermal  aspects  of  roost-site  selection  in  urban  Yellow-billed 
Magpies  ( Pica  nuttalli ) in  Sacramento,  California,  from  winter  2003  to  spring  2004.  Vegetation  used  for  roosting 
included  cultivated  species  such  as  glossy  privet  {Ligustrum  lucidum),  English  ivy  ( Hedera  helix),  and  white 
mulberry  ( Morns  alba),  and  native  species  such  as  interior  live  oak  ( Quercus  wislizeni),  valley  oak  ( Q . lobata), 
and  California  laurel  ( Umbellularia  calif omica).  Percent  canopy  cover  was  consistently  high  (mean  = 94%  ± 
1.9  SD).  Mean  roost  height  was  9.7  m ± 3.5  SD  and  the  mean  height  at  which  magpies  roosted  was  6.6  m ± 
2.0  SD.  Communal  roosts  were  generally  located  within  or  near  riparian  corridors.  Magpies  roosted  in  relatively 
warm  microhabitats,  but  they  did  not  appear  to  obtain  a thermal  advantage  by  roosting  communally.  The  timing 
of  roost  occupancy  was  restricted  primarily  to  times  when  the  roost  was  thermally  advantageous.  Received  22 
August  2005,  accepted  2 May  2006. 


The  Yellow-billed  Magpie  ( Pica  nuttalli)  is 
found  chiefly  in  the  Central  Valley  and  lower 
foothills  of  California  and  is  relatively  abun- 
dant in  the  residential  areas  of  Sacramento 
(Reynolds  1995).  The  roosting  behavior  of 
this  species  is  not  well  documented,  especially 
regarding  urban  populations.  However,  rural 
magpies  studied  at  and  near  Hastings  Natural 
History  Reserve  (HNHR)  in  Monterey  Coun- 
ty, California,  roost  almost  exclusively  in  live 
oaks  ( Quercus  spp.;  Verbeek  1973),  where 
roost  size  may  exceed  several  hundred  birds 
(Birkhead  1991). 

The  evolution  of  communal  roosting  has 
been  attributed  to  several  factors,  including  a 
decrease  in  predation  risk  (Pulliam  1973),  an 
increase  in  foraging  efficiency  (Marzluff  et  al. 
1996),  and  a reduction  in  thermoregulation  de- 
mands (Francis  1976).  The  thermoregulatory 
requirements  of  magpies  are  greatest  during 
the  winter  months  (Mugaas  and  King  1981), 
indicating  that  roost-site  selection  is  important 
to  energy  conservation  in  winter.  By  roosting 
in  dense  vegetation  or  cavities,  birds  can  re- 
duce heat  loss  and  gain  protection  from  wind 
and  rain  (Walsberg  1986).  Roosting  over  wa- 
ter or  moist  soil  also  may  moderate  extreme 


1 Dept,  of  Biological  Sciences,  California  State 
Univ.,  Sacramento,  CA  95819,  USA. 

2 North  State  Resources,  Inc.,  5000  Bechelli  Lane, 
Ste.  203,  Redding,  CA  96002,  USA. 

3 Current  address:  Univ.  of  California,  Davis,  Wild- 
life and  Ecology  Unit,  Veterinary  Genetics  Laboratory, 
One  Shields  Ave.,  Davis,  CA  95616,  USA. 

4 Corresponding  author;  e-mail: 
urbanmagpie@yahoo.com 


temperatures  (M0ller  1985).  Timing  of  roost 
occupancy  is  also  critical  to  energy  conser- 
vation: Black-billed  Magpies  {Pica  hudsonia) 
are  known  to  spend  relatively  more  time  at  the 
roost  when  faced  with  cold  temperatures 
(Reebs  1986). 

Our  goal  was  to  document  vegetative  and 
thermal  aspects  of  roost-site  selection  in  Yel- 
low-billed Magpies  inhabiting  urban  sites.  We 
hypothesized  that  urban  Yellow-billed  Mag- 
pies (1)  roost  in  a greater  number  of  plant  spe- 
cies than  magpies  in  rural  settings;  (2)  select 
roost  sites  that  provide  thermal  advantages 
such  as  high  percent  canopy  cover,  and  prox- 
imity to  water  and  other  places  where  tem- 
peratures may  be  moderated  by  nearby  sub- 
strates; (3)  may,  when  roosting  in  large 
groups,  increase  the  temperature  of  the  roost 
via  collective  production  of  body  heat;  and  (4) 
occupy  the  roost  only  when  its  temperature  is 
higher  than  that  of  the  surrounding  habitat. 

METHODS 

We  located  eight  Yellow-billed  Magpie 
communal  roosts  in  the  urban  (residential)  ar- 
eas of  Sacramento,  California  (Fig.  1,  Table 
1),  by  following  magpies  from  their  foraging 
grounds  to  their  roost  sites  and  by  querying 
the  local  ornithological  community.  Data  col- 
lection took  place  from  December  2003 
through  May  2004.  We  visited  each  roost  once 
per  week  during  morning  roost  departures  or 
evening  roost  arrivals  to  ascertain  roost  oc- 
cupancy and  determine  where  the  birds  slept. 
During  each  observation,  we  recorded  the 
number  of  birds  arriving  at,  or  departing  from, 


532 


Crosbie  et  al.  • URBAN  YELLOW-BILLED  MAGPIE  ROOSTS 


533 


FIG.  1.  Study  area  and  locations  of  Yellow-billed  Magpie  roosts  in  Sacramento,  California,  winter  2003 
through  spring  2004. 


the  roost  per  min  for  the  entire  period  of  roost 
entry  or  exodus.  On  average,  morning  obser- 
vation periods  lasted  75  min  and  evening  ob- 
servation periods  lasted  95  min.  Occasionally 
we  made  nighttime  visits  with  flashlights  to 
confirm  where  birds  roosted  (the  birds  were 
slightly  wary,  but  very  tolerant,  of  this  activ- 
ity). 

We  used  a densiometer  to  determine  the 


TABLE  1.  Latitude  and  longitude  coordinates  of 
urban  Yellow-billed  Magpie  communal  roost  sites  in 
Sacramento,  California,  2003-2004. 


Roost  no.  Latitude  Longitude 


1 

N 

38° 

40.021' 

W 

121° 

18.221' 

2 

N 

38° 

40.081' 

W 

121° 

18.241' 

3 

N 

38° 

40.05 1 ' 

w 

121° 

18.262' 

4 

N 

38° 

40.043' 

w 

121° 

18.207' 

5 

N 

38° 

37.324' 

w 

121° 

22.760' 

6 

N 

38° 

37.278' 

w 

121° 

22.666' 

7 

N 

38° 

37.278' 

w 

121° 

22.640' 

8 

N 

38° 

37.328' 

w 

121° 

22.696' 

mean  percent  canopy  cover  of  roosts  and  a 
clinometer  to  determine  mean  height  of  all 
trees/shrubs  comprising  each  roost  (roost 
height)  and  mean  height  of  each  group  of 
magpies  perched  in  their  roosts.  All  canopy 
cover  measurements  were  made  in  the  last  2 
weeks  of  May  to  ensure  that  our  estimates 
were  comparable  across  all  roosts.  We  mea- 
sured the  distance  from  each  roost  center  to 
the  closest  water  body  (always  a creek)  by 
using  a Garmin  eTrex  Legend  Global  Posi- 
tioning System  (Olathe,  Kansas). 

From  7 December  2003  through  13  Febru- 
ary 2004,  we  recorded  roost  temperatures  with 
Hobo  data-logging  thermometers  (Onset 
Computer  Corporation,  Bourne,  Massachu- 
setts). We  collected  paired  samples  at  20 
points  within  two  known  roosts  and  at  20 
points  within  eight  potential  roosts  (unoccu- 
pied vegetation)  that  were  located  within  a 
200-m  radius  of  a known  roost.  Potential 
roosts  were  selected  according  to  their  simi- 
larity to  known  roosts  in  terms  of  tree  or  shrub 


534 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  2.  Characteristics  of  urban  communal  roosts  of  Yellow-billed  Magpies  in  Sacramento,  California, 
2004. 


Roost  no. 

Mean  canopy 
cover  (%) 

Mean  height  of 
vegetation  used  (m) 

Mean  height  of 
magpie  perches  (m) 

Distance  to 
water  (m) 

Estimated  maximum 
no.  magpies 

1 

92.2  (n  = 

6) 

11.2  {n  = 

6) 

7.5  (n  = 

6) 

0 

879 

2 

95.9  {n  = 

19) 

8.5  ( n = 

19) 

5.4  ( n = 

19) 

106 

133 

3 

93.5  ( n = 

1) 

7.9  ( n = 

1) 

6.5  ( n = 

1) 

62 

14 

4 

91.0  {n  = 

1) 

17.6  {n  = 

1) 

11.0  (n  = 

1) 

29 

8 

5 

94.5  {n  = 

18) 

9.2  (n  = 

18) 

6.2  (n 

18) 

56 

818 

6 

95.8  {n  = 

2) 

8.6  (n  = 

2) 

6.0  ( n = 

2) 

89 

27 

7 

91.7  {n  = 

1) 

8.9  ( n = 

1) 

5.9  (n  = 

1) 

56 

12 

8 

94.8  {n  = 

3) 

5.7  ( n = 

3) 

4.2  (n  = 

3) 

13 

7 

Mean 

93.7 

9.7 

6.6 

51 

237 

SD 

1.9 

3.5 

2.0 

36 

380 

species  height,  percent  canopy  cover,  and 
proximity  to  water.  Within  a given  roost,  we 
used  a random  number  generator  to  select  a 
compass  bearing,  distance,  and  height  for  lo- 
cating the  tree  or  shrub  in  which  we  would 
place  the  data-loggers.  For  each  of  the  paired 
temperatures,  data-loggers  were  placed  at  sim- 
ilar heights  within  the  range  of  heights  at 
which  magpies  roosted  in  the  area.  Data-log- 
gers were  taped  to  the  upper  end  of  a 2-m- 
long  stick,  at  the  top  of  which  we  attached  a 
bent  coat  hanger  that  allowed  us  to  hang  the 
data-loggers  on  lateral  branches  between  the 
tree  or  shrub  center  and  the  outer  perimeter  of 
the  canopy  (where  magpies  roosted).  All 
paired  recordings  took  place  at  05:00  PST.  Us- 
ing SPSS  (1998),  we  conducted  a one-tailed, 
paired-sample  r-test  ( a = 0.05;  see  Zar  1999) 
to  determine  whether  the  roost  microhabitat 
was  significantly  warmer  than  the  nearby  po- 
tential roost  microhabitat. 

To  determine  whether  any  temperature  dif- 
ference in  occupied  versus  unoccupied  roosts 
was  due  to  the  birds’  presence,  we  recorded 
temperatures  at  45-min  intervals  in  two  ran- 
domly selected  locations:  one  within  a known 
roost  site  (Roost  1)  and  another  in  an  area  of 
unoccupied  vegetation  within  200  m of  Roost 
1 . We  recorded  temperatures  at  these  two  sites 
on  two  occasions:  once  on  a night  when  the 
known  roost  was  occupied  (by  317  individu- 
als) and  again  a week  later  (the  data-loggers 
were  left  in  place)  when  the  known  roost  was 
temporarily  unoccupied  (temporary  roost 
abandonment  was  a normal  phenomenon  re- 
lated to  the  birds’  seasonal  movements).  On 
the  evening  Roost  1 was  occupied,  we  record- 


ed the  time  at  which  magpies  arrived  and  sub- 
sequently departed  the  following  morning  to 
determine  whether  the  timing  of  roost  occu- 
pancy is  limited  to  when  the  roost  is  warmer 
than  its  surroundings. 

RESULTS 

A total  of  18  plant  species  were  used  for 
roosting.  Species  native  to  California  included 
(in  approximate  relative  order  of  usage)  inte- 
rior live  oak  ( Quercus  wislizeni),  valley  oak 
( Q . lobata ),  California  laurel  ( Umbellularia 
californica ),  boxelder  {Acer  negundo),  bishop 
pine  ( Pinus  muricata ) and  MacNab’s  cypress 
( Cupressus  macnabiana).  Species  not  native 
to  California  included  glossy  privet  ( Ligus - 
trum  lucidum),  English  ivy  {Hedera  helix ) that 
had  overtaken  trees,  an  undetermined  species 
(no  floral  structures  were  present)  of  bamboo 
(Bambusoideae),  white  mulberry  {Morus 
alba),  Japanese  cheesewood  {Pittosporum  to- 
bira),  Chinese  photinia  {Photinia  serrulata), 
dense  logwood  {Xylosma  congestion),  olean- 
der {Nerium  oleander),  Chinese  elm  {Ulmus 
parvifolia),  cherry  laurel  ( Primus  laurocera- 
sus),  pomegranate  ( Punica  granatum),  and 
southern  magnolia  ( Magnolia  grandiflora). 
Deciduous  species  were  only  occupied  when 
leafed  out.  Canopy  cover  at  roosts,  comprising 
leaves  and  dense  networks  of  branches,  was 
consistently  high  (>90%;  Table  2).  The  height 
of  occupied  vegetation  varied;  however,  mag- 
pies always  roosted  in  the  upper  third  of  the 
vegetation.  All  roosts  were  located  near  a 
creek  and  Roost  1 was  situated  almost  entirely 
over  a creek. 

The  microhabitat  of  known  roosts  was  sig- 


Crosbie  et  al.  • URBAN  YELLOW-BILLED  MAGPIE  ROOSTS 


535 


FIG.  2.  Morning  (05:00  PST)  temperatures  record- 
ed in  urban  communal  roosts  of  Yellow-billed  Magpies 
and  in  nearby  potential  roost  sites  (unoccupied  vege- 
tation), December  2003  through  February  2004,  Sac- 
ramento, California. 


nificantly  warmer  than  that  of  nearby  potential 
roosts  (mean  difference  = 0.72°  C ± 0.72, 
range  = 0.40-0.88°  C,  P < 0.001;  Fig.  2).  The 
45-min  interval  sampling  showed  that,  just  af- 
ter the  birds’  median  arrival  time,  known  roost 
temperature  exceeded  potential  roost  temper- 
ature (Fig.  3A).  About  25  minutes  before  the 
magpies  left  on  the  following  morning,  tem- 
perature in  the  known  roost  dropped  below 
that  of  the  potential  roost.  The  same  temper- 
ature inversion  occurred  a week  later  when  the 
known  roost  was  temporarily  unoccupied 
(Fig.  3B),  but  the  mean  temperature  difference 
was  greater  when  the  roost  was  unoccupied 
(0.65°  C ± 0.23  when  occupied;  1.54°  C ± 
0.41  when  unoccupied). 

DISCUSSION 

In  contrast  to  rural  magpies  roosting  at  and 
near  HNHR,  urban  magpies  in  our  study 
roosted  in  a variety  of  plant  species.  This  dif- 
ference is  undoubtedly  due  to  the  greater  di- 
versity of  plant  species  in  the  urban  setting 
that  provides  the  characteristics  necessary  for 
suitable  roost  sites.  However,  both  rural  and 
urban  populations  of  the  Yellow-billed  Mag- 
pie appear  to  roost  only  in  dense  evergreen 
vegetation  during  winter;  in  contrast,  some 
Black-billed  Magpie  and  Common  Magpie 
( Pica  pica ) populations  roost  in  deciduous 
vegetation  for  part  or  all  of  the  winter  (Mpller 
1985,  Reebs  1987).  Avoiding  wind  exposure 
has  been  identified  as  one  of  the  most  impor- 
tant factors  in  roost-site  selection  (Walsberg 
1986),  and  magpies  can  reduce  their  metabolic 


FIG.  3.  Temperatures  recorded  at  45-min  intervals 
in  (A)  an  urban  communal  roost  occupied  by  317  Yel- 
low-billed Magpies  and  in  nearby  unoccupied  vegeta- 
tion (an  interior  live  oak)  during  the  night  of  14-15 
December  2003,  Sacramento,  California,  and  (B)  in  a 
temporarily  unoccupied  urban  communal  roost  of  the 
Yellow-billed  Magpie  and  in  nearby  unoccupied  veg- 
etation (an  interior  live  oak)  during  the  night  of  25- 
26  December  2003,  Sacramento,  California. 

demand  substantially  by  roosting  in  dense 
vegetation  during  winter  (Mugaas  and  King 
1981).  Magpies  may  also  deter  predation 
events  by  roosting  in  dense  vegetation.  Coo- 
per’s ( Accipiter  cooperii ) and  Red-shouldered 
( Buteo  lineatus ) hawks  occasionally  prey 
upon  magpies  as  the  magpies  depart  from 
their  roosts  (Crosbie  2004). 

Whereas  magpies  at  and  near  HNHR  roost 
at  heights  >10-20  m (Reynolds  1995),  roost- 
ing height  in  this  study  was  lower;  this  was 
probably  due,  in  part,  to  the  fact  that  there  was 
no  taller  vegetation  that  provided  dense  cover. 
Similar  to  magpies  studied  in  Denmark 
(Mpller  1985)  and  Canada  (Reebs  1987), 
magpies  in  this  study  roosted  near  water,  like- 


536 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


ly  due  to  the  moderating  effect  that  water  or 
moist  soil  may  have  on  nighttime  tempera- 
tures (Mpller  1985). 

Wintering  blackbirds  studied  by  Francis 
(1976)  roosted  in  microhabitats  that  were  1.0 
to  1.5°  C warmer  than  their  surroundings, 
slightly  greater  than  the  range  of  difference 
observed  in  this  study.  However,  the  control 
site  used  by  Francis  (1976)  was  in  a clearing 
rather  than  in  vertical  vegetation,  as  was  the 
case  in  our  study.  The  temperature  difference 
between  magpie  roosts  and  nearby  potential 
roost  sites  was  greatest  on  the  coldest  nights 
(Fig.  2),  indicating  that  roosts  are  especially 
favorable  during  cold  spells.  Similarly,  mag- 
pies in  Teruel,  Spain,  prefer  thermally  advan- 
tageous roosts  when  temperatures  are  low 
(Miranda  and  Gonzalez  2000).  In  our  study, 
the  timing  of  roost  arrival  and  departure  con- 
formed almost  precisely  to  the  times  at  which 
the  temperature  of  the  roost  became  warmer 
or  cooler,  respectively,  than  the  surrounding 
habitat  (Fig.  3A).  However,  these  birds  may 
not  gain  any  thermoregulatory  benefit  by 
roosting  together:  the  temperature  difference 
between  the  roost  and  nearby  unoccupied  veg- 
etation was  not  greater  when  the  roost  was 
occupied  than  when  unoccupied  (Fig.  3). 

In  conclusion,  urban-dwelling  Yellow- 
billed Magpies  roosted  in  a variety  of  plant 
species.  Roost-site  selection  was  biased  to- 
ward habitat  structure  that  provided  thermal 
advantages,  such  as  a high  percent  of  over- 
head cover,  proximity  to  water,  and  warm  tem- 
peratures relative  to  surrounding  habitat.  It 
does  not  appear  that  magpies  gain  any  thermal 
benefit  via  collective  body  heat  production, 
but  the  timing  of  roost  occupancy  in  winter  is 
limited  primarily  to  times  when  the  roost  is 
thermally  advantageous.  The  habit  of  roosting 
near  water  may  be  detrimental  due  to  the  re- 
cent arrival  of  West  Nile  virus.  Further  study 
on  roost-site  selection,  mosquito  presence,  and 
management  options,  where  necessary,  is  war- 
ranted. 

ACKNOWLEDGMENTS 

This  paper  is  based  on  part  of  a thesis  presented  to 
the  Department  of  Biological  Sciences  at  California 
State  University,  Sacramento,  by  Scott  R Crosbie,  in 


partial  fulfillment  of  the  requirements  for  the  degree 
of  Master  of  Science.  We  thank  W.  E.  Avery,  M.  F. 
Baad,  L.  R.  Crosbie,  M.  D.  Reynolds,  H.  B.  Ernest, 
and  three  anonymous  referees  for  their  support,  guid- 
ance, and  review  of  this  manuscript.  We  are  grateful 
to  all  property  owners  who  allowed  us  access  to  roosts, 
especially  W.  D.  Shepard,  M.  Schlenker,  and  M.  Mor- 
ris. We  also  thank  the  California  State  University  Sac- 
ramento Foundation  for  a monetary  award  to  purchase 
equipment  used  in  this  study. 

LITERATURE  CITED 

Birkhead,  T.  R.  1991.  The  magpies;  the  ecology  and 
behaviour  of  Black-billed  and  Yellow-billed  mag- 
pies. Academic  Press,  San  Diego,  California. 
Crosbie,  S.  P.  2004.  The  communal  roosting  behavior 
of  urban  Yellow-billed  Magpies  ( Pica  nuttalli). 
M.Sc.  thesis,  California  State  University,  Sacra- 
mento. 

Francis,  W.  J.  1976.  Micrometeorology  of  a blackbird 
roost.  Journal  of  Wildlife  Management  40:132- 
136. 

Marzluff,  J.  M.,  B.  Heinrich,  and  C.  Marzluff. 
1996.  Raven  roosts  are  mobile  information  cen- 
ters. Animal  Behaviour  51:89-103. 

Miranda,  A.  P.  and  J.  S.  M.  Gonzalez.  2000.  Two 
factors  affecting  communal  roosting  in  magpies 
(Pica  pica):  human  disturbance  and  minimum 
temperatures. 

M0ller,  A.  P.  1985.  Communal  roosting  in  the  Magpie 
(Pica  pica).  Journal  of  Ornithology  126:405-419. 
Mugaas,  J.  N.  and  J.  R.  King.  1981.  Annual  variation 
of  daily  energy  expenditure  by  the  Black-billed 
Magpie:  a study  of  thermal  and  behavioral  ener- 
getics. Studies  in  Avian  Biology  5:1-78. 

Pulliam,  H.  R.  1973.  On  the  advantages  of  flocking. 

Journal  of  Theoretical  Biology  38:419-422. 
Reebs,  S.  G.  1986.  Influence  of  temperature  and  other 
factors  on  the  daily  roosting  times  of  Black-billed 
Magpies.  Canadian  Journal  of  Zoology  64:1614- 
1619. 

Reebs,  S.  G.  1987.  Roost  characteristics  and  roosting 
behaviour  of  Black-billed  Magpies,  Pica  pica , in 
Edmonton,  Alberta.  Canadian  Field-Naturalist 
101:519-525. 

Reynolds,  M.  D.  1995.  Yellow-billed  Magpie  (Pica 
nuttalli).  The  Birds  of  North  America,  no.  180. 
SPSS  Institute,  Inc.  1998.  SPSS  for  Windows,  ver. 

11.5.  SPSS  Institute,  Inc.,  Chicago,  Illinois. 
Verbeek,  N.  A.  M.  1973.  The  exploitation  system  of 
the  Yellow-billed  Magpie.  University  of  Califor- 
nia Publications  in  Zoology,  no.  99. 

Walsberg,  G.  E.  1986.  Thermal  consequences  of 
roost-site  selection:  the  relative  importance  of 
three  modes  of  heat  conservation.  Auk  103:1-7. 
Zar,  J.  H.  1999.  Biostatistical  analysis,  4th  ed.  Pren- 
tice-Hall, Upper  Saddle  River,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  1 18(4):537-546,  2006 


NESTING  SUCCESS  OF  GRASSLAND  AND  SAVANNA  BIRDS  ON 
RECLAIMED  SURFACE  COAL  MINES  OF  THE  MIDWESTERN 

UNITED  STATES 

EDWARD  W.  GALLIGAN,1’3 4 5  TRAVIS  L.  DeVAULT, 24  AND  STEVEN  L.  LIMA1 5 


ABSTRACT. — Reclaimed  surface  coal  mines  in  southwestern  Indiana  support  many  grassland  and  shrub/ 
savanna  bird  species  of  conservation  concern.  We  examined  the  nesting  success  of  birds  on  these  reclaimed 
mines  to  assess  whether  such  “unnatural”  places  represent  productive  breeding  habitats  for  such  species.  We 
established  eight  study  sites  on  two  large,  grassland-dominated  mines  in  southwestern  Indiana  and  classified 
them  into  three  categories  (open  grassland,  shrub/savanna,  and  a mixture  of  grassland  and  shrub/savanna)  based 
on  broad  vegetation  and  landscape  characteristics.  During  the  1999  and  2000  breeding  seasons,  we  found  and 
monitored  911  nests  of  31  species.  Daily  nest  survival  for  the  most  commonly  monitored  grassland  species 
ranged  from  0.903  (Dickcissel,  Spiza  americana)  to  0.961  (Grasshopper  Sparrow,  Ammodramus  savannarum). 
Daily  survival  estimates  for  the  dominant  shrub/savanna  nesting  species  ranged  from  0.932  (Brown  Thrasher, 
Toxostoma  rufum ) to  0.982  (Willow  Flycatcher,  Empidonax  trailin').  Vegetation  and  landscape  effects  on  nesting 
success  were  minimal,  and  only  Eastern  Meadowlarks  ( Sturnella  magna ) showed  a clear  time-of-season  effect, 
with  greater  nesting  success  in  the  first  half  of  the  breeding  season.  Rates  of  Brown-headed  Cowbird  ( Molothrus 
ater ) parasitism  were  only  2.1%  for  grassland  species  and  12.0%  for  shrub/savanna  species.  The  nesting  success 
of  birds  on  reclaimed  mine  sites  was  comparable  to  that  in  other  habitats,  indicating  that  reclaimed  habitats  on 
surface  mines  do  not  necessarily  represent  reproductive  traps  for  birds.  Received  1 August  2005,  accepted  10 
April  2006. 


Several  bird  species  have  benefited  in  re- 
cent decades  from  the  reclamation  of  surface 
coal  mines  in  the  midwestem  United  States 
(Bajema  et  al.  2001,  De Vault  et  al.  2002,  In- 
gold 2002).  The  Surface  Mining  Reclamation 
Act  of  1977  and  earlier  laws  led  (perhaps  un- 
intentionally) to  mine  reclamation  techniques 
that  favored  the  production  of  grasslands  rath- 
er than  forested  habitats  (Brothers  1990),  re- 
sulting in  hundreds  of  km2  of  newly  created 
grasslands.  These  “mine  grasslands”  harbor  a 
diverse  assemblage  of  grassland  birds,  many 
of  which  are  of  management  concern  at  state 
and  federal  levels.  Recent  studies  in  south- 
western Indiana,  covering  19  reclaimed  mines, 
suggest  that  populations  of  key  grassland  bird 
species,  such  as  Grasshopper  {Ammodramus 
savannarum)  and  Henslow’s  (A.  henslowii) 
sparrows,  are  quite  large  (Bajema  et  al.  2001, 
De  Vault  et  al.  2002).  Reclaimed  mines  also 


1 Dept,  of  Ecology  and  Organismal  Biology,  Indiana 
State  Univ.,  Terre  Haute,  IN  47809,  USA. 

2 Dept,  of  Forestry  and  Natural  Resources,  Purdue 
Univ.,  West  Lafayette,  IN  47907,  USA. 

3 Current  address:  Louisville  Metro  Health  Dept., 
400  E.  Gray  St.,  Louisville,  KY  40202,  USA. 

4 Current  address:  U.S.  Dept,  of  Agriculture,  Wild- 
life Services,  National  Wildlife  Research  Center,  5757 
Sneller  Rd.,  Brewerton,  NY  13029,  USA. 

5 Corresponding  author;  e-mail:  slima@indstate.edu 


contain  scattered  trees  (from  plantings  and 
natural  succession)  that  approximate  the  struc- 
ture of  savanna  habitat  to  a substantial  degree 
(Scott  et  al.  2002,  Scott  and  Lima  2004).  Ac- 
cordingly, these  reclaimed  mines  harbor  sev- 
eral savanna  bird  species  (De Vault  et  al.  2002) 
of  conservation  concern  (Davis  et  al.  2000, 
Hunter  et  al.  2001). 

The  size  of  reclaimed  mines  in  the  mid- 
westem United  States  is  one  of  their  most  im- 
portant characteristics — several  exceed  2,000 
ha  (Bajema  and  Lima  2001,  Ingold  2002). 
Many  grassland  bird  species  appear  to  be 
“area  sensitive”  in  that  usually  they  are  found 
only  in  grassland  fragments  of  a given  size  or 
greater  (Herkert  1994,  Walk  and  Warner  1999, 
Winter  and  Faaborg  1999;  but  see  Horn  et  al. 
2000,  Johnson  and  Igl  2001).  Most  studies 
suggest  that  grasslands  >50-100  ha  should 
contain  a full  complement  of  grassland  pas- 
serines. Virtually  all  grasslands  on  reclaimed 
mines  in  southwestern  Indiana  are  >100  ha 
(Bajema  and  Lima  2001).  Furthermore,  small 
grassland  size  may  be  associated  with  poor 
nesting  success,  reflecting  the  close  proximity 
of  habitat  edge,  which  can  lead  to  greater 
predator  densities  (Winter  et  al.  2000,  Herkert 
et  al.  2003)  and  greater  rates  of  Brown-headed 
Cowbird  {Molothrus  ater)  parasitism  (Johnson 


537 


538 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


and  Temple  1990).  These  effects  of  habitat 
size  are  similar  to  those  documented  for  many 
forest-nesting  passerines  (e.g.,  Donovan  et  al. 
1995,  Robinson  et  al.  1995). 

Even  though  large  reclaimed  coal  mines  in 
the  Midwest  harbor  a variety  of  breeding  bird 
species,  they  are  decidedly  unnatural  places  in 
terms  of  vegetation  (Scott  and  Lima  2004). 
Hence,  it  is  conceivable  that  reclaimed  mines 
function  as  giant  ecological  “traps”  that  di- 
vert breeding  birds  away  from  more  produc- 
tive habitats  {sensu  Gates  and  Gysel  1978). 
Even  though  grassland  birds  can  breed  suc- 
cessfully in  non-native  grasslands  (e.g.,  Warn- 
er 1994,  Best  et  al.  1997,  Kershner  and  Bol- 
linger 1998,  Robb  et  al.  1998,  Ingold  2002, 
Monroe  and  Ritchison  2005),  the  possibility 
that  they  represent  ecological  traps  is  not  triv- 
ial. For  example,  reclaimed  midwestern  mines 
often  are  dominated  by  tall  fescue  ( Festuca 
arundinacea;  Scott  et  al.  2002,  Scott  and  Lima 
2004),  which  often  is  infected  with  a symbi- 
otic fungal  endophyte  ( Neotyphodium  coeno- 
phialum).  Such  infected  fescue  is  associated 
with  declines  in  plant  diversity  and  lowered 
reproductive  success  of  herbivores  (vertebrate 
and  invertebrate;  Clay  and  Holah  1999).  Tall 
fescue  might  reduce  insect  production  and 
render  reclaimed  mine  grasslands  into  poor 
breeding  habitat.  Although  tall  fescue  also 
may  affect  the  breeding  prospects  of  savanna 
bird  species,  they  might  be  less  affected  than 
their  grassland  counterparts. 

There  are  few  data  available  for  assessing 
whether  birds  inhabiting  reclaimed  surface 
mines  are  nesting  successfully.  Thus,  our  goal 
in  this  study  was  to  investigate  patterns  of  avi- 
an nesting  success  within  reclaimed  surface 
coal  mines,  with  the  larger  goal  of  evaluating 
whether  reclaimed  mines  provide  productive 
breeding  habitats  for  grassland  and  savanna 
birds. 

METHODS 

Study  sites. — Our  work  spanned  the  1999 
and  2000  breeding  seasons.  In  both  years, 
field  work  began  in  late  April  and  continued 
through  the  1st  week  of  August.  Study  sites 
were  established  at  two  large  reclaimed  sur- 
face coal  mines  in  west-central  Indiana  within 
30  km  of  the  city  of  Terre  Haute.  Four  sites 
were  established  at  the  Chinook  Mine  (39° 
28'  N,  87°  13'  W;  2,000  ha)  in  Clay  and  Vigo 


counties  and  four  were  established  at  the  Uni- 
versal Mine  (39°  36'  N,  87°  28'  W;  3,450  ha) 
in  southern  Vermillion  County.  The  Chinook 
sites  ranged  in  size  from  39  to  67  ha,  whereas 
the  Universal  sites  were  smaller  (12  to  38  ha) 
due  to  constraints  imposed  by  cattle  and  hay- 
ing operations.  Chinook  Mine  comprised  61% 
undisturbed  grassland  and  18%  hayfields;  the 
remaining  2 1 % comprised  relatively  even  per- 
centages of  wetlands,  row  crops,  and  forests 
(Bajema  and  Lima  2001).  Universal  Mine  was 
33%  undisturbed  grassland  and  43%  hayfields 
and  cattle  pastures,  with  the  remaining  24% 
split  about  evenly  between  forest  and  lakes/ 
wetlands  (Bajema  and  Lima  2001). 

Study  sites  were  chosen  to  represent  the 
range  of  grassland-dominated  habitats  found 
in  the  reclaimed  surface  coal  mines  of  south- 
western Indiana.  Two  study  sites  (one  in  each 
mine)  were  classified  as  “open  grassland.” 
We  defined  open  grassland  sites  as  relatively 
undisturbed  areas  (no  mowing  for  >2  years, 
usually  many  more)  that  were  dominated  by 
grasses  (>95%,  by  area),  with  some  forbs  and 
very  few  saplings,  trees,  or  shrubs  (Scott  et  al. 
2002).  Open  grasslands  represented  the  most 
abundant  habitat  type  found  on  most  re- 
claimed surface  mines  (Bajema  and  Lima 
2001).  Nests  found  in  these  open  sites  were, 
on  average,  760  m from  the  nearest  mature 
forest  habitat,  with  many  nests  well  over 
1,000  m from  forest. 

Three  study  areas  were  classified  as  “shrub/ 
savanna”  sites  (one  at  Chinook  Mine  and  two 
at  Universal  Mine).  We  defined  shrub/savanna 
sites  as  predominantly  grassy  habitats  with 
many  scattered  young  trees  (4-8  m high,  gen- 
erally open  canopy)  and  shrubs,  often  repre- 
senting a transition  zone  between  grassland 
and  forested  areas.  Small  groves  of  trees  also 
were  associated  with  small  wetland  areas. 
Black  locusts  ( Robinia  pseudoacacia)  domi- 
nated in  shrub/savanna  sites,  although  signif- 
icant numbers  of  oaks  ( Quercus  spp.),  eastern 
cottonwoods  ( Populus  deltoides ),  and  mature 
autumn  olives  ( Elaeagnus  umbellata ) were 
found  in  some  areas.  “Shrubby”  species  in- 
cluded young  saplings  of  these  tree  species, 
along  with  hawthorn  ( Crataegus  spp.)  and 
multiflora  rose  bushes  ( Rosa  multiflora). 
Shrub/savanna  sites  were  adjacent  to  mature 
forest  (and  hence  were  mainly  on  the  edges  of 
the  reclaimed  mines).  The  average  distance 


Galligan  et  al.  • NESTING  SUCCESS  ON  RECLAIMED  SURFACE  MINES 


539 


between  nests  found  on  shrub/savanna  sites 
and  the  forest  edge  was  240  m.  Shrub/savanna 
sites  contained  significant  (30-60%,  by  area) 
open  grassland  habitat. 

Finally,  we  designated  the  remaining  three 
study  sites  (two  at  Chinook  Mine  and  one  at 
Universal  Mine)  as  “mixed”  sites.  Mixed 
sites  were  defined  as  mostly  open  grassland 
habitat  with  a few  areas  of  significant  shrub/ 
savanna  habitat.  In  general,  these  sites  were 
70-80%  open  grassland.  The  average  distance 
between  nests  found  on  mixed  sites  and  the 
nearest  mature  forest  habitat  was  430  m. 

Nest  location  and  monitoring. — Nest 
searches  were  conducted  daily  in  1999  and 
2000  from  early  morning  until  early  afternoon 
by  a team  of  three  to  five  field  workers.  Nests 
were  detected  by  (1)  rope  dragging,  (2)  fol- 
lowing adults  that  were  carrying  food  and 
nesting  material,  and  (3)  systematic  searches 
of  likely  nesting  sites  (Martin  and  Geupel 
1993).  During  the  2000  field  season,  we  also 
used  a thermographic  imager  to  aid  in  nest 
detection  (Galligan  et  al.  2003). 

When  a nest  was  located,  a small  colored 
flag  was  placed  10  m to  the  north  of  it  and  a 
small  piece  of  colored  tape  was  tied  to  vege- 
tation 5 m south  of  the  nest  (Picozzi  1975, 
Walk  2001).  The  accurate  alignment  of  flag, 
tape,  and  nest  allowed  workers  to  relocate 
nests  quickly  with  minimal  disturbance.  Spe- 
cies associated  with  each  nest  were  identified, 
and  nests  were  checked  only  every  3 to  4 days 
to  minimize  disturbance  (Bart  1977).  During 
each  nest  check,  we  recorded  the  presence  or 
absence  of  adults,  the  number  of  eggs  or 
chicks,  and,  if  appropriate,  the  developmental 
stage  of  the  chicks.  We  also  recorded  indica- 
tors of  nestling  mortality  or  cowbird  parasit- 
ism. 

Vegetation  and  landscape  variables. — We 
gathered  basic  information  on  the  physical  re- 
lationships between  nests,  the  surrounding 
vegetation,  and  major  landscape  features; 
however,  we  limited  analyses  of  these  vari- 
ables to  grassland  bird  species,  whose  nests 
were  located  in  greater  numbers  than  savanna 
species.  For  each  nest,  we  recorded  height 
above  ground,  species  and  height  of  the  veg- 
etation in  which  it  was  placed,  dominant  veg- 
etation and  vegetation  height  within  1 m of 
the  nest,  litter  depth  at  the  nest,  percent  cover 
of  litter  within  1 m of  the  nest,  distance  to  the 


nearest  forest  edge,  and  distance  to  the  nearest 
tree  (>1  m high).  We  used  GPS  units  to  re- 
cord the  location  of  all  nests  and  to  delineate 
nearby  forested  areas. 

Data  analyses. — We  estimated  the  daily 
probability  of  nest  survival  (DNS)  for  each 
species  according  to  the  Mayfield  method 
(Mayfield  1961,  1975).  We  assumed  that  any 
relevant  nesting  event  (e.g.,  hatching,  failure, 
fledging)  occurred  at  the  midpoint  of  the  in- 
terval between  nest  visits.  A nest  was  consid- 
ered successful  when  it  fledged  one  or  more 
young  (Mayfield  1961,  1975). 

Our  analyses  were  limited  primarily  to  uni- 
variate tests  of  vegetation,  landscape,  and 
temporal  variable  effects  on  DNS  or  the  fate 
of  individual  nests  (success  or  failure).  We 
tested  for  interactions  only  for  study  site  and 
time  of  season.  We  compared  DNS  estimates 
across  categorical  variables  (i.e.,  among  years, 
sites,  and  different  habitat  types)  by  using 
CONTRAST  (Hines  and  Sauer  1989).  CON- 
TRAST uses  a generalized  x2  statistic  that  al- 
lows multiple  comparisons  of  survival  rates 
from  different  time  periods  or  study  areas 
(Sauer  and  Williams  1989).  We  compared 
DNS  among  years  and  sites  for  all  species  list- 
ed in  Table  1.  Because  we  found  large  num- 
bers of  Field  Sparrow  ( Spizella  pusilla),  Am- 
modramus  spp.  (Henslow’s  and  Grasshopper 
sparrows,  combined),  Dickcissel  ( Spiza  amer- 
icana ),  Red-winged  Blackbird  ( Agelaius 
phoeniceus ),  and  Eastern  Meadowlark  ( Stur - 
nella  magna ) nests,  we  were  able  to  examine 
DNS  trends  within  breeding  seasons  (compar- 
ing DNS  between  the  first  and  second  halves 
of  the  breeding  seasons)  and  between  habitat 
types  for  these  species.  We  used  logistic  re- 
gression, with  the  fate  of  individual  nests 
(failure  or  success)  as  the  dependent  variable, 
to  evaluate  the  effects  of  various  continuous 
landscape  and  vegetation  variables  on  nesting 
success  (SPSS,  Norusis  1993).  Our  analyses 
were  applied  primarily  to  habitat  types  (open, 
mixed,  and  shrub/savanna)  because  they  were 
distinctly  different  from  the  surrounding  land- 
scape characteristics.  For  a given  habitat  type, 
we  limited  our  analyses  to  those  species  for 
which  we  had  adequate  numbers  of  nest-days 
(see  grassland  species  listed  in  Table  1).  The 
effects  of  various  factors  on  nest  survival 
were  analyzed  individually,  except  as  noted. 
Results  are  presented  as  means  and  standard 


540 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  1.  Daily  nest  survival  (DNS)  for  bird  species  inhabiting  grassland  and  shrub/savanna  on  reclaimed 
coal  mines  in  Indiana  during  1999  and  2000. 


n (no.  successful) 

NDa 

DNS 

SE 

Grassland  species 

Grasshopper  Sparrow 

41  (26) 

383 

0.961 

0.010 

Henslow’s  Sparrow 

21  (9) 

236 

0.949 

0.014 

Field  Sparrow 

90  (36) 

629 

0.919 

0.011 

Dickcissel 

47  (11) 

369 

0.903 

0.016 

Eastern  Meadowlark 

129  (58) 

1450 

0.951 

0.006 

Red- winged  Blackbird 

264  (74) 

2439 

0.923 

0.005 

Shrub/Savanna  species 

Mourning  Dove 

62  (33) 

816 

0.962 

0.007 

Willow  Flycatcher 

30  (22) 

440 

0.982 

0.006 

American  Robin 

33  (12) 

313 

0.933 

0.014 

Brown  Thrasher 

31  (14) 

251 

0.932 

0.016 

Yellow  Warbler 

21  (13) 

272 

0.971 

0.010 

a Number  of  nest-days  observed. 


errors;  the  level  of  significance  was  set  at 
0.05. 

RESULTS 

During  our  2-year  study,  we  found  9 1 1 ac- 
tive nests  of  31  species.  Of  these  nests,  465 
and  446  were  found  at  the  Chinook  and  Uni- 
versal mines,  respectively.  Red-winged  Black- 
birds, Eastern  Meadowlarks,  Field  Sparrows, 
Dickcissels,  Grasshopper  Sparrows,  and  Hen- 
slow’s  Sparrows  were  (in  that  order)  the  best 
represented  grassland  birds  (Table  1).  Nests  of 
other  grassland  species,  such  as  those  of  Ring- 
necked Pheasants  ( Phasianus  colchicus ), 
Sedge  Wrens  ( Cistothorus  platensis ),  and 
Bobolinks  ( Dolichonyx  oryzivorus ),  were  too 
few  in  number  for  analyses,  as  these  species 
are  relatively  rare  on  the  reclaimed  surface 
mines  (DeVault  et  al.  2002).  Among  the 
shrub/savanna  species,  nests  of  Mourning 
Doves  ( Zenaida  macroura),  Willow  Flycatch- 
ers ( Empidonax  traillii ),  American  Robins 
(Turdus  migratorius),  and  Brown  Thrashers 
(' Toxostoma  rufum)  were  found  most  frequent- 
ly (Table  1).  The  nests  of  other  savanna  spe- 
cies were  located  in  numbers  too  small  for 
analyses,  including  those  of  Eastern  Kingbird 
( Tyrannus  tyrannus ),  Bell’s  Vireo  ( Vireo  bel- 
lii ),  Song  Sparrow  ( Melospiza  melodia ),  Blue 
Grosbeak  ( Passerina  caerulea),  Indigo  Bun- 
ting ( Passerina  cyanea ),  Orchard  Oriole  ( Ic- 
terus spurius),  and  American  Goldfinch  (Car- 
due  lis  tristis). 


Daily  probability  of  nest  survival:  overall 
estimates. — The  overall  estimates  of  DNS  (all 
data  pooled)  showed  considerable  interspecif- 
ic variation.  Among  grassland  species  (Table 
1),  we  estimated  relatively  high  rates  of  DNS 
(near  0.950)  for  Grasshopper  Sparrows,  Hen- 
slow’s  Sparrows,  and  Eastern  Meadowlarks. 
Conversely,  we  estimated  DNS  of  <0.925  for 
Dickcissels  (the  lowest:  0.903),  Field  Spar- 
rows, and  Red-winged  Blackbirds.  Among  sa- 
vanna species.  Willow  Flycatchers  and  Yellow 
Warblers  (Dendroica  petechia)  experienced 
the  greatest  DNS  (0.982  and  0.971,  respec- 
tively); we  also  estimated  a high  DNS  for 
Mourning  Doves  (0.962),  and  our  DNS  esti- 
mate for  Brown  Thrasher  was  the  lowest 
(0.932). 

Predation  accounted  for  the  vast  majority  of 
nest  failures.  In  most  cases,  we  could  only 
guess  at  the  identity  of  the  predators  involved 
because  many  predators  do  not  leave  conclu- 
sive evidence  of  their  identities  at  depredated 
nests  (Thompson  et  al.  1999,  Maier  and 
DeGraaf  2000,  Pietz  and  Granfors  2000,  Bur- 
hans  et  al.  2002).  However,  many  snakes  were 
encountered  during  nest  searches,  mainly  rac- 
ers ( Coluber  constrictor)  and  black  rat  snakes 
(Elaphe  obsoleta  obsoleta);  we  also  encoun- 
tered smaller  numbers  of  garter  snakes  (Tham- 
nophis  spp.)  and  prairie  kingsnakes  (Lampro- 
peltis  calligaster).  Snakes  were  observed  con- 
suming eggs  or  chicks  on  two  occasions.  Only 
Red-winged  Blackbirds  appeared  to  suffer  any 


Galligan  et  al.  • NESTING  SUCCESS  ON  RECLAIMED  SURFACE  MINES 


541 


TABLE  2.  Daily  nest  survival 
savanna),  on  reclaimed  coal  mines 
CONTRAST. 

(DNS)  for  grassland  birds,  by  site  type  (open  grassland,  mixed,  and  shrub/ 
in  Indiana  during  1999  and  2000;  x2  statistics  were  determined  using  program 

Species 

Habitat  type 

n 

DNS 

SE 

Red- winged  Blackbird 

Open  grassland 

97 

0.914 

0.010 

Mixed 

154 

0.923 

0.007 

Shrub/savanna 

21 

0.949 

0.014 

X2  = 4.13, 

df  = 

2,  P = 0.13 

Eastern  Meadowlark 

Open  grassland 

62 

0.939 

0.010 

Mixed 

46 

0.962 

0.008 

Shrub/savanna 

23 

0.974 

0.010 

X2  = 6.39, 

df  = 

2,  P = 0.04 

Dickcissel 

Open  grassland 

23 

0.903 

0.022 

Mixed 

15 

0.916 

0.024 

Shrub/savanna 

9 

0.871 

0.043 

X2  = 0.86, 

df  = 

2,  P = 0.65 

Field  Sparrow 

Open  grassland 

33 

0.933 

0.017 

Mixed 

25 

0.938 

0.016 

Shrub/savanna 

34 

0.879 

0.024 

II 

45- 

o 

df  = 

2,  P = 0.11 

Ammodramus  spp. 

Open  grassland 

25 

0.977 

0.009 

Mixed 

20 

0.943 

0.017 

Shrub/savanna 

17 

0.928 

0.023 

X2  = 6.04, 

df  = 

2,  P = 0.05 

weather-induced  mortality  (nests  blown  over 
during  severe  thunderstorms),  and  then  only 
early  in  the  1999  breeding  season.  There  were 
no  indications  of  significant  nutritional  stress 
among  any  nestlings. 

Effects  of  time  and  site. — DNS  estimates 
(all  sites  pooled)  did  not  differ  between  years 
(X2:  all  P values  > 0.10)  for  any  grassland  or 
savanna  species  except  Brown  Thrasher  (x2  = 
5.70,  df  = 1,  P = 0.017).  Brown  Thrasher 
DNS  was  very  low  in  1999  (0.895  ± 0.027), 
but  was  much  greater  in  2000  (0.969  ± 
0.015).  For  American  Robin,  there  was  a sim- 
ilar across-year  trend  (x2  — 3.22,  df  = 1,  P = 
0.072)  in  DNS,  which  increased  from  0.885 
± 0.034  to  0.951  ± 0.014. 

We  found  a significant  time-of-season  ef- 
fect only  for  Eastern  Meadowlarks;  in  both 
years,  our  estimate  of  their  DNS  was  substan- 
tially greater  during  the  first  half  of  the  breed- 
ing season  than  in  the  second  half.  In  1999, 
their  DNS  decreased  from  0.974  ± 0.008  to 
0.919  ± 0.016  (x2  = 9.45,  df  = 1,  P = 0.005) 
and,  in  2000,  from  0.966  ± 0.009  to  0.934  ± 
0.014  (x2  = 3.70,  df  = 1,  P = 0.051).  When 
the  data  were  pooled  across  years,  DNS  in  the 
first  and  second  half  of  the  breeding  season 
differed  substantially  (0.970  ± 0.006  versus 


0. 926  ± 0.011,  respectively;  x2  — 12.33,  df  = 

1,  P < 0.001). 

Significant  differences  in  DNS  also  were 
observed  across  habitat  types  (Table  2).  DNS 
for  Eastern  Meadowlarks  was  greatest  in 
shrub/savanna  habitat  (0.974)  and  lowest  in 
the  open  habitats  (0.939;  x2  = 6.39,  df  = 2, 
P = 0.041).  DNS  of  Ammodramus  sparrows 
was  higher  in  the  increasingly  open  habitats 
(X2  = 6.04,  df  = 1,  P = 0.050).  For  Field 
Sparrows,  our  DNS  estimates  tended  to  be 
lower  in  the  shrub/savanna  habitats  (x2  = 
4.508,  df  = 1,  P = 0.11).  DNS  for  Dickcissels 
also  was  lowest  in  the  shrub/savanna  habitat 
(0.871),  but  not  significantly  so.  Logistic  re- 
gression analyses  of  these  data  produced  very 
similar  results,  indicating  no  significant  inter- 
actions between  habitat  type  and  time  of  sea- 
son, for  any  of  the  species  listed  in  Table  2 
(Wald  x2  tests:  all  P > 0.50).  DNS  did  not 
differ  between  mines  (x2:  all  P > 0.10;  pool- 
ing data  across  all  study  sites  within  a given 
mine)  for  any  species  listed  in  Table  1. 

Effects  of  vegetation  and  landscape  vari- 
ables.— Our  analyses  indicated  few  significant 
associations  between  DNS  and  vegetation  or 
landscape  features.  However.  DNS  for  Eastern 
Meadowlarks  increased  with  distance  to  forest 


542 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


in  the  shrub/savanna  sites  (logistic  regression: 
b = 0.019,  Wald  x2  = 3.95,  df  = 1,  P = 
0.047).  Increasing  nest  height  also  was  asso- 
ciated with  lower  DNS  for  Field  Sparrows,  but 
only  in  open  grassland  habitats  (b  = —0.487, 
Wald  x2  = 4.22,  df  = 1,  P = 0.040);  DNS 
was  lower  for  nests  in  low  shrubs  than  for 
those  on  the  ground.  For  Red- winged  Black- 
birds, height  of  vegetation  in  which  the  nest 
was  placed  was  positively  associated  with 
nesting  success,  but  only  in  the  mixed  habitat 
type  (b  = 0.051,  Wald  x2  = 6.42,  df  = 1,  P 
= 0.011).  Finally,  for  Grasshopper  Sparrows, 
height  of  the  dominant  vegetation  within  1 m 
of  the  nest  was  positively  associated  with 
nesting  success,  but  only  when  the  data  were 
pooled  across  all  habitat  types  (b  = 0.465, 
Wald  x2  = 4.14,  df  = 1,  P = 0.046).  It  is 
notable  that  tall  fescue  (either  as  the  vegeta- 
tion in  which  the  nest  was  placed  or  as  the 
dominant  vegetation  within  1 m of  the  nest) 
was  not  significantly  associated  with  the  DNS 
of  any  focal  species. 

Brood  parasitism. — Relatively  low  rates  of 
brood  parasitism  by  Brown-headed  Cowbirds 
were  observed  during  our  2-year  study.  Over- 
all, only  2.1%  of  grassland  bird  nests  were 
parasitized  by  cowbirds  (Table  3).  Field  Spar- 
rows were  the  most  heavily  parasitized 
(6.4%),  whereas  we  observed  no  parasitism  on 
Henslow’s  Sparrows  or  Eastern  Meadowlarks. 
Furthermore,  of  the  263  Red- winged  Black- 
bird nests  that  we  found,  only  four  were  par- 
asitized. Shrub/savanna  species  as  a group  (in- 
cluding all  species  monitored)  suffered  a 
greater  frequency  of  brood  parasitism  (12.0%; 
Table  3).  Of  the  savanna  species.  Orchard  Ori- 
oles and  Blue  Grosbeaks  were  most  heavily 
parasitized. 

DISCUSSION 

Daily  nest  survival. — Overall  estimates  of 
DNS  varied  considerably  across  species.  In 
general,  shrub/savanna  birds  experienced 
greater  rates  of  DNS  than  grassland  birds  (Ta- 
ble 1).  Among  grassland  birds.  Eastern  Mead- 
owlarks, Grasshopper  Sparrows,  and  Hen- 
slow’s  Sparrows  experienced  relatively  high 
rates  of  DNS,  whereas  Dickcissels,  Field 
Sparrows,  and  Red-winged  Blackbirds  expe- 
rienced lower  rates  of  DNS.  Among  shrub/sa- 
vanna species.  Mourning  Doves,  Willow  Fly- 
catchers, and  Yellow  Warblers  experienced 


TABLE  3.  Brown-headed  Cowbird  parasitism  of 
host  species  was  infrequent  on  reclaimed  coal  mines 
in  Indiana  during  1999  and  2000. 

Species3 

n 

No. 

Parasitized 

% 

Grassland 
Sedge  Wren 

l 

0 

0.0 

Red- winged  Blackbird 

263 

4 

1.5 

Bobolink 

1 

0 

0.0 

Eastern  Meadowlark 

131 

0 

0.0 

Dickcissel 

47 

2 

4.1 

Field  Sparrow 

93 

6 

6.4 

Grasshopper  Sparrow 

41 

1 

2.4 

Henslow’s  Sparrow 

21 

0 

0.0 

Total 

607 

13 

2.1 

Shrub/Savanna 

Eastern  Kingbird 

9 

1 

11.1 

Willow  Flycatcher 

30 

0 

0.0 

Bell’s  Vireo 

6 

1 

16.7 

Yellow  Warbler 

21 

3 

14.3 

Orchard  Oriole 

10 

4 

40.0 

Blue  Grosbeak 

6 

2 

33.3 

Indigo  Bunting 

4 

0 

0.0 

American  Goldfinch 

6 

0 

0.0 

Song  Sparrow 

8 

1 

12.5 

Total 

100 

12 

12.0 

3 Known  egg  rejectors  (e.g.,  American  Robins,  Brown  Thrashers)  and 
unsuitable  cowbird  hosts  (e.g..  Mourning  Doves)  were  not  included. 


relatively  high  rates  of  DNS,  whereas  Amer- 
ican Robins  and  Brown  Thrashers  experienced 
relatively  low  rates  of  DNS.  There  were  no 
significant  differences  in  DNS  across  the  two 
mines  studied,  despite  the  fact  that  these 
mines  encompass  the  range  of  land-use  pat- 
terns found  within  mines  (Bajema  and  Lima 
2001).  There  also  were  few  significant  differ- 
ences in  DNS  across  the  two  breeding  sea- 
sons, despite  the  fact  that  the  first  season 
(1999)  was  relatively  hot  and  dry,  and  the  sec- 
ond season  (2000)  was  cool  and  wet  (only 
Brown  Thrashers  and  American  Robins  had 
markedly  greater  DNS  in  2000  than  1999). 
Thus,  the  general  patterns  apparent  in  Table  1 
may  be  representative  of  the  long-term  situa- 
tions faced  by  birds  on  the  reclaimed  mines 
of  southwestern  Indiana. 

Ultimately,  the  variation  that  we  observed 
in  DNS  was  due  to  variation  in  nest  predation, 
the  primary  cause  of  nest  failure.  Among 
grassland  birds,  it  appears  that  open-cup, 
above-ground  nesters,  such  as  Field  Sparrows, 
Dickcissels,  and  Red-winged  Blackbirds,  suf- 
fered greater  predation  rates  than  ground-nest- 


Galligan  et  al.  • NESTING  SUCCESS  ON  RECLAIMED  SURFACE  MINES 


543 


ing  species  (Eastern  Meadowlarks  and  Am- 
modramus  sparrows;  Table  1 ).  During  both 
field  seasons,  we  estimated  greater  DNS  for 
Eastern  Meadowlarks  during  the  first  half  of 
the  breeding  season  than  during  the  second 
half.  This  time-of-season  effect  may  reflect 
the  fact  that  Eastern  Meadowlarks  began  nest- 
ing in  April  before  snakes  became  fully  ac- 
tive. No  other  temporal  patterns  in  DNS  were 
apparent  among  other  grassland  species. 

Open-cup  nesting  was  not  uniformly  asso- 
ciated with  greater  rates  of  nest  predation,  be- 
cause all  shrub/savanna  species  in  this  study 
are  open-cup  nesters,  and  many  experienced 
high  rates  of  DNS  (Table  1).  The  relatively 
low  rate  of  nesting  success  among  American 
Robins  and  Brown  Thrashers  was  due  to  ex- 
tremely high  levels  of  nest  predation  during 
1999  (which  was  not  observed  in  2000).  Why 
only  these  two  species  experienced  different 
levels  of  predation  across  years  is  not  clear; 
however,  because  robins  and  thrashers  nested 
in  very  similar  sites  in  the  shrub/savanna  hab- 
itat (interior  portions  of  larger  trees),  they 
likely  experienced  the  same  change  in  the 
predatory  environment  across  years. 

Significant  associations  between  DNS  and 
various  vegetation  and  landscape-level  fea- 
tures were  few,  and  provided  relatively  little 
insight  into  the  predation  processes  that  influ- 
enced DNS.  We  note,  however,  that  for  many 
species  we  located  too  few  nests  for  our  anal- 
yses to  detect  subtle  effects.  Regardless,  the 
significant  increase  in  DNS  with  increasing 
distance  from  the  forest — exhibited  only  in 
Eastern  Meadowlarks  in  the  shrub/savanna 
habitat — was  consistent  with  the  results  of 
other  studies  (e.g.,  Johnson  and  Temple  1990) 
that  implicated  forest-edge  predators  as  major 
agents  of  nest  failure  (recall  that  our  shrub/ 
savanna  sites  were  adjacent  to  forested  habi- 
tat). The  lack  of  an  effect  of  distance-to-forest 
in  the  open  grassland  and  mixed  study  sites 
may  reflect  the  relative  isolation  of  these  sites 
from  forested  habitat  (c/.  Paton  1994).  The 
relatively  high  rates  of  DNS  for  Ammodramus 
sparrows  in  the  open  grassland  habitats  (Table 
2)  also  may  reflect  the  isolation  from  forested 
habitat.  Nevertheless,  there  was  no  association 
between  distance-to-forest  and  DNS  for  any 
other  species  in  the  shrub-savanna  sites.  Fur- 
thermore, the  overall  nesting  success  of  East- 
ern Meadowlarks  was  actually  greater  in  the 


shrub/savanna  habitat  than  elsewhere  (Table 

2). 

Across  studies,  a consistent  picture  of  the 
effects  of  vegetation  and  landscape  variables 
on  nesting  success  of  many  grassland  species 
has  yet  to  emerge.  For  example,  Johnson  and 
Temple  (1990)  observed  increased  nest  pre- 
dation for  grassland  passerines  when  their 
nests  were  located  near  wooded  edges.  Winter 
et  al.  (2000)  found  that,  for  artificial  nests, 
fragment  size  and  vegetation  characteristics 
were  better  predictors  of  survival  than  dis- 
tance to  habitat  edge;  however,  Henslow’s 
Sparrow  nests  placed  within  50  m of  an  edge 
were  not  as  successful  as  those  at  greater  dis- 
tances from  forest  edge.  For  Dickcissels,  dis- 
tance to  habitat  edge  also  appeared  to  have 
little  effect  on  daily  survival  in  prairie  habitats 
(Hughes  et  al.  1999,  Winter  et  al.  2000).  Bur- 
hans  et  al.  (2002)  observed  that  Field  Spar- 
rows nesting  in  old  fields  had  greater  success 
when  nest  height  was  >3  m above  ground; 
however.  Best  (1978)  suggested  that  Field 
Sparrows  were  more  successful  when  nests 
were  near  the  ground  or  in  relatively  tall  veg- 
etation. Pribil  (1998)  did  not  detect  a relation- 
ship between  nest  success  and  vegetation  fea- 
tures for  Red-winged  Blackbirds. 

Brood  parasitism. — Brood  parasitism  was 
minimal  in  our  focal  species,  especially  when 
compared  with  the  high  frequency  of  parasit- 
ism reported  in  midwestem  forest  fragments 
(e.g.,  Robinson  et  al.  1995).  For  grassland 
birds,  only  2.1%  of  nests  were  parasitized. 
The  frequency  of  parasitism  for  Red-winged 
Blackbirds  at  our  reclaimed  surface  coal 
mines  (1.5%)  markedly  contrasts  with  the  par- 
asitism frequency  of  >30%  for  this  species  in 
other  habitats  and  areas  to  the  west  of  our 
study  sites  (Yasukawa  and  Searcy  1995,  Clot- 
felter  and  Yasukawa  1999).  Kershner  (2001) 
and  Walk  (2001)  reported  similarly  low  fre- 
quency of  parasitism  for  grassland  birds  nest- 
ing in  restored  prairies  in  nearby  eastern  Illi- 
nois (see  also  Robinson  and  Herkert  1997, 
Kershner  and  Bollinger  1998).  Perhaps  the 
frequency  of  grassland  bird  parasitism  is  gen- 
erally greater  well  to  the  west  of  Indiana 
(Johnson  and  Temple  1990,  Zimmerman 
1993,  Davis  2003;  but  see  Winter  1999,  Win- 
ter et  al.  2004).  In  any  case,  the  low  frequency 
of  cowbird  parasitism  for  grassland  birds  of 
western  Indiana  and  eastern  Illinois  supports 


544 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  4.  Daily  nest  survival  for  grassland  birds  on  reclaimed  coal  mines  in  Indiana  during  1999  and  2000 
was  similar  to  that  recorded  at  other  midwestern  grassland  sites. 


Species 

Kansas  CRP 
fields® 

Prairie 

(MO)b 

Big  Oaks 
NWR  (IN)C 

Iowa  CRP  fields 
(egg,  nestling  stage)d 

Restored 
prairie  (IL)e 

Reclaimed  coal 
mines  (IN)f 

Red-winged  Blackbird 

— 

— 

— 

0.943,  0.916 

0.954 

0.923 

Eastern  Meadowlark 

— 

0.940 

— 

— 

0.953 

0.951 

Dickcissel 

0.922 

0.940 

— 

0.951,  0.874 

0.941 

0.903 

Field  Sparrow 

— 

— 

0.919 

— 

0.955 

0.919 

Grasshopper  Sparrow 

— 

0.930 

— 

0.957,  0.937 

0.913 

0.961 

Henslow’s  Sparrow 

— 

0.950 

0.947 

— 

— 

0.949 

a Hughes  et  al.  (1999).  b Winter  and  Faaborg  (1999),  c Robb  et  al.  (1998),  d Patterson  and  Best  (1996),  e Kershner  (2001)  and  Walk  (2001);  f this  study. 


the  idea  that  cowbirds  in  the  eastern  United 
States  focus  on  forested  habitats  (Hahn  and 
Hatfield  1995).  Indeed,  Brown-headed  Cow- 
bird  is  among  the  rarest  passerine  species  in- 
habiting reclaimed  coal  mines  in  Indiana 
(De Vault  et  al.  2002). 

Shrub/savanna  species  underwent  greater 
rates  of  parasitism  than  grassland  species  (Ta- 
ble 3),  but  it  was  still  much  lower  than  that 
typically  observed  in  forested  habitats  across 
Indiana  and  Illinois  (e.g.,  Robinson  et  al. 
1995).  Among  the  focal  savanna  species  (Ta- 
ble 1),  only  Yellow  Warblers  were  parasitized 
to  a substantial  degree  (Table  3).  Note,  how- 
ever, that  three  of  our  focal  savanna  species 
are  either  inappropriate  cowbird  hosts 
(Mourning  Doves)  or  egg  rejectors  (American 
Robins  and  Brown  Thrashers).  Parasitism  ap- 
peared to  be  greater  for  some  non-focal  sa- 
vanna species  (e.g..  Orchard  Orioles  and  Blue 
Grosbeaks;  Table  3),  but  we  found  too  few 
nests  to  reach  a conclusion  concerning  their 
susceptibility  to  parasitism.  We  suspect  that 
greater  rates  of  cowbird  parasitism  in  our 
shrub/savanna  sites  reflected  their  proximity 
to  forested  habitat  (Hahn  and  Hatfield  1995). 

Conservation  implications. — Our  data  sug- 
gest that  reclaimed  surface  coal  mines  are  no 
more  likely  to  represent  reproductive  traps 
than  are  other  habitats  studied  to  date.  We 
base  this  view  on  a comparison  of  our  results 
with  those  from  comparable  studies  across  the 
midwestern  United  States.  DNS  within  re- 
claimed coal  mine  grasslands  at  our  study 
sites  is  broadly  comparable  to  that  in  other 
midwestern  grasslands  (Table  4).  The  most 
comparable  study  is  one  that  took  place  in 
large  blocks  of  restored  prairies  in  nearby 
eastern  Illinois  (Kershner  2001,  Walk  2001). 
DNS  of  Eastern  Meadowlarks  in  Illinois  was 


essentially  identical  to  that  observed  in  our 
reclaimed  mine  sites  (Table  4).  Dickcissels 
and  Field  Sparrows  experienced  greater  nest 
success  at  the  Illinois  sites  than  at  our  sites, 
whereas  Grasshopper  Sparrows  experienced 
greater  success  at  our  mine  sites  (few  Hen- 
slow’s  Sparrow  nests  were  found  at  the  Illi- 
nois site).  Similar  to  what  we  found  in  our 
study.  Red- winged  Blackbirds  in  Conserva- 
tion Reserve  Program  (CPR)  fields  of  Iowa 
experienced  poor  to  mediocre  nesting  success, 
Dickcissels  experienced  low  success  (with 
very  low  survival  in  the  nestling  stage),  and 
Grasshopper  Sparrows  had  relatively  high 
rates  of  success  (Patterson  and  Best  1996). 
Dickcissels  also  may  not  be  doing  well  in 
Kansas  or  Missouri  CRP  fields  (Hughes  et  al. 
1999,  Winter  and  Faaborg  1999).  Nesting  suc- 
cess of  Field  and  Henslow’s  sparrows  at  the 
Big  Oaks  National  Wildlife  Refuge  (formerly 
the  Jefferson  Proving  Ground)  in  southeastern 
Indiana  is  virtually  identical  to  that  of  birds 
nesting  on  reclaimed  surface  coal  mines 
(Robb  et  al.  1998).  Furthermore,  survival  es- 
timates for  Henslow’s  Sparrows  across  the 
three  relevant  studies  (Robb  et  al.  1998,  Win- 
ter and  Faaborg  1999;  this  study)  were  re- 
markably similar  and  relatively  high,  indicat- 
ing that  this  species  is  probably  doing  reason- 
ably well  where  it  is  still  nesting.  Similarly, 
Monroe  and  Ritchison  (2005)  reported  com- 
parable levels  of  nesting  success  for  Hen- 
slow’s Sparrows  on  reclaimed  mines  and  un- 
mined grasslands  in  western  Kentucky,  and 
suggested  that  reclaiming  surface  mines  could 
help  stabilize  the  population  decline  of  Hen- 
slow’s Sparrows.  We  suspect  that  similar  con- 
clusions also  could  be  drawn  for  some  savan- 
na species  on  reclaimed  mines,  but  compara- 


Galligan  et  al.  • NESTING  SUCCESS  ON  RECLAIMED  SURFACE  MINES 


545 


ble  data  are  not  yet  available  with  which  to 
make  analogous  comparisons. 

Reclaimed  mines  of  the  Midwest  provide  a 
unique  opportunity  in  avian  conservation,  es- 
pecially for  the  management  of  grassland 
birds.  Many  of  the  reclaimed  mines  are 
>2,000  ha,  larger  than  most  (if  not  all)  re- 
maining prairie  fragments  in  Indiana  and  Il- 
linois, and  contain  large  populations  of  several 
bird  species  of  concern  (Bajema  et  al.  2001, 
De Vault  et  al.  2002,  Ingold  2002).  The  nesting 
success  of  key  species  (e.g.,  Henslow’s  Spar- 
rows and  Grasshopper  Sparrows)  at  these  re- 
claimed mines  is  comparable  with  that  in  non- 
mined  grassland  habitats.  A feature  that 
should  make  reclaimed  midwestern  surface 
coal  mines  attractive  from  a management  per- 
spective is  that  they  are  usually  owned  by  a 
single  entity.  Furthermore,  most  reclaimed 
mines  are  typically  not  very  productive  as  ag- 
ricultural areas.  These  factors  combined  make 
possible  the  acquisition  or  management  of 
large  grassland-dominated  habitats.  Few  such 
opportunities  currently  exist  in  the  eastern 
United  States. 

ACKNOWLEDGMENTS 

The  U.S.  Fish  and  Wildlife  Service,  the  U.S.  Geo- 
logical Survey,  and  the  Ohio  River  Valley  Ecosystem 
Group  provided  financial  support  for  this  project.  We 
are  grateful  to  the  following  individuals  for  allowing 
us  full  access  to  reclaimed  mine  properties:  A.  Eicher 
and  S.  McGarvie  of  Peabody  Coal,  R.  Ronk  of  the 
Indiana  Department  of  Natural  Resources,  M.  Krieger 
at  Universal  Mine,  and  L.  Nelson  of  the  Midwest  Coal 
Company.  We  also  thank  G.  S.  Bakken,  M.  T.  Jackson, 
and  P.  E.  Scott  for  valuable  help  and  advice.  Special 
thanks  go  to  R.  Gushee,  J.  Mozingo,  C.  Roever,  B. 
Thomas,  and  A.  Worthington  for  their  competent  field- 
work and  keen  nest-finding  abilities.  We  are  grateful 
to  M.  D.  Carey  and  two  anonymous  reviewers  for 
helpful  comments  that  improved  this  manuscript. 

LITERATURE  CITED 

Bajema,  R.  A.,  T.  L.  DeVault,  P.  E.  Scott,  and  S.  L. 
Lima.  2001.  Large  reclaimed  coal  mine  grasslands 
and  their  significance  for  Henslow’s  Sparrows  in 
the  American  Midwest.  Auk  118:422-431. 
Bajema,  R.  A.  and  S.  L.  Lima.  2001.  Landscape-level 
analyses  of  Henslow’s  Sparrow  ( Ammodramus 
henslowii ) abundance  in  large,  reclaimed  coal 
mine  grasslands.  American  Midland  Naturalist 
145:288-298. 

Bart,  J.  1977.  Impact  of  human  visitations  on  nesting 
success.  Living  Bird  16:187-192. 


Best,  L.  B.  1978.  Field  Sparrow  reproductive  success 
and  nesting  ecology.  Auk  95:9-22. 

Best,  L.  B.,  H.  Campa,  III,  K.  E.  Kemp,  R.  J.  Robel, 
M.  R.  Ryan,  J.  A.  Savidge,  H.  P.  Weeks,  Jr.,  and 
S.  R.  W interstein . 1997.  Bird  abundance  and 
nesting  in  CRP  fields  and  cropland  in  the  Mid- 
west: a regional  approach.  Wildlife  Society  Bul- 
letin 25:864-877. 

Brothers,  T.  S.  1990.  Surface-mine  grasslands.  Geo- 
graphical Review  80:209-225. 

Burhans,  D.  E.,  D.  Dearborn,  F.  R.  Thompson,  III, 
and  J.  Faaborg.  2002.  Factors  affecting  predation 
at  songbird  nests.  Journal  of  Wildlife  Management 
66:240-249. 

Clay,  K.  and  J.  Holah.  1999.  Fungal  endophyte  sym- 
biosis and  plant  diversity  in  successional  fields. 
Science  285:1742-1744. 

Clotfelter,  E.  D.  and  K.  Yasukawa.  1999.  Impact 
of  brood  parasitism  by  Brown-headed  Cowbirds 
on  Red-winged  Blackbird  reproductive  success. 
Condor  101:105-114. 

Davis,  M.  A.,  D.  W.  Peterson,  P.  B.  Reich,  M.  Cro- 
zier,  T.  Query,  E.  Mitchell,  J.  Huntington,  and 
P.  Bazakas.  2000.  Restoring  savanna  using  fire: 
impact  on  the  breeding  bird  community.  Restora- 
tion Ecology  8:30-40. 

Davis,  S.  K.  2003.  Nesting  ecology  of  mixed-grass 
prairie  songbirds  in  southern  Saskatchewan.  Wil- 
son Bulletin  115:119-130. 

DeVault,  T.  L.,  P.  E.  Scott,  R.  A.  Bajema,  and  S.  L. 
Lima.  2002.  Breeding  bird  communities  of  re- 
claimed coal-mine  grasslands  in  the  American 
Midwest.  Journal  of  Field  Ornithology  73:268- 
275. 

Donovan,  T.  M.,  F.  R.  Thompson,  III,  J.  Faaborg,  and 
J.  R.  Probst.  1995.  Reproductive  success  of  mi- 
gratory birds  in  habitat  sources  and  sinks.  Con- 
servation Biology  9:1380-1395. 

Galligan,  E.  W.,  G.  S.  Bakken,  and  S.  L.  Lima.  2003. 
Using  a thermographic  imager  to  find  nests  of 
grassland  birds.  Wildlife  Society  Bulletin  31:865- 
869. 

Gates,  J.  E.  and  L.  W.  Gysel.  1978.  Avian  nest  dis- 
persion and  fledging  success  in  field-forest  eco- 
tones.  Ecology  59:871-883. 

Hahn,  D.  C.  and  J.  S.  Hatfield.  1995.  Parasitism  at 
the  landscape  scale:  cowbirds  prefer  forest.  Con- 
servation Biology  9:1415-1424. 

Herkert,  J.  R.  1994.  The  effects  of  habitat  fragmen- 
tation on  midwestern  grassland  bird  communities. 
Ecological  Applications  4:461-471. 

Herkert,  J.  R.,  D.  L.  Reinking,  D.  A.  Wiedenfeld, 
M.  Winter,  J.  L.  Zimmerman,  W.  E.  Jensen,  E.  J. 
Finck,  R.  R.  Koford,  D.  H.  Wolfe,  S.  K.  Sher- 
rod, et  al.  2003.  Effects  of  prairie  fragmentation 
on  the  nest  success  of  breeding  birds  in  the  mid- 
continental United  States.  Conservation  Biology 
17:587-594. 

Hines,  J.  E.  and  J.  R.  Sauer.  1989.  Program  CON- 
TRAST: a general  program  for  the  analysis  of  sev- 
eral survival  or  recovery  estimates.  U.S.  Fish  and 


546 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Wildlife  Service,  General  Technical  Report  24, 
Washington,  D.C. 

Horn,  D.  J.,  R.  J.  Fletcher,  Jr.,  and  R.  R.  Koford. 
2000.  Detecting  area  sensitivity:  a comment  on 
previous  studies.  American  Midland  Naturalist 
144:28-35. 

Hughes,  J.  P.,  R.  J.  Robel,  K.  E.  Kemp,  and  J.  L. 
Zimmerman.  1999.  Effects  of  habitat  on  Dickcis- 
sel  abundance  and  nest  success  in  Conservation 
Reserve  Program  fields  in  Kansas.  Journal  of 
Wildlife  Management  63:523-529. 

Hunter,  W.  C.,  D.  A.  Buehler,  R.  A.  Canterbury,  J. 
L.  Confer,  and  P.  B.  Hamel.  2001.  Conservation 
of  disturbance-dependent  birds  in  eastern  North 
America.  Wildlife  Society  Bulletin  29:440-455. 

Ingold,  D.  J.  2002.  Use  of  reclaimed  stripmine  by 
grassland  nesting  birds  in  east-central  Ohio.  Ohio 
Journal  of  Science  102:56-62. 

Johnson,  D.  H.  and  L.  D.  Igl.  2001.  Area  require- 
ments of  grassland  birds:  a regional  perspective. 
Auk  118:24-34. 

Johnson,  R.  G.  and  S.  A.  Temple.  1990.  Nest  preda- 
tion and  brood  parasitism  of  tallgrass  prairie  birds. 
Journal  of  Wildlife  Management  54:106-1 1 1 . 

Kershner,  E.  L.  2001.  Conservation  of  grassland  birds 
in  an  agricultural  landscape:  the  importance  of 
habitat  availability  and  demography.  Ph.D.  disser- 
tation, University  of  Illinois,  Urbana-Champaign. 

Kershner,  E.  L.  and  E.  K.  Bollinger.  1998.  Low 
incidence  of  cowbird  parasitism  of  grassland  birds 
on  Illinois  airports.  Transactions  of  the  Illinois 
Academy  of  Science  91:103-107. 

Maier,  T.  J.  and  R.  M.  DeGraaf.  2000.  Rhodamine- 
injected  eggs  to  photographically  identify  small 
nest-predators.  Journal  of  Field  Ornithology  71: 
694-701. 

Martin,  T.  E.  and  G.  R.  Geupel.  1993.  Nest-monitor- 
ing plots:  methods  for  locating  nests  and  moni- 
toring success.  Journal  of  Field  Ornithology  64: 
507-519. 

Mayfield,  H.  1961.  Nesting  success  calculated  from 
exposure.  Wilson  Bulletin  73:255-261. 

Mayfield,  H.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

Monroe,  M.  S.  and  G.  Ritchison.  2005.  Breeding  bi- 
ology of  Henslow’s  Sparrows  on  reclaimed  coal 
mine  grasslands  in  Kentucky.  Journal  of  Field  Or- 
nithology 76:143-149. 

Norusis,  M.  J.  1993.  SPSS  for  Windows:  base  system 
user’s  guide,  release  6.0.  SPSS  Inc.,  Chicago,  Il- 
linois. 

Paton,  P.  W.  C.  1994.  The  effect  of  edge  on  avian  nest 
success:  how  strong  is  the  evidence?  Conservation 
Biology  8:17-26. 

Patterson,  M.  P.  and  L.  B.  Best.  1996.  Bird  abun- 
dance and  nesting  success  in  Iowa  CRP  fields:  the 
importance  of  vegetation  structure  and  composi- 
tion. American  Midland  Naturalist  135:153-167. 

Picozzi,  N.  1975.  Crow  predation  on  marked  nests. 
Journal  of  Wildlife  Management  39:151-155. 


Pietz,  P.  J.  and  D.  A.  Granfors.  2000.  Identifying 
predators  and  fates  of  grassland  passerine  nests 
using  miniature  video  cameras.  Journal  of  Wildlife 
Management  64:71-87. 

Pribil,  S.  1998.  Reproductive  success  is  a misleading 
indicator  of  nest-site  preferences  in  the  Red- 
winged Blackbird.  Canadian  Journal  of  Zoology 
76:2227-2234. 

Robb,  J.  R.,  S.  A.  Miller,  T.  Vanosdol-Lewis,  and  J. 
P.  Lewis.  1998.  Productivity  of  interior  forest  and 
grassland  birds  on  Jefferson  Proving  Ground: 
1998  annual  report.  U.S.  Fish  and  Wildlife  Ser- 
vice, Jefferson  Proving  Ground,  Madison,  Indiana. 

Robinson,  S.  K.  and  J.  L.  Herkert.  1997.  Cowbird 
parasitism  in  different  habitats.  Illinois  Natural 
History  Reports  348:2-3. 

Robinson,  S.  K.,  F.  R.  Thompson,  III,  T.  M.  Donovan, 
D.  R.  Whitehead,  and  J.  Faaborg.  1995.  Re- 
gional forest  fragmentation  and  the  nesting  suc- 
cess of  migratory  birds.  Science  267:1987-1990. 

Sauer,  J.  R.  and  B.  K.  Williams.  1989.  Generalized 
procedures  for  testing  hypotheses  about  survival 
or  recovery  rates.  Journal  of  Wildlife  Management 
53:137-142. 

Scott,  P.  E.,  T.  L.  DeVault,  R.  A.  Bajema,  and  S.  L. 
Lima.  2002.  Grassland  vegetation  and  bird  abun- 
dances on  reclaimed  midwestern  coal  mines. 
Wildlife  Society  Bulletin  30:1006-1014. 

Scott,  P.  E.  and  S.  L.  Lima.  2004.  Exotic  grasslands 
on  reclaimed  midwestern  coal  mines:  an  ornitho- 
logical perspective.  Weed  Technology  18:151 8— 
1521. 

Thompson,  F.  R.,  Ill,  W.  Dijak,  and  D.  E.  Burhans. 
1999.  Video  identification  of  predators  at  songbird 
nests  in  old  fields.  Auk  116:259-264. 

Walk,  J.  W.  2001.  Nesting  ecology  of  grassland  birds 
in  an  agricultural  landscape.  Ph.D.  dissertation. 
University  of  Illinois,  Urbana-Champaign. 

Walk,  J.  W.  and  R.  E.  Warner.  1999.  Effects  of  hab- 
itat area  on  the  occurrence  of  grassland  birds  in 
Illinois.  American  Midland  Naturalist  141:339- 
344. 

Warner,  R.  E.  1994.  Agricultural  land  use  and  grass- 
land habitat  in  Illinois:  future  shock  for  midwest- 
ern birds?  Conservation  Biology  8:147-156. 

Winter,  M.  1999.  Nesting  biology  of  Dickcissels  and 
Henslow’s  Sparrows  in  southwestern  Missouri 
prairie  fragments.  Wilson  Bulletin  11:515-527. 

Winter,  M.  and  J.  Faaborg.  1999.  Patterns  of  area 
sensitivity  in  grassland-nesting  birds.  Conserva- 
tion Biology  13:1424-1436. 

Winter,  M.,  D.  H.  Johnson,  and  J.  Faaborg.  2000. 
Evidence  for  edge  effects  on  multiple  levels  in 
tallgrass  prairie.  Condor  102:256-266. 

Winter,  M.,  D.  H.  Johnson,  J.  A.  Shaffer,  and  W.  D. 
Svedarsky.  2004.  Nesting  biology  of  three  grass- 
land passerines  in  the  northern  tallgrass  prairie. 
Wilson  Bulletin  116:211-223. 

Yasukawa,  K.  and  W.  A.  Searcy.  1995.  Red-winged 
Blackbird  (Agelaius  phoeniceus ).  The  Birds  of 
North  America,  no.  184. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):547-55 1 , 2006 


DIFFERENTIAL  TIMING  OF  WILSON’S  WARBLER  MIGRATION 

IN  ALASKA 

ANNA-MARIE  BENSON,1 2  BRAD  A.  ANDRES,25  W.  N.  JOHNSON,3 
SUSAN  SAVAGE,4 5 6  AND  SUSAN  M.  SHARBAUGH1  6 


ABSTRACT. — We  examined  age-  and  sex-related  differences  in  the  timing  of  Wilson’s  Warbler  ( Wilsonia 
pusilla  pileolata ) migration  at  four  locations  in  Alaska:  Fairbanks,  Tok,  Mother  Goose  Lake,  and  Yakutat.  We 
captured  Wilson’s  Warblers  with  mist  nets  for  > 5 years  during  spring  (northbound)  and  autumn  (southbound) 
migration.  In  spring,  males  passed  through  our  two  northernmost  sites — Tok  and  Fairbanks — earlier  than  females. 
During  autumn,  timing  of  adult  migration  did  not  differ  by  sex,  but  immatures  passed  through  earlier  than  adults 
at  all  four  sites.  During  previous  studies  of  autumn  passage  sampled  at  lower  latitudes,  the  lack  of  age-related 
differences  in  migration  timing  could  be  attributed  to  adults  migrating  faster  than  immatures  (i.e.,  if  immatures 
from  higher  latitudes  began  migration  earlier  than  the  adults,  then  the  adults  may  have  caught  up  to  them  at 
lower  latitudes)  or  to  the  mixing  of  breeding  populations  from  different  locales.  Autumn  migration  of  adults  and 
immatures  netted  at  our  two  southernmost  sites,  both  coastal  locations,  preceded  migration  at  our  two  interior 
sites.  These  site-specific  differences  in  the  timing  of  autumn  migration  are  likely  the  result  of  our  coastal  stations 
sampling  birds  that  breed  farther  south  and  arrive  earlier  than  birds  breeding  in  more  northerly  regions  of  Alaska 
(and  sampled  at  our  interior  stations).  Early-arriving  populations  are  likely  able  to  complete  their  breeding  season 
activities  earlier  and,  subsequently,  initiate  their  autumn  migration  earlier.  Received  29  July  2005,  accepted  5 
May  2006. 


Age-  or  sex-related  differences  in  timing  of 
migrant  passage  have  been  documented  at 
several  locations  in  North  America  (see  re- 
views by  Gauthreaux  1982,  Woodrey  2000). 
Analyses  of  between-sex  variation  in  the  tim- 
ing of  spring  migration  have  shown  that  males 
of  several  North  American  passerine  species 
migrate  prior  to  females  (Francis  and  Cooke 
1986,  Yunick  1988,  Otahal  1995,  Yong  et  al. 
1998,  Swanson  et  al.  1999).  Studies  docu- 
menting age-class  differences  in  the  timing  of 
autumn  migration  have  revealed  varied  pat- 
terns. Immature  Wilson’s  Warblers  ( Wilsonia 
pusilla ) preceded  adults  by  9 days  in  south- 
western Idaho  (Carlisle  et  al.  2005a);  10  days 
at  Yakutat,  Alaska  (Andres  et  al.  2005);  and 
13  days  at  Fairbanks,  Alaska  (Benson  and 


1 Alaska  Bird  Observatory,  P.O.  Box  80505,  Fair- 
banks, AK  99708,  USA. 

2 U.S.  Fish  and  Wildlife  Service,  Migratory  Bird 
Management,  1011  E.  Tudor  Rd.,  Anchorage,  AK 
99503,  USA. 

3 Tetlin  National  Wildlife  Refuge,  P.O.  Box  779, 
Tok,  AK  99780,  USA. 

4 Alaska  Peninsula/Becharof  National  Wildlife  Ref- 
uge Complex,  P.O.  Box  277,  King  Salmon,  AK  99613, 
USA. 

5 Current  address:  U.S.  Fish  and  Wildlife  Service, 
P.O.  Box  25486,  DFC,  Denver,  CO  80225,  USA. 

6 Corresponding  author;  e-mail: 
ssharbaugh@alaskabird.org 


Winker  2001).  The  autumn  migration  timing 
of  adult  and  immature  Wilson’s  Warblers  did 
not  differ  in  South  Dakota  (Dean  et  al.  2004) 
or  in  the  riparian  forest  of  the  middle  Rio 
Grande  in  New  Mexico  (Yong  et  al.  1998). 

We  selected  the  Wilson’s  Warbler  to  ex- 
amine differential  migration  timing  because  it 
is  a relatively  abundant  migrant  and  is  sexu- 
ally dichromatic.  Wilson’s  Warblers  breed 
throughout  Alaska  and  winter  in  the  southern 
United  States,  Mexico,  and  Central  America 
(Ammon  and  Gilbert  1999).  W.  p.  pileolata  is 
the  only  subspecies  known  to  range  into  Alas- 
ka (American  Ornithologists’  Union  1957, 
Gibson  and  Kessel  1997). 

The  geographic  location  of  Alaska,  relative 
to  the  continental  landmass,  provides  an  op- 
portunity to  study  the  passage  of  migrants 
near  where  they  terminate  their  spring  migra- 
tion and  initiate  their  autumn  migration.  Our 
objectives  were  to  use  data  from  four  widely 
dispersed  migration  banding  stations  in  Alas- 
ka to  examine  differences  in  the  timing  of 
Wilson’s  Warbler  migration.  Our  specific  ob- 
jectives were  to  determine  (1)  between-sex 
differences  in  the  timing  of  spring  migration, 
(2)  between-age  differences  in  the  timing  of 
autumn  migration,  and  (3)  among-site  differ- 
ences in  the  timing  of  autumn  migration. 


547 


548  THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


FIG.  1 . Location  of  four  migration  monitoring  sta- 
tions in  Alaska:  (1)  Fairbanks,  (2)  Tok,  (3)  Mother 
Goose  Lake,  and  (4)  Yakutat,  1992-2000. 


METHODS 

Study  sites. — We  analyzed  data  from  four 
migration  stations  operated  for  >5  years  dur- 
ing 1992-2000.  Fairbanks  and  Tok  were  op- 
erated in  spring  and  autumn,  and  Yakutat  and 
Mother  Goose  Lake  were  operated  only  in  the 
autumn.  The  Fairbanks  banding  station,  op- 
erated by  the  Alaska  Bird  Observatory  on  the 
Creamer’s  Field  Migratory  Waterfowl  Refuge 
(64°  50'  N,  147°  50'  W),  and  the  Tok  banding 
station  (63°  22' N,  143°  12'  W),  operated  by 
the  Tetlin  National  Wildlife  Refuge,  are  lo- 
cated in  interior  Alaska  in  the  Tanana  River 
Valley  (Fig.  1).  The  Yakutat  station,  operated 


by  the  U.S.  Fish  and  Wildlife  Service,  is  on 
the  Gulf  of  Alaska  coastline  —300  km  north- 
west of  Juneau  (59°  30' N,  139°  40' W;  Fig. 
1).  The  Mother  Goose  Lake  station  (57°  11' 
N,  157°  15'  W),  operated  by  the  Alaska  Pen- 
insula/Becharof  National  Wildlife  Refuge 
Complex,  lies  west  of  the  Aleutian  Mountain 
Range  in  southwestern  Alaska,  —165  km 
southwest  of  King  Salmon  (Fig.  1). 

We  used  2.6-  X 12-m  nets  with  30-mm 
mesh  at  all  stations;  specific  operation  details 
are  provided  in  Table  1.  The  netting  period  at 
all  stations  spanned  the  entire  duration  of  Wil- 
son’s Warbler  migration.  Our  studies  were  de- 
signed to  capture  a wide  suite  of  passerine 
species,  many  of  which  pass  through  study 
sites  earlier  and  depart  later  than  Wilson’s 
Warblers. 

Ageing  and  sexing. — At  all  locations  during 
fall  migration,  age  was  determined  by  degree 
of  skull  ossification  (Pyle  1997).  During 
spring  at  Fairbanks  and  Tok,  and  during  fall 
at  Yakutat  and  Mother  Goose  Lake,  birds  were 
sexed  by  plumage  and  morphometric  charac- 
teristics (Pyle  1997).  During  autumn  at  Fair- 
banks and  Tok,  birds  were  sexed  using  the  fol- 
lowing discriminant  function,  developed  from 
known-age  Alaskan  birds  (Weicker  and  Wink- 
er 2002),  whereby  96%  of  known-age  birds 
were  classified  correctly: 

D = 0.9189  cap  category 
+ 0.1800  cap  length 
+ 0.0977  tail  length 
+ 0.0938  wing  chord 
- 13.9426, 

where  D is  the  discriminant  function,  cap  cat- 
egory separates  caps  into  one  of  four  classes 
(ranging  from  solid  olive-green  to  solid 


TABLE  1.  Spring  and  autumn  mist-netting  efforts  to  capture  migrant  Wilson’s  Warblers  at  four  banding 
stations  in  Alaska,  1992-2000. 


Station 

Season 

Years 

Period 

No.  nets 

Time 

Total  net  hr 

Fairbanks 

Spring 

1992-2000 

25  Apr- 15  Jun 

22-50 

06:00-13:00 

81,736 

Autumn 

1992-2000 

15  Jul-30  Sep 

22-50 

sunrise  4-  7 hr 

114,053 

Tok 

Spring 

1994-1998 

late  Apr— early  Jun 

20-24 

sunrise  + 6 hr 

22,707 

Autumn 

1993-2000 

early  Aug-late  Sep 

20-24 

sunrise  + 6 hr 

49,322 

Mother  Goose  Lake 

Autumn 

1994-2000 

1 Aug-22  Sep 

10-15 

sunrise  + 6 hra 

11,018 

Yakutat 

Autumn 

1994-1999 

1 Aug-5  Oct 

10-15 

sunrise  + 6 hr 

23,256 

a Nets  were  opened  0.5  hr  after  sunrise. 


Benson  et  al.  • WILSON’S  WARBLER  MIGRATION  IN  ALASKA 


549 


TABLE  2.  Median  passage  dates  of  Wilson’s  Warbler  at  four  locations  in  Alaska:  Fairbanks  (1992-2000), 
Tok  (1993-2000),  Yakutat  (1994-1999),  and  Mother  Goose  Lake  (1994-2000). 

Season  Site 

Adult  between-sex  differences 

Between- 

age-class  differences 

Males 

Females 

z 

Immatures 

Adults 

z 

Date8 

» 

Date3 

n 

Date3 

Date3 

n 

Spring 

Fairbanks 

143 

105 

148 

143 

4.40**b 

Tok 

142 

771 

150 

450 

18.33** 

Autumn 

Fairbanks 

243 

58 

253 

28 

1.56 

230 

1,009 

243 

105 

9.52** 

Tok 

242 

195 

240 

36 

1.29 

230 

1,185 

241 

616 

17.71** 

Yakutat 

228 

73 

228 

38 

0.70 

222 

374 

228 

111 

5.60** 

Mother  Goose  234 

160 

234 

50 

0.32 

225 

10,481 

235 

287 

17.29** 

Lake 

3 Median  Julian  date  of  passage. 
b Double  asterisk  indicates  P < 0.001. 


black),  and  cap  length  is  the  extent  of  black 
feathers  from  the  front  to  the  back  of  the  head. 
For  our  analyses,  we  included  only  records 
with  >75%  probability  that  individuals  were 
sexed  correctly. 

Definition  of  migrants. — In  analyses  for  all 
sites,  we  included  only  first  captures  of  birds. 
Based  on  two  criteria,  we  eliminated  individ- 
uals that  may  not  have  been  migrating  at  the 
time  of  capture:  (1)  birds  recaptured  >7  days 
after  first  capture  and  (2)  locally  fledged  birds 
(i.e.,  birds  retaining  >60%  of  their  juvenal 
plumage).  We  did  not  specifically  remove  fe- 
males with  brood  patches  because  this  could 
potentially  bias  the  retention  of  males  and 
elimination  of  females,  and  affect  our  be- 
tween-sex  comparisons.  No  females  with 
brood  patches  were  captured  at  Fairbanks, 
Tok,  or  Yakutat,  and  only  nine  such  individ- 
uals were  captured  and  included  in  the  data 
set  from  Mother  Goose  Lake.  It  is  possible 
that  birds  not  migrating  at  the  time  of  capture 
were  included  in  our  analyses,  resulting  in  an 
early-biased  median  date  of  autumn  passage. 
However,  considering  the  relatively  few  birds 
netted  in  summer  compared  to  the  vast  num- 
bers captured  during  the  brief  and  intense  mi- 
gration pulse,  we  suspect  the  numbers  of 
breeding  birds  included  in  these  analyses  were 
small.  If  some  non-migratory  birds  were  in- 
cluded in  these  analyses,  they  likely  affected 
the  data  from  each  station  and,  therefore, 
should  not  have  affected  our  among-site  com- 
parisons. 

Data  analysis. — We  tested  for  age-,  sex-. 


and  site-related  differences  in  median  passage 
dates  by  using  Mann-Whitney  U- tests.  For 
two  reasons,  we  did  not  standardize  by  unit  of 
netting  effort.  First,  standardizing  by  unit  of 
effort  can  artificially  inflate  or  deflate  sample 
sizes,  which,  in  turn,  can  affect  the  power  of 
a test  (see  examples  in  Benson  and  Winker 
2001).  Second,  standardizing  by  unit  of  effort 
was  not  necessary  in  these  analyses  because 
even  in  Fairbanks,  where  there  were  some  net- 
ting-effort  inconsistencies  in  earlier  years,  net 
hr  over  a given  season  had  a uniform  distri- 
bution when  all  years  were  combined  (see 
Benson  and  Winker  2001). 

RESULTS 

During  spring  migration,  males  preceded 
females  by  5 days  at  Fairbanks  (Z  = 4.40,  n 
= 248,  P < 0.001;  Table  2)  and  by  8 days  at 
Tok  (Z  = 18.33,  n = 1,221,  P < 0.001;  Table 
2).  In  autumn,  we  found  no  between-sex  dif- 
ference in  the  timing  of  adult  migration  at  any 
location  (Table  2).  However,  immatures  con- 
sistently preceded  adults  at  all  locations:  by 
13  days  at  Fairbanks  (Z  = 9.52,  n = 1,114,  P 
< 0.001),  11  days  at  Tok  (Z  = 17.71,  n = 
1,801,  P < 0.001),  6 days  at  Yakutat  (Z  = 
5.60,  n = 485,  P < 0.001),  and  10  days  at 
Mother  Goose  Lake  (Z  = 17.29,  n = 10,768, 
P < 0.001;  Table  2).  Passage  of  both  adults 
and  immatures  was  significantly  earlier  at  the 
two  coastal  sites  than  at  the  two  interior  sites 
(all  Z > 7.84,  P < 0.001).  Wilson’s  Warblers 
also  passed  through  Yakutat  significantly  ear- 
lier than  they  did  at  Mother  Goose  Lake  (all 


550 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  4,  December  2006 


Z > 5.23,  P < 0.001).  There  was  no  signifi- 
cant difference  between  the  passage  dates  at 
Fairbanks  and  Tok. 

DISCUSSION 

Basic  patterns  in  the  timing  of  migration 
were  similar  at  all  four  migration  stations  in 
Alaska.  In  spring,  the  earlier  passage  of  male 
Wilson’s  Warblers,  compared  with  females, 
was  similar  to  results  found  by  Francis  and 
Cooke  (1986)  and  Yong  et  al.  (1998).  These 
results  were  expected  because  of  the  selective 
pressures  that  favor  males  to  arrive  early  and 
obtain  a high-quality  territory,  whereas  fe- 
males likely  benefit  by  arriving  later  when  re- 
sources are  more  predictable  (see  review  by 
Francis  and  Cooke  1986). 

Immature  Wilson’s  Warblers  migrate  south- 
ward from  Alaska  significantly  earlier  than 
adults,  most  likely  because  they  do  not  un- 
dergo the  full  prebasic  molt  that  adults  must 
complete  before  migration  (Dwight  1900). 
Adults,  however,  compensate  for  their  later 
migration  by  migrating  with  greater  mass  and 
fat  stores  (Andres  et  al.  2005,  Benson  and 
Winker  2005).  The  differences  in  age-related 
migration  timing  among  Wilson  Warblers  in 
fall  may  not  be  detectable  at  lower  latitudes 
(e.g.,  Yong  et  al.  1998,  Dean  et  al.  2004)  be- 
cause immatures  may  migrate  at  slower  rates 
due  to  their  inability  to  forage  as  efficiently  as 
adults.  During  fall  migration  in  New  Mexico, 
immature  Wilson’s  Warblers  had  lower  fat 
scores  than  adults,  but  age-class  differences  in 
mass  and  rates  of  mass  gain  have  not  been 
detected  at  other  locations  for  this  species 
(Jones  et  al.  2002,  Carlisle  et  al.  2005b). 

The  among-site  differences  in  median  dates 
of  autumn  passage  were  not  surprising.  The 
onset  of  winter  can  vary  substantially 
throughout  the  large  and  mountainous  state  of 
Alaska,  and  populations  originating  from  re- 
gions with  briefer  summers  are  likely  to  de- 
part earlier.  Stopover  ecology  of  Wilson’s 
Warblers  is  also  influenced  by  habitat  (Hutto 
1985,  Skagen  et  al.  1998),  but  we  did  not 
measure  the  effect  of  this  variable  at  the  lo- 
cations studied. 

We  currently  lack  sufficient  information  for 
defining  the  breeding  ranges  of  populations 
sampled  at  our  four  study  sites;  however,  we 
hypothesized  that  samples  from  interior  sites 
represented  different  populations  than  those 


sampled  at  coastal  sites  because  large  moun- 
tain ranges  separate  the  southern  coast  of 
Alaska  from  the  state’s  interior.  Isotopic  ratios 
of  Wilson’s  Warblers  breeding  in  western 
North  America  indicate  that  coastal  breeders 
overwinter  in  western  Mexico  and  those  that 
breed  farther  inland  and  at  higher  elevations 
overwinter  in  eastern  Mexico  (Clegg  et  al. 
2003).  However,  a few  recoveries  of  birds 
banded  at  Mother  Goose  Lake  indicate  that 
birds  occurring  at  that  site  may  represent  pop- 
ulations that  winter  in  both  eastern  and  west- 
ern locations. 

ACKNOWLEDGMENTS 

Funding  for  this  project  was  provided  by  the  Alaska 
Bird  Observatory,  Alaska  Department  of  Fish  and 
Game,  Alaska  Peninsula/Becharof  National  Wildlife 
Refuge,  Earthwatch  Institute,  Tetlin  National  Wildlife 
Refuge,  and  U.S.  Fish  and  Wildlife  Service  (Region  7) 
Migratory  Bird  Management,  and  Alaska  Department 
of  Natural  Resources.  These  studies  would  not  have 
been  possible  without  the  many  staff  and  volunteers 
that  contributed  countless  hours  capturing  and  banding 
birds.  We  especially  thank  B.  Browne,  T.  J.  Doyle,  A. 
R.  Ajmi,  C.  Adler,  T.  Burke,  K.  Convery,  C.  R.  Davis, 
D.  Dewhurst,  N.  DeWitt,  R.  C.  Egan,  T.  Eskelin,  J. 
Foster,  N.  Gregory,  R.  I.  Frey,  J.  Klima,  A.  L.  Lan- 
caster, K.  W.  Larson,  M.  Margulies,  R.  C.  Means,  J. 
Melton,  R.  Moore,  H.  Moore,  R.  K.  Papish,  T.  H.  Pog- 
son,  G.  Ruhl,  M.  Sardy,  D.  Shaw,  K.  M.  Sowl,  S.  K. 
Springer,  H.  K.  Timm,  L.  Wells,  and  D.  L.  Williams. 
We  also  thank  G.  Collins  for  preparing  the  map  and  P. 
J.  Heglund  for  reviewing  an  earlier  version  of  the  man- 
uscript. Many  thanks  to  D.  L.  Swanson  and  an  anon- 
ymous reviewer  for  comments  that  improved  the  man- 
uscript. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1957.  Check-list 
of  North  American  birds,  5th  ed.  American  Or- 
nithologists’ Union,  Washington,  D.C. 

Ammon,  E.  M.  and  W.  M.  Gilbert.  1999.  Wilson’s 
Warbler  ( Wilsonia  pusilla ).  The  Birds  of  North 
America,  no.  478. 

Andres,  B.  A.,  B.  T.  Browne,  and  D.  L.  Brann.  2005. 
Composition,  abundance,  and  timing  of  post- 
breeding landbirds  at  Yakutat,  Alaska.  Wilson 
Bulletin  117:270-279. 

Benson,  A.  M.  and  K.  Winker.  2001.  Timing  of 
breeding  range  occupancy  among  high-latitude 
passerine  migrants.  Auk  118:513-519. 

Benson,  A.  M.  and  K.  Winker.  2005.  Fat  deposition 
strategies  among  high-latitude  passerine  migrants. 
Auk  122:544-557. 

Carlisle,  J.  D.,  G.  S.  Kaltenecker,  and  D.  L.  Swan- 
son. 2005a.  Molt  strategies  and  age  differences  in 


Benson  et  al.  • WILSON’S  WARBLER  MIGRATION  IN  ALASKA 


551 


migration  timing  among  autumn  Iandbird  migrants 
in  Southwestern  Idaho.  Auk  122:1070-1085. 

Carlisle,  J.  D.,  G.  S.  Kaltenecker,  and  D.  L.  Swan- 
son. 2005b.  The  stopover  ecology  of  autumn  land- 
bird  migrants  in  the  Boise  foothills  of  Southwest- 
ern Idaho.  Condor  107:244-258. 

Clegg,  S.  M.,  J.  F.  Kelly,  M.  Kimura,  and  T.  B. 
Smith.  2003.  Combining  genetic  markers  and  sta- 
ble isotopes  to  reveal  population  connectivity  and 
migration  patterns  in  a Neotropical  migrant,  Wil- 
son’s Warbler  ( Wilsonia  pusillo ).  Molecular  Ecol- 
ogy 12:819-830. 

Dean,  K.  L.,  H.  A.  Carlisle,  and  D.  L.  Swanson. 
2004.  Age  structure  of  Neotropical  migrants  dur- 
ing fall  migration  in  South  Dakota:  is  the  northern 
Great  Plains  region  an  inland  “coast”?  Wilson 
Bulletin  116:295-303. 

Dwight,  J.,  Jr.  1900.  The  sequence  of  plumages  and 
moults  of  passerine  birds  of  New  York.  Academy 
of  Sciences  13:73-360. 

Francis,  C.  M.  and  F.  Cooke.  1986.  Differential  tim- 
ing of  spring  migration  in  wood  warblers  (Paru- 
linae).  Auk  103:548-556. 

Gauthreaux,  S.  A.,  Jr.  1982.  The  ecology  and  evo- 
lution of  avian  migration  systems.  Avian  Biology 
6:93-168. 

Gibson,  D.  D.  and  B.  Kessel.  1997.  Inventory  of  the 
species  and  subspecies  of  Alaska  birds.  Western 
Birds  28:45-95. 

Hutto,  R.  L.  1985.  Seasonal  changes  in  the  habitat 
distribution  of  insectivorous  birds  in  southeastern 
Arizona:  competition  mediated?  Auk  102:120- 
132. 


Jones,  J.,  C.  M.  Francis,  M.  Drew,  S.  Fuller,  and 
M.  W.  S.  Ng.  2002.  Age-related  differences  in 
body  mass  and  rates  of  mass  gain  of  passerines 
during  autumn  migratory  stopover.  Condor  104: 
49-58. 

Otahal,  C.  D.  1995.  Sexual  differences  in  Wilson’s 
Warbler  migration.  Journal  of  Field  Ornithology 
66:60-69. 

Pyle,  P.  P.  1997.  Identification  guide  to  North  Ameri- 
can birds.  Slate  Creek  Press,  Bolinas,  California. 

Skagen,  S.  K.,  C.  P.  Melcher,  W.  H.  Howe,  and  F.  L. 
Knopf.  1998.  Comparative  use  of  riparian  corri- 
dors and  oases  by  migrating  birds  in  Southeast 
Arizona.  Conservation  Biology  12:896-909. 

Swanson,  D.  L.,  E.  T.  Liknes,  and  K.  L.  Dean.  1999. 
Differences  in  migratory  timing  and  energetic 
condition  among  sex/age  classes  in  migrant  Ruby- 
crowned  Kinglets.  Wilson  Bulletin  111:61-69. 

Weicker,  J.  J.  and  K.  Winker.  2002.  Sexual  dimor- 
phism in  birds  from  southern  Veracruz,  Mexico, 
and  other  localities.  III.  Wilson’s  Warbler  ( Wilson- 
ia pusilla).  Journal  of  Field  Ornithology  73:62- 
69. 

Woodrey,  M.  S.  2000.  Age-dependent  aspects  of  stop- 
over biology  of  passerine  migrants.  Studies  in 
Avian  Biology  20:43-52. 

Yong,  W.,  D.  M.  Finch,  F.  R.  Moore,  and  J.  F.  Kelly. 
1998.  Stopover  ecology  and  habitat  use  of  migra- 
tory Wilson’s  Warblers.  Auk  115:829-842. 

Yunick,  R.  P.  1988.  Differential  spring  migration  of 
Dark-eyed  Juncos.  Journal  of  Field  Ornithology 
59:314-320. 


The  Wilson  Journal  of  Ornithology  1 18(4):552— 557,  2006 


NESTING  SUCCESS  OF  WESTERN  BLUEBIRDS 
c SIALIA  MEXICANA)  USING  NEST  BOXES  IN  VINEYARD  AND 
OAK-SAVANNAH  HABITATS  OF  CALIFORNIA 

CRAIG  M.  FIEHLER,1  24  WILLIAM  D.  TIETJE,1 2 3 4 *  AND  WILLIAM  R.  FIELDS13 


ABSTRACT. — Loss  of  oak  woodlands  to  vineyard  development  in  California  is  a growing  concern  to  con- 
servationists. Analyzing  breeding  performance  of  birds  that  nest  in  and  around  vineyards  versus  those  that  nest 
in  nearby  native  habitat  can  provide  information  on  the  suitability  of  vineyard  environments  to  birds.  We  placed 
predator-protected  nest  boxes  in  vineyard  and  oak-savannah  habitats  and  monitored  nest-box  occupancy,  nesting 
success,  and  life  history  characteristics  of  Western  Bluebirds  ( Sialia  mexicana ) that  used  the  boxes.  Western 
Bluebirds  were  common  occupants  in  both  habitats,  occupying  >50%  of  available  nest  boxes.  Analysis  using 
program  MARK  revealed  that  nest  survival  was  not  associated  with  habitat  type;  however,  clutch  size  was 
greater  and  nests  were  initiated  earlier  in  vineyard  than  in  oak-savannah  habitat.  Our  results  suggest  that  when 
naturally  occurring  nest  sites  are  limiting,  vineyards  could  be  converted  to  good  breeding  habitat  for  Western 
Bluebirds  with  the  addition  of  nest  boxes.  Nest  boxes,  however,  should  not  be  viewed  as  a remedy  for  the 
chronic  problem  of  habitat  loss  and  degradation.  Received  27  June  2005,  accepted  5 May  2006. 


The  loss  of  oak  woodland  habitat  to  vine- 
yard expansion  is  a growing  concern  in  Cali- 
fornia (Zack  2002).  More  than  100  bird  spe- 
cies breed  in  California’s  oak  woodlands  (Ver- 
ner  1980),  making  the  loss  and  degradation  of 
this  habitat  particularly  problematic.  In  San 
Luis  Obispo  County,  California,  land  used  for 
viticulture  increased  from  4,008  to  10,851  ha 
between  1996  and  2000  (Mummert  et  al. 
2002).  Conservationists  generally  view  vine- 
yards as  sub-optimal  habitat  for  birds  due  to 
the  potential  impacts  of  pesticides  and  herbi- 
cides, habitat  fragmentation,  attraction  of  non- 
native bird  species  and  predators,  loss  of  wild- 
life shelter  and  forage,  and  changes  to  the  na- 
tive plant  community.  The  ecological  conse- 
quences of  this  large-scale  habitat  conversion, 
however,  are  not  well  understood. 

The  addition  of  nest  boxes  has  been  found 
to  augment  nesting  success  and  breeding  den- 
sities of  secondary  cavity-nesting  bird 
(SCNB)  species  in  altered  habitats  (Brawn  and 
Baida  1988,  Twedt  and  Henne-Kerr  2001, 
LeClerc  et  al.  2005).  In  golf  course  habitats. 


1 Dept,  of  Environmental  Science.  Policy,  and  Man- 
agement, 145  Mulford  Hall,  Univ.  of  California, 
Berkeley.  CA  94720,  USA. 

2 Current  address:  California  State  Univ.,  Stanislaus, 
Endangered  Species  Recovery  Program,  PO.  Box 
9622,  Bakersfield,  CA  93389,  USA. 

3 Current  address:  North  Carolina  State  Univ.,  Dept, 
of  Zoology,  Box  7617,  Raleigh.  NC  27695,  USA. 

4 Corresponding  author;  e-mail: 

cfiehler@esrp.csustan.edu 


Le  Clerc  et  al.  (2005)  found  that  nest  boxes 
provide  high-quality  nesting  habitat  for  East- 
ern Bluebirds  ( Sialia  sialis ).  Little  is  known, 
however,  about  the  nesting  success  of  SCNB 
species  that  breed  in  vineyards  compared  to 
those  that  breed  in  native  oak  woodland,  and 
it  is  unknown  whether  vineyards  that  feature 
nest  boxes  provide  adequate  breeding  habitat 
for  the  closely  related  Western  Bluebird  ( Sia- 
lia mexicana ).  The  main  objective  of  our 
study  was  to  compare  breeding  performance 
and  life  history  characteristics  of  Western 
Bluebirds  using  nest  boxes  in  a minimum-im- 
pact vineyard  with  bluebirds  using  nest  boxes 
in  native  oak-woodland  habitat. 

METHODS 

Study  site  and  study  species. — We  studied 
Western  Bluebirds  on  the  Santa  Margarita 
Ranch,  approximately  25  km  north  of  San 
Luis  Obispo  in  central  coastal  California,  dur- 
ing the  breeding  seasons  of  2003  and  2004. 
This  privately  owned,  5,700-ha  property  sur- 
rounding the  town  of  Santa  Margarita  (35° 
23.39' N,  120°  36.55' W)  features  a working 
cattle  operation  and  1,000  acres  comprising 
the  Cuesta  Ridge  Vineyard.  The  dominant  tree 
species  on  the  study  area  are  valley  oak 
( Quercus  lobata ),  blue  oak  ( Q . douglasii), 
coast  live  oak  ( Q . agrifolia ),  California  foot- 
hill pine  ( Pinus  sabiniana ),  and  willow  ( Salix 
spp.).  The  understory  is  predominantly  open 
and  consists  primarily  of  annual  grasses  and 
forbs,  including  ryegrass  ( Lolium  spp.),  wild 


552 


Fiehler  et  al.  • WESTERN  BLUEBIRD  NEST  BOX  USE 


553 


oat  ( Avena  spp.),  brome  ( Bromus  spp.),  milk- 
weed ( Asclepias  spp.),  and  exotic  weeds  such 
as  star-thistle  ( Centaurea  spp.)  and  other  this- 
tles ( Cirsium  spp.).  Unlike  typical  California 
vineyards,  which  comprise  large,  contiguous 
tracts  of  trellised  vines,  the  Cuesta  Ridge 
Vineyard  is  a minimum-impact  vineyard  char- 
acterized by  smaller  planted  areas  that  follow 
contours  of  the  surrounding  hills  and  the  re- 
tention of  relict  oak  trees  ( Quercus  spp.)  in, 
and  adjacent  to,  the  vineyard. 

The  Western  Bluebird  is  the  most  common 
SCNB  species  on  the  study  area.  It  is  migra- 
tory, returning  in  late  winter  and  initiating 
nest  building  in  early  March.  This  insectivo- 
rous species  is  monogamous  and  is  known  to 
rear  one  to  two  broods  over  the  spring  and 
summer,  with  both  parents  caring  for  the 
young  (Guinan  et  al.  2000).  Other  SCNB  spe- 
cies on  the  study  area  included  Tree  Swallow 
( Tachycineta  bicolor ),  Violet-green  Swallow 
( Tachycineta  thalassina ),  Ash-throated  Fly- 
catcher ( Myiarchus  cinerascens ),  and  House 
Wren  ( Troglodytes  aedon). 

Nest  boxes. — During  January  and  February 
2003,  we  placed  120  nest  boxes  in  each  of  two 
habitat  types  on  the  Santa  Margarita  ranch: 
oak-savannah  and  vineyard.  The  oak-savan- 
nah habitat  was  open  oak  woodland  (<10% 
canopy  coverage)  characterized  by  grassland 
and  scattered  oak  trees.  We  placed  vineyard 
nest  boxes  ^12  m outside  of  the  vineyard 
edge  because  placing  nest  boxes  in  the  middle 
of  a vineyard  matrix  would  have  interfered 
with  daily  vineyard  management.  To  reduce 
anthropogenic  disturbance  and  minimize 
home-range  overlap  between  bluebird  pairs 
nesting  in  vineyard  versus  oak-savannah  hab- 
itats, we  placed  oak-savannah  nest  boxes 
^300  m from  any  vineyard  edge. 

Boxes  were  constructed  of  rough-cut  cedar 
fence  board  using  a plan  developed  by  the 
North  American  Bluebird  Society  and  fea- 
tured in  Berger  (2000).  The  boxes  were  mod- 
ified such  that  they  opened  from  the  top  in- 
stead of  from  the  side.  In  each  habitat  type, 
we  randomly  selected  30  points  that  were  then 
used  as  starting  points  for  lines  of  four  nest 
boxes.  Each  line  featured  two  nest  boxes  with 
large-diameter  entrance  holes  (3.9  cm)  and 
two  boxes  with  small-diameter  entrance  holes 
(3.2  cm).  Entrance  hole  sizes  were  chosen  to 
promote  nesting  by  native  SCNBs  and  to  pre- 


vent nesting  by  nonnative  cavity  nesters,  such 
as  European  Starlings  ( Sturnus  vulgaris ) and 
House  Sparrows  ( Passer  domesticus).  Using 
metal  hose  clamps,  we  mounted  two  boxes  of 
different  entrance  hole  sizes  back-to-back  on 
a single  2.4-m-high  T-post;  the  other  two  box- 
es were  mounted  singly  on  two  separate  T- 
posts.  To  minimize  the  chances  of  nest  pre- 
dation, we  used  bailing  wire  to  fasten  a 61- 
cm-long,  5.1-cm-diameter  PVC  pipe  to  each 
T-post  directly  under  the  nest  box.  Foam  seal- 
ant was  injected  into  the  core  of  the  PVC  pipe 
to  prevent  snakes  and  small  mammals  from 
climbing  between  the  post  and  the  PVC.  The 
mounted  boxes  were  then  placed  in  lines  of 
three  T-posts  spaced  100  m apart  to  decrease 
nest-site  competition  between  Western  Blue- 
bird pairs  (Perren  1994).  The  four  boxes  were 
placed  such  that  two  entrance  holes  faced  east 
and  two  faced  west.  Box  placement  (paired  or 
single)  and  direction  (east  or  west)  were  as- 
signed randomly. 

Nest  box  monitoring. — In  2003,  we  moni- 
tored nest  boxes  every  7-14  days  throughout 
the  nesting  season,  which  was  sufficient  for 
accurately  determining  rates  of  nest-box  oc- 
cupancy but  not  nest  stages  and  fates.  From 
March  to  May  2004,  we  inspected  each  nest 
box  at  least  every  7-10  days.  Once  we  found 
a nest  box  with  signs  of  nesting  activity,  we 
determined  the  initiation  date  and  monitored 
the  nest  box  at  3-4  day  intervals  to  determine 
its  status;  when  stage  transitions  (e.g.,  onset 
of  incubation,  hatching,  and  fledging)  were 
expected,  we  monitored  nests  every  1-2  days 
(Ralph  et  al.  1993,  Martin  et  al.  1997).  To 
reduce  the  possibility  of  forced  fledging  (Key- 
ser  et  al.  2004),  we  did  not  open  nest  boxes 
after  Western  Bluebird  nestlings  were  14  days 
old.  For  nest  boxes  with  bluebird  nestlings 
older  than  14  days,  we  evaluated  the  nest  sta- 
tus by  observing  parental  behavior  and  listen- 
ing for  nestlings  in  the  box.  We  monitored 
each  Western  Bluebird  nest  until  all  young 
had  fledged  or  the  nest  had  failed.  We  consid- 
ered a nest  successful  if  it  was  empty  within 
2 days  of  the  calculated  fledging  date  and 
there  was  no  sign  of  predation  and/or  if  we 
observed  fledglings  in  the  area  (Martin  et  al. 
1997).  We  checked  each  nest  1-2  days  after 
the  calculated  fledging  date  to  confirm  the 
presence  of  a family  group  in  the  area. 

Habitat  measurements. — In  2004,  we  mea- 


554 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


sured  nine  habitat  variables  at  each  nest  box 
after  the  young  fledged  or  the  nest  failed. 
Many  of  the  measurements  were  based  on 
those  used  in  the  BBIRD  protocol  (Martin  et 
al.  1997).  Variables  included  slope,  aspect, 
and  orientation  of  the  nest-box  entrance,  dis- 
tance to  the  nearest  vines,  and  the  distance  to 
and  the  height  and  DBH  of  the  nearest  tree. 
Within  10  m of  the  nest  box,  we  used  a spher- 
ical densiometer  to  measure  percent  canopy 
cover  and  we  visually  estimated  the  percent 
cover  of  shrubby,  downed  woody  material, 
forbs,  and  grasses.  We  defined  “distance  to 
nearest  perch”  as  the  distance  to  the  nearest 
tree  in  oak-savannah  habitat  and  distance  to 
nearest  vines  in  vineyard  habitat.  This  variable 
provided  an  index  of  perch-site  availability  in 
the  two  habitats. 

We  measured  the  interior  temperature  of 
four  nest  boxes  in  2004  (two  in  vineyard  and 
two  in  oak-savannah  habitat)  by  fastening  a 
HOBO  H8  (Onset  Computer  Corp.,  Bourne, 
Massachusetts)  temperature  data  logger  to  the 
T-post  and  extending  a thermocouple  inside 
the  nest  box.  For  each  box,  temperature  read- 
ings were  recorded  every  15  min  during  the 
entire  nestling  stage  (37-39  days). 

Statistical  analyses. — We  used  a x2  good- 
ness-of-fit  test  (Zar  1996)  to  compare  ob- 
served versus  expected  nest-box  occupancy  in 
oak-savannah  and  vineyard  habitat.  We  used 
the  nest  survival  model  in  program  MARK 
(White  and  Burnham  1999)  to  model  effects 
of  biologically  relevant  factors,  such  as  habitat 
(vineyard  and  oak-savannah)  on  daily  survival 
rate  (Dinsmore  et  al.  2002).  Model  A included 
nest  survivorship  as  a function  of  the  grouping 
variable  (habitat),  and  model  B assumed  con- 
stant survivorship  over  time.  We  used  Akai- 
ke’s  Information  Criterion  corrected  for  small 
sample  size  (AICc)  to  compare  the  set  of  a 
priori  candidate  models  (Burnham  and  An- 
derson 1998).  The  best  model  was  selected  by 
evaluating  the  degree  of  support  for  each 
model  using  the  AICc  values  and  normalized 
Akaike  weights  (w,;  Burnham  and  Anderson 
1998).  The  Akaike  weight  evaluates  the 
strength  of  evidence  for  each  model;  the  high- 
er the  weight,  the  stronger  the  model  (Bum- 
ham  and  Anderson  1998).  We  examined  the 
relationship  between  mean  clutch  size  and  ini- 
tiation date  using  a linear  regression  and  test- 


FIG.  1 . Nest-box  occupancy  (%)  of  120  nest  boxes 
used  by  secondary  cavity-nesting  bird  species  on  the 
Santa  Margarita  Ranch,  San  Luis  Obispo  County,  Cal- 
ifornia, in  2003  and  2004. 


ed  the  significance  of  the  regression  with  an 
F-test  (Zar  1996). 

We  used  a Shapiro- Wilk  statistic  (SPSS  In- 
stitute, Inc.  2003)  to  test  all  variables  for  nor- 
mality. We  then  used  Mann- Whitney  F-tests 
(Zar  1996)  to  test  for  habitat-based  differences 
in  clutch  initiation  date,  clutch  size,  number 
of  eggs  hatched,  number  of  young  fledged, 
slope,  percent  canopy  cover,  and  distance  to 
the  nearest  perch. 

RESULTS 

Nest  box  occupancy. — Western  Bluebirds 
were  the  most  common  nest  box  occupants 
across  habitats  and  years  (Fig.  1).  Western 
Bluebirds  occupied  27.9%  and  33.6%  of  all 
nest  boxes  in  2003  (n  = 240)  and  2004  (n  = 
208),  respectively  (Fig.  1).  Nest  boxes  with 
the  smaller  diameter  entrance  hole  were  un- 
available to  bluebirds;  therefore,  considering 
only  available  boxes,  bluebirds  occupied 
55.8%  of  the  boxes  in  2003  and  67.3%  in 
2004.  In  2004,  Western  Bluebirds  used  nest 
boxes  in  oak-savannah  and  vineyard  habitats 
in  proportion  to  their  availability  (x2  — 0.91, 
df  =1,  P = 0.34). 

Nesting  success. — In  2004,  we  monitored 
70  Western  Bluebird  nests  (n  = 39  in  vineyard 
and  n = 31  in  oak-savannah).  In  program 
MARK,  model  A (habitat)  estimated  daily 
nest  survival  for  the  nesting  period  (i.e.,  egg- 


Fiehler  et  al.  • WESTERN  BLUEBIRD  NEST  BOX  USE 


555 


TABLE  1.  Variables  (mean  ± SE)  describing  nesting  success 
Ranch,  San  Luis  Obispo  County,  California,  2004. 

of  Western  Bluebirds  at  the  Santa  Margarita 

Variable 

Vineyard 

Habitat 

Oak-savannah 

P-value 

Number  of  nests 

39 

31 

. 

Clutch  size 

5.28  ± 0.08 

4.97  ±0.12 

0.040 

Number  of  nestlings  per  nest 

4.90  ± 0.14 

4.63  ± 0.21 

0.465 

Number  of  fledglings  per  nest 

4.69  ± 0.14 

4.63  ± 0.24 

0.799 

Initiation  date  (days  since  1 January) 

88.61  ± 1.56 

92.58  ± 1.48 

0.053 

laying  to  fledging)  at  0.995,  and  model  B 
(constant  survivorship)  estimated  it  at  0.998. 
Furthermore,  AICc  values  for  model  A 
(100.162)  and  model  B (100.729)  were  simi- 
lar, indicating  that  habitat  type  did  not  affect 
the  survival  of  Western  Bluebird  nests  on  the 
Santa  Margarita  Ranch.  Of  the  70  nests,  10 
(14%)  failed,  including  only  two  (3%)  prob- 
able predation  events:  one  nest  appeared  to  be 
depredated  during  the  nestling  stage  by  a 
snake,  and  ants  swarmed  the  other  during  the 
incubation  stage.  The  other  eight  (11%)  failed 
nests  contained  either  dead  chicks  or  cold 
eggs,  and  we  assumed  that  they  were  aban- 
doned. At  least  one  chick  fledged  from  each 
of  the  remaining  60  (86%)  nests. 

Life-history  characteristics. — Clutch  size 
for  many  avian  species  has  been  found  to  de- 
cline over  the  course  of  the  breeding  season 
(Perrins  and  McCleery  1989,  Hochachka 
1990,  Winkler  and  Allen  1996).  In  2004,  there 
was  not  a significant  relationship  between 
mean  clutch  size  and  initiation  date  for  West- 
ern bluebird  nests  across  treatments  (r2  = 
0.11,  df  = 5,  FlA  = 0.51,  P = 0.51).  Clutch 
sizes  were  larger  in  the  vineyard  than  in  oak- 
savannah  (5.28  ± 0.08  versus  4.97  ± 0.12; 
Mann-Whitney  U = 461.00,  P = 0.040)  and 
nests  were  initiated  significantly  earlier  in 
vineyard  habitat  than  in  oak-savannah  (Mann- 
Whitney  U = 400.50,  P = 0.036;  Table  1). 
However,  we  found  no  statistically  significant 
difference  in  number  of  nestlings  (Mann- 
Whitney  U — 473.50,  P = 0.47)  and  number 
of  fledglings  (Mann-Whitney  U = 416.00,  P 
— 0.80)  for  nests  in  vineyard  versus  oak-sa- 
vannah in  2004  (Table  1). 

Habitat  measurements. — Mean  percent  can- 
opy cover  around  the  nest  boxes  did  not  differ 
by  habitat  (5.73  ± 3.44  in  oak-savannah  ver- 
sus 6.28  ± 3.14  in  vineyard;  Mann-Whitney 


U = 604.00,  P = 0.95).  We  found  a difference 
in  mean  distance  to  perch  site  (Mann-Whitney 
U = 84.5,  P < 0.001)  between  nests  in  vine- 
yard and  oak-savannah;  on  average,  perch 
sites  were  closer  to  nest  boxes  in  the  vineyard 
(11.44  ± 0.39)  than  in  the  oak-savannah 
(35.64  ± 4.21)  habitat.  Mean  maximum  tem- 
perature in  nest  boxes  was  28.50°  C ± 0.63  in 
oak-savannah  and  28.53°  C ± 0.65  in  vine- 
yard habitat.  Mean  minimum  temperature  in 
nest  boxes  was  6.22°  C ± 0.27  in  oak-savan- 
nah and  6.14°  C ± 0.27  in  vineyard  habitat. 
Mean  maximum  temperature  (f-test:  t = 
0.042,  df  = 74,  P = 0.97)  and  mean  minimum 
temperature  ( t = —0.232,  df  = 74,  P = 0.82) 
inside  the  nest  box  over  the  nestling  period 
did  not  differ  between  habitat  types. 

DISCUSSION 

The  results  of  this  study  indicate  that  vine- 
yard habitat,  with  its  limited  availability  of 
naturally  occurring  nest  sites,  could  be  con- 
verted to  good  breeding  habitat  for  Western 
Bluebirds  with  the  addition  of  nest  boxes.  In 
the  two  habitat  types.  Western  Bluebirds  were 
the  most  common  nest-box  occupants 
(>55%).  In  2004,  nest  survival  was  high 
across  habitats;  at  least  one  chick  fledged  from 
86%  of  the  nests.  It  should  be  noted,  however, 
that  predator  guards  were  included  on  all  of 
our  nest  boxes,  as  they  are  a common  com- 
ponent of  many  commercially  available  nest- 
box  designs,  and  the  high  nest  survival  and 
fledging  rate  that  we  observed  could  have 
been  an  effect  of  the  predator  guards.  Thus, 
the  high  rate  of  nest  survival  that  we  report 
should  be  interpreted  cautiously. 

Clutch  initiation  date  and  clutch  size  dif- 
fered between  bluebirds  nesting  in  vineyard 
versus  oak-savannah  habitat.  Bluebirds  nest- 
ing in  the  vineyard  initiated  nesting  earlier  and 


556 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


laid  larger  clutches  than  those  in  oak-savan- 
nah habitat.  Habitat  differences  in  food  supply 
have  been  shown  to  affect  the  timing  of  egg 
laying  and  clutch  size  among  passerines 
(Blondel  et  al.  1993,  Siikamaki  1995),  and  the 
predictable  water  supply  provided  by  daily  ir- 
rigation at  Cuesta  Ridge  Vineyard  may  have 
supported  a larger  insect  population  in  the 
vineyard.  In  turn,  this  could  have  allowed  fe- 
male bluebirds  to  start  laying  earlier  and  to 
lay  more  eggs.  There  was  no  significant  dif- 
ference, however,  between  the  two  habitats  in 
terms  of  number  of  nestlings  or  young 
fledged. 

Nest  boxes  in  both  vineyard  and  oak-savan- 
nah habitats  did  not  differ  with  respect  to  per- 
cent canopy  cover  or  interior  nest-box  tem- 
peratures. However,  the  Cuesta  Ridge  vine- 
yard was  structurally  different  from  the  ma- 
jority of  vineyards  in  San  Luis  Obispo 
County:  it  was  composed  of  smaller  areas  of 
vines  that  encompassed  large  valley  oaks  ad- 
jacent to  large  patches  of  native  oak  wood- 
land. Therefore,  our  results  may  not  be  rep- 
resentative of  conditions  in  other  vineyards  in 
the  area.  Additional  research  is  needed  in  the 
more  traditional  vineyards,  which  are  typical- 
ly characterized  by  large,  flat  expanses  of 
vines  and  a lack  of  large  trees. 

Adding  nest  boxes  to  certain  habitats  has 
been  found  to  increase  the  breeding  densities 
of  several  species  of  SCNBs  (Brawn  and  Bai- 
da 1988,  Newton  1994,  Twedt  and  Henne- 
Kerr  2001).  However,  density  can  be  a mis- 
leading indicator  of  habitat  quality  (Van 
Home  1983).  Therefore,  adding  nest  boxes  to 
vineyard  habitats  may  enhance  those  habitats 
so  that  they  serve  as  population  sources  that 
could  stem  the  decline  of  Western  Bluebirds; 
conversely,  such  vineyards  could  be  function- 
ing as  “ecological  traps”  (Delibes  et  al.  2001, 
Mand  et  al.  2005),  population  sinks  that  yield 
no  net  reproduction.  It  is  important  to  note 
that  our  survival  and  productivity  results 
come  from  a single  breeding  season  and  from 
a minimum-impact  vineyard;  also,  nestling 
condition  and  post-fledging  survival  were  not 
quantified.  Additional  research  investigating 
post-fledging  survival  and  nest-site  fidelity  are 
needed  in  vineyards  with  nest  boxes  to  clarify 
their  role  as  population  sources  or  sinks. 

Though  our  data  indicate  that  vineyards 
with  nest  boxes  provide  suitable  breeding  hab- 


itat for  bluebirds,  nest  boxes  in  vineyards 
should  not  be  viewed  as  a remedy  for  the 
chronic  problem  of  habitat  degradation  and 
loss  of  oak  woodlands.  Mpller  (1989)  and 
Purcell  et  al.  (1997)  also  warned  against  using 
nest  boxes  as  a cure-all  for  declining  popula- 
tions. Whereas  nest  boxes  may  be  an  effec- 
tive, short-term  conservation  tool  for  enhanc- 
ing or  maintaining  populations  of  SCNBs — 
Western  Bluebirds  in  particular — they  do  not 
mitigate  the  effects  of  chronic  habitat  loss  for 
the  many  species  that  occupy  oak  woodland 
habitats  in  California. 

ACKNOWLEDGMENTS 

We  thank  three  anonymous  reviewers  for  their  in- 
valuable comments  on  this  manuscript.  We  thank  A. 
Prevel  and  K.  Vincent  for  their  help  erecting  and  mon- 
itoring nest  boxes  in  2003;  M.  Battany  for  assistance 
with  the  dataloggers;  D.  Tempel,  D.  Winslow,  S. 
Bremner-Harrison,  and  B.  Cypher  for  comments  on  the 
manuscript;  and  D.  Filiponi  and  D.  John  for  allowing 
us  access  to  the  Santa  Margarita  Ranch  and  for  logis- 
tical support  in  and  around  the  Cuesta  Ridge  Vineyard. 
The  University  of  California  (UC)  Integrated  Hard- 
wood Range  Management  Program  Grant  # 00-4  sup- 
ported this  research;  the  Morro  Coast  and  the  Santa 
Barbara  Audubon  societies  provided  supplemental 
funding.  The  UC  Cooperative  Extension  Office,  San 
Luis  Obispo,  provided  logistical  support. 

LITERATURE  CITED 

Berger,  C.  2000.  Attracting  birds:  feathers  in  your 
nest  box.  Birder’s  World  14:43. 

Blondel,  J.,  P.  C.  Dias,  M.  Maistre,  and  P.  Perret. 
1993.  Habitat  heterogeneity  and  life-history  vari- 
ation of  Mediterranean  Blue  Tits  ( Parus  caeru- 
leus ).  Auk  110:511-520. 

Brawn,  J.  D.  and  R.  P.  Balda.  1988.  Population  bi- 
ology of  cavity  nesters  in  northern  Arizona:  do 
nest  sites  limit  breeding  densities?  Condor  90:61- 
71. 

Burnham,  K.  P.  and  D.  R.  Anderson.  1998.  Model 
selection  and  inference:  a practical  information- 
theoretic  approach.  Springer- Verlag,  New  York. 
Delibes,  M.,  P.  Gaona,  and  P.  Ferreras.  2001.  Effects 
of  an  attractive  sink  leading  into  maladaptive  hab- 
itat selection.  American  Naturalist  158:277-285. 
Dinsmore,  S.  J.,  G.  C.  White,  and  F.  L.  Knopf.  2002. 
Advanced  techniques  for  modeling  avian  nest  sur- 
vival. Ecology  83:3476-3488. 

Guinan,  J.  A.,  P.  A.  Gowaty,  and  E.  K.  Eltzroth. 
2000.  Western  Bluebird  ( Sialia  mexicana ).  The 
Birds  of  North  America,  no.  510. 

Keyser,  A.  J..  M.  T.  Keyser,  and  D.  E.  L.  Promislow. 
2004.  Life-history  variation  and  demography  in 
Western  Bluebirds  ( Sialia  mexicana)  in  Oregon. 
Auk  121:118-133. 


Fiehler  et  al.  • WESTERN  BLUEBIRD  NEST  BOX  USE 


557 


LeClerc,  J.  E.,  J.  R K.  Che,  J.  P.  Swaddle,  and  D.  A. 
Cristol.  2005.  Reproductive  success  and  devel- 
opmental stability  of  eastern  bluebirds  on  golf 
courses:  evidence  that  golf  courses  can  be  pro- 
ductive. Wildlife  Society  Bulletin  33:483-493. 

Mand,  R.,  V.  Tilgar,  A.  Lohmus,  and  A.  Leivits. 
2005.  Providing  nest  boxes  for  hole-nesting 
birds — does  habitat  matter?  Biodiversity  and  Con- 
servation 14:1823-1840. 

Martin,  T.  E.,  C.  R.  Paine,  C.  J.  Conway,  W.  M.  Ho- 
chachka,  P.  Allen,  and  W.  Jenkins.  1997. 
BBIRD  field  protocol.  Montana  Cooperative 
Wildlife  Research  Unit,  University  of  Montana, 
Missoula. 

M0ller,  A.  P.  1989.  Parasites,  predators  and  nest  box- 
es: facts  and  artifacts  in  nest  box  studies  of  birds? 
Oikos  56:421-423. 

Mummert,  D.  R,  L.  Baines,  and  W.  D.  Tietje.  2002. 
Cavity-nesting  bird  use  of  nest  boxes  in  vineyards 
of  central-coast  California.  Pages  335-340  in  Pro- 
ceedings of  a symposium  on  oak  woodlands:  oaks 
in  California’s  changing  landscape.  General  Tech- 
nical Report  PSW-GTR-184,  USDA  Forest  Ser- 
vice, Pacific  Southwest  Research  Station,  Albany, 
California. 

Newton,  I.  1994.  The  role  of  nest  sites  in  limiting  the 
numbers  of  hole-nesting  birds:  a review.  Biologi- 
cal Conservation  70:265-276. 

Perren,  S.  G.  1994.  Orientation  and  spacing  of  nesting 
boxes  used  by  Eastern  Bluebirds  and  Tree  Swal- 
lows. Sialia  16:127-129. 

Perrins,  C.  M.  and  R.  H.  McCleery.  1989.  Laying 
dates  and  clutch  size  in  the  Great  Tit.  Wilson  Bul- 
letin 101:236-253. 

Purcell,  K.  L.,  J.  Verner,  and  L.  W.  Oring.  1997.  A 


comparison  of  the  breeding  ecology  of  birds  nest- 
ing in  boxes  and  tree  cavities.  Auk  1 14:646-656. 

Ralph,  C.  J.,  G.  R.  Geupel,  P.  Pyle,  T.  E.  Martin, 
and  D.  F.  DeSante.  1993.  Handbook  of  field 
methods  for  monitoring  landbirds.  General  Tech- 
nical Report  PSW-GTR-144,  USDA  Forest  Ser- 
vice, Pacific  Southwest  Research  Station,  Albany, 
California. 

Siikamaki,  P.  1995.  Habitat  quality  and  reproductive 
traits  in  the  Pied  Flycatcher:  an  experiment.  Ecol- 
ogy 76:308-312. 

SPSS  Institute,  Inc.  2003.  SPSS  for  Windows,  ver. 
12.0.  SPSS  Institute,  Inc.,  Chicago,  Illinois. 

Twedt,  D.  J.  and  J.  L.  Henne-Kerr.  2001.  Artificial 
cavities  enhance  breeding  bird  densities  in  man- 
aged cottonwood  forests.  Wildlife  Society  Bulletin 
29:680-687. 

Van  Horne,  B.  1983.  Density  as  a misleading  indi- 
cator of  habitat  quality.  Journal  of  Wildlife  Man- 
agement 47:893-901. 

Verner,  J.  1980.  Birds  of  California  oak  habitats: 
management  implications.  Pages  246-264  in  Pro- 
ceedings of  the  symposium  on  the  ecology,  man- 
agement, and  utilization  of  California  Oaks.  Gen- 
eral Technical  Report  GTR-PSW-44,  USDA  For- 
est Service,  Pacific  Southwest  Research  Station, 
Albany,  California. 

White,  G.  C.  and  K.  P.  Burnham.  1999.  Program 
MARK:  survival  estimation  from  populations  of 
marked  animals.  Bird  Study  46:120-139. 

Zack,  S.  2002.  The  oak  woodland  bird  conservation 
plan:  a strategy  for  protecting  and  managing  oak 
woodland  habitats  and  associated  birds  in  Califor- 
nia. California  Oak  Foundation,  Oakland. 

Zar,  J.  H.  1996.  Biostatistical  analysis,  3rd  ed.  Pren- 
tice Hall,  Upper  Saddle  River,  New  Jersey. 


The  Wilson  Journal  of  Ornithology  1 18(4):558— 562,  2006 


SEXUAL  DIMORPHISM,  DISPERSAL  PATTERNS,  AND  BREEDING 
BIOLOGY  OF  THE  TAIWAN  YUHINA: 

A JOINT-NESTING  PASSERINE 

HSIAO-WEI  YUAN,1 5 SHENG-FENG  SHEN,23  AND  HISN-YI  HUNG1  4 


ABSTRACT. — We  studied  the  breeding  ecology  of  Taiwan  Yuhinas  ( Yuhina  brunneiceps ) at  the  Highlands 
Experiment  Farm  at  Meifeng,  National  Taiwan  University,  in  1995  and  from  1997-2002.  The  Taiwan  Yuhina  is 
a joint-nesting,  cooperatively  breeding  species  endemic  to  Taiwan.  Males  had  significantly  longer  wing  chords 
and  tail  lengths  than  females,  probably  due  to  sexual  selection.  Males  also  had  a longer  residence  time  at  Meifeng 
than  their  female  mates,  which  could  be  explained  by  philopatry  being  greater  in  males.  Alpha  males  had  a 
significantly  longer  residence  time  at  Meifeng  than  beta  males,  but  this  was  not  the  case  for  females,  because 
females  did  not  remain  in  the  same  group  as  males  did  after  their  mates  disappeared.  The  breeding  season  was 
approximately  6 months  long  and  multiple  brooding  was  common.  Nest  building  took  3 days,  egg  laying  occurred 
over  3—4  days,  the  average  incubation  period  was  14  days,  and  the  nestling  period  was  12  days.  Breeding 
success  did  not  decrease  later  in  the  breeding  season.  Maximum  longevity  was  12  years,  and  the  estimate  of 
average  annual  overwinter  survival  rate  for  adults  at  Meifeng  was  74%.  Received  3 August  2005,  accepted  3 
May  2006. 


The  Taiwan  Yuhina  ( Yuhina  brunneiceps ), 
a Timaliine  babbler,  is  a resident  bird  species 
endemic  to  subtropical  Taiwan  (Clements 
2000).  Male  and  female  yuhinas  are  morpho- 
logically indistinguishable  in  the  field.  Joint- 
nesting behavior  in  yuhinas  was  first  de- 
scribed by  Yamashina  (1938).  Recently  our 
group  reported  the  social  system  (including 
joint  nesting)  and  reproductive  success  (Yuan 
et  al.  2004),  incubation  behavior  (Yuan  et  al. 
2004,  2005),  and  habitat  selection  (Lee  et  al. 
2005)  of  yuhinas  from  a 7-year  intensive 
study.  Yuhinas  formed  breeding  groups  of 
2-7  individuals;  group-size  mode  was  four. 
The  yuhina  is  the  only  known  passerine  spe- 
cies to  adopt  a joint-nesting  strategy  for  a 
large  proportion  of  its  nests  (Vehrencamp  and 
Quinn  2004). 

The  majority  (69%)  of  passerine  species 
have  been  considered  sexually  monomorphic 
(Barraclough  and  Harvey  1995);  however,  for 
many  avian  species  there  are  subtle  sexual  dif- 
ferences in  plumage  color  and  morphology 
(Mays  et  al.  2006).  Animals  that  live  in 


1 School  of  Forestry  and  Resource  Conservation, 
National  Taiwan  Univ.,  Taipei,  106,  Taiwan. 

2 Inst,  of  Ecology  and  Evolutionary  Biology,  Na- 
tional Taiwan  Univ.,  Taipei,  106,  Taiwan. 

3 Current  address:  Dept,  of  Neurobiology  and  Be- 
havior, Cornell  Univ.,  Ithaca,  NY  14853,  USA. 

4 Current  address:  Taipei  Zoo,  Taipei,  116,  Taiwan. 

5 Corresponding  author;  e-mail: 
hwyuan@  ntu.edu. tw 


groups  usually  establish  hierarchies,  and 
members  of  different  hierarchical  levels  often 
differ  in  terms  of  body  size  and  age.  There- 
fore, morphological  and  age  comparisons  be- 
tween individuals  of  different  sexes  and  hi- 
erarchical levels  will  shed  light  on  the  extent 
of  sexual  selection  and  the  process  of  group 
formation.  In  this  paper  we  describe  the  mor- 
phological differences  between  male  and  fe- 
male yuhinas,  residence  times  of  different  sex- 
es and  hierarchies,  breeding  chronology,  lon- 
gevity, and  adult  survival  rate. 

METHODS 

We  studied  a population  of  yuhinas  at  the 
Highlands  Experiment  Farm  at  Meifeng,  Na- 
tional Taiwan  University,  in  central  Taiwan 
(24°  05'  N,  121°  10'  E;  2,150-m  elevation) 
during  1995  and  from  1997-2002.  The  study 
area  is  described  in  detail  elsewhere  (Yuan  et 
al.  2004). 

During  this  study,  we  color-banded  252 
adult  yuhinas.  We  measured  bill,  head  (from 
the  back  of  the  cranium  to  the  upper  bill  tip), 
tarsus,  relaxed  wing  chord,  flattened  wing 
chord,  and  tail  length;  crest  height  (from  the 
base  of  the  bill  to  the  tip  of  the  longest  crest 
feather)  and  width  (above  the  eyes);  and  the 
weight  of  each  captured  adult.  A 20-  to  70- 
|jiL  blood  sample  was  collected  from  the  bra- 
chial vein  of  each  adult  and  juvenile.  Each 
sample  was  transferred  into  500-|jlL  Queen’s 
lysis  buffer  (Seutin  et  al.  1991)  and  frozen  at 


558 


Yuan  et  al.  • BREEDING  BIOLOGY  OF  YUHINAS 


559 


— 20°C  until  analyzed.  Sex  was  tentatively  as- 
signed in  the  field  based  on  observations  of 
singing  and  copulation  and  later  verified  using 
sex-specific  genetic  markers  (Fridolfsson  and 
Ellegren  1999). 

We  defined  a breeding  group  as  a set  of 
individuals  exhibiting  parental  behavior  to- 
ward the  young  of  a single  nest.  Within  each 
group,  there  was  a linear  hierarchy  of  socially 
monogamous  pairs.  Dominance  hierarchies 
were  easily  determined  by  observing  chasing 
and  displacement  behavior  among  group 
members  (Yuan  et  al.  2004).  We  monitored 
the  breeding  chronology  of  4,  6,  10,  11,  and 
13  groups  in  1997,  1998,  1999,  2000,  and 
2001,  respectively.  Mayfield  nest  survival 
rates  (Mayfield  1961,  1975)  for  different 
months  were  ascertained  by  intensively  mon- 
itoring 13  breeding  groups  in  2001.  Nest  sta- 
tus was  checked  at  2-  to  10-day  intervals  in 
different  years.  Predation  events  were  deter- 
mined by  checking  whether  there  were  eggs, 
remains  of  eggs,  or  nestlings  left  in  the  nest. 
We  assumed  that  there  was  no  partial  preda- 
tion at  yuhina  nests,  which  was  reasonable  be- 
cause the  eggs  and  nestlings  are  rather  small 
compared  to  those  of  their  predators.  We  con- 
firmed this  assumption  later  by  video-moni- 
toring nests. 

To  estimate  the  adult  overwinter  survival 
rate,  we  monitored  the  fate  of  125  banded  in- 
dividuals. For  the  years  1997-1998,  1998- 
1999,  1999-2000  and  2000-2001,  we  divided 
the  number  of  banded  birds  that  survived  to 
the  second  year  by  the  number  of  banded 
birds  present  in  the  first  year.  Following  Veh- 
rencamp  et  al.  (1988),  we  identified  six  cate- 
gories of  disappearance:  one  of  a mated  pair; 
a dominant  mated  pair;  an  unmated  bird;  a 
non-breeding  bird;  a bird  of  uncertain  status; 
and  an  entire  group.  We  only  counted  the  first 
two  categories  as  mortalities;  the  others  were 
more  likely  to  have  dispersed. 

In  1990,  10  adult  yuhinas  were  banded  at 
Meifeng  as  part  of  a previous  study  (C.-W. 
Yen  pers.  comm.).  Recaptures  of  these  birds 
were  used  to  estimate  long-term  survival.  Be- 
cause most  birds  were  banded  as  adults,  we 
could  not  determine  their  exact  ages.  Instead, 
we  calculated  minimum  residence  time  at 
Meifeng.  For  banded  birds  present  in  2000 
and  2001,  we  determined  the  number  of  years 
in  residence  from  the  date  of  banding.  Birds 


present  in  both  years  were  counted  only  once 
(in  2001). 

Statistical  analyses  were  performed  using 
SAS  software,  ver.  8 (SAS  Institute,  Inc. 
2000).  The  morphological  characteristics  and 
residence  times  of  mated  males  versus  fe- 
males, and  of  alpha  versus  beta  males  and  al- 
pha versus  beta  females,  were  compared  using 
unpaired  or  paired  (as  appropriate)  t-tests  to 
determine  whether  there  were  significant  dif- 
ferences between  groups.  Means  are  repre- 
sented as  ± SD. 

RESULTS 

The  behavior  of  118  individuals  was  ob- 
served in  the  field  and  their  sexes  were  deter- 
mined by  genetic  markers.  We  correctly  iden- 
tified the  sex  of  all  paired  individuals  in  the 
field,  including  53  males  and  47  females. 
However,  the  sex  of  unpaired  individuals  was 
difficult  to  determine  solely  by  field  obser- 
vation. Of  18  unpaired  birds,  including  10 
males  and  8 females,  the  sex  of  only  6 males 
(and  no  females)  was  successfully  determined 
by  behavioral  observation.  Wing  chord  and 
tail  length  of  males  were  significantly  greater 
than  those  of  females,  but  we  detected  no  sta- 
tistically significant  differences  in  any  other 
morphological  variables  (Table  1).  Males  also 
had  a longer  residence  time  than  their  mates 
(3.2  ± 2.2  versus  2.4  ± 1.7  years;  paired 
r-test:  tl6  = 2.36,  P = 0.033).  In  addition,  we 
found  that,  for  a given  group,  alpha  males  had 
longer  residence  times  than  beta  males  (4.3  ± 
1.7  versus  2.8  ±1.3  years;  paired  r-test,  tn  — 
2.92,  P = 0.014).  We  found  no  difference  in 
residence  times  of  alpha  versus  beta  females 
(3.2  ± 1.9  versus  2.5  ± 1.1  years;  paired 
t- test,  tu  = 1.10,  P = 0.30). 

The  breeding  season  lasted  approximately  6 
months,  beginning  in  March  or  April  and  end- 
ing in  August  or  September.  Weather  and  pre- 
dation were  the  two  major  causes  of  nest  fail- 
ure. In  2000  and  2001,  strong  winds  and 
heavy  rains  during  typhoons  and  afternoon 
thunderstorms  destroyed  58%  (n  12)  and 
21%  ( n = 42)  of  the  nests,  respectively.  Pred- 
ators caused  the  failure  of  21%  (2000)  and 
55%  (2001)  of  nests.  Confirmed  predators  of 
yuhina  eggs  and  nestlings  were  Eurasian  Jay 
( Garrulus  glandarius ) and  Taiwan  Sibia  ( Het - 
erophasia  auricularis). 

Nest  success  did  not  decrease  as  the  season 


560 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


TABLE  1 . Morphological  measurements  (mm,  except  for  weight)  of  male  and  female  Taiwan  Yuhinas  from 
17  groups  studied  in  1995  and  from  1997-2001  in  Meifeng,  Taiwan  (24°  05'  N,  121°  10'  E).  Significant  between- 
gender  differences  are  boldfaced. 


Male 


Female 


Measurement  Mean  ± SD  (n)  Mean  ± SD  ( n ) t P 


Bill 

11.9 

4- 

0.3  (25) 

11.9 

0.6  (28) 

0.19 

0.85 

Head 

29.7 

4- 

0.6  (30) 

29.6 

± 

1.3  (27) 

-0.29 

0.78 

Tarsus 

17.9 

4- 

0.6  (25) 

17.8 

+ 

0.6  (28) 

0.81 

0.42 

Wing  chord 

Relaxed 

62.0 

4- 

1.6  (27) 

59.9 

+ 

1.4  (31) 

5.26 

<0.001 

Flattened 

63.5 

± 

1.6  (26) 

61.9 

+ 

1.6  (31) 

3.86 

<0.001 

Crest  height 

27.4 

1.5  (13) 

26.3 

-+- 

1.2  (13) 

1.76 

0.10 

Crest  width 

10.7 

4- 

0.7  (13) 

10.6 

± 

1.2  (9) 

0.22 

0.83 

Tail 

46.3 

4- 

1.9  (27) 

45.3 

1.3  (31) 

3.20 

<0.001 

Weight  (g) 

12.5 

4- 

0.8  (22) 

12.3 

+ 

0.8  (26) 

0.98 

0.33 

progressed  in  2001  (linear  regression,  FXA  = 
0.001,  P = 0.89;  Fig.  1),  and  multiple-brood- 
ing was  common  among  yuhinas  at  Meifeng. 
In  1997,  1998,  and  2000,  at  least  3 of  26 
groups  successfully  raised  chicks  to  fledging 
in  three  consecutive  broods,  and  at  least  4 
groups  produced  two  successful  broods  each. 
In  2001,  one  group  made  nine  nesting  at- 
tempts after  prior  attempts  were  destroyed  ei- 
ther by  inclement  weather  or  predators.  In 
2000  and  2001,  we  found  one  and  two  cases, 
respectively,  in  which  group  members  were 
building  a new  nest  while  still  feeding  fledged 


1.0 


> 

I 

0.8 

_>% 

CD 

T3 

1 0-7 

% 

CD 

2 0.8 


0.5  1 1 1 1 1 1 1 

23456789 

Month 

FIG.  1.  Mayfield  daily  survival  of  Taiwan  Yuhinas 
in  different  months  of  2001  at  Meifeng,  Taiwan  (24° 
05'  N,  121°  10'  E).  Survival  did  not  decrease  later  in 
the  breeding  season.  Sample  size  (nests)  is  shown 
above  each  point;  month  number  corresponds  to  month 
sequence  in  a calendar  year  (i.e.,  2 = February,  3 = 
March,  etc.). 


young  from  their  previous  brood.  Nest  build- 
ing took  approximately  3 days  and  egg  laying 
occurred  over  3—4  days.  Incubation  averaged 
14.3  ± 1.9  days  ( n = 21)  and  the  nestling 
period  was  11.6  ± 2.0  days  ( n = 19).  Re- 
nesting attempts  were  usually  initiated  within 
17.5  ± 2.6  days  ( n = 7)  of  fledging  from  the 
first  nest  if  the  nest  was  successful  and  within 
5.8  ± 3.5  days  ( n = 49)  if  the  nest  failed. 

Of  the  10  adult  yuhinas  banded  in  1990,  we 
recaptured  four  in  1998  (i.e.,  they  were  >9  yr 
old).  Only  one  of  the  four  was  seen  in  1999, 
and  this  individual  was  seen  again  in  2001 
(>12  yr  old).  The  estimated  average  annual 
adult  overwinter  survival  was  74  ± 5%  ( n = 
4 yr  and  125  individual-yr). 

DISCUSSION 

At  Meifeng,  breeding  males  had  longer  res- 
idence times  than  did  the  females.  Alpha 
males  had  longer  residence  times  than  beta 
males,  but  female  dominance  was  not  corre- 
lated with  residence  time.  The  longer  residen- 
cy of  alpha  males  is  likely  because  males  need 
to  queue  into  the  groups  to  become  dominants 
(Kokko  and  Johnstone  1999).  The  difference 
in  male  and  female  residence  times  could  be 
explained  by  our  observation  that  female  sta- 
tus depended  on  the  status  of  their  mates: 
when  paired  alpha  females  disappeared,  most 
of  their  mates  retained  their  alpha  status  and 
found  a new  mate,  but,  when  paired  alpha 
males  disappeared,  few  of  their  mates  retained 
their  dominant  status  (Yuan  et  al.  2004).  Fe- 
males had  shorter  residence  times  than  their 


Yuan  et  al.  • BREEDING  BIOLOGY  OF  YUHINAS 


561 


mates,  possibly  because  females  dispersed  far- 
ther and  searched  for  mates  in  larger  areas, 
which  might  have  increased  their  chances  of 
encountering  available  dominant  males.  Alter- 
natively, females  might  be  forced  to  disperse 
when  their  mates  die.  Males  remained  in  a 
group  and  queued  for  better  breeding  status 
for  comparatively  longer  periods  of  time.  An- 
other explanation  for  the  difference  in  male 
and  female  residence  times  might  be  different 
survival  rates  between  males  and  females.  Ad- 
ditional data  on  the  relationship  between  age 
structure  and  group  composition  are  needed, 
especially  as  they  relate  to  sex  and  domi- 
nance. 

The  size  difference  between  breeding  male 
and  female  yuhinas  could  indicate  that  sexual 
selection  has  been  occurring  in  this  species. 
Larger  body  size  is  related  to  a better  ability 
to  compete  for  resources  (Pusey  and  Packer 
1997).  Indeed,  the  body  size  of  higher  ranking 
male  yuhinas  was  greater  than  that  of  lower- 
ranking  males,  but  there  was  no  such  differ- 
ence in  females  (Yuan  et  al.  2004).  Because  a 
female  yuhina’s  status  is  dependent  upon  that 
of  her  mate,  larger  males  might  have  an  ad- 
vantage because  they  can  maintain  higher 
breeding  status  and  more  easily  attract  mates. 

Given  that  we  did  not  find  any  evidence  for 
a seasonal  decline  in  nest  success,  and  be- 
cause harsh  weather  and  predation  were  the 
main  causes  of  nest  failure,  we  reasoned  that 
the  combined  effects  of  weather  and  predation 
pressure  were  consistent  within  a given  breed- 
ing season.  Therefore,  the  ability  to  renest 
faster  and  more  frequently  is  probably  one  of 
the  main  determinants  of  the  yuhina’s  seasonal 
fecundity.  As  we  have  shown,  yuhinas  could 
make  up  to  nine  nesting  attempts  and  were 
able  to  fledge  multiple  broods  in  a season. 
This  result  supports  the  recent  argument  that 
the  number  of  nesting  attempts  made  by  song- 
birds is  usually  greater  than  formerly  assumed 
(Farnsworth  and  Simons  2001,  Grzybowski 
and  Pease  2005).  A seasonal  trend  in  clutch 
size  could  have  been  another  important  factor 
affecting  seasonal  fecundity  of  yuhinas  (e.g., 
Winkler  and  Allen  1996),  although  we  did  not 
have  enough  data  to  evaluate  this  possibility. 
Because  yuhinas  are  too  small  to  mob  most 
of  their  predators  and  can  renest  faster  in  larg- 
er groups,  we  suggest  that  the  joint-nesting 
behavior  is  a bet-hedging  strategy  to  cope 


with  the  yuhina’s  highly  variable  environment, 
such  as  frequent  typhoons  and  a high  risk  of 
predation;  yuhinas  invest  less  in  single  at- 
tempts and  renest  faster  to  permit  more  nest- 
ing attempts  (Yuan  et  al.  2004). 

ACKNOWLEDGMENTS 

We  thank  H.  L.  Mays,  Jr.,  D.  B.  Burt,  P.  F.  Coulter, 
and  three  anonymous  reviewers  for  valuable  comments 
on  an  earlier  draft  of  our  manuscript.  We  thank 
M.-C.  Tsai,  K.-Z.  Lin,  and  workers  at  Meifeng  Farm 
for  logistical  support.  We  greatly  appreciate  the  vol- 
unteers from  the  National  Taiwan  University  Nature 
Conservation  Students’  Club  and  School  of  Forestry 
and  Resource  Conservation,  in  particular  M.  Liu,  I.-H. 
Chang,  and  K.-D.  Zhong  for  their  help  in  the  field  and 
lab.  Our  research  was  supported  by  grants  from  the 
National  Science  Council,  Taiwan. 

LITERATURE  CITED 

Barraclough,  T.  G.  and  P.  H.  Harvey.  1995.  Sexual 
selection  and  taxonomic  diversity  in  passerine 
birds.  Proceedings  of  the  Royal  Society  of  Lon- 
don, Series  B 259:211-215. 

Clements,  J.  F.  2000.  Birds  of  the  world:  a checklist. 

Ibis  Publishing  Company,  Vista,  California. 
Farnsworth,  G.  L.  and  T.  R.  Simons.  2001.  How 
many  baskets?  Clutch  sizes  that  maximize  annual 
fecundity  of  multiple-brooded  birds.  Auk  118: 
973-982. 

Fridolfsson,  A.  K.  and  H.  Ellegren.  1999.  A simple 
and  universal  method  for  molecular  sexing  of  non- 
ratite  birds.  Journal  of  Avian  Biology  30:1 16-121. 
Grzybowski,  J.  A.  and  C.  M.  Pease.  2005.  Renesting 
determines  seasonal  fecundity  in  songbirds:  what 
do  we  know?  What  should  we  assume?  Auk  122: 
280-291. 

Kokko,  H.  and  R.  A.  Johnstone.  1999.  Social  queuing 
in  animal  societies:  a dynamic  model  of  repro- 
ductive skew.  Proceedings  of  the  Royal  Society  of 
London,  Series  B 266:571-578. 

Lee,  P.-F,  S.-F.  Shen,  T.-S.  Ding,  C.-R.  Chiou,  and 
H.-W.  Yuan.  2005.  Habitat  selection  of  the  co- 
operative breeding  Taiwan  Yuhina  ( Yuhina  brun- 
neiceps)  in  a fragmented  forest  habitat.  Zoological 
Studies  44:494-504. 

Mayfield,  H.  1961.  Nesting  success  calculated  from 
exposure.  Wilson  Bulletin  73:255-261. 

Mayfield,  H.  1975.  Suggestions  for  calculating  nest 
success.  Wilson  Bulletin  87:456-466. 

Mays,  H.  L.,  Jr.,  S.  Doucet,  C.-T.  Yao,  and  H.-W. 
Yuan.  2006.  Sexual  dimorphism  and  dichroma- 
tism in  Steere’s  Liocichla  ( Liocichla  steerii).  Jour- 
nal of  Field  Ornithology  77:1-7. 

Pusey,  A.  E.  and  C.  Packer.  1997.  The  ecology  of 
relationships.  Pages  254-283  in  Behavioural  ecol- 
ogy, 4th  ed.  (J.  R.  Krebs  and  N.  B.  Davies,  Eds.). 
Blackwell  Scientific  Publications,  Oxford,  United 
Kingdom. 


562 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  4,  December  2006 


SAS  Institute,  Inc.  2000.  SAS/STAT  user’s  guide, 
ver.  8.  SAS  Institute,  Inc.,  Cary,  North  Carolina. 

Seutin,  G.,  B.  N.  White,  and  P.  T.  Boag.  1991.  Pres- 
ervation of  avian  blood  and  tissue  samples  for 
DNA  analyses.  Canadian  Journal  of  Zoology  69: 
82-90. 

Vehrencamp,  S.  L.,  R.  R.  Koford,  and  B.  S.  Bowen. 
1988.  The  effect  of  breeding-unit  size  on  fitness 
components  in  groove-billed  anis.  Pages  291-304 
in  Reproductive  success:  studies  of  individual  var- 
iation in  contrasting  breeding  systems  (T.  H.  Clut- 
ton-Brock,  Ed.).  University  of  Chicago  Press,  Chi- 
cago. 

Vehrencamp,  S.  L.  and  J.  S.  Quinn.  2004.  Joint  laying 
systems.  Pages  177-196  in  Cooperative  breeding 
in  birds:  recent  research  and  new  theory  (W.  D. 


Koenig  and  J.  Dickinson,  Eds.).  Cambridge  Uni- 
versity Press,  Cambridge,  United  Kingdom. 

Winkler,  D.  W.  and  P.  E.  Allen.  1996.  The  seasonal 
decline  in  Tree  Swallow  clutch  size:  physiological 
constraint  or  strategic  adjustment?  Ecology  77: 
922-932. 

Yamashina,  M.  1938.  A sociable  breeding  habit 
among  timaliine  birds.  Proceedings  of  the  ninth 
International  Ornithological  Congress  9:453-456. 

Yuan,  H.-W.,  M.  Liu,  and  S.-F.  Shen.  2004.  Joint  nest- 
ing in  Taiwan  Yuhinas:  a rare  passerine  case.  Con- 
dor 106:867-872. 

Yuan,  H.-W.,  S.-F.  Shen,  K.-Y.  Lin,  and  P.-F.  Lee. 
2005.  Group-size  effects  and  parental  investment 
strategies  during  incubation  in  joint-nesting  Tai- 
wan Yuhinas  ( Yuhina  brunneiceps).  Wilson  Bul- 
letin 117:306-312. 


Short  Communications 


The  Wilson  Journal  of  Ornithology  1 1 8(4):563 — 566,  2006 


Ant  Presence  in  Acacias:  An  Association  That  Maximizes  Nesting 

Success  in  Birds? 


Adan  Oliveras  de  Ita1 3 and  Octavio  R.  Rojas-Soto1 2 3 


ABSTRACT. — Nest  predation  is  the  main  cause  of 
reproductive  failure  in  birds,  yet  the  factors  that  drive 
predation  pressure,  as  well  as  the  avian  strategies  to 
minimize  it,  are  poorly  understood.  There  is  a well- 
known  commensal  relationship  between  ants  and  birds 
nesting  in  acacia  trees,  but  the  direct  benefit  in  terms 
of  avian  reproductive  success  has  not  been  tested  prop- 
erly. We  used  artificial  nests  to  compare  success  and 
survival  probability  of  nests  placed  in  Hinds’  acacia 
trees  ( Acacia  hindsii)  associated  with  ants  ( Pseudo - 
myrmex  spp.)  with  those  of  nests  placed  in  trees  with- 
out ants.  Nesting  success  and  the  probability  of  daily 
survival  were  greater  in  acacias  than  in  antless  trees. 
All  cases  of  nest  failure  were  due  to  egg  predation,  but 
none  resulted  from  wren  activities,  as  has  been  re- 
ported in  previous  studies.  The  results  of  this  experi- 
mental study  indicate  that  the  presence  of  ants  in  aca- 
cias may  enhance  avian  reproductive  success  by  re- 
ducing the  probability  of  nest  predation.  Received  30 
June  2005,  accepted  28  June  2006. 


Several  bird  species  of  the  families  Formi- 
cariidae,  Tyrannidae,  Troglodytidae,  and  Em- 
berizidae  prefer  to  establish  their  nests  in  aca- 
cias with  which  Pseudomyrmex  spp.  ants  as- 
sociate (Janzen  1969,  Young  et  al.  1990,  Flas- 
pohler  and  Laska  1994).  The  relationship 
between  birds  nesting  in  acacias  inhabited  by 
ants  seems  to  be  commensal,  because  ants  that 
protect  acacias  against  herbivores  also  offer 
protection  against  avian  nest  predators 
(Skutch  1945,  Janzen  1983,  Flaspohler  and 
Laska  1994).  On  the  other  hand,  birds  do  not 
seem  to  provide  any  benefit  to  acacias  or  ants 
(Gilardi  and  Von  Kugelgen  1991). 


1 Centro  de  Investigaciones  en  Ecosistemas  (CIE- 
CO),  Univ.  Nacional  Autonoma  de  Mexico,  Antigua 
Carretera  a Patzcuaro  No.  8701,  C.R  58190,  Morelia, 
Michoacan,  Mexico. 

2 Instituto  de  Ecologfa,  A.C.,  Depto.  de  Biologfa  Ev- 
olutiva,  km  2.5  Carretera  Antigua  a Coatepec  No.  351, 
Congregacion  el  Haya,  C.R  91070,  Xalapa,  Veracruz, 
Mexico. 

3 Corresponding  author;  e-mail: 
oliveras  @ laneta.apc.org 


It  has  not  been  proven,  however,  that  a myr- 
mecophytic  association  confers  greater  breed- 
ing success  to  birds.  A study  conducted  in 
Costa  Rica  (Young  et  al.  1990)  revealed  a 
36%  failure  rate  of  artificial  nests  {n  = 50) 
placed  in  myrmecophyte  acacias,  but,  in  ant- 
less tress,  only  18%  ( n = 49)  of  the  nests 
failed  (Young  et  al.  1990).  Of  the  failed  nests, 
72%  of  those  located  in  acacias  and  44%  of 
those  located  in  antless  trees  failed  due  to  egg 
destruction  by  Rufous-naped  Wrens  {Campy  - 
lorhynchus  rufinucha). 

We  conducted  an  experiment  on  the  Pacific 
coast  of  Mexico  using  artificial  nests  to  deter- 
mine whether  the  myrmecophytic  association 
confers  a benefit  to  birds  in  terms  of  greater 
nesting  success.  We  also  examined  whether 
nesting  failure  at  our  study  site  was  related  to 
egg  destruction  by  species  ecologically  equiv- 
alent to  the  Rufous-naped  Wren  (Ehrlich  et  al. 
1988,  Dion  et  al.  2000) — Sinaloa  Wren  ( Thry - 
othorus  sinaloa ),  Happy  Wren  (T.  felix ),  and 
White-bellied  Wren  {Uropsila  leucogastra). 

METHODS 

We  conducted  our  study  during  September 
2004  in  the  Chamela-Cuixmala  Biosphere  Re- 
serve on  the  Pacific  coast  of  Mexico  (19°  30' 
N,  105°  0.3'  W).  Tropical  dry  deciduous  forest 
is  the  dominant  vegetation,  and  acacias  gen- 
erally occur  as  secondary  growth  in  locally 
distributed  sites  near  the  coast.  We  collected 
data  at  two  sites  characterized  by  similar  veg- 
etation: Careyes  and  Negritos,  situated  south- 
east and  northeast,  respectively,  of  the  Biolog- 
ical Station.  We  randomly  selected  a 1-km 
transect  at  each  site  and  placed  28  artificial 
nests  along  each  transect:  14  in  Hinds’  acacia 
trees  {Acacia  hindsii ) and  14  in  antless  trees. 
The  cup-shaped  nests  were  placed  1.7— 2.2  m 
above  ground  and  wired  to  the  tree  trunks.  In 
each  nest,  we  placed  three  hand-made  eggs 
(20-mm  length) — made  of  white  plasticine 
and  sprayed  with  varnish — to  resemble  eggs 


563 


564 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


of  the  Social  Flycatcher  ( Myiozetetes  similis). 
Social  Flycatchers  are  common  breeders  in 
the  area  and  reportedly  nest  in  acacias  (Pettin- 
gill  1942).  Predators  readily  left  marks  in  the 
plasticine,  thus  allowing  us  to  identify  preda- 
tor species  and  the  impact  of  wrens  on  nesting 
success,  if  any  (Major  1991,  Major  and  Ken- 
dal 1996,  Dion  et  al.  2000,  Zanette  and  Jen- 
kins 2000). 

Nests  were  exposed  to  predators  for  6 days. 
We  recorded  egg  condition  every  2 days  and 
removed  those  nests  in  which  eggs  showed 
evidence  of  predation.  Based  on  previous  re- 
ports (Kennedy  and  White  1996,  Hannon  and 
Cotterill  1998),  wren  species  usually  peck 
small  holes  in  the  eggs  of  other  species.  To 
determine  whether  wrens  were  responsible  for 
nest  “failure,”  we  compared  marks  on  the 
plasticine  eggs  recovered  from  depredated 
nests  with  those  we  made  using  the  bills  of 
museum  specimens  representing  the  three 
wren  species  that  occurred  in  our  study  area: 
Sinaloa  Wren,  Happy  Wren,  and  White-bellied 
Wren. 

The  percentage  of  nests  in  which  no  eggs 
showed  damage  by  the  end  of  our  experiment 
was  our  measure  of  nesting  success.  To  deter- 
mine differences  in  failure  probabilities  be- 
tween sites  and  tree  type  in  which  nests  were 
located,  we  analyzed  the  data  with  a linear 
generalized  model  (GENMOD),  assuming  a 
binominal  distribution  and  a logit  function 
(SAS  Institute,  Inc.  2000).  The  independent 
categorical  variables  were  our  two  sites  (Car- 
eyes  and  Negritos)  and  the  two  tree  types 
(myrmecophyte  acacia  or  antless  tree);  in  both 
cases  the  dependent  variable  was  the  proba- 
bility of  nest  failure. 

We  calculated  daily  survival  rate  (DSR),  by 
tree  type,  using  the  daily  probability  of  nest 
survival.  Survival  rate — the  most  reliable 
measure  of  nesting  success  (Ralph  et  al. 
1996) — was  calculated  with  the  MAYFIELD 
program  (Hines  1996)  based  on  the  method 
proposed  by  Mayfield  (1961,  1975)  and  re- 
vised by  Bart  and  Robson  (1982).  Differences 
in  DSR  means  were  assessed  with  a Z-test  us- 
ing variances  obtained  from  the  MAYFIELD 
program.  Means  are  reported  ± SE. 

RESULTS 

Nest  success  was  similar  at  both  sites  (39% 
at  Careyes  and  43%  at  Negritos;  x2  — 0. 15,  P 


= 0.70,  df  = 1).  However,  nest  success  was 
greater  for  nests  placed  in  acacias  (64.3%) 
than  those  placed  in  antless  trees  (17.8%;  x2 
= 13.06,  P < 0.001,  df  = 1).  Because  there 
was  no  site  effect,  we  pooled  our  data  for  cal- 
culating DSR  estimates.  DSR  was  greater  for 
nests  located  in  acacias  (0.944  ± 0.017,  n = 
28)  than  it  was  for  those  located  in  antless 
trees  (0.808  ± 0.036,  n = 28;  Z = 10.73,  P 
= 0.010).  Overall  nest  survival  (6  days  of  ex- 
posure) was  70.5%  {n  = 28)  in  acacias,  and 
28%  (n  = 28)  in  antless  trees.  All  nest  failures 
were  due  to  predation;  however,  based  on  our 
observations  of  marks  left  on  the  plasticine 
eggs,  no  eggs  were  destroyed  by  wrens. 

DISCUSSION 

Our  results  indicate  that  the  type  of  tree 
where  nests  were  placed  (acacias  versus  ant- 
less) affected  the  probability  of  nest  success. 
Probability  of  survival  was  greater  for  nests 
placed  in  acacias,  which  may  be  related  to  the 
presence  of  ants.  This  supports  Skutch’s 
(1945)  hypothesis,  which  suggests  that  nests 
in  acacias  have  a higher  probability  of  surviv- 
al due  to  the  ants  that  associate  with  them, 
despite  the  minimal  cover  that  acacias  provide 
for  nest  concealment  (Young  et  al.  1990).  The 
results  of  previous  studies  with  artificial  nests 
of  other  species  indicate  that  egg  predation 
may  be  greater  where  canopy  cover  is  mini- 
mal (Crabtree  et  al.  1989,  Sullivan  and  Dins- 
more  1990,  Mankin  and  Warnen  1992,  Martin 
1992;  but  see  Gottfried  and  Thompson  1978). 
Although  we  did  not  measure  canopy  cover 
around  the  nests,  egg  predation  was  not  great- 
er under  the  poor  canopy  cover  that  charac- 
terizes Acacia  spp.  Indeed,  low  rates  of  egg 
predation  in  acacias — despite  their  minimal 
foliage  cover — underscores  the  potential  role 
of  ants  in  providing  protection  against  nest 
predators. 

In  Costa  Rica,  the  success  rate  of  artificial 
nests  placed  in  acacias  (64%;  Young  et  al. 
1990)  was  similar  to  the  rate  we  detected  at 
Chamela  (64.3%),  but  the  percentage  of  suc- 
cessful nests  in  antless  trees  was  much  greater 
(81.6%)  than  it  was  at  Chamela  (17.8%).  In 
addition,  we  found  no  evidence  of  wren  pre- 
dation on  eggs,  though  longer  observation  pe- 
riods may  be  necessary  to  confirm  this  pat- 
tern. The  low  rates  of  success  that  we  ob- 
served for  nests  placed  in  antless  trees  (en- 


SHORT  COMMUNICATIONS 


565 


tirely  due  to  predation)  suggest  that,  in  the 
absence  of  Rufous-naped  Wrens,  acacias  with 
which  ants  associate  increases  the  probability 
of  avian  nest  survival,  despite  of  the  presence 
of  other  wren  species. 

Previous  researchers  have  proposed  that 
birds  reduce  the  probability  of  nesting  failure 
by  minimizing  parental  activity  around  the 
nest  (Martin  et  al.  2000);  producing  smaller 
clutches  to  minimize  parental  activity  (Skutch 
1949,  1976)  or  to  save  energy  for  a second 
brood  (Slagsvold  1982);  evolving  shorter  in- 
cubation periods  (Ricklefs  1969;  but  see  Mar- 
tin 2002);  and/or  nesting  at  the  end  of  the  dry 
season  (Morton  1971,  Poulin  et  al.  1992).  Jan- 
zen  (1969)  and  Young  et  al.  (1990)  found  that 
several  species  were  more  likely  to  nest  in 
acacias  than  in  antless  trees.  Consistent  with 
these  observations,  our  results  indicate  that  ar- 
tificial nests  located  in  acacias  with  ants  have 
greater  probabilities  of  nest  survival.  Thus,  we 
propose  that  this  may  be  yet  another  strategy 
for  maximizing  nest  success. 

Unfortunately,  no  antless  acacias  were 
available  at  our  study  sites;  evaluations  of  nest 
success  in  antless  acacias  will  be  necessary  to 
confirm  the  role  of  ants  in  discouraging  pre- 
dation. In  addition,  evaluating  the  effects  of 
different  acacia  species,  canopy  cover,  and  the 
possible  influence  of  different  ant  species  on 
nest  success  will  provide  better  insights  into 
the  mechanisms  behind  enhanced  nesting  suc- 
cess in  acacias  with  which  ants  associate. 

ACKNOWLEDGMENTS 

M.  Quesada,  I.  Herrerfas,  P.  Cuevas,  and  G.  Sanchez 
supported  the  planning  and  development  of  our  proj- 
ect. K.  Renton  provided  material  and  ideas  for  devel- 
oping the  study.  B.  Mila,  F.  Lopez,  C.  Gonzalez  and 
two  anonymous  reviewers  provided  comments  that  im- 
proved the  manuscript.  We  thank  P.  Mosig  and  E.  Silva 
for  helping  us  with  manuscript  translation.  Estacion  de 
Biologfa  Chamela  of  the  Chamela-Cuixmala  Biosphere 
Reserve  provided  field  facilities. 

LITERATURE  CITED 

Bart,  J.  and  D.  S.  Robson.  1982.  Estimating  survi- 
vorship when  the  subjects  are  visited  periodically. 
Ecology  63:1078-1090. 

Crabtree,  R.  L.,  L.  S.  Broome,  and  M.  L.  Wolfe. 
1989.  Effects  of  habitat  characteristics  on  Gadwall 
nest  predators  and  nest-site  selection.  Journal  of 
Wildlife  Management  53:129-137. 

Dion,  N.,  K.  A.  Hobson,  and  S.  Lariviere.  2000.  In- 
teractive effects  of  vegetation  and  predators  on  the 


success  of  natural  and  simulated  nests  of  grassland 
songbirds.  Condor  102:629-634. 

Ehrlich,  P.  R.,  D.  S.  Dobkin,  and  D.  Wheye.  1988. 
The  birder’s  handbook.  Simon  and  Schuster,  New 
York. 

Flaspohler,  D.  J.  and  M.  S.  Laska.  1994.  Nest  site 
selection  in  acacia  trees  in  a Costa  Rican  dry  de- 
ciduous forest.  Wilson  Bulletin  106:162-165. 

Gilardi,  J.  D.  and  K.  Von  Kugelgen.  1991.  Bird/ant/ 
acacia  symbiosis  in  a mature  Neotropical  forest. 
Wilson  Bulletin  103:711-712. 

Gottfried,  B.  M.  and  C.  F.  Thompson.  1978.  Exper- 
imental analysis  of  nest  predation  in  an  old-field 
habitat.  Auk  95:304-312. 

Hannon,  S.  J.  and  S.  E.  Cotterill.  1998.  Nest  pre- 
dation in  aspen  woodlots  in  an  agricultural  area  in 
Alberta:  the  enemy  from  within.  Auk  115:16-25. 

Hines,  J.  E.  1996.  MAYFIELD  Software  U.S.  Geological 
Survey-Patuxent  Wildlife  Research  Center,  www. 
mbr-p wrc . usgs . go v/software/may field . html  (accessed 
3 February  2006). 

Janzen,  D.  H.  1969.  Birds  and  the  ant  X acacia  inter- 
action in  Central  America.  Evolution  20:248-275. 

Janzen,  D.  H.  1983.  Costa  Rican  natural  history.  Uni- 
versity of  Chicago  Press,  Chicago,  Illinois. 

Kennedy,  E.  D.  and  D.  W.  White.  1996.  Interference 
competition  from  House  Wrens  as  a factor  in  the 
decline  of  Bewick’s  Wrens.  Conservation  Biology 
10:281-284. 

Major,  R.  E.  1991.  Identification  of  nest  predators  by 
photography,  dummy  eggs,  and  adhesive  tape. 
Auk  108:190-195. 

Major,  R.  E.  and  C.  E.  Kendal.  1996.  The  contri- 
bution of  artificial  nest  experiments  to  understand- 
ing avian  reproductive  success:  a review  of  meth- 
ods and  conclusions.  Ibis  138:298-307. 

Mankin,  P.  C.  and  R.  E.  Warner.  1992.  Vulnerability 
of  ground  nests  to  predation  on  an  agricultural 
habitat  island  in  east-central  Illinois.  American 
Midland  Naturalist  128:281-291. 

Martin,  T.  E.  1992.  Breeding  productivity  consider- 
ations: what  are  the  appropriate  habitat  features 
for  management?  Pages  455-473  in  Ecology  and 
conservation  of  Neotropical  migrant  landbirds  (J. 
M.  Hagan,  III,  and  D.  W.  Johnston,  Eds.).  Smith- 
sonian Institution  Press,  Washington,  D.C. 

Martin,  T.  E.  2002.  A new  view  of  avian  life-history 
evolution  tested  on  an  incubation  paradox.  Pro- 
ceedings of  the  Royal  Society  of  London,  Series 
B 269:309-316. 

Martin,  T.  E.,  J.  Scott,  and  C.  Menge.  2000.  Nest 
predation  increases  with  parental  activity:  sepa- 
rating nest  site  and  parental  activity  effects.  Pro- 
ceedings of  the  Royal  Society  of  London,  Series 
B 267:2287-2293. 

Mayfield,  H.  F.  1961.  Nesting  success  calculated  from 
exposure.  Wilson  Bulletin  73:255-261. 

Mayfield,  H.  F.  1975.  Suggestions  for  calculating 
nesting  success.  Wilson  Bulletin  87:456-466. 

Morton,  E.  S.  1971.  Nest  predation  affecting  the 


566 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


breeding  season  of  the  Clay-colored  Robin.  Sci- 
ence 181:920-921. 

Pettingill,  O.  S.,  Jr.  1942.  The  birds  of  a bull’s  horn 
acacia.  Wilson  Bulletin  54:89-96. 

Poulin,  B.,  G.  Lefebvre,  and  R.  McNeil.  1992.  Trop- 
ical avian  phenology  in  relation  to  abundance  and 
exploitation  of  food  resources.  Ecology  73:2295- 
2309. 

Ralph,  C.  J.,  G.  R.  Geupel,  P.  Pyle,  T.  E.  Martin,  D. 
F.  DeSante,  and  B.  Mila.  1996.  Manual  de  me- 
todos  de  campo  para  el  monitoreo  de  aves  terres- 
tres.  General  Technical  Report  PSW-GTR  159, 
USDA  Forest  Service,  Pacific  Southwest  Research 
Station,  Albany,  California. 

Ricklefs,  R.  E.  1969.  An  analysis  of  nesting  mortality 
in  birds.  Smithsonian  Contributions  to  Zoology  9: 
1-48. 


SAS  Institute,  Inc.  2000.  SAS  statistics,  release  8.02. 
SAS  Institute,  Inc.,  Cary,  North  Carolina. 

Skutch,  A.  F.  1945.  The  most  hospitable  tree.  Science 
Monthly  60:5-17. 

Skutch,  A.  F.  1976.  Parental  birds  and  their  young. 
University  of  Texas  Press,  Austin. 

Slags vold,  T.  1982.  Clutch  size  variation  in  passerine 
birds:  the  nest  predation  hypothesis.  Oecologia  54: 
159-169. 

Sullivan,  B.  D.  and  J.  J.  Dinsmore.  1990.  Factors 
affecting  egg  predation  by  American  Crows.  Jour- 
nal of  Wildlife  Management  54:433-437. 

Young,  E.  B.,  M.  Kaspari,  and  T.  E.  Martin.  1990. 
Species-specific  nest  site  selection  by  birds  in  aca- 
cia trees.  Biotropica  2:310-315. 

Zanette,  L.  and  B.  Jenkins.  2000.  Nesting  success 
and  nest  predators  in  forest  fragments:  a study  us- 
ing real  and  artificial  nests.  Auk  117:445-454. 


The  Wilson  Journal  of  Ornithology  1 18(4):566-569,  2006 

Pair  Roosting  of  Nesting  Carolina  Wrens  ( Thryothorus  ludovicianus ) 

Ronald  F.  Labisky1 2 and  John  E.  Arnett,  Jr.1 2 


ABSTRACT. — Carolina  Wrens  ( Thryothorus  ludo- 
vicianus),  which  maintain  lifetime  pair  bonds  and  year- 
round  territories,  huddle  in  pair  or  communal  roosts 
during  the  non-breeding  season,  particularly  during 
cold  winter  nights.  Pair  roosting  during  the  nesting 
season,  however,  is  not  known  to  occur.  Here,  we  re- 
port huddled  pair  roosting  by  Carolina  Wrens  in  Flor- 
ida. The  dates  of  pair  roosting  took  place  during  nest 
construction  through  laying  of  the  first  egg  (9-20 
March  2004),  and  also  on  the  date  the  fourth  egg  was 
laid  in  a clutch  of  five  (24  March).  The  wrens  roosted 
in  a hanging  flower  basket  located  2.4  m from  their 
nest  site.  Although  huddled  pair  roosting  by  wrens  dur- 
ing periods  of  low  ambient  temperatures  in  the  non- 
breeding season  likely  achieves  thermal  conservation, 
the  benefits  derived  during  the  breeding  season  remain 
unclear.  We  discuss  the  possible  thermoregulatory  and 
pair-bond  maintenance  functions  of  pair  roosting.  Re- 
ceived 6 September  2005,  accepted  5 July  2006. 


Roosting  by  two  or  more  birds  has  been 
hypothesized  to  ameliorate  the  energetic  cost 
of  thermoregulation  during  cold  temperatures, 
lower  the  risk  of  predation,  and  improve  for- 
aging efficiency  (Beauchamp  1999).  Numer- 


1  Dept,  of  Wildlife  Ecology  and  Conservation,  Univ. 
of  Florida,  P.O.  Box  110430,  Gainesville,  FL  32611, 
USA. 

2 Corresponding  author;  e-mail:  labiskyr@ufl.edu 


ous  researchers  have  examined  pair,  commu- 
nal, or  huddled  roosting  during  the  non-breed- 
ing season  (in  cavities:  du  Plessis  and  Wil- 
liams 1994;  in  dormitory  nests:  Sharrock 
1980,  Gill  and  Stutchbury  2005;  in  foliage: 
Baida  et  al.  1977).  Yet,  the  occurrence  and 
function  of  these  types  of  roosts  during  the 
breeding  season  remains  a poorly  understood 
aspect  of  avian  behavior. 

The  Carolina  Wren  ( Thryothorus  ludovici- 
anus) is  the  only  Thryothorus  wren  whose 
range  extends  beyond  tropical  latitudes  (Mor- 
ton 1982).  In  contrast  to  wren  species  with 
which  it  is  sympatric  in  North  America,  Car- 
olina Wrens  form  lifetime  pair  bonds  and  de- 
fend a territory  throughout  the  year  (Morton 
and  Shalter  1977).  They  also  roost  in  a variety 
of  natural  and  anthropogenic  structures  (Hag- 
gerty and  Morton  1995)  and  are  known  to 
roost  in  pairs  during  the  non-breeding  season 
(Brooks  1932,  Tamar  1980).  Whereas  some 
tropical  wrens  form  communal  or  pair  roosts 
throughout  the  year  (Skutch  1940,  Robinson 
et  al.  2000,  Gill  and  Stutchbury  2005),  to  our 
knowledge  there  are  no  reports  of  pair  roost- 
ing during  the  breeding  season  for  tropical  or 
temperate  populations  of  Carolina  Wrens.  Las- 
key (1948)  assumed  that  both  members  of  a 


SHORT  COMMUNICATIONS 


567 


pair  of  Carolina  Wrens  she  observed  during 
the  egg-laying  phase  were  roosting  together, 
but  she  did  not  confirm  this.  Here,  we  confirm 
huddled  pair  roosting  by  Carolina  Wrens  dur- 
ing the  egg-laying  phase  of  the  nesting  season 
in  northern  Florida. 

Observations  were  made  in  an  urban  setting 
(residence  of  RFL)  in  Gainesville,  Florida 
(29°  40'  N,  82°  24'  W).  From  5 to  17  March 
2004,  a pair  of  Carolina  Wrens  carried  nest 
material  to  the  base  of  a potted  bromeliad  on 
an  east-facing  ledge,  1.2  m above  the  floor  of 
a covered  patio  deck.  On  9 March,  approxi- 
mately 5 min  after  sunset,  the  pair  flew  di- 
rectly to  the  rim  of  an  open-topped  hanging 
plant  basket  (devoid  of  plants)  2.4  m from  the 
nest  site  and,  within  seconds,  dropped  down 
to  roost  in  the  slightly  cupped  depression  on 
the  peat/soil  surface.  From  10-15  March,  the 
pair  exhibited  similar  roosting  behavior,  both 
birds  arriving  at  the  roost  site  at  the  same 
time.  On  16  March,  just  after  sunset,  one  of 
the  pair  went  to  roost  in  the  hanging  basket, 
and  emitted  soft  “cheeps”  until  the  second 
wren  joined  it  4 min  later.  This  roosting  pat- 
tern was  repeated  in  a similar  fashion  from 
17-19  March. 

The  first  egg  was  deposited  in  the  nest 
shortly  after  sunrise  on  20  March  and,  on  this 
date,  the  pair  again  roosted  together.  On  21 
and  22  March,  the  second  and  third  eggs  were 
laid,  and  one  bird  (presumably  the  female) 
roosted  on  the  nest  while  the  other  roosted  in 
the  hanging  basket.  On  23  March,  however, 
when  the  fourth  egg  was  laid,  both  wrens 
roosted  in  the  hanging  basket.  This  date  was 
the  last  on  which  both  birds  were  observed 
roosting  together.  On  24  March,  when  the  fifth 
and  final  egg  of  the  clutch  was  laid,  one  bird 
roosted  on  the  nest  and  the  other  in  the  hang- 
ing basket.  On  25  March,  only  the  bird  roost- 
ing on  the  nest  was  observed;  however,  on  the 
following  night,  one  of  the  pair  roosted  in  the 
hanging  basket  and  the  other  on  the  nest.  After 
26  March,  no  further  roosting  in  the  hanging 
basket  was  observed. 

This  pair  of  Carolina  Wrens  roosted  togeth- 
er in  the  hanging  basket  for  a period  of  12 
days  (9-20  March),  which  spanned  the  period 
of  nest  construction  and  deposition  of  the  first 
egg.  They  roosted  together  again  only  on  23 
March,  the  day  on  which  the  female  laid  the 
fourth  egg  of  the  five-egg  clutch.  Observa- 


tions on  4 of  the  13  nights  during  which  the 
pair  roosted  together  revealed  that  the  two 
birds  were  always  in  contact  with  one  another 
(huddled),  with  one  wren  positioned  slightly 
in  front  of  the  other.  The  roosting  birds  always 
departed  from  the  roost  site  shortly  after  day- 
break. The  eggs  hatched  on  9 April,  and  four 
young  fledged  on  18  April  with  both  adults 
present. 

We  discuss  two  alternative,  but  not  mutu- 
ally exclusive,  explanations  for  these  obser- 
vations: thermoregulation  (Beauchamp  1999) 
and  pair-bond  maintenance  (Kellam  2003). 
Small  birds  lose  heat  rapidly,  even  in  tropical 
climates  (Merola-Zwartjes  1998),  and  the  en- 
ergetic cost  of  thermoregulation  is  high  (Fer- 
guson et  al.  2002).  At  low  ambient  tempera- 
tures in  winter,  Carolina  Wrens  in  the  temper- 
ate region  can  experience  high  mortality 
(Brooks  1936,  Tamar  1980).  A possible  neg- 
ative relationship  between  temperature  and  di- 
urnal foraging  time  for  Carolina  Wrens  (Strain 
and  Mumme  1988)  could  further  limit  the  en- 
ergy available  for  nocturnal  thermoregulation. 
Given  that  low  temperatures  increase  the  en- 
ergetic requirements  of  birds,  and  that  the  en- 
ergetic requirements  of  female  birds  increase 
before  and  during  laying  (Nager  and  van 
Noordwijk  1992),  a laying  female  may  display 
behaviors  that  would  mitigate  thermoregula- 
tory losses  resulting  from  low  nocturnal  am- 
bient temperatures  (Weeks  1994).  Pair  roost- 
ing by  altricial  passerines  may  create  a micro- 
climate that  ameliorates  the  energetic  costs  of 
thermoregulation  (Merola-Zwartjes  1998)  and 
mitigates  the  effects  of  low  temperature  on  de- 
creased egg  volume  (Nager  and  van  Noord- 
wijk 1992)  and  on  interrupted  egg  laying 
(Yom-Tov  and  Wright  1993). 

Nocturnal  temperatures  during  the  period 
(5-26  March)  of  our  observations  generally 
ranged  between  7 and  10°C  (http://weather. 
herald.com/auto/miamiherald/history/airport/ 
KGNV/2004/3/26/DailyHistory.html).  Mini- 
mum temperatures  during  the  nights  when  the 
pair  roosted  together  averaged  2°  C colder 
than  the  other  nights  during  March  2004.  The 
wrens  roosted  together  on  8 of  the  10  coldest 
nights  of  the  month,  and  only  on  1 of  the  10 
warmest  nights  of  the  month.  The  roosting 
birds  fluffed  their  head,  back,  and  rump  feath- 
ers— typical  of  sleeping  wrens  (Williams 
1941.  Haggerty  and  Morton  1995).  Feather 


568 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


erection  not  only  facilitates  convective  cool- 
ing of  birds  in  hot  climates  (Ferns  1992),  but 
also  reduces  the  thermal  conductance  of  plum- 
age, thus  providing  insulation  (Ferguson  et  al. 
2002)  in  cold  climates.  If  thermoregulation 
best  explains  pair  roosting  by  Carolina  Wrens 
during  egg  laying,  both  parents  may  benefit 
via  enhanced  egg  volume  and  uninterrupted 
laying.  However,  if  roosting  in  cavities  and 
roost  nests  evolved  as  an  anti-predator  behav- 
ior (Merola-Zwartjes  1998),  any  thermoregu- 
latory benefit  might  be  only  coincidental. 

Pair  roosting  before  and  during  egg  laying 
may  reinforce  the  pair  bond  and  prevent  di- 
vorce in  Carolina  Wrens.  Behaviors  that  pro- 
mote contact,  achieve  breeding  synchrony, 
and  demonstrate  commitment  may  serve  to 
maintain  avian  pair  bonds  (Hall  2000).  For  ex- 
ample, some  males  of  a tropical  congener  spe- 
cies that  forms  permanent  pair  bonds  may  ini- 
tiate duets  in  order  to  limit  extra-pair  mating 
and  divorce  (Gill  and  Stutchbury  2005),  and, 
in  some  passerine  species  that  form  lifetime 
pair  bonds,  both  sexes  may  actively  guard 
their  mates  (Hall  2000,  Gill  2003).  Carolina 
Wrens  are  genetically  monogamous  and  rarely 
divorce  (Haggerty  et  al.  2001);  thus,  we  might 
expect  at  least  one  sex  to  actively  limit  extra- 
pair mating.  Due  to  the  rigors  of  fledgling  care 
and  providing  food  to  their  mates,  Haggerty 
et  al.  (2001)  doubted  that  male  Carolina 
Wrens  could  prevent  females  from  engaging 
in  extra-pair  mating;  however,  this  explanation 
does  not  preclude  males  from  mate  guarding 
during  the  relatively  less  intense  nest-building 
and  egg-laying  phases. 

Paired  female  Carolina  Wrens  may  have  a 
higher  probability  of  year-round  survival  than 
solitary  females  (Haggerty  et  al.  2001).  Mor- 
ton and  Shalter  (1977)  speculated  that  because 
individual  male  Carolina  Wrens  can  maintain 
a territory,  whereas  individual  females  cannot, 
females  may  actively  reinforce  the  lifetime 
pair  bond  as  a safeguard  against  divorce.  Ac- 
cordingly, the  female  would  likely  initiate  pair 
roosting  during  the  nesting  season.  In  our  ob- 
servations, both  members  of  the  pair  arrived 
at  the  roost  simultaneously  during  nest  build- 
ing, but,  as  laying  approached,  the  birds  ar- 
rived separately  and  one  bird  (sex  unknown), 
called  to  the  other  from  the  roost.  Of  the  Car- 
olina Wren  pair  that  she  observed,  Laskey 
(1948)  noted  that  the  male  arrived  first  at  the 


roost  site  and  called  to  the  female  from  there. 
This  anecdotal  evidence  suggests  that  pair 
roosting  during  nest  construction  and  egg  lay- 
ing is  initiated  by  the  male.  Because  the 
wren’s  short  period  of  fertility  represents  the 
most  advantageous  time  for  opportunistic 
males  to  mate  with  other  females  (Gill  2003), 
mate  guarding  by  males  during  egg  laying 
seems  plausible. 

In  this  paper,  we  have  reported  huddled  pair 
roosting  by  Carolina  Wrens  during  the  nesting 
season,  and  we  have  discussed  two  possible 
mechanisms,  thermoregulatory  benefits  and 
pair-bond  maintenance,  to  explain  this  behav- 
ior. The  possibility  that  this  behavior  was  that 
of  a non-breeding  pair  continuing  their  winter 
roosting  into  the  early  part  of  the  nesting  sea- 
son is  most  unlikely  for  two  reasons;  (1)  the 
pair  roosting  that  we  observed  spanned  the 
duration  of  nest  construction  and  egg  laying, 
and  (2)  other  physiological  and  behavioral 
changes  occur  concomitantly  during  this 
phase  of  the  breeding  season.  Consequently, 
the  evidence  suggests  that  we  documented  a 
previously  unconfirmed  behavior.  Whereas  the 
functions  of  huddled  pair  and  communal 
roosting  during  the  non-breeding  season  have 
been  studied  in  detail,  more  study  is  needed 
to  identify  the  function  of  pair  roosting  during 
the  breeding  season  by  birds  that  form  lifetime 
pair-bonds,  and  which  sex,  if  either,  typically 
initiates  pair  roosting. 

ACKNOWLEDGMENTS 

Earlier  versions  of  this  manuscript  were  improved 
thanks  to  comments  from  T.  A.  Contreras,  S.  A.  Gill, 
K.  E.  Sieving,  and  two  anonymous  reviewers.  This  pa- 
per is  a contribution  (Journal  Series  No.  R- 11007)  of 
the  Florida  Agricultural  Experiment  Station,  Gaines- 
ville, Florida. 

LITERATURE  CITED 

Balda,  R.  P.,  M.  L.  Morrison,  and  T.  R.  Bement. 
1977.  Roosting  behavior  of  the  Pinon  Jay  in  au- 
tumn and  winter.  Auk  94:494-504. 

Beauchamp,  G.  1999.  The  evolution  of  communal 
roosting  in  birds:  origin  and  secondary  losses.  Be- 
havioral Ecology  10:675-687. 

Brooks,  M.  1932.  Carolina  Wrens  roosting  in  aban- 
doned hornets  nests.  Auk  49:223-224. 

Brooks,  M.  1936.  Winter  killing  of  Carolina  Wrens. 
Auk  53:449. 

du  Plessis,  M.  A.  and  J.  B.  Williams.  1994.  Com- 
munal cavity-roosting  in  cooperatively-breeding 
Green  Woodhoopoes:  consequences  for  energy 


SHORT  COMMUNICATIONS 


569 


expenditure  and  the  seasonal  pattern  of  mortality. 
Auk  111:292-299. 

Ferguson,  J.  W.  H.,  M.  J.  M.  Nijland,  and  N.  C.  Ben- 
nett. 2002.  Simple  roost  nests  confer  large  ener- 
getic savings  for  sparrow- weavers.  Journal  of 
Comparative  Physiology  B 172:137-143. 

Ferns,  P.  N.  1992.  Thermoregulatory  behavior  of  Rock 
Doves  roosting  in  the  Negev  Desert.  Journal  of 
Field  Ornithology  63:57-63. 

Gill,  S.  A.  2003.  Timing  and  duration  of  egg  laying 
in  duetting  Buff-breasted  Wrens.  Journal  of  Field 
Ornithology  74:31-36. 

Gill,  S.  A.  and  B.  J.  M.  Stutchbury.  2005.  Nest 
building  is  an  indicator  of  parental  quality  in  the 
monogamous  Neotropical  Buff-breasted  Wren 
(' Thryothorus  leucotis).  Auk  122:1169-1181. 

Haggerty,  T.  M.  and  E.  S.  Morton.  1995.  Carolina 
Wren  ( Thryothorus  ludovicianus).  The  Birds  of 
North  America,  no.  188. 

Haggerty,  T.  M.,  E.  S.  Morton,  and  R.  C.  Fleischer. 
2001.  Genetic  monogamy  in  Carolina  Wrens 
{Thryothorus  ludovicianus).  Auk  118:215-219. 

Hall,  M.  L.  2000.  The  function  of  duetting  in  magpie- 
larks:  conflict,  cooperation,  or  commitment?  An- 
imal Behaviour  60:667-677. 

Kellam,  J.  S.  2003.  Pair  bond  maintenance  in  Pileated 
Woodpeckers  at  roost  sites  during  autumn.  Wilson 
Bulletin  115:186-192. 

Laskey,  A.  R.  1948.  Some  nesting  data  on  the  Caro- 
lina Wren  at  Nashville,  Tennessee.  Bird-Banding 
19:101-121. 

Merola-Zwartjes,  M.  1998.  Metabolic  rate,  temper- 
ature regulation,  and  the  energetic  implications  of 
roost  nests  in  the  Bananaquit  ( Coereba  flaveola). 
Auk  115:780-786. 


Morton,  E.  S.  1982.  Grading,  discreteness,  redundan- 
cy, and  motivation-structural  rules.  Pages  1 83 — 
212  in  Acoustic  communication  in  birds,  vol.  1 
(D.  E.  Kroodsma  and  E.  H.  Miller,  Eds.).  Academ- 
ic Press,  New  York. 

Morton,  E.  S.  and  M.  D.  Shalter.  1977.  Vocal  re- 
sponse to  predators  in  pair-bonded  Carolina 
Wrens.  Condor  79:222-227. 

Nager,  R.  G.  and  A.  J.  van  Noordwijk.  1992.  Ener- 
getic limitation  in  the  egg-laying  period  of  Great 
Tits.  Proceedings  of  the  Royal  Society  of  London, 
Series  B 249:259-263. 

Robinson,  T.  R.,  W.  D.  Robinson,  and  E.  C.  Edwards. 
2000.  Breeding  ecology  and  nest-site  selection  of 
Song  Wrens  in  central  Panama.  Auk  1 17:345-354. 

Sharrock,  J.  T.  R.  1980.  Wren  apparently  building 
winter  roosting-nest.  British  Birds  73:106-107. 

Skutch,  A.  F.  1940.  Social  and  sleeping  habits  of  Cen- 
tral American  wrens.  Auk  57:293-312. 

Strain,  J.  G.  and  R.  L.  Mumme.  1988.  Effects  of  food 
supplementation,  song  playback,  and  temperature 
on  vocal  territorial  behavior  of  Carolina  Wrens. 
Auk  105:11-16. 

Tamar,  H.  1980.  Carolina  Wren  winter  roosting  box. 
Indiana  Audubon  Quarterly  58:97-102. 

Weeks,  H.  P,  Jr.  1994.  Pre-laying  nest  roosting  in  the 
Eastern  Phoebe:  an  energy-conserving  behavior? 
Journal  of  Field  Ornithology  65:52-57. 

Williams,  L.  1941.  Roosting  habits  of  the  Chestnut- 
backed  Chickadee  and  the  Bewick  Wren.  Condor 
43:274-285. 

Yom-Tov,  Y.  and  J.  Wright.  1993.  Effect  of  heating 
nest  boxes  on  egg  laying  in  the  Blue  Tit  {Parus 
caeruleus).  Auk  110:95-99. 


The  Wilson  Journal  of  Ornithology  1 18(4):569-570,  2006 

Bald  Eagle  Kills  Crow  Chasing  a Hawk 

Bruce  D.  Ostrow1 


ABSTRACT. — I report  predation  of  an  American 
Crow  ( Corvus  brachyrhyncos ) by  a Bald  Eagle  {Hal- 
iaeetus  leucocephalus ) in  Washington  state.  The  crow 
was  attacked  and  killed  while  it  was  chasing  a Red- 
tailed Hawk  {Buteo  jamaicensis).  To  the  best  of  my 
knowledge,  this  is  the  first  report  of  a bird  of  one  spe- 
cies killing  a bird  of  a second  species  that  was  chasing 
a bird  of  a third  species.  Received  15  September  2005, 
accepted  5 May  2006. 


On  8 August  2005,  along  with  five  other 


1 Dept,  of  Biology,  Grand  Valley  State  Univ.,  Allen- 
dale, MI  49401,  USA;  e-mail:  ostrowb@gvsu.edu 


observers,  I was  observing  a mature  Bald  Ea- 
gle ( Haliaeetus  leucocephalus ) at  Hammer- 
sley  Inlet  (47°  12'  N,  122°  56'  W)  near  Arca- 
dia in  Mason  County,  Washington,  while  in  a 
boat  drifting  in  the  middle  of  the  narrow  inlet. 
I was  using  an  8 X 30  monocular  to  observe 
the  eagle,  which  was  perched  in  a tree  on  the 
southeast  side  of  the  inlet,  ~ 1 00  m away  from 
the  boat. 

At  15:22  PST,  I noticed  a Red-tailed  Hawk 
{Buteo  jamaicensis ) and  an  American  Crow 
{Corvus  brachyrhyncos ) fly  out  of  the  trees  on 
the  northwest  side  of  the  inlet.  The  crow  was 
chasing  the  hawk  and  repeatedly  attacking  the 


570 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  4,  December  2006 


hawk’s  tail  from  above  with  its  bill  and  feet 
in  typical  mobbing  behavior.  The  hawk  and 
crow  were  flying  southeast  across  the  inlet  di- 
rectly toward  the  eagle.  When  the  hawk  and 
crow  were  halfway  across  the  inlet,  —50  m 
from  my  position,  the  eagle  flew  directly  at 
the  pair.  Just  as  the  eagle  reached  them,  the 
hawk  dived  out  of  the  way,  but  the  crow  did 
not  have  time  to  evade  the  eagle.  The  eagle 
grabbed  the  crow  head-on  with  its  talons,  kill- 
ing it  instantly.  The  hawk  flew  away  quickly 
to  the  northeast,  and  the  eagle  took  the  dead 
crow  to  the  southeast  bank  below  its  initial 
perch.  The  entire  sequence  of  events  occurred 
in  —10  sec. 

Within  1 min  of  landing,  the  eagle  flew 
away  to  the  northeast,  leaving  the  crow’s  car- 
cass on  the  bank.  I was  unable  to  ascertain 
whether  the  eagle  ate  any  of  the  crow  because 
the  carcass  was  hidden  from  view  and  the  ea- 
gle did  not  return  within  the  time  I remained 
in  the  area  (2  min).  I do  not  believe  that  the 
presence  of  our  boat  of  observers  influenced 
the  birds’  behaviors.  Their  flight  paths  were 
direct  and  they  were  actively  engaged  with 
each  other.  Also,  I doubt  that  our  presence 
scared  away  the  eagle  because  the  boat  was 
drifting  silently  and  was  out  of  sight  of  the 
eagle  when  the  bird  was  on  the  bank. 

Mobbing  is  a common  avian  response  used 
to  drive  away  larger  predators,  including  Bald 
Eagles  (Hayward  et  al.  1977).  Mobbing  can- 
not take  place  without  risk,  however,  as  some- 
times the  mobbing  bird  (including  crows)  is 
killed  by  the  bird  it  is  harassing  (reviewed  by 


Sordahl  1990).  To  the  best  of  my  knowledge, 
this  is  the  first  report  of  a bird  of  one  species 
killing  a bird  of  a second  species  that  was 
chasing  a bird  of  a third  species.  Southern 
(1970)  reported  a Northern  Harrier  {Circus  cy- 
aneus ) chasing  away  eight  crows  that  were 
mobbing  a Great  Horned  Owl  {Bubo  virgini- 
anus ),  but  none  of  the  crows  were  harmed. 
Rudebeck  (1951),  however,  reported  a Pere- 
grine Falcon  {Falco  peregrinus ) capturing  a 
Northern  Lapwing  {Vanellus  vanellus ) that 
had  been  harassing  the  author.  My  observa- 
tion, along  with  these  other  observations,  sug- 
gests that  a mobbing  bird  may  be  a relatively 
easy  target  for  predators,  as  it  is  otherwise 
preoccupied. 

ACKNOWLEDGMENTS 

I would  like  to  thank  M.  R Lombardo,  T.  A.  Sordahl, 
W.  E.  Southern,  and  an  anonymous  reviewer  for  crit- 
ically reading  this  manuscript.  O.  M.  Ostrow,  F.  Q. 
Ostrow,  G.  L.  Sass,  R.  P.  Royal,  and  A.  B.  Royal  also 
observed  this  event. 

LITERATURE  CITED 

Hayward,  J.  L.,  Jr.,  W.  H.  Gillett,  C.  J.  Amlaner, 
Jr.,  and  J.  F.  Stout.  1977.  Predation  on  gulls  by 
Bald  Eagles  in  Washington.  Auk  94:375. 
Rudebeck,  G.  1951.  The  choice  of  prey  and  modes  of 
hunting  of  predatory  birds  with  special  reference 
to  their  selective  effect.  Oikos  3:200-231. 
Sordahl,  T.  A.  1990.  The  risks  of  avian  mobbing  and 
distraction  behavior:  an  anecdotal  review.  Wilson 
Bulletin  102:349-352. 

Southern,  W.  E.  1970.  Marsh  Hawk  chases  crows 
mobbing  owl.  Wilson  Bulletin  82:98-99. 


SHORT  COMMUNICATIONS 


571 


The  Wilson  Journal  of  Ornithology  1 18(4):57 1—572,  2006 

Rapid  Beak-Swinging  Locomotion  in  the  Puerto  Rican  Spindalis 

Ernest  H.  Williams,  Jr.1’3  and  Lucy  Bunkley-Williams2 


ABSTRACT. — We  observed  a Puerto  Rican  Spin- 
dalis ( Spindalis  portoricensis,  Thraupidae)  rapidly 
move  through  an  area  of  dense  vines  by  grasping  vines 
in  its  beak  and  swinging  from  vine-to-vine  without  the 
use  of  its  wings  or  feet.  This  behavior  appears  to  be 
unique  in  birds.  Received  8 August  2005,  accepted  24 
April  2006. 


The  Puerto  Rican  Spindalis  {Spindalis  por- 
toricensis, Thraupidae)  is  a moderate-sized 
(16.5  cm)  tanager  endemic  to  Puerto  Rico  and 
its  eastern  islands.  It  occurs  commonly,  but 
rather  sporadically  (Bunkley-Williams  and 
Williams  2000),  in  forests  and  woodlands  at 
all  elevations  throughout  Puerto  Rico  (Raffae- 
le  1989,  American  Ornithologists’  Union 
1998). 

At  10:00  AST  on  1 1 April  2005,  EHW  ob- 
served an  adult  female  Puerto  Rican  Spindalis 
on  the  outskirts  of  the  University  of  Puerto 
Rico  campus  in  Mayagtiez,  Puerto  Rico  (18° 
12.85'  N,  67°  08.35'  W;  elevation  37  m).  The 
bird  flew  into  a large  grove  of  trumpet  trees 
( Cecropia  schreberiana,  Cecropiaceae)  <3  m 
away  from  the  observer;  because  the  ground 
sloped  downward  steeply  towards  and  into  the 
grove  and  the  bird  flew  from  upslope,  the  bird 
entered  the  trees  at  a height  of  approximately 
6 m without  changing  its  altitude.  It  flew  into 
an  area  (—1.5  X 2 m)  of  densely-packed  (— 2— 
10  cm  apart),  fine-stemmed  (4-7  mm  in  di- 
ameter) pudding  vines  {Cissus  verticillata , Vi- 
taceae)  hanging  from  a trumpet  tree.  The  vines 
were  denuded  of  leaves  due  to  a 2-month-long 
drought.  Without  slowing,  landing,  or  hover- 
ing, the  bird  grasped  one  of  the  vines  in  its 
beak,  ceased  flying,  and  its  momentum  swung 
it  into  the  dense  vines.  Then  it  released  the 
first  vine  and,  dropping  a few  centimeters, 


1 Dept,  of  Marine  Sciences,  Univ.  of  Puerto  Rico  at 
Mayagiiez,  P.O.  Box  908,  Lajas,  Puerto  Rico  00667- 
0908. 

2 Caribbean  Aquatic  Animal  Health  Project,  Dept,  of 
Biology,  Univ.  of  Puerto  Rico,  P.O.  Box  9012,  Ma- 
yagiiez,  Puerto  Rico  00861-9012. 

3 Corresponding  author;  e-mail: 
ewilliams@uprm.edu 


grasped  a second  vine.  The  bird  repeated  this 
action  moving  to  a third,  and  then  a fourth, 
vine.  In  this  manner,  it  passed  completely 
through  a 1.5-m-wide  area  of  densely  packed 
vines  in  less  than  4 sec  without  flapping  its 
wings  or  using  its  feet  to  grasp  the  vines. 
Without  hesitating  or  stopping,  the  bird  then 
flew  further  into  the  grove  of  trees. 

Rapid,  beak-swinging  locomotion  apparent- 
ly has  not  been  described  for  this  species,  or 
for  any  other  species  that  we  have  been  able 
to  determine.  Leek  (1972)  did  not  report  this 
behavior  while  observing  Puerto  Rican  Spin- 
dalis in  trumpet  trees  in  Puerto  Rico,  and  Isler 
and  Isler  (1987)  did  not  note  it  in  any  of  their 
tanager  accounts.  However,  Garrido  et  al. 
(1997)  suggested  that  very  little  is  known 
about  the  behavior  of  Spindalis  spp. 

The  described  behavior  allowed  the  bird  to 
move  through  densely  packed  vines  where 
wings  could  not  be  used  for  support  or  loco- 
motion. The  bird  did  not  appear  to  feed  on 
anything  within  the  vines,  was  not  being  pur- 
sued by  a predator,  and  did  not  collect  any 
nesting  material.  The  behavior  did  not  appear 
to  be  a mechanism  of  accident  avoidance  (i.e., 
crashing  into  the  dense  vines),  as  it  was  too 
rapid,  smoothly  coordinated,  and  complicated. 

Birds  will  sometimes  use  their  beaks  to  aid 
locomotion  on  land  (e.g.,  Turkey  Vultures: 
Vogel  1950;  Red-tailed  Tropicbirds  and 
White-tailed  Tropicbirds:  Lee  and  Walsh- 
McGehee  1998).  Birds  are  also  able  to  support 
their  body  weight  with,  and  swing  from,  their 
beak  while  grasping  onto  something  with  it 
(e.g.,  Law  1926,  Brazil  2002).  Birds  that  hang 
from  perches  (chickadees  and  titmice,  Paridae; 
cockatoos,  Cacatuidae;  kinglets,  Sylviidae;  lo- 
ries, Loriidae;  parrots,  Psittacidae)  are  well 
known  to  use  their  bill  as  a “third  foot”  to 
assist  in  climbing,  but  unlike  what  we  ob- 
served, it  is  a relatively  slow  action  (Zeefer 
and  Lindhe  Norberg  2002)  and  the  feet  are 
used.  Although  it  has  been  established  that 
birds  may  exhibit  a rapid,  swinging  locomo- 
tion with  the  aid  of  their  wings  and  feet  (e.g., 
Potter  2003),  our  observation  should  alert  oth- 


572 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


ers  to  look  for  additional  cases  of  swinging 
locomotion  without  use  of  the  wings  and  feet, 
in  both  the  Puerto  Rican  Spindalis  and  in  oth- 
er species. 

ACKNOWLEDGMENTS 

We  thank  G.  J.  Breckon  for  identifying  vegetation 
mentioned  in  this  manuscript,  and  A.  R.  Lewis  for  re- 
viewing this  note.  The  comments  of  three  anonymous 
reviewers  also  improved  the  paper. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1998.  Checklist  of 
North  American  birds,  7th  ed.  American  Orni- 
thologists’ Union,  Washington,  D.C. 

Brazil,  M.  2002.  Common  Raven  Corvus  corax  at 
play;  records  from  Japan.  Ornithological  Science 
1:150-152. 

Bunkley-Williams,  L.  and  E.  H.  Williams,  Jr.  2000. 
Unusual  nesting  and  occurrence  records  for  Gua- 
ma,  Puerto  Rico,  1975-1999.  El  Pitirre  12:92-94. 
Garrido,  O.  H.,  K.  C.  Parkes,  and  R.  Sutton.  1997. 


Taxonomy  of  the  Striped-headed  Tanager,  genus 
Spindalis  (Aves:  Thraupidae)  of  the  West  Indies. 
Wilson  Bulletin  109:561-594. 

Isler,  M.  L.  AND  P.  R.  Isler.  1987.  The  tanagers:  nat- 
ural history,  distribution,  and  identification. 
Smithsonian  Institution  Press,  Washington,  D.C. 

Law,  J.  E.  1926.  Green-tailed  Towhee  qualifies  in  in- 
telligence test.  Condor  28:133-134. 

Leck,  C.  F.  1972.  Observations  of  birds  at  cecropia 
trees  in  Puerto  Rico.  Wilson  Bulletin  84:498-500. 

Lee,  D.  S.  and  M.  Walsh-McGehee.  1998.  White- 
tailed Tropicbird  ( Phaethon  lepturus).  The  Birds 
of  North  America,  no.  353. 

Potter,  E.  F.  2003.  Male-female  interactions  by  Yel- 
low-bellied Sapsuckers  on  wintering  grounds. 
Chat  67:107-109. 

Raffaele,  H.  A.  1989.  A guide  to  the  birds  of  Puerto 
Rico  and  the  Virgin  Islands,  rev.  ed.  Princeton 
University  Press,  Princeton,  New  Jersey. 

Vogel,  H.  H.,  Jr.  1950.  Observations  on  social  behav- 
ior in  Turkey  Vultures.  Auk  67:210-267. 

Zeefer,  A.  and  U.  M.  Lindhe  Norberg.  2002.  Leg 
morphology  and  locomotion  in  birds:  require- 
ments for  force  and  speed  during  ankle  flexion. 
Journal  of  Experimental  Biology  206:1085-1097. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):572— 573,  2006 

American  Crow  Caches  Rabbit  Kits 

Justin  J.  Shew1’2 


ABSTRACT. — For  corvids,  the  decision  to  cache  is 
a complex  behavior  likely  influenced  by  many  inter- 
acting factors.  On  8 April  2004,  I observed  an  Amer- 
ican Crow  ( Corvus  brachyrhynchos ) caching  eastern 
cottontail  ( Sylvilagus  floridanus ) kits  taken  from  a rab- 
bit nest  on  the  Missouri  State  University  campus  in 
Springfield,  Missouri.  The  crow  cached  at  least  three 
kits  and  flew  away  with  at  least  one  other.  Caches  were 
covered  with  dead  leaves  and  landscape  mulch.  During 
the  ensuing  3-day  period,  some  caches  disappeared, 
were  partially  eaten,  or  were  moved  to  a different  near- 
by location.  To  my  knowledge,  this  is  the  first  docu- 
mented case  of  caching  numerous  rabbit  kits  from  a 
single  nest,  and  it  is  one  of  the  few  documented  cases 
of  cache-moving  by  American  Crows.  Received  29 
July  2005,  accepted  24  April  2006. 


Many  different  factors  influence  caching 
behavior  in  American  Crows  ( Corvus  bra- 
chyrhynchos),  including  food  value,  handling 


1 Dept,  of  Biology,  Missouri  State  Univ.,  Spring- 
field,  MO  65804,  USA. 

2 Current  address:  104  Bell  Canyon  Rd.,  Trabuco 
Canyon,  CA  92679,  USA;  e-mail: 
jjshew@hotmail.com 


time,  time  of  day,  perishability,  and  klepto- 
parasitism  (Cristol  2001).  American  Crows 
are  known  to  cache  various  nuts,  prey  (inver- 
tebrate and  vertebrate),  eggs,  dung,  and  car- 
rion items  for  later  consumption  (Phillips 
1978,  Conner  and  Williamson  1984,  Kilham 
1989,  Verbeek  and  Caffrey  2002).  Caches  are 
sometimes  covered  with  debris,  substrate,  or 
leaves  (Phillips  1978,  Conner  and  Williamson 
1984,  Kilham  1989). 

On  8 April  2004  at  approximately  17:00  CST 
( 1 8°  C)  while  walking  across  the  Missouri  State 
University  campus  in  Springfield,  Missouri  (37° 
11'  N,  93°  16'  W),  I observed  the  caching  be- 
havior of  an  American  Crow.  I heard  animal 
distress  calls,  which  came  from  an  almost  hair- 
less baby  mammal  that  the  crow  (approximately 
20-30  m away)  was  handling  in  its  bill.  Al- 
though this  bill-manipulation  period  was  short 
(—5-10  sec),  it  seemed  to  injure  the  animal  se- 
verely and  silence  its  distress  calls.  The  crow 
was  handling  the  prey  while  perched  on  top  of 
a small  concrete  sign  (—  1 m tall,  —25  cm  wide) 
on  a campus  lawn.  I slowly  approached  the 
crow  to  within  —5-8  m,  and  it  dropped  to  the 


SHORT  COMMUNICATIONS 


573 


ground,  quickly  picked  up  surrounding  dead 
leaves  and  sticks,  and  placed  them  over  the  prey 
item  (cache  #1).  I uncovered  the  cache  and  de- 
termined that  the  mammal  was  a rabbit  kit.  I re- 
covered the  cache,  leaving  it  in  its  original  lo- 
cation, and  continued  to  watch  the  crow  from 
approximately  30-40  m away. 

The  crow  flew  —20  m and  attended  a kit 
apparently  cached  earlier  (cache  #2)  in  a 
mulch  pile  under  a landscape  tree.  The  crow 
then  moved  this  cache  to  another  mulch  pile 
about  5-10  m away,  where  it  carefully  picked 
up  individual  pieces  of  mulch  and  laid  them 
over  the  cache.  Subsequently,  the  crow  pecked 
around  within  0-2  m of  the  cache  while  pick- 
ing up  other  bits  of  mulch  and  quickly  drop- 
ping them.  The  crow  then  flew  back  to  the 
concrete  sign,  probed  into  the  ground  with  its 
bill,  and  pulled  out  an  eastern  cottontail  ( Syl - 
vilagus  floridanus ) from  a rabbit  nest.  From 
there,  the  crow  flew  a few  meters  as  the  kit 
gave  distress  calls;  once  the  kit  became  silent, 
the  crow  cached  it  (cache  #3)  in  another 
mulch  pile  by  covering  it  with  mulch  and  de- 
bris. Soon  the  crow  flew  back  to  the  cottontail 
nest,  pulled  out  another  kit,  and  flew  north- 
west beyond  my  view.  After  a few  minutes,  a 
crow  flew  from  the  southwest  to  the  rabbit 
nest,  pulled  out  another  kit,  and  flew  off  in  the 
same  direction  as  before. 

After  another  few  minutes  had  passed,  a crow 
flew  to  the  rabbit  nest  again  and  probed  the  nest 
several  times,  pulling  out  only  nesting  material 
(dead  grass).  From  there,  it  went  to  the  first  kit 
(cache  #1),  uncovered  it,  and  began  tearing  up 
and  eating  the  prey.  At  approximately  17:20, 
this  crow  flew  away  and  no  crows  returned  for 
—5  min.  I then  confirmed  the  locations  of  cach- 
es #2  and  #3,  finding  that  kits  in  both  caches 
were  still  alive  and  thoroughly  covered  with 
mulch.  I also  searched  other  mulch  piles  in  the 
area,  but  found  no  other  caches.  At  18:45  the 
same  day,  the  two  caches  were  still  in  the  same 
locations. 

On  9 April  at  1 1 :00,  I returned  to  the  site  to 
verify  the  locations  of  caches  #2  and  #3.  The 
kits  in  caches  #3  (closest  to  the  cottontail  nest) 
and  #2  were  gone.  I scanned  other  nearby  mulch 
piles  and  found  a cached  kit  with  a majority  of 


its  posterior  missing.  This  half-eaten  cache  was 
5-10  m away  from  cache  #2.  At  14:00,  the  half- 
eaten  kit  was  in  the  same  location,  but  on  1 1 
April,  the  kit  remains  were  gone. 

To  my  knowledge,  this  is  the  first  obser- 
vation of  an  American  Crow  caching  eastern 
cottontail  kits  and  one  of  the  few  documented 
observations  of  a cache  being  stored  at  mul- 
tiple locations  (cache  #2).  The  kits  were  10 
cm  long  and  may  have  represented  valuable 
prey  items  for  a crow,  particularly  given  the 
cottontail  litter  size  of  four  to  five  kits  (Whi- 
taker 1996).  Similar  sightings  have  entailed  a 
crow  in  Florida  that  moved  a cached  snake 
(Kilham  1989)  and  a crow  in  Tennessee  that 
cached  four  live  gizzard  shad  ( Dorosoma  ce- 
pedianum ) in  beach  sand  (Phillips  1978).  Also 
similar  to  my  observations  was  that  of  crows 
on  a Texas  university  campus  caching  pecans 
and  then  tearing  up  the  nearby  grass  after  hid- 
ing the  caches  (Conner  and  Williamson  1984). 
The  purpose  of  these  post-caching  behaviors 
remains  unclear;  possibilities  include  creation 
of  landmarks  that  help  individuals  locate  their 
caches,  or  it  may  serve  to  disguise  caching 
behavior  from  potential  kleptoparasites.  My 
observation  illustrates  some  of  the  complexi- 
ties of  crow  behavior,  and  indicates  that  more 
field  studies  are  needed  to  determine  factors 
that  lead  to  and  affect  caching  behavior. 

ACKNOWLEDGMENTS 

I thank  C.  M.  Smith  for  encouraging  submission  of 
this  short  communication  and  three  anonymous  refer- 
ees whose  comments  improved  the  manuscript. 

LITERATURE  CITED 

Conner,  R.  N.  and  J.  H.  Williamson.  1984.  Food  stor- 
ing by  American  Crows.  Bulletin  of  the  Texas  Or- 
nithological Society  17:13-14. 

Cristol,  D.  A.  2001.  American  Crows  cache  less  pre- 
ferred walnuts.  Animal  Behaviour  62:331-336. 
Kilham,  L.  1989.  The  American  Crow  and  Common  Ra- 
ven. Texas  A&M  University  Press,  College  Station. 
Phillips,  R.  A.  1978.  Common  crow  observed  catch- 
ing living  fish.  The  Migrant  49:85-86. 

Verbeek,  N.  A.  M.  and  C.  Caffrey.  2002.  American 
Crow  ( Corvus  brachyrhynchos).  The  Birds  of 
North  America,  no.  647. 

Whitaker,  J.  O.,  Jr.  1996.  National  Audubon  Society 
field  guide  to  North  American  mammals.  Alfred 
A.  Knopf,  New  York. 


574 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


The  Wilson  Journal  of  Ornithology  1 18(4):574- 576,  2006 


First  Nesting  Record  of  the  Gray-crowned  Yellowthroat 
(■ Geothlypis  poliocephala)  in  the  United  States  since  1894 

Stephan  Lorenz,14  Chris  Butler,1 2  and  Jimmy  Paz3 4 


ABSTRACT. — A Gray-crowned  Yellowthroat 
{Geothlypis  poliocephala ) nest  was  discovered  in  Tex- 
as during  June  2005,  providing  the  first  documentation 
of  nesting  in  the  United  States  since  1894.  The  nest 
was  located  within  the  Sabal  Palm  Grove  Audubon 
Center  and  Sanctuary  in  Cameron  County,  but  was 
depredated  within  4 days  of  discovery.  Gray-crowned 
Yellowthroats  are  fairly  common  breeders  in  north- 
eastern Mexico,  but  are  currently  listed  as  accidental 
in  Texas.  The  future  of  this  species  in  the  United  States 
is  uncertain.  Received  7 November  2005,  accepted  22 
April  2006. 


The  Gray-crowned  Yellowthroat  {Geothlyp- 
is poliocephala ) is  a resident  species  ranging 
from  central  Sinaloa  and  south-central  Tamau- 
lipas,  Mexico,  to  western  Panama  (American 
Ornithologists’  Union  1998).  It  is  found  in 
open,  grassy  habitats,  often  with  scattered 
bushes  and  scrub  (Howell  and  Webb  1995). 
Before  the  turn  of  the  19th  century,  it  was  a 
fairly  common  breeding  species  in  extreme 
southern  Texas,  including  Cameron  and  Hi- 
dalgo counties  (Oberholser  1974,  American 
Ornithologists’  Union  1998).  From  May  1890 
through  May  1894,  for  example,  at  least  34 
specimens  were  collected  near  Brownsville, 
Texas  (Lockwood  and  Freeman  2004),  and  the 
population  may  have  persisted  into  the  late 
1920s  (Lockwood  and  Freeman  2004). 

Currently,  the  species  is  listed  as  accidental 
in  Texas  (Bryan  et  al.  2003),  as  the  last  doc- 
umented breeding  record  in  the  United  States 
dates  back  to  1894  in  Cameron  County,  Texas 
(Oberholser  1974).  Since  then,  however,  the 
species  has  been  reported  from  Cameron  and 
Hidalgo  counties  with  increasing  frequency. 


1 Dept,  of  Biology,  Univ.  of  Texas  at  Tyler,  3900 
University  Blvd.,  Tyler,  TX  75799,  USA. 

2 Dept  of  Biology,  Univ.  of  Central  Oklahoma,  100 
N.  University  Dr.,  Edmond,  OK  73034,  USA. 

3 Sabal  Palm  Audubon  Center  & Sanctuary,  P.O. 
Box  5169.  Brownsville,  TX  78523,  USA. 

4 Corresponding  author;  e-mail:  slorenz@mail.com 


Oberholser  (1974)  listed  records  from  1956, 
1959,  and  1965;  more  recently,  Kutac  (1998) 
and  Lockwood  (2000,  2001,  2005)  listed  re- 
cords from  1988,  1989,  1999,  2000  and  2005. 
In  1997,  a possible  breeding  pair  of  Gray- 
crowned  Yellowthroats  was  found  in  Webb 
County,  Texas  (Woodin  et  al.  1998).  Despite 
recent  sightings  of  singing  males,  however, 
breeding  had  not  been  confirmed  (Brush 
2005). 

The  reasons  for  the  species’  disappearance 
from  Texas  are  unclear.  Habitat  similar  to  that 
currently  occupied  by  breeding  Gray-crowned 
Yellowthroats  in  Mexico  and  Central  America 
is  still  available  in  the  Lower  Rio  Grande  Val- 
ley (Brush  2005).  Oberholser  (1974)  cites 
habitat  reduction  caused  by  development, 
shifts  in  agricultural  practices,  and  disappear- 
ance of  large  freshwater  marshes  as  possible 
reasons  for  the  species’  range  contraction.  Sa- 
bal Palm  Grove  Audubon  Center  and  Sanc- 
tuary in  Cameron  County,  Texas  (21°5LN, 
97°  25'  W),  is  a 213-ha  preserve  along  the  Rio 
Grande  that  protects  one  of  the  last  remaining 
stands  of  Rio  Grande  palmettos  {Sabal  mexi- 
cana ).  The  site  provides  habitat  for  a variety 
of  bird  species  at  the  northern  terminus  of 
their  ranges  in  the  Lower  Rio  Grande  Valley 
of  Texas. 

From  8 February  (Lockwood  2004)  through 
August  2004  (pers.  obs.),  a male  Gray- 
crowned  Yellowthroat  was  frequently  ob- 
served at  the  Sabal  Palm  Sanctuary.  After  Au- 
gust, the  bird  apparently  left  the  area,  but  re- 
turned on  8 December  2004  (Lockwood  et  al. 
2005)  and  remained  at  the  sanctuary  at  least 
through  July  2005  (pers.  obs.).  On  the  evening 
of  25  June  2005,  a Gray-crowned  Yellow- 
throat was  heard  singing  at  the  sanctuary  and, 
the  next  morning,  a Gray-crowned  Yellow- 
throat (presumably  the  male)  was  observed 
carrying  food  items  to  a nest  hidden  in  dense 
grass.  Another  bird  (presumably  the  female) 
was  flushed  from  the  nest  when  an  observer 


SHORT  COMMUNICATIONS 


575 


FIG.  1 . Male  Gray-crowned  Yellowthroat  ( Geothlypis  poliocephala ) captured  in  a mist  net  at  Sabal  Palm 
Sanctuary  in  Cameron  County,  Texas,  29  June  2005  (photograph  by  C.  Butler). 


approached  the  nest  site.  Later,  a Gray- 
crowned  Yellowthroat  was  again  flushed  from 
the  nest,  after  which  it  gave  sharp  chips  from 
nearby.  On  the  same  date,  both  birds  were  ob- 
served repeatedly  carrying  food  items  to  the 
nest.  During  5 hr  of  observation,  the  male 
sang  continuously  while  foraging,  primarily 
near  ground  level  or  in  dense  understory.  The 
song,  a musical  warble  without  a clear  pattern, 
was  reminiscent  of  a bunting  ( Passerina  spp.) 
song  and  decidedly  different  from  that  of  a 
Common  Yellowthroat  ( Geothlypis  trichas ). 
The  second  bird  was  observed  less  often, 
probably  because  it  was  on  the  nest. 

At  one  point,  extended  study  of  the  birds’ 
field  marks  was  possible  when  both  birds 
landed  near  the  grass  clump  that  concealed  the 
nest.  Both  were  medium-sized  warblers,  larger 
and  bulkier  than  Common  Yellowthroats  and 
with  longer  tails.  Their  culmens  were  curved 
and  their  lower  mandibles  were  flesh-colored 
(Fig.  1).  The  birds’  upper  sides  were  an  even, 
greenish-olive,  the  wings  lacked  any  pattern- 
ing or  wing  bars,  and  the  crowns  and  auricu- 
lars  were  washed  with  a slate-gray.  Their  un- 
der parts  were  predominantly  yellow,  brightest 
in  the  throat  area  and  faded  along  the  flanks, 
and  their  bellies  were  whitish.  Observers  also 
noted  that  the  birds  had  broken  eye  rings  and 


black  lores,  the  black  extending  slightly  onto 
the  face  and  creating  a black  smudge.  Gray- 
crowned  Yellowthroats  exhibit  only  limited 
age-  or  sex-related  plumage  dimorphism  (Sib- 
ley 2000)  and  the  only  variation  noticed  be- 
tween the  two  birds  was  the  amount  of  black 
extending  from  the  lores  onto  the  face.  The 
presumed  male  had  slightly  more  black  ex- 
tending up  and  above  the  eye,  obscuring  half 
of  the  upper  eye-ring  arc. 

The  birds’  nest  was  located  along  the  edge 
of  a dry  mesquite  ( Prosopis  glandulosa ) 
grassland  near  a tree-lined  resaca.  It  was  0.3 
m above  ground  on  the  base  of  a dense  clump 
of  grass  ( Panicum  sp.)  and  constructed  mainly 
of  grasses,  which  is  consistent  with  published 
descriptions  of  the  species’  nesting  habits 
(Oberholser  1974,  Howell  and  Webb  1995, 
Dunn  and  Garrett  1997,  Martinez  et  al.  2004). 
Baicich  and  Harrison  (1997)  describe  the  spe- 
cies’ nest  as  a stout  cup  of  dry  grasses  and 
dead  leaves  built  atop  a grass  tussock.  When 
discovered,  the  Sabal  Palm  Sanctuary  nest 
contained  four  recently  hatched  nestlings,  rep- 
resenting a clutch  size  typical  for  Gray- 
crowned  Yellowthroats  (3-5  eggs,  usually  4; 
Oberholser  1974,  Baicich  and  Harrison  1997). 
The  hatchlings  had  blackish  down  on  top  of 
their  heads  and  their  eyes  were  still  closed. 


576 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  4,  December  2006 


On  29  June,  a Gray-crowned  Yellowthroat 
was  inadvertently  caught  in  a mist  net  set  up 
as  part  of  an  ongoing  study  on  the  population 
biology  of  “Brownsville”  Common  Yellow- 
throats  ( Geothlypis  trichas  insperata ) and  lo- 
cated approximately  10  m from  the  Gray- 
crowned  Yellowthroat  nest  found  at  Sabal 
Palm  Sanctuary.  The  bird  was  identified  as  a 
male,  based  on  a pronounced  cloacal  protu- 
berance and  more  extensive  black  on  the 
lores,  and  was  believed  to  be  the  male  of  the 
nesting  pair  (Fig.  1).  The  plumage  character- 
istics were  consistent  with  those  of  an  after- 
hatch-year bird.  Wing  length  was  56  mm  and 
mass  was  12.9  g,  both  somewhat  greater  than 
the  mean  wing  length  (53.6  ± 0.5  SE,  n = 9) 
and  mean  mass  (10.6  ± 0.3  SE,  n = 9)  of 
“Brownsville”  Common  Yellowthroats  (n  = 
9;  CB  unpubl.  data). 

On  the  morning  of  30  June,  the  nest  was 
checked  again,  but  apparently  it  had  been  dep- 
redated, as  all  nestlings  were  gone.  The  nest 
was  intact,  but  identification  of  the  predator 
species  would  be  purely  speculative. 

Identifying  Gray-crowned  Yellowthroats  in 
the  Rio  Grande  Valley  is  difficult  due  to  the 
possible  occurrences  of  Gray-crowned  X 
Common  Yellowthroat  hybrids.  A male  bird 
present  at  San  Ygnacio  in  Zapata  County, 
Texas,  from  1995  through  1996  was  appar- 
ently a hybrid,  and  he  paired  with  a female 
Common  Yellowthroat  (Dunn  and  Garrett 
1997).  On  several  visits  to  the  Sabal  Palm 
Sanctuary  in  March  and  April  2005,  Common 
Yellowthroats  had  been  observed  within  the 
area  used  by  the  pair  of  Gray-crowned  Yel- 
lowthroats; however,  no  interactions  between 
the  two  species  were  observed.  Although  we 
cannot  completely  rule  out  the  possibility  that 
either  of  the  Gray-crowned  Yellowthroats  at 
Sabal  Palm  Sanctuary  was  a hybrid,  the  field 
marks  and  song  indicated  that  both  birds  were 
“pure”  Gray-crowned  Yellowthroats. 

This  documents  the  first  Gray-crowned  Yel- 
lowthroat nest  detected  in  the  United  States 
since  1894.  The  current  breeding  site  deserves 
careful  monitoring  to  determine  the  species’ 
residency  status  and  prevent  human  distur- 
bance. Prescribed  burns  in  suitable  areas  (Ob- 
erholser  1974),  along  with  further  habitat  ac- 
quisition and  protection,  could  facilitate  the 


return  of  a breeding  population  to  the  United 
States. 

ACKNOWLEDGMENTS 

We  thank  C.  Cavazos  for  assisting  in  the  nest  search 
and  T.  Brush  and  J.  C.  Arvin  for  their  correspondence. 
We  would  also  like  to  thank  D.  W.  Pogue,  M.  W.  Lock- 
wood,  and  R.  L.  Gutberlet  for  their  help  with  this  man- 
uscript. Additionally,  we  thank  C.  E.  Shackelford  and 
two  anonymous  reviewers,  who  provided  helpful  com- 
ments. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1998.  Check-list 
of  North  American  birds,  7th  ed.  American  Or- 
nithologists’ Union,  Washington,  D.C. 

Baicich,  P.  J.  and  C.  J.  O.  Harrison.  1997.  A guide 
to  the  nests,  eggs,  and  nestlings  of  North  Ameri- 
can birds.  Academic  Press,  San  Diego,  California. 
Brush,  T.  2005.  Nesting  birds  of  a tropical  frontier,  the 
Lower  Rio  Grande  Valley  of  Texas.  Texas  A&M 
University  Press,  College  Station. 

Bryan,  K.,  T.  Gallucci,  G.  Lasley,  M.  Lockwood, 
and  D.  H.  Riskind.  2003.  A checklist  of  Texas 
birds.  Texas  Parks  & Wildlife  Press,  Austin. 
Dunn,  J.  L.  and  K.  L.  Garrett.  1997.  A field  guide 
to  warblers  of  North  America.  Houghton  Mifflin 
Company,  New  York. 

Howell,  S.  N.  G.  and  S.  Webb.  1995.  A guide  to  the 
birds  of  Mexico  and  northern  Central  America. 
Oxford  University  Press,  New  York. 

Kutac,  E.  A.  1998.  A birder’s  guide  to  Texas.  Gulf 
Publishing,  Houston,  Texas. 

Lockwood,  M.  W.  2000.  Texas  Bird  Records  Com- 
mittee report  for  2000.  Bulletin  of  the  Texas  Or- 
nithological Society  34:1-4. 

Lockwood,  M.  W.  2001.  Texas  Bird  Records  Com- 
mittee report  for  2001.  Bulletin  of  the  Texas  Or- 
nithological Society  35:1-10. 

Lockwood,  M.  W.  2004.  Texas.  North  American  Birds 
58:250-254. 

Lockwood,  M.  W.  2005.  Texas  Bird  Records  Com- 
mittee report  for  2004.  Bulletin  of  the  Texas  Or- 
nithological Society  38:21-28. 

Lockwood,  M.  W.  and  B.  Freeman.  2004.  TOS  hand- 
book of  Texas  birds.  Texas  A&M  University 
Press,  College  Station. 

Lockwood,  M.  W.,  R.  Pinkston,  and  W.  Sekula. 

2005.  Texas.  North  American  Birds  59:291-295. 
Martinez,  W.  E.,  V.  D.  Piaskowski,  and  M.  Teul. 
2004.  Reproductive  biology  of  the  Gray-crowned 
Yellowthroat  ( Geothlypis  poliocephala)  in  central 
Belize.  Ornitologia  Neotropical  15:155-162. 
Oberholser,  H.  C.  1974.  The  bird  life  of  Texas.  Uni- 
versity of  Texas  at  Austin,  Austin. 

Sibley,  D.  A.  2000.  The  Sibley  guide  to  birds.  Alfred 
A.  Knopf,  New  York. 

Woodin,  M.  C.,  M.  K.  Skoruppa,  and  G.  C.  Hickam. 
1998.  Breeding  bird  surveys  at  the  Galvan  Ranch, 
Webb  County,  Texas.  Ed  Rachel  Foundation,  Cor- 
pus Christi,  Texas. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):577-579,  2006 


Once  Upon  a ‘ Time  in 

Samuel  Hearne  (Fig.  1)  was  born  in  Lon- 
don, England,  in  1745.  In  1766  he  joined  the 
Hudson’s  Bay  Company  as  a seaman  and  mate 
of  the  Charlotte , sailing  out  of  Churchill  on 
Hudson  Bay,  Canada.  In  1771  he  was  the  first 
European  to  reach  the  Arctic  coast  of  North 
America,  traveling  on  foot  with  a group  of 
Chipewyan  Indians  from  Churchill  to  the 
mouth  of  the  Coppermine  River.  In  1774  he 
founded  the  first  inland  trading  post  of  the 
Hudson’s  Bay  Company  at  Cumberland 
House,  now  Saskatchewan’s  oldest  settlement. 

Ironically,  only  the  historians  appear  to 
have  appreciated  what  a great  naturalist  Hearne 
was.  In  his  introduction  to  the  1958  reprint  of 
Hearne ’s  book,  A Journey  from  Prince  of 
Wales ’s  Fort  in  Hudson ’s  Bay  to  the  Northern 
Ocean  (MacMillan  Company,  Toronto,  Ontar- 
io, 1958),  the  editor,  Richard  Glover,  correctly 
recognized  that  “Samuel  Hearne  was,  of 
course,  another  first  class  observer  and  re- 
porter . . . head  and  shoulders  superior  to  ev- 
ery other  North  American  naturalist  who  pre- 
ceded Audubon.” 

An  observer,  not  a collector,  Hearne  was  the 
first  to  give  a recognizable  description  of  the 
Ross’s  Goose,  named  Anser  rossii  by  John 
Cassin  some  80  years  later: 


{American  Ornithology 


FIG.  1.  This  portrait  of  Samuel  Hearne,  repro- 
duced with  permission  from  Stuart  Houston  (Houston, 
S.,  T.  Ball,  and  M.  Houston.  2003.  Eighteenth-Century 
Naturalists  of  Hudson  Bay.  McGill-Queen’s  University 
Press,  Montreal,  Quebec),  first  appeared  in  The  Euro- 
pean Magazine  in  1797  (original  artist  unknown). 


HORNED  WAVEY.  This  delicate  and  diminutive  species  of  the  Goose  is  not 
much  larger  than  the  Mallard  Duck.  Its  plumage  is  delicately  white,  except  the 
quill-feathers,  which  are  black.  The  bill  is  not  more  than  an  inch  long,  and  at  the 
base  is  studded  round  with  little  knobs  about  the  size  of  peas,  but  more  remarkably 
so  in  the  males.  Both  the  bill  and  feet  are  of  the  same  colour  with  those  of  the 
Snow  Goose.  The  species  is  very  scarce  at  Churchill  River,  and  I believe  are  never 
found  at  any  of  the  Southern  settlements;  but  about  two  or  three  hundred  miles  to 
the  North  West  of  Churchill,  I have  seen  them  in  as  large  flocks  as  the  Common 
Wavey,  or  Snow  Goose.  The  flesh  of  this  bird  is  exceedingly  delicate,  but  they  are 
so  small,  that  when  I was  on  my  journey  to  the  North  I eat  [ate]  two  of  them  one 
night  for  supper. 


As  the  quotation  above  illustrates,  many  of 
Hearne ’s  observations  were  practical  in  na- 
ture. Many  species  were  numerous  at  that 
time.  Similarly,  Hearne  noted  that  one  Indian 
could  kill  twenty  Spruce  Grouse  in  a day  with 
his  bow  and  arrow  and  some  would  kill  up- 
wards of  a hundred  Snow  Geese  in  a day, 


whereas  the  most  expert  of  the  English  hunt- 
ers would  think  it  a good  day’s  work  to  kill 
thirty.  At  Albany  Fort  in  one  season,  sixty 
hogsheads  (i.e.,  220-245  liters  each)  of  geese 
were  salted  away  for  winter  consumption. 
Hearne  also  mentioned  that  Arctic  Terns, 
which  he  ranked  as  being  among  “the  elegant 


577 


578 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


part  of  the  feathered  creation,”  occurred  in 
flocks  of  hundreds;  bushels  of  their  eggs  were 
taken  on  a tiny  island. 

Heame  once  saw  a flock  of  more  than  400 
Willow  Ptarmigan  near  the  Churchill  River. 
The  Indians  had  put  framed  nets  on  stakes  and 
placed  them  over  gravel  bait  to  entice  ptar- 
migan to  gather  under  the  net.  The  stake  was 
then  pulled  to  drop  the  net  on  top  of  the  birds. 
Using  this  method,  3 people  could  catch  up  to 
300  birds  in  1 morning;  in  the  winter  of  1786, 
Mr.  Prince  at  Churchill  caught  204  with  two 
separate  pulls.  Ptarmigan  feathers  made  ex- 
cellent beds;  the  feathers  sold  for  three  pence 
per  pound.  The  smaller  Rock  Ptarmigan 
would  not  go  under  nets,  but  up  to  120  could 
be  shot  in  a few  hours. 

In  Heame’s  time,  cranes,  curlews,  and  Pas- 
senger Pigeons  also  were  regularly  shot  for 


food;  the  latter  flew  in  large  flocks  in  the  in- 
terior near  Cumberland  House  where  Heame 
saw  12  killed  at  one  shot.  Whooping  Cranes, 
only  occasionally  seen,  most  often  occurred  in 
pairs.  He  indicated  that  this  largest  crane  was 
good  eating,  and  its  wing  bones  were  so  long 
and  large  that  they  were  sometimes  made  into 
flutes.  Heame  was  the  first  to  recognize  two 
different  species  of  curlew,  the  Hudsonian  and 
the  Eskimo.  He  also  provided  invaluable  in- 
formation concerning  the  northern  edge  of  the 
Eskimo  Curlew’s  breeding  range — Egg  River, 
on  the  west  coast  of  Hudson  Bay  at  59°  30' 
N,  about  150  miles  north  of  Churchill. 

Heame  combined  keen  powers  of  observa- 
tion with  a deep  appreciation  for  the  natural 
world.  His  observations  of  the  Ruffed  Grouse, 
although  precise  and  accurate,  also  convey  a 
real  sense  of  awe  and  wonder: 


THE  RUFFED  GROUSE.  This  is  the  most  beautiful  of  all  [grouse].  . . . They 
always  make  their  nests  on  the  ground,  generally  at  the  root  of  a tree,  and  lay  to 
the  number  of  twelve  or  fourteen  eggs.  . . . There  is  something  very  remarkable  in 
those  birds,  and  I believe  peculiar  to  themselves,  which  is  that  of  clapping  their 
wings  with  such  a force,  that  at  half  a mile  distance  it  resembles  thunder.  I have 
frequently  heard  them  make  that  noise  near  Cumberland  House  in  the  month  of 
May,  but  it  was  always  before  Sun-rise,  and  a little  after  Sun-set. 


Hearne  did  not,  however,  restrict  his  atten- 
tion to  edible  birds;  he  also  described  small 
birds,  such  as  the  chickadee,  or  the  ground 
nest  of  a White-crowned  Sparrow  at  the  root 
of  a dwarf  willow  or  a gooseberry.  He  under- 
stood the  concept  of  bird  migration,  describ- 
ing the  Trumpeter  Swan  as  the  first  species  of 
waterfowl  to  return  each  spring,  sometimes  as 
early  as  late  March,  and  frequenting  the  open 
waters  of  falls  and  rapids.  He  also  named 
year-round  residents,  such  as  the  Willow  Ptar- 
migan and  Arctic  Hare.  Hearne ’s  understand- 
ing of  sexual  dimorphism  showed  in  his  re- 
mark that  the  male  Willow  Ptarmigan  was 
larger  than  the  female.  His  description  of  the 
body-size  range  among  ptarmigans  demon- 
strates his  understanding  of  what  was  later  to 
be  described  as  Gaussian  distribution. 

Hearne  noted  that  the  pouch  at  the  base  of 
the  pelican’s  beak  had  a capacity  of  three 
quarts  and  that,  in  the  1770s  as  well  as  today, 
muskrat  houses  were  favorite  nesting  sites  for 
Canada  Geese.  He  evidently  was  the  first  to 
dissect  the  “windpipe”  of  an  adult  Trumpeter 
Swan,  noting  that  the  convoluted  trachea 


passed  into  the  broad  and  hollow  breastbone 
of  the  swan  and,  after  passing  the  length  of 
the  sternum,  returned  into  the  chest  to  join  the 
lungs.  He  also  dissected  a Tundra  Swan  but 
failed  to  appreciate  its  lack  of  the  extra  per- 
pendicular hump  in  the  trachea  that  is  present 
in  the  larger  Trumpeter  Swan. 

While  in  England  during  the  winter  of 
1782-1783,  Heame  met  Thomas  Pennant  and 
gave  him  a copy  of  his  natural  history  sight- 
ings, a dozen  years  in  advance  of  their  post- 
humous publication.  Pennant  incorporated  a 
number  of  Heame’s  observations  into  Arctic 
Zoology  (in  3 volumes,  Robert  Faulder,  Lon- 
don, 1792).  Five  years  after  retiring  to  Eng- 
land in  1787,  Hearne  sold  his  manuscript,  A 
Journey  from  Prince  of  Wales’s  Fort  in  Hud- 
son's Bay  to  the  Northern  Ocean,  to  a pub- 
lishing firm  in  London  (A.  Strahan  and  T. 
Cadell)  for  the  unprecedented  sum  of  £200. 
Only  a month  later,  when  only  47  years  old, 
Heame  died  “of  the  dropsy.”  His  book,  one 
of  the  greatest  travel  narratives  ever  written, 
appeared  in  print  posthumously  in  1795. 

From  my  point  of  view,  Heame’s  account 


ONCE  UPON  A TIME  IN  AMERICAN  ORNITHOLOGY 


579 


of  the  large  subspecies  of  Canada  Goose 
( Branta  canadensis  maxima)  best  reveals  his 
scientific  bent.  He  met  these  very  large  geese 
on  the  Barren  Grounds,  but  he  did  not  call 
them  Barren  Geese  because  they  summered 
there;  rather,  he  named  them  after  dissecting 
them  and  discovering  an  “exceeding  small- 
ness of  their  testicles.”  Heame’s  observation 
of  the  unusually  large  race  of  geese  with  small 
testicles  was  confirmed  more  than  a century 
and  a half  later  in  Harold  C.  Hanson’s  book, 
The  Giant  Canada  Goose  (Southern  Illinois 


University  Press,  Carbondale,  1965).  The 
book  detailed  how,  in  the  1960s,  Giant  Can- 
ada Geese  were  captured  and  banded  as  flight- 
less young  in  Rochester,  Minnesota,  southern 
Manitoba,  and  southern  Saskatchewan,  after 
which  they  traveled  north  1,600  km  to  molt 
(thus  arriving  later  in  the  year  than  the  breed- 
ing individuals).  Because  the  geese  were  too 
young  to  breed,  they  had  small  testicles.  This 
confirmed  the  phenomenon  that  Samuel  Heame, 
truly  one  of  the  most  talented  of  the  early 
North  American  naturalists,  noted  with  such 
insight: 


BARREN  GEESE.  These  are  the  largest  of  all  the  species  of  Geese  that  frequent 
Hudson’s  Bay,  as  they  frequently  weigh  sixteen  or  seventeen  pounds.  They  differ 
from  the  Common  Grey  Goose  in  nothing  but  size,  and  in  the  head  and  breast  being 
tinged  with  a rusty  brown.  They  never  make  their  appearance  in  the  Spring  till  the 
greatest  part  of  the  other  species  of  Geese  are  flown  Northward  to  breed,  and  many 
of  them  remain  near  Churchill  River  the  whole  summer.  This  large  species  are 
generally  found  to  be  male,  and  from  the  exceeding  smallness  of  their  testicles, 
they  are,  I suppose,  incapable  of  propagating  their  species. 


The  original  reference  for  this  piece  is  S. 
Heame,  1795,  A Journey  from  Prince  of  Wales's 
Fort  in  Hudson’s  Bay  to  the  Northern  Ocean, 
A.  Strahan  and  T.  Cadell,  London.  The  mod- 
em reference  is  S.  Houston,  T.  Ball,  and  M. 


Houston,  2003,  Eighteenth-Century  Naturalists 
of  Hudson  Bay,  McGill-Queen’s  University 
Press,  Montreal,  Quebec. — C.  STUART 
HOUSTON;  e-mail:  houstons@duke.usask. 
ca 


The  Wilson  Journal  of  Ornithology  1 18(4):580-585,  2006 


Ornithological  Literature 

Compiled  by  Mary  Gustafson 


FIRE  AND  AVIAN  ECOLOGY  IN 
NORTH  AMERICA.  By  Victoria  A.  Saab  and 
Hugh  D.  W.  Powell  (Eds.).  Studies  in  Avian 
Biology  no.  30,  Cooper  Ornithological  Soci- 
ety, Camarillo,  California.  2005:  vii  + 193 
pp.,  20  tables,  12  maps,  8 other  figs.  ISBN: 
0943610648.  $18.00  (paper). — Formerly  the 
purview  of  agency  personnel  and  a handful  of 
academics,  over  the  last  30  years  wildland  fire 
management  has  entered  the  mainstream  con- 
sciousness as  a topic  of  debate  and  interest. 
This  has  been  accompanied  by  a correspond- 
ing increase  in  attention  paid  by  ornithologists 
to  topics  on  fire  ecology.  This  volume  adds  to 
the  ever-growing  list  of  fire-related  papers  and 
books,  in  this  case  providing  a well  edited  and 
useful  literature  review  specifically  concerned 
with  the  effects  of  fire  and  fire  exclusion  on 
birds  and  their  habitats. 

This  work  is  largely  the  result  of  a Partners 
In  Flight  symposium  (held  in  2002)  that  fo- 
cused on  patterns  in  human  alteration  of  fire 
regimes  and  the  consequences  on  bird  popu- 
lations and  habitats.  The  introductory  chapter 
provides  an  overall  summary,  highlights  pat- 
terns, and  suggests  future  research  needs. 
While  not  a definitive  treatment  of  all  avian 
habitats  found  in  North  America,  discussion 
of  more  than  40  North  American  ecosystems 
provides  ample  opportunity  for  the  emergence 
of  some  broad  patterns  in  fire  regimes  and  avi- 
an responses.  For  example,  habitats  with  nat- 
urally long  fire-free  periods  have  been  less  af- 
fected by  fire  exclusion  practices  because  the 
period  of  fire  exclusion  is  not  markedly  dif- 
ferent from  the  normal  fire-return  interval. 

Ten  chapters  summarize  the  current  state  of 
knowledge  regarding  fire  and  birds  in  the 
southwestern  United  States,  California’s  oak 
woodlands,  the  maritime  Pacific  Northwest, 
sagebrush  habitats,  the  Rocky  Mountains,  the 
boreal  forests  of  Canada,  central  tallgrass  prai- 
ries, eastern  deciduous  forests,  grasslands  and 
shrublands  in  New  England,  and  southeastern 
pine  savannas  and  native  prairies.  Many  au- 
thors point  out  the  lack  of  fire-effects  data  for 
particular  habitats,  and  base  projected  fire  ef- 


fects on  what  is  known  about  general  avian 
habitat  associations  and  responses  to  habitat 
change,  or  on  the  results  of  fire  studies  in  sim- 
ilar habitats.  For  example,  although  fire  is  rel- 
atively common  in  California’s  oak  wood- 
lands, only  one  study  has  focused  on  the  ef- 
fects of  an  actual  fire  on  birds  in  that  system. 

In  total,  the  responses  of  more  than  200  bird 
species  to  fire  are  discussed,  with  some  pre- 
dictable outcomes.  For  example,  it  is  clear  that 
frequent  burning  creates  less  favorable  con- 
ditions for  forest  birds  that  nest  low  or  on  the 
ground,  and  that  foliage  gleaners  prefer  un- 
bumed  habitats.  The  predictability  of  a given 
species’  response,  however,  may  not  be 
straightforward:  it  may  vary  by  region  or  with 
differences  in  fire  size,  intensity,  frequency, 
and  seasonal  timing.  In  the  case  of  Greater 
Sage-Grouse  ( Centrocercus  urophasianus),  an 
objective  analysis  suggests  that  prescribed 
fire — although  often  touted — may  not  have 
been  overly  successful  as  a management  tool. 

Although  not  part  of  the  typical  Studies  in 
Avian  Biology  format,  an  index  summarizing 
the  effects  of  fire  on  different  species  would 
have  been  useful  to  workers  concentrating  on 
one  or  a few  bird  species.  All  1 1 chapters  are 
well-referenced,  as  evidenced  by  more  than 
900  sources  listed  in  the  Literature  Cited  sec- 
tion. Such  a hefty  Literature  Cited  section  on 
the  relatively  narrow  topic  of  fire  and  birds 
further  increases  this  work’s  utility  as  a ref- 
erence. 

Several  recurring  themes  appear  in  the 
chapters,  including  a call  for  additional  re- 
search— especially  experimental  work  on  fire 
effects,  which  makes  for  good  science  and  is 
entirely  feasible  in  many  prescribed  fire  sce- 
narios. Response  variables  should  focus  on 
avian  demographics,  rather  than  on  bird  abun- 
dance, as  is  the  case  in  many  previous  bird- 
fire  studies.  Well-stated  was  the  premise  that 
“understanding  past  fire  regimes  is  of  less 
practical  value  than  investigating  how  present- 
day  fires  fit  into  the  landscape  and  how  they 
can  be  used  to  achieve  management  objec- 
tives.” Given  the  clear  need  for  more  fire  on 


580 


ORNITHOLOGICAL  LITERATURE 


581 


the  landscape,  many  of  the  authors  suggest  an 
approach  to  using  prescribed  fire  that  does  not 
involve  burning  all  the  available  acres  in  a 
short  time  period,  but  rather  at  a variety  of 
temporal  and  spatial  scales  to  produce  a mo- 
saic of  different  habitat  and  age  classes.  This 
well-reasoned  approach  to  maintaining  varia- 
tion in  the  landscape  might  contrast  with  some 
practices,  such  as  the  large-scale  application 
of  frequent  understory  fires  (as  is  typical  in 
southwestern  pine  forests)  in  the  Rocky 
Mountains,  where  a stand-replacing  fire  might 
be  an  objective. 

Like  most  treatises  on  fire  ecology,  this  one 
makes  the  obligatory  call  for  less  fire  sup- 
pression with  statements  like  “.  . . it  clearly 
seems  reactive  to  continue  battling  naturally 
ignited  fires  burning  within  historic  ranges  of 
severity.”  Although  understandable,  such 
statements  fail  to  appreciate  the  current  im- 
practicality  of  letting  most  wildfires  bum,  con- 
sidering that  modem  wildlands  comprise  a 
complex  mix  of  fire-adapted  vegetation,  small 
remnant  patches  of  vulnerable  special  habitats 
(e.g.,  riparian  and  stands  of  old-growth  forest), 
areas  of  increased  flammability  due  to  the 
presence  of  exotic  plants  and  other  buildups 
of  fuels,  and  at-risk  investments  (e.g.,  conifer 
plantations  and  other  anthropogenic  improve- 
ments). Such  a landscape,  combined  with  dy- 
namic weather  patterns,  a political  atmosphere 
driven  by  special  interest  groups  (e.g.,  pro- 
ponents of  scenic  values  for  tourism),  public 
health  (e.g.,  smoke  management)  and  safety 
concerns,  and  an  increasingly  litigious  society 
make  risk-averse  decision  makers  unlikely  to 
push  too  hard  for  expanded  let-bum  policies 
any  time  soon.  While  many  authors  call  for 
expanded  prescribed  burning  programs,  large- 
scale  application  of  fire  as  the  primary  fuels 
treatment  could  only  be  done  with  massive 
(and  seemingly  unlikely)  increases  in  pre- 
scribed fire  budgets.  Thus,  although  fire  is  an 
appealing  treatment  for  ecosystem  restoration 
and  management,  it  seems  likely  that  mechan- 
ical thinning,  livestock  grazing,  and  other 
treatments  intended  as  surrogates  for  fire  will 
provide  land  managers  with  solutions  over  the 
short  run,  so  researchers  should  probably  look 
a bit  harder  at  such  options.  However,  since 
much  of  the  discussion  in  this  volume  deals 
with  responses  of  birds  to  habitat  change,  not 
necessarily  their  responses  to  fire,  per  se,  the 


information  provided  will  facilitate  planning 
for,  and  implementation  of,  a range  of  habitat 
treatments. 

In  light  of  the  ongoing  public  debate  re- 
garding forest  health  and  fire,  especially  wel- 
come was  a statement  contrasting  the  effects 
of  fuels  treatments  involving  commercial  har- 
vest of  large  trees  with  those  treatments  in- 
tended to  remove  highly  combustible,  small- 
diameter  fuels.  We  can  only  hope  that  forest 
managers  also  heed  the  cautions  provided  by 
many  authors  on  post-fire  salvage  logging, 
which  can  easily  reverse  any  benefits  the  bum 
may  have  provided  to  certain  groups  of  birds, 
especially  cavity-nesters. 

Fire  and  Avian  Ecology  in  North  America 
will  be  an  interesting  and  useful  addition  to 
the  reference  libraries  of  agency  biologists, 
fire  managers,  ecologists,  and  others  involved 
in  fire  and  fuels  issues.  I recommend  this 
book.— JOHN  E.  HUNTER,  U.S.  Fish  and 
Wildlife  Service,  Areata,  California;  e-mail: 
John_E_Hunter@fws.gov 


BIRDS  OF  WESTERN  AFRICA.  By  Nik 
Borrow  and  Ron  Demey.  Princeton  University 
Press,  Princeton,  New  Jersey.  2004:  512  pp., 
147  color  plates,  3,000+  color  illustrations. 
ISBN:  0691123217.  $40.00  (paper ).— Birds  of 
Western  Africa,  by  N.  Borrow  and  R.  Demey, 
was  originally  published  in  2001  by  Christo- 
pher Helm,  London  (hard  cover),  whereas  this 
volume  was  released  as  part  of  the  Princeton 
Field  Guide  series  (soft  cover).  This  magnifi- 
cent field  guide  covers  all  1,285  species  of 
birds  found  within  the  present  region  of  West- 
ern Africa,  which  the  authors  define  as  ex- 
tending from  Senegal  and  southern  Mauritania 
east  to  Chad  and  the  Central  Africa  Republic, 
and  south  to  Congo,  including  Cape  Verde  and 
the  Gulf  of  Guinea  islands.  A color-shaded 
map  shows  the  location  of  each  country. 

The  introduction  provides  information  on 
changes  to  scientific  and  common  names,  in- 
cluding standardizations  of  English  names, 
made  since  the  2001  publication.  Name 
changes  are  those  recommended  by  David  and 
Gosselin  (David,  N.  and  M.  Gosselin.  2002. 
Gender  agreement  of  avian  species  names. 
Bulletin  of  the  British  Ornithology  Club  122: 
257-282)  (David,  N.  and  M.  Gosselin.  2002. 


582 


THE  WILSON  JOURNAL  OL  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


The  grammatical  gender  of  avian  genera.  Bul- 
letin of  the  British  Ornithology  Club  122:14- 
49).  The  introduction  is  followed  by  an  ex- 
cellent review  of  the  climate,  topography,  hab- 
itats, and  restricted  ranges  of  certain  species; 
a glossary  of  terms;  and  excellent  illustrations 
and  descriptions  for  morphological  terminol- 
ogy. Western  Africa  has  no  fewer  than  87  re- 
stricted-range species  occurring  in  7 areas  of 
avian  endemism,  including  the  Cape  Verde  Is- 
lands, Annobon,  Sao  Tome,  Principe,  Upper 
Guinea  forests,  Cameroon  and  Gabon  low- 
lands, and  Cameroon  mountains.  Another  four 
species  are  considered  confined  to  restricted- 
range  areas  in  the  Upper  Niger  valley,  south- 
west Nigeria,  the  Lower  Niger  valley,  and  the 
Gabon-Cabinda  coastal  area.  For  the  regions 
noted  above,  the  authors  list  the  species  that 
are  highly  threatened. 

For  each  species,  the  authors  provide  a col- 
or distribution  map  and  authoritative  descrip- 
tions of  distinctive  characteristics  needed  to 
identify  the  species.  For  nearly  all  species, 
they  also  provide  color  illustrations  of  the  spe- 
cies. All  the  illustrations  in  this  compact  field 
guide  were  rendered  by  the  same  acclaimed 
bird  artist,  Nik  Borrow,  and  their  layout  is 
similar  to  that  of  the  Peterson  Field  Guides; 
however,  they  lack  Peterson’s  arrows  pointing 
out  distinctive  species  characteristics  that 
would  have  made  it  easier  to  identify  species 
in  the  field.  A unique  feature  of  this  book  is 
the  set  of  black  and  white  plates  illustrating 
nest  construction  for  20  species  of  weaver 
birds. 

This  is  the  first  field  guide  to  cover  Western 
Africa  exclusively,  and  it  should  enable  bird- 
ers to  identify  any  species  found  within  the  23 
countries  and  territories  covered  within  the 
text.  The  book  is  a concise,  authoritative,  and 
reasonably  priced  guide  available  from  a lead- 
ing university  publisher  that  employs  a critical 
review  system.  We  highly  recommend  this 
must-have  reference  for  anyone  interested  in 
the  birds  of  Western  Africa  or  concerned  with 
ornithology  on  a worldwide  basis,  and/or  for 
those  who  wish  to  augment  their  field  guide 
collection.  The  cover  design  is  attractive,  and 
should  catch  the  eye  of  bird  lovers.  The  pub- 
lishers should  be  commended  for  producing 
another  excellent,  reasonably  priced  mono- 
graph.—HARLAN  D.  WALLEY  and  PATRI- 
CIA A.  RUBACK,  Department  of  Biology, 


Northern  Illinois  University,  DeKalb,  Illinois; 
e-mail:  hdw@niu.edu  and  pattyruback@ 

hotmail.com 


RAPTORS  AND  OWLS  OF  GEORGIA. 
By  Rafael  A.  Galvez,  Lexo  Gavashelishvili, 
and  Zura  Javakhishvili.  Georgian  Centre  for 
the  Conservation  of  Wildlife  and  Buneba  Print 
Publishing,  Tibilsi,  Georgia.  Distributed  by 
NHBS,  United  Kingdom.  2005:  128  pp.,  47 
color  maps,  447  color  illustrations.  ISBN: 
9994077 18X.  £14.99  (paper).  [In  English  and 
Georgian] — This  is  the  first  field  guide  to  cov- 
er all  the  raptors  and  owls  recorded  in  Geor- 
gia, and  a first  for  the  Caucasus  region.  It  de- 
scribes the  45  raptor  species  recorded  in  the 
country,  including  the  breeding  species,  sea- 
sonal residents,  migrants,  and  rare  visitors. 
The  status  of  each  species  is  color-coded  on 
an  accompanying  distribution  map  of  Georgia. 

The  field  guide  has  a foreword  by  the  au- 
thors and  an  introductory  chapter  comprising 
several  sections,  the  first  of  which  is  a short 
explanation  of  raptor  classification.  This  is  un- 
usual in  that  it  includes  silhouettes  of  the  ma- 
jor families  of  raptors  and  owls  and  explains 
how  to  distinguish  them  in  the  field.  The  sec- 
tion on  “wing  attitudes”  is  especially  inter- 
esting because  it  shows  the  novice  what  to 
expect  in  the  field  under  different  weather 
conditions.  The  next  section  presents  a short 
description  of  raptor  migration  and  Georgia’s 
role  in  the  Palearctic  flyways.  There  is  also  a 
section  on  the  conservation  status  of  nocturnal 
and  diurnal  raptors  from  a continental  per- 
spective, with  a brief  history  of  Georgia  and 
a map  showing  the  locations  of  Georgia’s  27 
protected  areas.  The  section  on  how  to  use  the 
book  should  be  read  carefully  to  gain  a better 
understanding  of  the  maps  and  accompanying 
symbols  used  in  the  species  accounts. 

Following  the  introductory  chapter  are  the 
45  species  accounts.  Each  species  is  allocated 
a minimum  of  two  facing  pages.  Provided  on 
the  left  (text)  page  of  each  account  is  the  spe- 
cies’ common  name  (alternative  additional 
names  are  listed  parenthetically)  and  Latin 
names  (including  subspecies  inhabiting  Geor- 
gia), biometric  data  (body  length,  wingspan, 
and  body  mass),  and  the  known  or  extrapo- 
lated number  of  breeding  pairs  in  the  country. 


ORNITHOLOGICAL  LITERATURE 


583 


The  text  also  briefly  describes  the  species’  di- 
agnostic identification  features.  Here  the  au- 
thors have  been  very  innovative:  they  have 
emphasized  the  most  prominent  features  by 
underlining  them  and  pointing  to  them  in  the 
species’  illustration  on  the  facing  page;  a short 
comparison  with  potentially  confusing  species 
is  also  provided.  Additional  text  provides  an 
aid  to  a better  understanding  of  the  behaviors 
and  habitats  occupied  by  the  species.  Other 
natural  history  information  provided  includes 
the  species’  foods,  nest  characteristics,  clutch 
size,  egg  size  and  laying  period,  and  the  num- 
ber of  days  in  the  incubation  and  nestling  pe- 
riods; also  mentioned  is  how  many  years  it 
takes  an  individual  to  reach  sexual  maturity. 
Lastly,  the  authors  discuss  the  species’  con- 
servation status  and  population  trend  in  Geor- 
gia. A color-coded  map  shows  the  species’ 
year-round  distribution. 

The  facing  (illustration)  page  depicts  the 
species.  I found  it  very  instructive  that  the  au- 
thors chose  to  show  each  of  the  sexes  in  sep- 
arate columns  and,  where  relevant,  they  illus- 
trated different  morphs  at  different  ages.  Lines 
point  to  the  most  diagnostic  features  to  look 
for  during  field  observation.  I especially  en- 
joyed the  sketches  that  show  habitats  in  which 
the  species  should  be  found,  or  engaged  in 
some  unique  behavior,  and  the  fact  that — in- 
terspersed between  the  species  accounts — 
there  are  two  pages  of  field  drawings  of  spe- 
cies addressed  in  the  previous  pages.  These 
drawings  illustrate  habitats,  behaviors,  prey, 
inter-  and  intra-specific  interactions,  and  nest 
structures  and  locations. 

I greatly  appreciate  this  compact  field 
guide.  It  will  be  a good  companion  for  raptor 
watchers  who  will  find  that  it  is  relevant  not 
only  to  Georgia  but  also  to  most  of  the  neigh- 
boring countries  (i.e.,  all  of  the  Caucasus  re- 
gion). The  only  flaws  I found  in  the  book  were 
in  the  illustrations.  A few  of  the  drawings  con- 
tain errors,  including  some  that  do  not  cor- 
rectly depict  the  raptor’s  exact  “jizz”  and  pos- 
ture; examples  of  this  problem  may  be  found 
on  page  83  in  the  drawings  of  Honey  Buz- 
zards. I also  found  the  plates  too  dark.  I have 
handled  hundreds  of  raptors  every  year  for 
more  than  a decade  and  know  these  birds  up 
close — the  colors  of  most  are  not  as  dark  as 
they  are  in  the  illustrations.  This  criticism, 
however,  should  not  put  off  raptorphiles  or 


birdwatchers  that  need  a good  raptor  identifi- 
cation guide  for  that  part  of  the  world.  Fur- 
thermore, proceeds  from  the  sales  of  this  book 
are  donated  to  the  Georgian  Centre  for  the 
Conservation  of  Wildlife  and  to  conservation 
efforts  within  the  region.  On  the  whole,  this 
is  a worthwhile  undertaking  by  local  ornithol- 
ogists whose  worthy  endeavors  within  the  re- 
gion deserve  recognition. — REUVEN  YOSEF, 
International  Birding  and  Research  Center,  Eilat, 
Israel;  e-mail:  ryosef@eilatcity.co.il 


BIRDS  OF  MEXICO  AND  CENTRAL 
AMERICA.  By  Ber  Van  Perlo.  Princeton  Uni- 
versity Press,  Princeton,  New  Jersey.  2006: 
336  pp..  98  color  plates.  ISBN:  0691120706. 
$29.95  (paper). — The  format  of  this  newest 
guide  in  the  Princeton  Illustrated  Checklist  se- 
ries is  best  described  as  an  abbreviated  field 
guide  format.  The  guide  covers  Mexico  and 
all  of  Central  America  to  Panama — a vast  area 
containing  a huge  number  of  species  (1,574) 
to  illustrate  in  a single  guide.  The  98  color 
plates  have  thumbnail  illustrations  of  the  birds 
and  a brief  text  (on  the  facing,  or  a nearby, 
page).  Maps  showing  geographic  distributions 
follow  the  color  plates.  Other  than  an  index, 
table  of  contents,  and  brief  introductory  pages, 
that  is  the  total  extent  of  this  guide.  This  book 
should  not  be  viewed  as  a replacement  for 
books  like  Howell  and  Webb’s  excellent,  com- 
prehensive guide  (Howell,  S.  N.  G.  and  S. 
Webb.  1995.  A Guide  to  the  Birds  of  Mexico 
and  Northern  Central  America.  Oxford  Uni- 
versity Press,  New  York),  which  provides  a 
much  more  complete  account  for  each  species, 
including  in-depth  coverage  of  identification, 
distribution,  taxonomy,  vocalizations,  and 
more.  This  is  a compact  and  useful  guide  to 
tote  in  the  field,  however  it  only  complements 
rather  than  replaces  handbooks  like  Howell 
and  Webb’s  guide. 

The  plates  are  generally  well  done  and  il- 
lustrate all  species  found  in  the  area,  including 
hypothetical  or  rare  species,  whereas  the 
Howell  and  Webb  guide  omits  illustrations  of 
many  North  American  migratory  passerines 
and  provides  only  black  and  white  drawings 
for  some  waterbirds.  The  plates  in  Princeton’s 
Illustrated  Checklist,  however,  do  not  depict 
all  the  plumages  essential  for  identification; 


584 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


for  example,  immature  plumages  of  Great 
Blue  Heron  (Ardea  herodias).  Cooper’s  Hawk 
( Accipiter  cooperii),  Cedar  Waxwing  ( Bom - 
bycilla  cedrorum),  Loggerhead  ( Lanius  ludo- 
vicianus ) and  Northern  (L.  excubitor ) shrikes, 
Scissor-tailed  Flycatcher  ( Tyrannus  forfica- 
tus ),  Prairie  Warbler  ( Dendroica  discolor ), 
and  Chestnut-sided  Warbler  {Dendroica  pen- 
sylvanica ) are  not  included.  The  plates  and 
text  will  not  help  with  more  difficult  identifi- 
cations; even  adult  Glossy  {Plegadis  falcinel- 
lus ) and  White-faced  (P.  chihi ) ibis,  or  Rusty 
{Euphagus  carolinus ) and  Brewer’s  ( E . cyan- 
ocephalus ) blackbirds  will  be  unidentifiable  if 
only  this  guide  is  used.  The  immature  White- 
tailed Hawk  {Buteo  albicaudatus ) is  labeled  in 
the  text  as  not  identifiable,  and  the  illustration 
does  not  show  one  representative  age,  making 
correct  identification  impossible. 

Unlike  many  Latin  American  guides  that 
include  North  American  migrants,  the  plates 
of  these  species  are  fairly  well  done.  I could 
quibble  with  illustrations  like  that  of  the  Pine 
Siskin  ( Carduelis  pinus),  but  in  general  the 
illustrations  are  accurate.  Indications  of 
changes  in  scale  within  a given  plate  were  not 
provided;  for  example,  plate  77  depicts  Green 
Shrike- Vireo  ( Vireolanius  pulchellus ) and 
gnatcatchers  as  being  the  same  size.  On  an- 
other plate,  the  Red-winged  Blackbird  (Age- 
laius  phoeniceus ) male  and  female  are  the 
same  size  and  are  illustrated  as  larger  than  the 
Yellow-headed  Blackbird  {Xanthocephalus 
xanthocephalus).  Among  the  illustrations  of 
neotropical  species,  some  could  have  been 
better  portrayed  or  benefitted  from  another 
view.  For  example,  the  unique  tail  pattern  of 
the  Olivaceous  Piculet  ( Picumnus  olivaceus ) 
is  not  illustrated  or  described,  and  the  tuft  on 
the  Tufted  Flycatcher  ( Mitrephanes  phaeocer- 
cus ) is  very  weak  and  the  illustration  does  not 
look  much  like  the  species.  The  text  accom- 
panying the  illustrations  is  concise  and  pro- 
vides codes  for  range,  status  (endemic,  hy- 
pothetical, rare,  etc.),  and  seasonality.  There 
are  several  problems,  however,  including  a re- 
versed caption  or  plate  number  (Baltimore  [ Ic- 
terus galbula ] and  Orchard  [/.  spurious ] ori- 
oles), and  inappropriate  abbreviations  of  com- 
mon names  (e.g.,  “Grosbeak”  for  Blue  Gros- 
beak, Passerina  caerulea ).  Most  common 
names  and  taxonomy  follow  the  American  Or- 
nithologists’ Union,  but  there  are  exceptions. 


including  the  use  of  the  common  name  Gray 
Plover  for  Black-bellied  Plover  {Pluvials 
squatarola ) and  the  split  of  Stephen’s  (Mexi- 
can) Whip-poor-will  ( Caprimulgus  arizonae ) 
from  Whip-poor-will  {Caprimulgus  vocifer- 
us). 

The  text  for  each  plate  often  extends  to  the 
next  page,  adjacent  to  the  following  plate,  the 
text  for  which  then  also  runs  over  to  the  next 
page,  and  so  on  until  half  the  text  on  any  one 
page  may  pertain  to  the  current  plate  and  half 
to  the  preceding  plate.  Eventually  it  evens  out 
(or  additional  textual  pages  are  included),  but 
this  makes  the  guide  more  difficult  to  use  (al- 
beit slightly).  The  maps  are  understandably 
small,  as  there  are  27  maps  per  page,  each 
including  the  species’  name,  plate  number, 
and  the  species’  number  on  the  plate  to  aid 
cross-referencing  between  the  maps  and 
plates.  A neat  innovation  is  that  the  maps 
show  the  species’  detectabilities  (common  to 
frequent,  frequent  to  uncommon,  uncommon 
to  rare;  or  a percent  likelihood  of  detectabili- 
ty) and  status  (resident,  transient,  present  in 
northern  winter  or  northern  summer).  This  al- 
lows the  maps  to  convey  more  information 
than  just  presence/absence  for  a given  loca- 
tion, making  them  extremely  useful.  Locations 
of  rarities  or  isolated  populations  are  identi- 
fied with  cross  hairs  or  stars. 

I have  quibbled  over  some  issues  in  this 
guide,  but  I am  very  pleased  to  have  it  avail- 
able and  I will  give  it  the  greatest  complement 
I can  give  to  a field  guide;  I will  use  it.  I will 
carry  this  guide  in  the  field  and  leave  both  A 
Guide  to  the  Birds  of  Mexico  and  Northern 
Central  America  and  The  Sibley  Guide  to 
Birds  (Sibley,  D.  A.  2000.  The  Sibley  Guide 
to  Birds.  Alfred  A.  Knopf,  New  York.)  in  the 
car.  This  guide  will  be  especially  useful  for 
those  unfamiliar  with  the  North  American  mi- 
grants and  who  want  illustrations  of  the  mi- 
grant and  resident  birds  in  one  small  volume. 
If  this  guide  were  to  be  translated  into  Span- 
ish, it  would  become  the  standard  guide  for 
use  in  Mexico  and  Latin  America;  thus,  pub- 
lication of  a Spanish  version  should  be  a high 
priority  to  benefit  conservation  and  education 
in  the  region.- — MARY  GUSTAFSON,  Rio 
Grande  Joint  Venture,  Texas  Parks  and  Wild- 
life Department,  Mission.  Texas;  e-mail: 
mary.gustafson@tpwd.  state,  tx.  us 


ORNITHOLOGICAL  LITERATURE 


585 


FALCONRY  IN  LITERATURE.  By  David 
Horobin.  Hancock  House,  Surry,  British  Co- 
lumbia, Canada.  2004:  240  pp.,  1 color  draw- 
ing, numerous  line  drawings  and  sketches 
from  older  books,  21  black  & white  photo- 
graphs. ISBN:  0888395477.  $50.00  (cloth).— 
I am  not  a practicing  falconer,  nor  do  I have 
much  experience  in  falconry  as  a hobby,  a 
sport,  or  a trade.  I have  always  had  an  interest 
in  falconry,  however,  because  I have  been 
aware  of  its  historical  role  and  was  exposed 
to  it  by  some  of  the  most  respected  conser- 
vationists in  the  field. 

Falconry  today  is  a controversial  subject. 
This  is  especially  so  because  we  are  aware  of 
the  dangers  that  wild  populations  face,  and 
their  related  conservation  status  is  endangered 
by  those  who  have  the  financial  resources  to 
acquire  raptors.  The  high  prices  that  certain 
raptors  bring  in  falconry  circles,  and  the  trade 
in  eggs,  young,  and  birds  taken  from  the  wild, 
are  raising  a lot  of  questions  about  the  validity 
of  continuing  the  practice  of  falconry.  Few  are 
the  countries  where  falconry  is  regulated  by 


legislative  authorities  that  understand  the  sub- 
ject. 

Having  said  this,  I was  fascinated  by  this 
book.  It  brings  to  the  reader  writings  by  Eu- 
ropean poets  and  dramatists  of  the  Medieval 
and  Renaissance  periods.  The  book  opens  a 
window  to  how  falconry  was  perceived  in  the 
past  and  the  infatuation  of  the  aristocratic 
classes  with  birds  of  prey.  This  book  is  a clas- 
sical English  literature  review  of  texts  ranging 
“from  Chaucer  to  Marvell”  and  explores  the 
meaning  (and  confusion,  for  that  matter)  of 
falconry.  This  is  a book  for  the  intellect  that 
is  able  to  see  beyond  the  sport  of  flying  one’s 
raptor  and  provides  a perspective  on  the  his- 
tory in  which  the  sport  is  steeped.  The  au- 
thor’s knowledge  of  birds  and  their  natural 
history  is  presented  in  a very  scholarly  man- 
ner. I strongly  recommend  this  book  for  those 
practicing  falconers  who  like  a good  evening 
read  in  the  armchair — for  me  it  certainly  was 
a pleasant  change  from  the  current  television 
programming. — REUVEN  YOSEF,  Interna- 
tional Birding  & Research  Center,  Eilat,  Isra- 
el; e-mail:  ryosef@eilatcity.co.il 


The  Wilson  Journal  of  Ornithology  1 18(4):586— 592,  2006 


PROCEEDINGS  OF  THE  EIGHTY-SEVENTH  ANNUAL  MEETING 

SARA  R.  MORRIS,  SECRETARY 


The  eighty-seventh  annual  meeting  of  the  Wilson 
Ornithological  Society  (WOS)  was  held  Tuesday,  3 
October,  through  Saturday,  7 October  2006,  at  the 
World  Trade  Center  in  Veracruz,  Mexico,  in  joint  ses- 
sion with  the  American  Ornithologists’  Union;  Asso- 
ciation of  Field  Ornithologists;  Seccion  Mexicana  de 
Consejo  Internacional  para  la  Preservacion  de  las 
Aves,  A.  C.;  Cooper  Ornithological  Society;  Raptor 
Research  Foundation;  Society  of  Canadian  Ornitholo- 
gists/Societe  des  Ornithologistes  du  Canada;  and  Wa- 
terbird  Society.  This  joint  meeting,  the  fourth  quadren- 
nial meeting  of  professional  North  American  ornitho- 
logical societies,  was  called  the  IV  North  American 
Ornithological  Conference  (NAOC).  The  conference 
was  themed,  “Wings  Without  Borders/Alas  Sin  Fron- 
teras.’’  The  steering  committee  was  co-chaired  by 
Charles  M.  Frances  and  Jose  L.  Alcantara  and  included 
Bonnie  S.  Bowen,  Eduardo  E.  Inigo-Elias,  M.  Ross 
Lein,  Cecilia  Riley,  Betty  Ann  Schreiber,  and  Doris 
Watt.  Juan  E.  Martinez  Gomez  and  Ernesto  Ruelas  In- 
zunza  co-chaired  the  local  committee.  The  Conference 
Administration/Finance  Committee  co-chairs  were 
Bonnie  S.  Bowen,  Frank  B.  Gill,  and  Helen  Schneider 
Lemay;  the  Fundraising  Committee  co-chairs  were 
Frank  B.  Gill  and  Eduardo  E.  Inigo-Elias.  The  meeting 
was  co-hosted  by  the  Instituto  de  Ecologia,  A.C.;  Ve- 
racruz Visitors  and  Conventions  Bureau;  Consejo  Re- 
gulador  del  Cafe  Veracruz,  A.C.;  Universidad  Veracru- 
zana;  Island  Endemics  Foundation/Endemicos  Insula- 
res,  A.C.;  Municipality  of  Boca  del  Rio;  and  Consejo 
de  Promocion  Turfstica  de  Mexico. 

The  Council  met  from  13:33  to  17:43  CDT  in  the 
Centro  de  Negocios-2  room  of  the  Hotel  Galena  Plaza 
on  Monday,  2 October.  On  Tuesday,  3 October,  Hotel 
Mocambo  hosted  an  opening  reception  from  18:00  to 
22:00  on  the  terraces  and  around  the  pool.  Each  of  the 
next  four  mornings  began  with  a plenary  and  presen- 
tation of  different  society  awards  in  the  World  Trade 
Center  Ulua  Rooms  1-3.  Scientific  papers  were  pre- 
sented during  eight  concurrent  sessions  held  in  the  late 
mornings  and  afternoons  in  the  World  Trade  Center 
Ulua  and  Olmeca  Rooms.  Business  meetings  of  the 
individual  societies  were  conducted  in  the  early  eve- 
nings beginning  at  17:35.  Poster  sessions  were  held 
from  19:30  to  22:00  on  Wednesday,  4 October,  and 
Friday,  6 October. 

The  scientific  program  committee  was  co-chaired  by 
John  R.  Faaborg,  Juan  Francisco  Ornelas,  and  Maria 
del  Coro  Arizmendi.  The  U.S.  members  of  the  scien- 
tific program  committee  were  A1  Dufty,  Elizabeth  A. 
Schreiber,  George  Wallace,  Beth  Wallace,  Peter 
Lowther,  and  Steven  C.  Latta;  Mexican  members  of 
the  committee  were  Octavio  Rojas,  Carlos  Lara,  Flor 
Rodriguez,  Adolfo  G.  Navarro  S.,  Alejandro  Espinosa 
de  los  Monteros,  and  J.  Fernando  Villasenor  G.  The 


scientific  program  consisted  of  a total  of  1 ,239  presen- 
tations, including  4 plenary  talks,  336  oral  papers  con- 
tributed to  24  symposia,  368  oral  papers  in  38  general 
sessions,  and  531  poster  presentations  split  between 
two  poster  sessions,  each  of  which  was  divided  into 
46  different  topics.  Additionally,  there  were  17  work- 
shops organized  in  conjunction  with  the  conference. 
On  Thursday,  5 October,  Jed  Burtt  introduced  the  Mar- 
garet Morse  Nice  lecture,  which  was  the  conference 
plenary  on  that  day.  Jed  presented  the  biography  of 
Margaret  Morse  Nice  in  Spanish  and  introduced  the 
speaker,  Gary  Stiles,  in  English.  After  the  lecture.  Pres- 
ident Doris  Watt  presented  Gary  Stiles  with  the  Mar- 
garet Morse  Nice  medal. 

The  Student  Affairs  Committee — co-chaired  by  An- 
drea Cruz-Angon  and  James  W.  Rivers  and  including 
Eben  Paxton,  Doug  Robinson,  Julie  Garvin,  Jose  Luis 
Rangel-Salazar,  Vicki  Garcia,  Lori  Blanc,  Jackie 
Nooker,  and  Jean-Michel  DeVink — organized  a num- 
ber of  activities  for  students.  A Grant  Proposal  Work- 
shop was  held  on  Tuesday,  3 October.  The  professional 
societies,  including  WOS,  contributed  financial  sup- 
port for  a student-professional  ornithologist  social  on 
Thursday  evening.  The  social  was  followed  by  a Jeop- 
ardy-style quiz  bowl  for  nine  teams  of  three  students 
each,  which  was  played  energetically  and  boisterously, 
to  the  entertainment  of  all  assembled.  Students  also 
were  given  the  opportunity  to  be  matched  with  pro- 
fessional ornithologists  in  a student  mentoring  program 
that  provided  one-on-one  interaction  between  students 
and  researchers  in  their  areas  of  interest. 

A variety  of  field  trips  before,  during,  and  after  the 
conference  delighted  participants  with  opportunities  to 
see  Mexican  resident  and  Neotropical  migratory  birds. 
Daily  trips  during  the  conference  took  participants  to 
the  State  Park  Arroyo  Moreno  to  see  the  mangroves 
and  to  Cardel  and  Chichicaxtle  to  see  migrating  rap- 
tors. Four-day,  pre-  and  post-conference  tours  included 
birding,  cultural,  and  archaeological  sites;  birds  and 
butterflies  of  lowlands  and  highlands  in  Central  Vera- 
cruz; Catemaco  rainforest;  Veracruz  coffee  plantations 
and  highlands;  and  Veracruz  highlands.  One-day  trips 
before  the  conference  were  made  to  coastal  habitats 
and  lowland  tropical  forest  at  La  Mancha  and  Quia- 
huiztlan,  mangroves  and  wetlands  of  Alvarado  and 
Tlacotalpan.  and  transition  zones  between  lowland  and 
cloud  forests  at  El  Mirador  and  Las  Canadas.  After  the 
conference,  day  trips  included  visits  to  conifer  forest 
and  cloud  forest  of  Las  Minas  and  Los  Humeros,  trop- 
ical rain  forest  of  Los  Tuxtlas,  and  cloud  forest  and 
Isthmus  plateau  in  Oaxaca,  the  state  bordering  Vera- 
cruz. 

The  conference  was  closed  by  a Fiesta  Jarocha — 
with  a social  hour,  a seated  dinner,  entertainment  by 
the  Universidad  Veracruzana,  including  Ballet  Folk- 


586 


ANNUAL  REPORT 


587 


lorico  and  music,  and  the  announcement  of  students 
receiving  student  presentation  awards  and  honorable 
mentions.  Although  the  final  announcements  were 
completed  at  21:35,  the  music  and  dancing  continued 
into  the  night. 

BUSINESS  MEETING 

President  Doris  Watt  called  the  business  meeting  to 
order  at  17:59  on  4 October  in  the  Olmeca-5  Room  of 
the  World  Trade  Center.  She  recognized  a quorum  and 
thanked  those  assembled  for  attending. 

Secretary  Morris  presented  a summary  of  the  Coun- 
cil meetings,  which  were  held  Saturday,  18  March,  at 
Hawk  Mountain  in  Pennsylvania  and  Monday,  2 Oc- 
tober, in  Veracruz.  As  of  September  2006,  the  Wilson 
membership  stood  at  1,937,  which  includes  268  stu- 
dents and  166  new  members.  We  also  have  417  insti- 
tutional subscriptions  to  the  Wilson  Journal  of  Orni- 
thology, which  is  down  from  463  last  year.  As  part  of 
the  Ornithological  Societies  of  North  America  (OSNA) 
report.  Council  learned  of  several  WOS  members  who 
passed  away  during  the  last  year,  and  Secretary  Morris 
asked  those  assembled  to  stand  while  she  read  the  fol- 
lowing names:  Stanley  H.  Anderson  (Laramie,  WY), 
Carl  N.  Becker  (St.  Petersburg,  FL),  Herbert  L.  Cilley 
(Center  Strafford,  NH),  James  F.  Clements  (Temecula, 
CA),  Abbot  S.  Gaunt  (Columbus,  OH),  A.  Durand 
Jones  (Estes  Park,  CO),  Frank  J.  Ligas  (Naples,  FL), 
Karl  H.  Maslowski  (Cincinnati,  OH),  Richard  T.  Paul 
(Tampa,  FL),  Mario  A.  Ramos  (Washington,  DC), 
Clayton  G.  Rudd  (Moose,  WY),  Haven  H.  Spencer 
(Dover,  MA),  Mardi  Stoffel  (Rochester  Hills,  MI),  and 
Jeff  Swinebroad  (Montgomery  Village,  MD). 

After  members  were  seated.  Secretary  Morris  com- 
mented that  the  Schneider  Group  continues  to  manage 
the  membership  and  executive  director  duties  for 
OSNA.  Membership  renewal  was  much  smoother  this 
year  and  the  renewal  notices  for  next  year  were  mailed 
recently,  but  please  let  one  of  the  Council  officers 
know  if  you  are  experiencing  difficulty  with  your 
membership.  Council  thanked  the  Investing  Trustees 
for  their  excellent  work  in  managing  the  investments, 
and  directed  them  to  continue  managing  the  WOS 
portfolio  for  total  return. 

The  Council  approved  offering  a free  membership 
to  students  who  are  not  currently  members  of  the  So- 
ciety and  who  attend  and  present  a paper  or  poster  at 
a WOS  annual  meeting  (one  that  is  not  held  in  con- 
junction with  the  American  Ornithologist  Union  and 
Cooper  Ornithological  Society).  Council  also  in- 
creased the  funds  allotted  for  student  travel  from 
$5,000  to  $10,000  for  the  North  American  Ornitholog- 
ical Conference  (NAOC),  which  funded  25  students  at 
$400  each.  Additionally,  Council  approved  a one-time 
contribution  of  $7,500  to  the  Ornithological  Council 
for  revisions  to  the  Guidelines  for  the  Use  of  Wild 
Birds  in  Research. 

The  Council  elected  Clait  Braun  as  editor  of  The 
Wilson  Journal  of  Ornithology  for  Volume  119.  Coun- 
cil expressed  sincere  gratitude  for  Jim  Sedgwick’s 


work  in  getting  the  Wilson  Bulletin  back  on  its  publi- 
cation schedule  and  steering  changes  that  resulted  in 
the  new  Wilson  Journal  of  Ornithology,  an  updated 
and  revitalized  journal.  Council  accepted  a recommen- 
dation to  appoint  associate  editors  for  the  journal. 
Council  also  approved  archiving  The  Wilson  Bulletin 
and  its  successor  The  Wilson  Journal  of  Ornithology, 
in  JSTOR  (Journal  Storage,  The  Scholarly  Journal  Ar- 
chive) and  approved  a licensing  agreement  with  EB- 
SCO  Information  Services  to  include  The  Wilson  Bul- 
letin and  its  successor.  The  Wilson  Journal  of  Orni- 
thology, in  their  database.  There  is  a three-year  lag 
between  publication  and  availability  on  JSTOR  and 
EBSCO.  President  Watt  has  established  a new  ad  hoc 
Web  site  Committee,  chaired  by  Bob  Curry,  to  spear- 
head an  updated  Web  presence  for  the  Society. 

At  the  2007  annual  meeting,  the  Society  will  present 
the  first  Klamm  Awards:  the  William  and  Nancy 
Klamm  Service  Award  and  the  Klamm  Outstanding 
Undergraduate  Student  Paper  Awards  (one  for  the  best 
oral  paper  and  a second  for  the  best  poster). 

The  Council  approved  the  following  future  meet- 
ings: 2007  in  the  Boston,  Massachusetts,  area,  hosted 
by  Massachusetts  Audubon;  2008  in  southern  Missis- 
sippi, hosted  by  Frank  Moore;  and  2009  in  Pittsburgh, 
Pennsylvania,  co-hosted  by  the  National  Aviary  and 
Powdermill  Avian  Research  Center  of  the  Carnegie 
Museum  of  Natural  History.  Council  also  approved  in- 
volvement in  the  planning  of  the  next  NAOC  and  Dale 
Kennedy  will  be  the  WOS  representative  on  the  plan- 
ning committee. 

Treasurer,  Melinda  Clark,  presented  her  Treasurer’s 
Report  and  Doris  Watt  presented  highlights  of  the  2005 
Editor’s  Report  from  Jim  Sedgwick  and  an  update 
from  Clait  Braun  on  the  establishment  of  the  new  ed- 
itorial office. 

Doris  Watt  presented  the  report  of  the  Nominating 
Committee,  chaired  by  Robert  C.  Beason  and  includ- 
ing Mary  Bomberger  Brown,  Sara  R.  Morris,  and  Tim- 
othy J.  O’Connell.  The  committee  recommended  the 
following  slate  of  candidates:  President,  Doris  J.  Watt; 
First  Vice-President,  James  D.  Rising;  Second  Vice- 
President,  E.  Dale  Kennedy;  Secretary,  John  Small- 
wood and  W.  Herbert  Wilson;  Treasurer,  Melinda  M. 
Clark;  and  Members  of  Council  (2006-2009),  Carla  J. 
Dove,  Greg  H.  Farley,  and  Mia  R.  Revels.  President 
Watt  thanked  the  nominating  committee  and  asked  for 
any  nominations  from  the  floor.  Hearing  none,  she 
closed  nominations  following  a motion  by  Jerry  Jack- 
son,  seconded  by  Peter  Stettenheim.  Judy  McIntyre 
moved  that  the  Secretary  cast  a single  ballot  for  the 
slate  of  unopposed  candidates;  Bob  Curry  seconded 
that  motion,  which  passed  unanimously.  Secretary 
Morris  cast  the  ballot,  electing  those  officers  and  coun- 
cil members.  John  Smallwood  was  elected  Secretary 
by  paper  ballots  of  the  membership. 

The  Society’s  awards  (see  below)  were  announced 
during  the  business  meeting  (except  for  the  student 
presentation  awards,  which  were  announced  at  the  ban- 
quet). President  Doris  Watt  announced  the  Edwards 
Prize  recipients  for  2004  and  2005.  Secretary  Sara 


588 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  4,  December  2006 


Morris  announced  the  recipients  of  the  research 
awards.  Bob  Curry  presented  a commendation,  which 
was  approved  by  the  membership  by  acclamation  fol- 
lowing a motion  by  Chan  Robbins,  seconded  by  Mary 
Bomberger  Brown. 

President  Watt  adjourned  the  meeting  at  18:30  after 
a motion  from  Tim  O’Connell,  which  Jerry  Jackson 
seconded. 

MARGARET  MORSE  NICE  MEDAL 

(for  the  2006  WOS  plenary  lecture) 

Dr.  E Gary  Stiles,  “Ornithology  in  a troubled  coun- 
try: progress,  problems,  and  recent  work  on  nec- 
tar-feeding birds.” 

EDWARDS  PRIZE 

(for  the  best  major  article  in  volume  116  of 
The  Wilson  Bulletin ) 

Carolyn  B.  Meyer,  Sherri  L.  Miller,  and  C.  John 
Ralph,  “Stand-scale  habitat  associations  across  a 
large  geographic  region  of  an  old-growth  special- 
ist, the  Marbled  Murrelet,”  Wilson  Bulletin  116: 
197-210. 

EDWARDS  PRIZE 

(for  the  best  major  article  in  volume  117  of 
The  Wilson  Bulletin ) 

J.  Daniel  Lambert,  Kent  P.  McFarland,  Christopher 
C.  Rimmer,  Steven  D.  Faccio,  and  Jonathan  L. 
Atwood,  “A  practical  model  of  Bicknell’s  Thrush 
distribution  in  the  Northeastern  United  States,” 
Wilson  Bulletin  117:1-11. 

LOUIS  AGASSIZ  FUERTES  AWARD 

Chris  Merkord,  University  of  Missouri-Columbia, 
“Altitudinal  migration  in  the  Andes  of  southeast- 
ern Peru.” 

PAUL  A.  STEWART  AWARDS 

Kathleen  Coates,  Purdue  University,  “Swamp  Spar- 
row ( Melospiza  georgiana ) population  dynamics 
and  breeding  bird  communities  at  restored  and 
natural  marshes.” 

Kristen  M.  Covino,  University  of  Maine-Orono, 
“The  influence  of  an  ecological  barrier  on  direc- 
tional decisions  of  nocturnal  migrants.” 

Ana  Maria  Gabela,  University  of  Massachusetts— 
Amherst,  “Site  fidelity  and  human  impact  on  the 
Medium  Ground  Finch  (Geospiza  fortis)  on  Santa 
Cruz,  Galapagos  Islands.” 

Harry  R.  Jageman,  University  of  Idaho.  “Habitat 
use  and  ecology  of  Northern  Pygmy  Owls  ( Glau - 
cidium  gnoma ),” 

Alex  Jahn,  University  of  Florida,  “Testing  proxi- 
mate hypotheses  of  bird  migration  in  a forgotten 
migratory  system.” 

Jason  Townsend,  SUNY  College  of  Environmental 
Science  and  Forestry,  State  University  of  New 
York,  “The  role  of  sexual  segregation  in  the  win- 
ter ecology  of  the  Bicknell’s  Thrush.” 


GEORGE  A.  HALL/ 

HAROLD  F.  MAYFIELD  AWARD 

(formerly  the  Margaret  Morse  Nice  Award) 

Karla  Kinstler,  “Vocal  repertoire  of  the  Great 
Homed  Owl." 

Selection  committee  for  the  Nice  Medal — Charles 
Blem  (Chair).  Dale  Kennedy,  James  Rising,  and  Doris 
Watt;  for  the  Edwards  Prize — James  A.  Sedgwick 
(Chair);  for  the  Fuertes  and  Stewart  Awards — James 
D.  Rising  (Chair),  Clait  Braun.  Richard  Conner,  Dale 
Kennedy.  Alex  Mills.  David  Podlesak,  Craig  Rudolph, 
and  Doug  White;  and  for  the  Wilson  and  Lynds  Jones 
Prizes — Kevin  Omland  and  Katherine  Renton  (co- 
chairs  of  the  NAOC  Student  Presentation  Awards 
Committee  that  awarded  15  unranked  best  student  pa- 
pers at  the  conference).  The  recipients  of  the  WOS 
travel  Awards  were  chosen  by  the  NAOC  Student 
Awards  Committee,  composed  of  Matthias  Leu,  Mike 
Webster.  Patricia  Escalante.  Kim  Sullivan,  and  Tom 
Sherry. 

ALEXANDER  WILSON  PRIZE 

(for  a student  oral  paper,  one  of  15  unranked 
best  student  papers  presented  at  the  NAOC) 

Corey  E.  Tar  water.  University  of  Illinois  at  Urbana- 
Champaign,  “Life  history  implications  of  the 
post-fledging  period  in  a Neotropical  passerine./ 
Implicaciones  del  periodo  posterior  al  empluma- 
miento  para  la  historia  de  vida  de  un  ave  paserina 
Neotropical.” 

LYNDS  JONES  PRIZE 

(for  a student  poster,  one  of  15  unranked 
best  student  posters  presented  at  the  NAOC) 

Chris  J.  Clark,  University  of  Califomia-Berkeley, 
“Are  hummingbird  dive-noises  vocal  or  non-vo- 
cal ?/Los  ruidos  del  vuelo  en  picada  de  los  coli- 
bries  ^son  vocales  o no  vocales?” 

COMMENDATION 

WHEREAS,  the  WOS  held  its  2006  annual  meeting  in 
Veracruz,  Mexico,  as  part  of  the  fourth  NAOC;  and 
RECOGNIZING  that  the  conference  represents  one  of 
the  most  significant  ornithological  gatherings  in  his- 
tory, offering  members  of  the  WOS  opportunities  to 
socialize  and  share  scientific  information  about  birds 
with  ornithologists  from  throughout  North  America 
and  beyond;  and 

RECOGNIZING  that  this  unprecedented  event  has 
been  made  possible  only  by  the  dedicated  efforts  of 
a large,  international  group  of  ornithologists  and 
friends; 

THEREFORE  BE  IT  RESOLVED  that  the  WOS 
thanks  Juan  E.  Martinez  Gomez  and  Ernesto  Ruelas 
Inzunza,  the  rest  of  the  local  committee,  the  NAOC 
Steering  Committee  and  other  committees,  and  the 
Veracruz  community  for  making  the  conference  an 
extraordinarily  valuable  and  enjoyable  event. 


ANNUAL  REPORT 


589 


REPORT  OF  THE  TREASURER 

OPERATING  BUDGET  FOR  FISCAL  YEAR  2006  AND  2007 

2006  Budget  Amended  and  Approved  at  Council  Meeting,  18  March  2006 

2007  Budget  Amended  and  Approved  at  Council  Meeting,  2 October  2006 


2007 

Proposed 

Budget 

2006 

Annual 

Budget 

2005 

Actual 

Budget 

2005 

Annual 

Budget 

Revenue 

Contributions 

$ 

1,200 

$ 

1,000 

$ 

1,289 

$ 

— 

Student  Travel  Research  Fund 

— 

— 

126 

Van  Tyne  Library  Book  Fund 

— 

Sales — Back  Issues 

518 

— 

563 

Sales — Books  (Van  Tyne  Library) 

500 

900 

921 

Subscriptions 

17,317 

18,000 

18,769 

10,000 

Page  Charges 

15,506 

16,750 

16,615 

8,000 

Royalties 

3,409 

1,600 

1,688 

1,000 

BioOne  Electronic  Licensing 

10,760 

10,055 

10,055 

10,055 

Mailing  List  Rental  Income 

660 

500 

652 

Memberships 

31,332 

40,000 

37,499 

46,000 

Other  Income 

— 

2,000 

— 

4,000 

Total  Revenue  from  Operations 

$ 

81,202 

$ 

90,805 

$ 

88,176 

$ 

79,055 

Expenses 

Journal  Publication  Expenses 

Editorial  Honorarium 

$ 

4,000 

$ 

— 

$ 

— 

$ 

— 

Editor  Travel/Supplies 

1,000 

230 

226 

— 

Editorial  Assistance 

25,000 

55,000 

53,373 

55,000 

Copyright  Expense 

48 

50 

48 

— 

Printing — Journal 

64,400 

65,000 

64,336 

60,000 

Printing  Color  Plates 

2,400 

2,500 

2,472 

— 

Total  Journal  Expenses 

$ 

96,848 

$ 

122,780 

$ 

120,455 

$ 

115,000 

Operating  Expenses 

Postage  and  Mailing — Back  Issues 

$ 

440 

$ 

320 

$ 

312 

$ 

— 

Storage — Back  Issues 

680 

1,400 

1,379 

2,000 

Van  Tyne  Library  Expenses 

1,500 

1,500 

1,451 

4,000 

OSNA  Management  Services 

21,000 

21,000 

20,428 

25,000 

Credit  Card  Fees 

1,100 

1,200 

1,138 

— 

Travel  Expenses — OSNA  Representative 

1,500 

1,800 

1,758 

— 

Travel  Expenses — General 

450 

5,000 

2,465 

5,000 

Travel  Expenses — Ornithological  Council 

200 

900 

873 

— 

Meeting  Expenses 

1,000 

1,500 

10,170 

15,152 

Accounting  Fees 

4,500 

4,500 

3,627 

5,580 

Insurance — D&O 

1,425 

1,500 

1,401 

1,200 

Office  Supplies 

570 

300 

292 

1,000 

Postage — General 

260 

260 

254 

— 

Other  Expenses 

— 

250 

250 

— 

Filing  Fees 

5 

5 

5 

— 

Discretionary  Expenses 

3,000 

3,500 

— 

4,000 

Total  Operating  Expenses 

$ 

37,630 

$ 

44,935 

$ 

45,802 

$ 

62,932 

Awards 

Hall/Mayfield 

$ 

1,000 

$ 

1,000 

$ 

— 

$ 

1,000 

Stewart 

3,000 

3,000 

2,000 

2,000 

Fuertes 

2,500 

2,500 

2,500 

2,500 

Wilson,  Lynds  Jones,  Klamm 

1,200 

1,200 

500 

500 

Student  Travel  Grants 

5,000 

10,000 

2,600 

5,000 

Nice  Award  Expenses 

3,000 

6,800 

2,893 

5,800 

Total  Awards  Expenses 

$ 

8,000 

$ 

24,500 

$ 

10,493 

$ 

16,800 

590 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Contributions 


Support — Ornithological  Council 
Support — Ornithological  Council 

$ 

9,000 

$ 

9,000 

$ 

9,000 

$ 

9,000 

(restricted  to  revision  costs) 

7,500 

— 

— 

— 

American  Bird  Conservancy  Dues 
American  Association  for  Zoological 

250 

250 

— 

250 

Nomenclature  Dues 

250 

250 

— 

250 

Total  Contributions 

$ 

17,000 

$ 

9,500 

$ 

9,000 

$ 

9,500 

Total  Expenses 

$ 

159,478 

$ 

201,715 

$ 

185,750 

$ 

203,982 

Expenses  in  Excess  of  Revenue  Before 

Investment  Income 

$ 

(78,276) 

$ (110,910) 

$ 

(97,574) 

$ 

(124,927) 

Investment  Activity 

Revenue 

Investment  earnings  (budgeted) 

$ 

— 

$ 

70,000 

$ 

— 

$ 

126,718 

Realized  gain/loss — Merrill  Lynch 

23,612 

62,904 

Realized  gain/loss — Howland 

18,968 

47,045 

Realized  gain/loss — Sutton 

5,812 

13,034 

Unrealized  gain/loss — Merrill  Lynch 

36,722 

(51,548) 

Unrealized  gain/loss — Howland 

29,887 

(69,590) 

Unrealized  gain/loss — Sutton 

4,794 

(9,339) 

Investment  earnings — Merrill  Lynch 

20,000 

25,564 

Investment  earnings — Howland 

25,000 

46,575 

Investment  earnings — Sutton 

4,200 

3,731 

Total  Revenue  from  Investment 

Activity 

$ 

168,995 

$ 

70,000 

$ 

68,376 

$ 

126,718 

Investment  Fees 

25,091 

22,000 

21,660 



Investment  Revenue  in  Excess  of 

Expenses 

$ 

143,904 

$ 

48,000 

$ 

46,716 

$ 

126,718 

Total  Revenue  in  Excess  of  Expenses 

$ 

65,628 

$ 

(62,910) 

$ 

(50,858) 

$ 

1,791 

Investment  Principal  Needed  to  Cover 

Deficits 

62,910 

STATEMENT  OF  FINANCIAL  POSITION 

31  December  2005 


Assets 

Cash  Investments 

Merrill  Lynch — Cash  

Coamerica — Van  Tyne  Checking  

Van  Tyne  University  Michigan  Account  . 

Sutton  Fund — Cash  Equivalents  

Howland  Management — Cash  Equivalent 
Total  Cash  and  Cash  Equivalents 

Other  Investments 

Merrill  Lynch — Common  Stocks  

Merrill  Lynch — Corporate  Bonds  

Merrill  Lynch — Mutual  Funds  

Sutton  Fund — Equities 

Sutton  Fund — Corporate  Bonds  

Howland  Management — Equities 

Howland  Management — Fixed  Income  . . 

Total  Other  Investments  

Total  Assets  


$ (2,427) 

1,354 

353 

7,557 

118,397 

$ 125,233 


$ 689,356 

63,461 
26,982 
125,415 
10,033 
1,131,130 
301,914 

$ 2,348,291 

$ 2,473,524 


ANNUAL  REPORT 


591 


Fund  Balances 

Restricted  Funds — Sutton  Fund 

Unrestricted  Funds  

Net  Income  

Fund  Balance — Klamm  

Total  Fund  Balances  


143,005 

829,937 

(50,858) 

779,079 

1,551,441 

$ 2,473,524 


Melinda  Clark,  Treasurer 


EDITOR  S REPORT— 2005 

The  Wilson  Bulletin  Editorial  Office  received  a total 
of  162  manuscripts  during  2005  (compared  with  135 
in  2004  and  130  in  2003).  All  papers  received  three 
peer  reviews,  except  in  rare  instances  when  a referee 
failed  to  complete  and  return  a review  (<5%  of  cases). 
Correspondence  from  authors  and  referees  was  han- 
dled promptly  (usually  within  3 days  of  receipt).  I ac- 
cepted 18%  and  rejected  24%  of  manuscripts  received 
in  2005,  and  returned  the  remainder  (58%)  to  authors 
for  extensive  revision  or  revision  and  re-review.  Vol- 
ume 117  consisted  of  41  major  papers  and  20  short 
communications,  totaling  403  pages  (456  total  journal 
pages);  each  issue  had  a color  frontispiece.  Beginning 
with  the  June  2005  issue,  the  journal  has  been  pub- 
lished on  time.  The  median  time  from  receipt  to  pub- 
lication for  manuscripts  published  in  volume  117  was 
374  days.  The  dates  of  publication  for  the  issues  of 
volume  1 17  were  19  April,  21  June,  14  September,  and 
15  December  2005.  Except  for  the  original  submission 
of  manuscripts,  most  correspondence  and  document 
transmittal  between  The  Wilson  Bulletin  Editorial  Of- 
fice and  authors,  reviewers,  and  Allen  Press  was  elec- 
tronic. Design  changes  for  the  new  Wilson  Journal  of 
Ornithology  were  completed  in  2005. 

I am  grateful  to  Clait  Braun,  Richard  Conner,  Kath- 
leen Beal,  and  Karl  Miller,  who  served  on  the  Editorial 
Board  and  reviewed  numerous  manuscripts.  Kathy 
Beal  offered  statistical  critiques  of  several  manuscripts 
and  compiled  the  index.  Editorial  assistants  Beth  Dil- 
lon, Alison  Goffredi,  and  Cynthia  Melcher  performed 
essential  editorial  office  operations  including  mainte- 
nance of  the  e-mail  correspondence  tracking  system 
and  the  author/referee/manuscript  database;  corre- 
sponding with  authors  and  reviewers;  copy  editing; 
and  consulting  with  Allen  Press,  frontispiece  artists, 
and  other  editors.  I thank  Allen  Press,  especially  Karen 
Ridgway  and  Keith  Parsons,  for  guidance  and  helpful 
advice  on  the  final  stages  of  the  editorial  and  printing 
process.  The  U.S.  Geological  Survey  Fort  Collins  Sci- 
ence Center  has  continued  to  be  instrumental  in  its 
support  of  the  editorial  office. 

The  editorial  office  expenses,  publication  costs,  and 
income  for  volume  117  (2005)  were  as  follows:  (1) 
Editorial  Office  expenses  were  $56,197  (salaries: 
$52,919;  Editor’s  honorarium:  $2,000;  miscellaneous 
[office  supplies,  mailing]:  $394;  Editor’s  travel  to  2005 
WOS  meeting:  $883);  (2)  Publication  costs  (Allen 


Press)  were  $45,937.79;  and  (3)  Income:  authors  paid 
$2,427.36  in  page  charges  (403  manuscript  pages  were 
published  for  a mean  author  contribution  of  $6.02/ 
page). 

James  A.  Sedgwick,  Editor 
The  reports  of  the  standing  committees  are  as  follows: 


REPORT  OF  THE  JOSSELYN  VAN 
TYNE  MEMORIAL  LIBRARY 
COMMITTEE 

I am  very  pleased  to  submit  this  report  of  the  activ- 
ities at  the  Josselyn  Van  Tyne  Memorial  Library.  The 
following  library  transactions  occurred  over  the  past 
calendar  year: 

Loans: 

Loans  of  library  materials  to  members  involved  44 
transactions  to  13  members;  these  included  7 books 
loaned  and  131  articles  copied  and  scanned. 

Acquisitions: 

Exchanges:  A total  of  135  publications  were  re- 
ceived by  exchange  from  110  organizations  or  indi- 
viduals. 

Gifts:  We  received  28  publications  from  25  organi- 
zations. 

Subscriptions:  We  also  received  34  publications 
from  23  subscriptions.  We  spent  a total  of  $1,128.42 
on  subscriptions  in  2005. 

Donations:  Members  and  friends  donated  95  items. 
These  donations  included  1 book,  92  journal  issues, 
and  1 translation. 

Donors'.  The  four  members  and  friends  donating 
materials  include  Joseph  Jehl,  Jr.,  Sharon  Johnson,  Ed- 
ward H.  Miller,  and  Tim  Smart. 

Purchases:  New  items  purchased  for  $290.50  in- 
cluded 3 books  and  54  journal  issues. 

Dispersals: 

Gifts  to  other  institutions:  A total  of  19  journal  is- 
sues were  donated  to  The  Edward  Grey  Institute  for 
Field  Ornithology,  Oxford,  UK;  1,572  journal  issues 
were  sent  to  The  Peregrine  Fund  library,  for  the  cost 
of  postage;  and  147  journal  issues  were  sent  to  the 


592 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Point  Reyes  Bird  Observatory  library,  California,  for 
the  cost  of  postage. 

Back  issues:  We  sent  out  79  back  issues  of  The  Wil- 
son Bulletin  for  only  the  cost  of  postage. 

Duplicates:  We  sold  21  duplicate  books  for  $560.93. 

Accessibility  on  the  Web: 

Web  site:  The  Web  site  (http://www.ummz.lsa. 
umich.edu/birds/wos.html)  continues  to  provide  access 
to  the  library.  Journals  currently  received  are  listed  on 
the  site  as  well  as  instructions  for  accessing  the  Uni- 
versity of  Michigan’s  online  catalogue,  which  interest- 
ed people  can  use  to  check  holdings. 

Books  for  sale:  Our  Web  site  lists  duplicate  books 
for  sale. 

Journals  for  trade:  Also  listed  on  the  Web  site  are 
journals  available  for  sale  or  trade. 

Thank  Yous: 

Many  thanks  to  our  secretary,  Janet  Bell,  for  keep- 
ing the  library  loan  records  and  our  work-study  stu- 
dent, Rebecca  Carter,  for  copying  and  scanning  arti- 


cles, keeping  the  library  running,  and  mailing  out  back 
issues  of  The  Wilson  Bulletin. 

Janet  Hinshaw,  Librarian 

REPORT  OF  THE  CONSERVATION 
COMMITTEE 

In  response  to  a request  by  WOS  President,  Doris 
J.  Watt,  the  Conservation  Committee  was  re-estab- 
lished in  February  2006.  Committee  members  current- 
ly include  Daniel  Klem,  Jr.,  Joan  L.  Morrison,  John  A. 
Smallwood,  and  Douglas  W.  White.  The  committee 
will  assess  conservation  issues,  including  those 
brought  to  it  by  Council,  the  membership,  and  the  pub- 
lic at  large.  To  accomplish  this  charge,  the  committee 
expects  to  solicit,  as  needed,  input  from  those  with 
expertise  relevant  to  particular  issues.  The  committee 
looks  forward  to  working  closely  with  the  WOS  Res- 
olutions Committee,  and  to  making  recommendations 
for  consideration  by  the  WOS  Council. 

John  Smallwood,  Chair 
The  list  of  papers  and  posters  presented  at  the  NAOC 
meeting  will  be  published  in  a supplement  to  The  Auk, 
volume  124  (2007). 


The  Wilson  Journal  of  Ornithology  1 18(4):593-594,  2006 


REVIEWERS  FOR  VOLUME  1 1 8 


Referees  play  a critical  role  in  the  editorial  process.  Thoughtful,  incisive  reviews  are  paramount  in  the  main- 
tenance of  high  scientific  standards  and  journal  quality.  The  following  individuals  completed  and/or  agreed  to 
complete  a review  for  me  between  1 September  2005  and  31  August  2006  (referees  who  contributed  two  or 
more  reviews  appear  in  boldface).  The  Wilson  Ornithological  Society  and  the  editorial  staff  of  The  Wilson 
Journal  of  Ornithology  are  deeply  grateful  to  them  for  their  assessments  and  recommendations. — James  A. 
Sedgwick,  Editor. 


K.  Abraham,  P.  H.  Albers,  J.  C.  Alonso,  F.  K. 
Ammer,  E.  Ammon,  D.  E.  Andersen,  D.  J.  An- 
derson, G.  Angehr,  G.  W.  Archibald,  V.  Bag- 
lione,  F.  Bairlein,  R.  P.  Baida,  J.  Baribura,  J. 
Barlow,  J.  Bart,  L.  M.  Bautista,  K.  S.  Bawa, 
R.  C.  Beason,  A.  Bechet,  J.  C.  Bednarz,  M. 
A.  Belisle,  J.  R.  Belthoff,  D.  Berezanski,  K. 
Berg,  T.  M.  Bergin,  P.  Berthold,  R.  O.  Bier- 
regaard,  K.  L.  Bildstein,  C.  A.  Bishop,  J.  D. 
Bland,  R.  E.  Bleiweiss,  C.  E.  Bock,  W.  E. 
Boles,  S.  H.  Borges,  C.  Bosque,  F.  Botella,  M. 
Boulet,  J.  Boylan,  M.  J.  Braun,  J.  D.  Brawn, 

R.  M.  Brigham,  D.  J.  Brightsmith,  L.  Brotons, 

C.  R.  Brown,  M.  B.  Brown,  S.  T.  Buckland, 
A.  Buckley,  N.  J.  Buckley,  D.  A.  Buehler,  T. 
Bugnyar,  E.  L.  Bull,  L.  W.  Burger,  D.  E.  Bur- 
hans,  D.  Busby,  R.  W.  Butler,  B.  E.  Byers,  D. 
F.  Caccamise,  B.  Cade,  C.  D.  Cadena,  C.  L. 
Caffrey,  T.  W.  Campbell,  R.  J.  Cannings,  R. 
A.  Canterbury,  S.  W.  Cardiff,  M.  D.  Carey,  J. 
H.  Carter,  III,  J.  F.  Cavitt,  F.  Chavez-Ramirez, 

C.  Cicero,  D.  A.  Cimprich,  A.  P.  Clausen,  A. 
Cockburn,  M.  L.  Cody,  M.  Cohn-Haft,  N.  J. 
Collar,  J.  A.  Collazo,  M.  A.  Colwell,  S.  Co- 
nant,  J.  L.  Confer,  R.  N.  Conner,  C.  J.  Con- 
way, W.  C.  Conway,  S.  J.  Cooper,  W.  E.  Coo- 
per, N.  J.  Cordeiro,  J.  C.  Coulson,  M.  C.  Coul- 
ter, K.  A.  Crandall,  D.  A.  Cristol,  J.  P.  Croxall, 

L.  Cruz-Martinez,  P.  Cry  an,  S.  M.  Cutler,  T. 

D.  Dahmer,  A.  Datta,  C.  A.  Davis,  S.  K.  Da- 
vis, D.  K.  Dawson,  R.  D.  Dawson,  J.  B.  de 
Almeida,  D.  C.  Dearborn,  S.  L.  Deem,  T. 
De  Vault,  D.  R.  Diefenbach,  J.  J.  Dinsmore, 

S.  J.  Dinsmore,  P.  F.  Doherty,  Jr.,  A.  S.  Dol- 
by, S.  Droege,  K.  W.  Dufour,  K.  M.  Dugger, 

E.  H.  Dunn,  P.  O.  Dunn,  G.  Dutson,  J.  M. 
Eadie,  S.  D.  Emslie,  S.  Engel,  T.  K.  Engstrom, 

T.  C.  Erdman,  P.  Escalante,  D.  Evans,  W.  R. 
Evans,  D.  Evans-Mack,  J.  G.  Ewen,  J.  Faa- 
borg,  B.  C.  Faircloth,  A.  Farmer,  G.  L.  Farns- 
worth, P.  T.  Fauth,  J.  R.  Fellowes,  G.  Fernan- 
dez, C.  E.  Filardi,  R.  J.  Fisher,  J.  W.  Fitzpat- 
rick, R.  C.  Fleischer,  R.  J.  Fletcher,  Jr.,  M.  S. 


Foster,  J.  D.  Fraser,  P.  C.  Frederick,  M.  Gal- 
etti,  J.  Garcia-Moreno,  S.  A.  Gauthreaux,  F.  R. 
Gehlbach,  D.  D.  Gibson,  H.  G.  Gilchrist,  S. 
A.  Gill,  M.  E.  Gonzalez,  T.  P.  Good,  C.  E. 
Gordon,  P.  A.  Gowaty,  J.  B.  Grace,  M.  Green, 
J.  E.  Gross,  T.  C.  Grubb,  Jr.,  C.  G.  Gugliel- 
mo,  J.  A.  Guinan,  F.  S.  Guthery,  R.  J.  Gu- 
tierrez, J.  Ha,  J.  Haffer,  J.  C.  Hagar,  T.  M. 
Haggerty,  A.  J.  Hansen,  G.  M.  Haramis,  R. 
E.  Harness,  D.  A.  Haukos,  J.  Haydock,  S.  E. 
Hayslette,  J.  L.  Hayward,  P.  Heeb,  R.  Hen- 
geveld,  J.  R.  Herkert,  S.  K.  Herzog,  M.  R.  J. 
Hill,  K.  A.  Hobson,  R.  L.  Holberton,  R.  T. 
Holmes,  W.  H.  Howe,  G.  R.  Hunt,  L.  D.  Igl, 
W.  Iko,  M.  J.  Imber,  D.  J.  Ingold,  I.  Izhaki,  F. 

M.  Jaksic,  J.  M.  Jawor,  R.  K.  B.  Jenkins,  W. 
Jetz,  D.  H.  Johnson,  J.  A.  Jones,  J.  J.  Kap- 
pes,  Jr.,  G.  Katzir,  L.  F.  Keller,  J.  F.  Kelly,  B. 
Kempenaers,  P.  Kerlinger,  D.  I.  King,  T.  D. 
King,  S.  Kitamura,  F.  L.  Knopf,  W.  D.  Koenig, 
R.  R.  Koford,  P Koleff,  N.  Komar,  M.  Koop- 
man,  A.  W.  Kratter,  W.  B.  Kristan,  J.  A.  Kush- 
lan,  R.  Lanctot,  D.  B.  Lank,  M.  A.  Larson,  S. 
C.  Latta,  L.  Lefebvre,  D.  W.  Leger,  E.  Leh- 
koinen,  G.  Leonardi,  C.  A.  Lepczyk,  D.  J. 
Levey,  C.  A.  Lindell,  B.  C.  Livezey,  C.  Loeh- 
le,  B.  A.  Loiselle,  P.  E.  Lowther,  B.  C.  Lu- 
bow,  J.  R.  Lucas,  P.  M.  Lukacs,  G.  Luna-Jor- 
quera,  J.  J.  Lusk,  B.  E.  Lyon,  A.  D.  C. 
MacColl,  B.  F.  J.  Manly,  J.  S.  Marks,  K.  Mar- 
tin, L.  B.  Martin,  J.  M.  Marzluff,  M.  Massaro, 
R.  A.  Mauck,  H.  L.  Mays,  D.  G.  McAuley,  J. 
P.  McCarty,  W.  C.  McComb,  K.  G.  Mc- 
Cracken, W.  B.  McGillivray,  K.  J.  McGowan, 

N.  McIntyre,  J.  A.  McNeely,  S.  B.  McRae,  S. 
R.  McWilliams,  T.  D.  Meehan,  E.  H.  Miller, 
J.  R.  Miller,  K.  E.  Miller,  B.  Millsap,  D.  E. 
Mitchell,  D.  S.  Mizrahi,  A.  P.  Moller,  F.  R. 
Moore,  Y.  Mori,  S.  Morris,  R.  I.  G.  Morrison, 
E.  S.  Morton,  C.  Moskat,  M.  J.  Mossman,  A. 
M.  Mostrom,  C.  E.  Moulton,  L.  R.  Nagy,  K. 
Naoki,  S.  Naurin,  S.  A.  Nesbitt,  G.  L.  Neuch- 
terlein,  D.  L.  Neudorf,  K.  R.  Newlon,  W.  L. 


593 


594 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Nicholson,  V.  Nijman,  I.  C.  T.  Nisbet,  E.  Nol, 

F.  Olmos,  S.  Oppel,  L.  W.  Oring,  G.  W.  Page, 
W.  E.  Palmer,  K.  C.  Parsons,  M.  A.  Patten,  B. 
S.  Pedersen,  B.  D.  Peer,  C.  J.  Pennycuick,  N. 

G.  Perlut,  C.  M.  Perrins,  B.  G.  Peterjohn,  D. 

R.  Petit,  L.  J.  Petit,  M.  J.  Petrie,  P J.  Pietz,  B. 
Pinshow,  M.  A.  Pizo,  J.  H.  Plissner,  P.  Poon- 
swad,  R.  Poulin,  L.  A.  Powell,  T.  D.  Price,  K. 
L.  Purcell,  J.  S.  Quinn,  M.  G.  Raphael,  L.  M. 
Ratcliffe,  J.  T.  Ratti,  J.  M.  Reed,  S.  Reid,  J. 

V.  Remsen,  C.  Rengifo,  L.  M.  Renjifo,  M.  D. 
Reynolds,  T.  Z.  Riley,  C.  C.  Rimmer,  J.  D. 
Rising,  C.  S.  Robbins,  M.  B.  Robbins,  R.  J. 
Robel,  R.  J.  Robertson,  S.  Robinson,  R.  F. 
Rockwell,  N.  L.  Rodenhouse,  P.  G.  Rode- 
wald,  J.  A.  Rodgers,  Jr.,  A.  Rodriguez,  F.  C. 
Rohwer,  S.  A.  Rohwer,  J.  Rolstad,  S.  Roos,  S. 

S.  Rosenstock,  G.  V.  Roslik,  R.  R.  Roth,  S.  I. 
Rothstein,  A.  Roulin,  J.  M.  Ruth,  V.  A.  Saab, 
A.  Salinas-Melgoza,  D.  W.  Sample,  J.  A.  San- 
chez-Zapata,  F.  J.  Sanders,  J.  H.  Sarasola,  J. 
R.  Sauer,  J.-P  Savard,  R.  R.  Schaefer,  M. 
Schaub,  K.  A.  Schmidt,  T.  Schowalter,  E.  A. 
Schreiber,  M.  A.  Schroeder,  K.  L.  Schuch- 
mann,  T.  S.  Schulenberg,  S.  H.  Schweitzer, 

W.  A.  Searcy,  N.  Seddon,  B.  Semel,  F.  Sergio, 
C.  A.  Shackelford,  S.  Sharp,  W.  M.  Shields, 


W.  G.  Shriver,  D.  Shutler,  J.  G.  Sidle,  K.  E. 
Sieving,  K.  M.  Silvius,  S.  K.  Skagen,  T.  Slags- 
vold,  J.  A.  Smallwood,  N.  G.  Smith,  N.  F.  R. 
Snyder,  J.  J.  Soler,  T.  A.  Sordahl,  W.  E. 
Southern,  R.  Spaar,  T.  H.  Sparks,  J.  R.  Speak- 
man,  D.  A.  Spector,  J.  A.  Spendelow,  J.  R. 
Squires,  T.  R.  Stanley,  H.  Stein,  L.  Stemp- 
niewicz,  J.  A.  Stratford,  B.  M.  Strausberger, 
A.  Strong,  B.  J.  M.  Stutchbury,  D.  L.  Swan- 
son, T.  Swem,  C.  Swennen,  P.  A.  Szczys,  J. 
Y.  Takekawa,  K.  A.  Tarvin,  P.  B.  Taylor,  D.  R. 
Thompson,  R.  Thorstrom,  J.  M.  Tirpak,  D. 
Tome,  J.  Torok,  R.  Torres,  P.  Tryjanowski,  Y. 
Turcotte,  W.  Turner,  F.  Valera,  S.  van  Balen, 
F.  G.  Van  Dyke,  C.  van  Riper,  III,  E.  A. 
VanderWerf,  D.  Varland,  N.  A.  M.  Verbeek, 
P.  D.  Vickery,  F.  J.  Vilella,  P.  A.  Vohs,  N.  T. 
Vy,  Y.  Wang,  D.  M.  Watson,  P J.  Weather- 
head,  W.  C.  Webb,  A.  A.  Weller,  K.  S.  Wells, 

A.  D.  West,  D.  F.  Westneat,  N.  T.  Wheel- 
wright, C.  J.  Whelan.  C.  M.  White,  L.  A. 
Whittingham.  P.  Widen.  D.  S.  Wilcove,  J.  W. 
Wiley,  R.  H.  Wiley,  M.  F.  Willson,  W.  H.  Wil- 
son, M.  Winter,  M.  C.  Witmer,  S.  Wolf,  S. 
Woltmann,  M.  S.  Woodrey,  J.  T.  Wootton,  M. 

B.  Wunder,  R.  H.  Yahner,  S.  A.  Yaremych,  R. 
Ydenberg,  L.  Young,  C.  B.  Zavalaga,  M.  C. 
Zicus,  G.  S.  Zimmerman,  R.  M.  Zink. 


The  Wilson  Journal  of  Ornithology  1 1 8(4):595— 610,  2006 


Index  to  Volume  118,  2006 

Compiled  by  Rita  A.  Janssen  and  James  A.  Sedgwick 


This  index  includes  references  to  genera,  species,  authors,  and  key  words  or  terms.  In  addition  to  avian  species, 
references  are  made  to  the  scientific  names  of  all  vertebrates  mentioned  within  the  volume  and  other  taxa 
mentioned  prominently  in  the  text.  Nomenclature  follows  the  American  Ornithologists’  Union  Check-list  of 
North  American  Birds  (1998)  and  its  supplements.  Reference  is  made  to  books  reviewed  and  announcements  as 
they  appear  in  the  volume. 


A 

abundance 

effect  of  habitat  variables  in  southern  Appalachian 
wetlands  on,  399 

effect  of  understory  composition  on,  461 
of  Black- throated  Blue  Warbler,  461 
Acacia  spp.,  563 
Accipiter  badius,  50 
brevipes,  50,  476 
cooperii,  535 
faciatus,  cf.  307 

acoustic  components,  of  Greater  Sage-Grouse,  36 
Acrocephalus  scirpaceus,  371 
schoenobaenus,  191,  371 
Actitis  macularius,  221 
activity,  pre-migratory,  187 
adaptive  value  of  eggshell  removal,  59 
Aegolius  acadicus,  411-413 
funereus,  4 1 1 
age 

effect  on  singing  behavior  in  male  Setophaga  ruti- 
cilla,  439 

ratio,  effect  of  understory  composition  on,  461 
Agelaius  phoeniceus,  158,  331,  391-398,  416,  539 
aggregation,  164,  364 

Aguirre,  Ray,  see  Metz,  Steve  T.,  Kyle  B.  Melton, 

, Bret  A.  Collier,  T.  Wayne  Schwertner, 

Markus  J.  Peterson,  and  Nova  J.  Silvy 
Aimophila  aestivalis,  131-280,  138-144 
Aix  sponsa,  102 
Alkodon  spp.,  95 

Allen,  Deborah,  see  DeCandido,  Robert,  and 

allometry,  173 

Alvarez  A.,  Jose,  see  Lane,  Daniel  F.,  Thomas  Valqui 

H.,  , Jessica  Armenta,  and  Karen  Eck- 

hardt 

Amazona  aestiva,  233 
albifrons,  225 
autumnalis,  231 
barbadensis,  233 
finschi,  240 

leucocephala  bahamensis,  233 
ochrocephala  panamensis,  225-236 
viridigenalis,  225 
vittata,  233 

American  Woodcock,  see  Scolopax  minor 
Ammodramus  savannarum,  414 
floridanus,  539 


spp.,  539 

Anas  bahamensis,  215 
clypeata,  156 
crecca,  156 
cyanoptera,  415 
discors,  156 
platyrhynchos,  156,  424 

Anderson,  David  J.,  and  Peter  T.  Boag,  No  extra-pair 
fertilization  observed  in  Nazca  Booby  ( Sula 
grand)  broods,  244—247 

Andres,  Brad  A.,  see  Benson,  Anna-Marie, , W. 

N.  Johnson,  Susan  Savage,  and  Susan  M.  Shar- 
baugh 

Ankney,  C.  Davison,  see  Lavers,  Jennifer  L.,  Jonathan 
E.  Thompson,  Cynthia  A.  Paszkowski,  and 


Antbird,  see  Percnostola  arenarum 
Bicolored,  see  Gymnopithys  leucaspis 
Hairy-crested  see  Rhegmatorhina  melanosdcta 
Scale-backed,  see  Hylophylax  poecilinota 
White-Masked,  see  Pithys  castaneus 
Whiteplumed,  see  Pithys  albifrons  peruvianas 
Anthony,  Robert  G.,  see  Loegering,  John  R,  and 


anti-predator  function,  59 
ants  in  acacias,  563 

army,  see  Eciton  burchelli  and  Labidus  praedator 
Aphrastura  spinicauda,  252 
Aplonis  santovestris,  307 
sp.  undescribed,  307 
zelandicus,  307 

Appalachia,  wetland  habitats  of  southern,  399 
Applegate,  Roger  D.,  see  Pitman,  James  C.,  Christian 
A.  Hagen,  Brent  E.  Jamison,  Robert  J.  Robel, 

Thomas  M.  Loughin,  and 

Apus  apus,  425 
Aquila  clanga,  50 
pomarina,  50 
Ara  militaris,  231 
Ardea  alba,  103,  215 
cinerea,  113 
herodias,  112-113 
Argentina,  251 

Armenta,  Jessica,  see  Lane,  Daniel  F.,  Thomas  Valqui 

H.,  Jose  Alvarez  A.,  , and  Karen  Eck- 

hardt 

Arnett,  John  E.,  see  Labi  sky,  Ronald  R.,  and 

Arredondo,  Juan  A.  see  Hernandez,  Fidel,  , 


595 


596 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Froylan  Hernandez,  Fred  C.  Bryant,  and  Leo- 
nard A.  Brennan 

Artamus  leucorhynchus  tenuis,  295-308 

Artemisia  spp.,  23,  36—41 

aspen,  quaking,  see  Populus  tremuloides 

Asturina  nitidus,  42 

Athene  cunicularia,  83,  88 

Auk,  Great 

in  Once  Upon  a Time  in  American  Ornithology,  427 
see  Pinguinus  impennis 
Aythya  americana,  415 

B 

badger,  see  Taxidea  taxus 
Baeolophus  bicolor,  107 

Bailie,  James  Little,  in  Once  Upon  a Time  in  American 
Ornithology,  427 
balsam  fir,  density  of,  461 
Baltic  Coast,  364 
Bananaquit,  see  Coereba  flaveola 
Bare-eye,  Reddish-winged,  see  Phlegopsis  erythrop- 
tera 

Barg,  Jennifer  J.,  Jason  Jones,  M.  Katharine  Girvan, 
and  Raleigh  J.  Robertson,  Within-pair  interac- 
tions and  parental  behavior  of  Cerulean  War- 
blers breeding  in  eastern  Ontario,  316-325 
Barlow,  Clive,  and  Tim  Wacher,  A field  guide  to  the 
birds  of  the  Gambia  and  Senegal,  reviewed, 
433-434 

bat,  little  brown,  see  Myotis  lucifugus 
red,  see  Lasiurus  borealis 
beak-swinging,  by  Puerto  Rican  Spindalis,  571 
beaver,  399 
behavior 

flocking,  164 

male  Greater  Sage-Grouse  strut,  36 
migration,  471 

nest  defense,  by  Northern  Flickers,  452 
parental,  251,  309,  316 
stopover,  364 

Bell,  Douglas  A.,  see  Crosbie,  Scott  P,  , and 

Ginger  M.  Bolen 

Benson,  Anna-Marie,  Brad  A.  Andres,  W.  N.  Johnson, 
Susan  Savage,  and  Susan  M.  Sharbaugh,  Dif- 
ferential timing  of  Wilson’s  Warbler  migration 
in  Alaska,  547-55 1 
benthic  invertebrates,  152 

Bertran,  Joan,  and  Antoni  Margalida,  Reverse  mount- 
ing and  copulation  behavior  in  polyandrous 
Bearded  Vulture  ( Gypaetus  barbatus)  trio, 
254-256 

Bildstein,  Keith,  see  Careau,  Vincent,  Jean-Frangois 
Therrien,  Pablo  Porras,  Don  Thomas,  and 


bird  trade,  225 

bison.  North  American,  see  Bison  bison 
Bison  bison,  81,  399 
Bittern,  Least,  see  Ixobrychus  exilis 
Blackbird,  see  Turdus  merula 

Red- Winged,  see  Agelaius  phoeniceus 
Yellow-headed,  see  Xanthocephalus  xanthocephalus 


Blackcaps,  see  Sylvia  atricapilla 
Blem,  Charles  R.,  and  Leann  B.  Blem,  Variation  in 
mass  of  female  Prothonotary  Warblers  during 
nesting,  3-12 

Blem,  Leann  B.,  see  Blem,  Charles  R.,  and 

Blomdahl,  Anders,  Bertil  Breife,  and  Niklas  Holms- 
trom,  Flight  identification  of  European  sea- 
birds, reviewed,  124-125 
Bluebird,  Eastern,  see  Sialia  sialis 
Mountain,  see  Sialia  currucoides 
Western,  see  Sialia  mexicana 
Boa  constrictor,  232 

Boag,  Peter  T.,  see  Anderson,  David  J.,  and 

Boal,  Clint  W.,  Fred  C.  Sibley,  Tracy  S.  Estabrook,  and 
James  Lazell,  Insular  and  migrant  species,  lon- 
gevity records,  and  new  species  records  on 
Guana  Island,  British  Virgin  Islands,  218-224 
Bobolink,  see  Dolichonyx  oryzivorus 
Bobwhite,  Northern,  see  Colinus  virginianus 
Boiga  irregularis,  309 

Bolen,  Ginger  M.,  see  Crosbie,  Scott  R,  Douglas  A. 

Bell,  and 

Bombycilla  cedrorum,  454,  522 
Booby,  Blue-footed,  see  Sula  nebouxii 
Nazca,  see  Sula  grand 
boreal  forest,  164 

Borrow,  Nik,  and  Ron  Demey,  Birds  of  western  Africa, 
reviewed,  581-582 
Bouton,  Jeffrey,  review  by,  275-276 
Branta  canadensis,  1 1 4 
maxima,  579 
breeding 

a new  record  of.  White-winged  Nightjar,  109 
on  a coastal  barrier  island  by  Black  Tern,  104 
productivity,  of  Bachman’s  Sparrow,  131 
range,  of  Northern  Saw-whet  Owl,  41 1 
status  of  Setophaga  ruticilla,  439 
success,  of  Taiwan  Yuhina,  558 
territory,  of  San  Clemente  Loggerhead  Shirkes,  333 
breeding  biology 

of  Amazona  ochrocephala,  225 
of  Sporophila  cearulescens,  85 
breeding  ecology 

cooperative,  of  Taiwan  Yuhina,  558 
of  Aimophila  aestivalis,  131 
of  avifauna  on  Vanuatu,  295 
of  Dendroica  cerulea,  145 
of  Fulica  americana,  208 
of  F.  caribaea,  208 

breeding  population  estimates,  of  Semipalmated  Sand- 
piper, 478 

Brennan,  Leonard  A.,  see  Hernandez,  Fidel,  Juan  A. 
Arredondo,  Froylan  Hernandez,  Fred  C.  Bry- 
ant, and 

British  Virgin  Islands,  218 

Bronze  Cuckoo,  Shining,  see  Chrysococcyx  lucidus 
layardi 

brood  parasite,  99 
brush  cutting,  353 

Brush,  Timothy,  Nesting  birds  of  a tropical  frontier: 


INDEX  TO  VOLUME  1 1 8 


597 


the  lower  Rio  Grande  Valley  of  Texas,  re- 
viewed, 270-271 

Bryant,  Fred  C.,  see  Hernandez,  Fidel,  Juan  A.  Arre- 
dondo, Froylan  Hernandez,  , and  Leo- 

nard A.  Brennan 
Bubulcus  ibis,  255 
Bucephala  albeola,  173-177 
islandica,  173-177 
budgets 
diet,  380 
energy,  380 
time,  380 

Buecking,  Jeff  A.,  review  by,  431-433 
Bufflehead,  see  Bucephala  albeola 
Buidin,  Christophe,  Yann  Rochepault,  Michel  Savard, 
and  Jean-Pierre  L.  Savard,  Breeding  range  ex- 
tension of  the  Northern  Saw-whet  Owl  in  Que- 
bec, 411-413 

Bunkley-Williams,  Lucy,  see  Williams,  Earnest  H.,  Jr., 
and 

Bunting,  Indigo,  see  Passerina  cyanea 

Burhans,  Dirk  E.,  see  Furey,  Maria  A.,  and 

Burnett,  J.  Alexander,  A passion  for  wildlife:  the  his- 
tory of  the  Canadian  Wildlife  Service,  re- 
viewed, 121-122 
burning,  353 

Buteo  albicaudatus,  91-98 
buteo,  42 

galapagoensis,  44,  195 
hemilasius,  42 

jamaicensis,  147,  569-570 
borealis,  43 
harlani,  43 
lagopus,  42—52 
lineatus,  42,  535 
platypterus,  471-477 
polyosoma,  42 
regalis,  42,  83 
rufinus,  42 

swainsoni,  42-52,  472 

Butler,  Chris,  see  Lorenz,  Stephan, , and  Jimmy 

Paz 

Buzzard 

Common,  see  Buteo  buteo 
Long-legged,  see  Buteo  rufinus 
Red-backed,  see  Buteo  polyosoma 
Upland,  see  Buteo  hemilasius 

c 

cache-moving,  by  American  Crows,  572 
caching,  of  rabbits  by  American  Crows,  572 
Cacomantis  pyrrhophanus,  307 

Caldwell,  Sarah  S.,  and  Alexander  M.  Mills,  Compar- 
ative spring  migration  arrival  dates  in  the  two 
morphs  of  White-throated  Sparrow,  326-332 
Calidridini,  478-484 
Calidris  alpina,  479 
himantopus,  479 
mauri,  478 
melanotos,  156 
minutilla,  156,  479 


pusilla,  478-484 
California,  178,  256 
Callipepla  californica,  256-259 
gambelii,  256 
Calomys  tener,  95 
Calypte  anna,  425 
Calyptorhynchus 

baudinii  latirostris,  233 
funereus  latirostris,  234 
magnificus,  233 

Campephilus  magellanicus,  251-254 
Camptorhynchus  labradorius,  427 
Campylorhynchus  rufinucha,  563-566 
Canada 

Aegolius  acadicus  breeding  range  in  Quebec,  41 1 
body  molt  of  wood  warblers  in  Ontario,  374 
Haliaeetus  leucocephalus  foraging  in  British  Co- 
lumbia, 380 

multispecies  feeding  flocks  in  boreal  forests  of  west- 
ern, 164 

Zonotrichia  albicolis  in  southern  Ontario,  326 
Canis  latrans,  23,  27 
cannibalism,  101 
Capella  gallinago,  425 
Capra  hircus,  333 
capture-mark-recapture,  513 
Caraduellis  tristis,  540 

Cardinal,  Northern,  see  Cardinalis  cardinalis 
Cardinalis  cardinalis,  75 
Carduelis  tristis,  457 

Careau,  Vincent,  Jean-Frangois  Therrien,  Pablo  Porras, 
Don  Thomas,  and  Keith  Bildstein,  Soaring  and 
gliding  flight  of  migrating  Broad-winged 
Hawks:  behavior  in  the  Nearctic  and  Neotrop- 
ics compared,  471-477 

Carib,  Green-throated,  see  Eulampis  holosericeus 

Caribbean,  194,  218 

Carpodacus  mexicanus,  413—415 

Carter,  William  A,  see  Wood,  Douglas  R.,  and 


Castillo-Guerrero,  Jose  Alfredo,  and  Eric  Mellink, 
Maximum  diving  depth  in  fledging  Blue-footed 
Boobies:  skill  development  and  transition  to  in- 
dependence, 527-531 
Castor  canadensis,  399-410 
cat,  domestic,  see  Felis  cattus 
Catbird,  Gray,  see  Dumetella  carolinensis 
Cathartes  aura,  53,  147,  473 
guttatus,  522 
minimus,  522 
ustulatus,  222,  522 
Catharus  fuscescens,  461-470 

cavity  conditions,  of  surrogate  Cuban  Parrot  nest,  508 
cavity  nesting,  by  a Blue  Grosbeak,  107 
Centrocercus  minimus,  36-41 
urophasianus,  31,  36—41 
Certhia  americana,  164-172 
Cervus  elaphus,  399 
Chaetura  peliagica,  425 
vauxi,  424-426 

Chalcophaps  indica  sandwichensis,  295-308 


598 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Charadriiformes,  152-163 
Charadrius  montanus,  59-63,  81-84 
vociferous,  115,  156 
wilsonia,  215,  222 
Charmosyna  palmarum,  295—308 
Chartier,  Allen  T.,  and  Jerry  Ziarno,  A birder’s  guide 
to  Michigan,  reviewed,  431-432 
Chazarreta,  M.  Laura,  see  Ojeda,  Valerie  S.,  and 


Chernetsov,  Nikita,  and  Andrey  Mukhin,  Spatial  be- 
havior of  European  Robins  during  migratory 
stopovers:  a telemetry  study,  364-373 
Chickadee,  Black-capped,  see  Poecile  atricapillus 
Boreal,  see  Poecile  hudsonica 
Chicken,  see  Gallus  domesticus 
Chihuahua,  237 

Chilidonias  niger  surinamensis,  104-106 
chipmunk,  eastern,  see  Tamias  striatus 
Chlamydophila  psittaci,  in  Galapagos  Doves,  195 
Chordeiles  minor,  425 
Choristoneura  fumiferana,  1 64- 1 72 
Chough,  Alpine,  or  Yellow-billed,  see  Pyrrhocorax 
graculus 

Chrysococcyx  caprius,  99-101 
klaas,  99-101 
spp.,  99 

Cichlornis  whitneyi,  307 
Ciconia  ciconia,  380 
Cinclus  cinclus,  291 
mexicanus,  281-294 
Circus  aeruginosus,  50 
approximans,  295—308 
macrourus,  50 
pygargus,  50 
spionotus,  50 
Cistothorus  palustris,  416 
platensis,  341,  540 
Cladorhynchus  leucocephalus,  478 
Clark,  William  S.,  and  Christopher  C.  Witt,  First 
known  specimen  of  a hybrid  Buteo : Swainson’s 
Hawk  ( Buteo  swainsoni ) X Rough-Legged 
Hawk  ( B . lagopus)  from  Louisiana,  42-52 
Cleptornis  marchei,  308 
Clomba  vitiensis  leopoldi,  295-308 
clutch  size,  23,  70,  225 

Clytorhynchus  pachycephaloides  grisescens,  295—308 
Coccyzus  americanus,  55 

Cockatoo,  see  Calyptorhynchus  funereus  latirostris 

Coereba  flaveola,  219 

Colaptes  auratus,  452-460 

Colinus  virginianus,  27,  114-116,  259 

Collared-Dove,  Eurasian,  see  Streptopelia  decaocto 

Collier,  Brett  A.,  see  Metz,  Steve  T.,  Kyle  B.  Melton, 

Ray  Aguirre,  , T.  Wayne  Schwertner, 

Markus  J.  Peterson,  and  Nova  J.  Silvy 
Collocalia  esculenta  uropygialis,  295-308 
vanikorensis  vanikorensis,  295-308 
coloniality,  in  Semipalmated  Sandpiper,  478 
Coluber  constrictor,  540 
Columba  livia,  55,  195 
Columbina  passerine,  222 


communal  relationships,  563 
roosting,  532 
roosts,  566 

communities,  upland  bird,  295 
Conepatus  semistriatus,  88 
Coot,  American,  see  Fulica  americana 
Caribbean,  see  Fulica  caribaea 
Coracina  caledonica  thilenii,  295-308 
Corman,  Troy  E.,  and  Cathryn  Wise-Gervais,  Arizona 
Breeding  Bird  Atlas,  reviewed,  268-270 
Corvus  brachyrhynchos,  150,  357,  380,  569-570, 
572-573 
corax,  380 
cryptoleucus,  32 
hawaiiensis,  79 
Coturnix  coturnix,  88 
japonic  a,  60 

Covino,  Kristin  M.,  see  Morris,  Sara  R.,  Amanda  M. 

Larracuente,  , Melissa  S.  Mustillo, 

Kathryn  E.  Mattern,  David  A.  Liebner,  and  H. 
David  Sheets 

Cowbird,  Brown-headed,  see  Molothrus  ater 
coyote,  see  Canis  latrans 

Craik,  Shawn  R.,  Rodger  D.  Titman,  Amelie  Rousseau, 
and  Michael  J.  Richardson,  First  report  of 
Black  Terns  breeding  on  a coastal  barrier  is- 
land, 104-106 

Crane,  Common,  see  Grus  grus 
Creeper,  Brown,  see  Certhia  americana 
Cringan,  Alexander  T.,  Once  Upon  a Time  in  Ameri- 
can Ornithology,  427-429 

Crosbie,  Scott  P,  Douglas  A.  Bell,  and  Ginger  M.  Bol- 
en, Vegetative  and  thermal  aspects  of  roost-site 
selection  in  urban  Yellow-billed  Magpies,  532- 
536 

Crow,  American,  see  Corvus  brachyrhynchos 
Hawaiian,  see  Corvus  hawaiiensis 
crowing,  256 

Cruz  Nieto,  Miguel  A.,  see  Gonzales  Rojas,  Jose  I., 

, Oscar  Ballesteros  Medrano,  and  Irene 

Ruvalcaba  Ortega 

Cruz-Nieto,  Javier,  see  Monterrubio-Rico,  Tiberio  C., 

, Ernesto  Enkerlin-Hoeflich,  Diana  Ve- 

negas-Holguin,  Lorena  Tellez-Garcia,  and  Con- 
suelo  Marin-Togo 

Cuckoo,  Diederik,  see  Chrysococcyx  caprius 
Fan-tailed,  see  Cacomantis  pyrrhophanus 
Guira,  see  Guira  guira 
Klaas,  see  Chrysococcyx  klaas 
Old  World,  see  Chrysococcyx  spp. 

Squirrel,  see  Piaya  cayana 
Yellow-billed,  see  Coccyzus  americanus 
Cuckoo-Dove,  Mackinlay’s,  see  Macropygia  m.  mack- 
inlayi 

Cuckoo-Shrike,  Melanesian,  see  Coracina  caledonica 
thilenii 
cuckoos 

feeding  conspecific  young,  99 
fledgling  provisioning,  99 
Curlew,  Long-billed,  see  Numenius  americanus 
Cyanocitta  cristata,  150,  321,  357 


INDEX  TO  VOLUME  1 18 


599 


Cyanocorax  chrysops,  88 
Cynomys  ludovicianus,  81 

D 

Davis,  Craig  A.,  see  Graber,  Allen  E.,  , and 

David  M.  Leslie,  Jr. 

Davis,  William  E.,  Jr.,  review  by,  121-122 
Debruyne,  Christine  A.,  Janice  M.  Hughes,  and  David 
J.  T.  Hussell,  Age-related  timing  and  patterns 
of  prebasic  body  molt  in  Wood  Warblers  (Pa- 
rulidae),  374-379 

DeCandido,  Robert,  and  Deborah  Allen,  Nocturnal 
Hunting  by  Peregrine  Falcons  at  the  Empire 
State  Building,  New  York  City,  53-58 
Deconychura  longicauda,  17 
Dedrocincla  merula,  17 

del  Hoyo,  Josep,  Andrew  Elliott,  and  David  Christie 
(Eds.),  Handbook  of  the  birds  of  the  world,  vol- 
ume 9:  Cotingas  to  Pipits  and  Wagtails,  re- 
viewed, 430-431 
Delichon  urbica,  178 
urbicum,  178 

DeLong,  John  P,  Pre-migratory  fattening  and  mass 
gain  in  Flammulated  Owls  in  central  New 
Mexico,  187-193 
Dendrocolaptes  certhia,  17 
Dendroica  caerulescens,  149,  322,  461-470 
castanea,  164-172,  322 
cerulea,  145-151,  249,  316-325 
chrysoparia,  247-251 
coronata,  164-172,  322,  521 
discolor,  249,  357-358,  377 
fusca,  168,  170,  322 
magnolia,  168,  170,  222,  322,  523 
nigrescens,  249 
occidentalis,  249,  377 
pensylvanica,  168,  170,  322 
petechia,  164-172,  322,  374-379,  414,  540 
striata,  221,  322,  523 
tigrina,  164-172 
townsendi,  249,  378 
virens,  164-172,  249,  322 
density,  nest-site,  237,  478 
Deroptyus  accipitrinus,  20 

De Vault,  Travis  L.,  see  Galligan,  Edward  W., , 

and  Steven  L.  Lima 
Didelphis  spp.,  72,  88 

diet,  of  White-tailed  Hawk  in  southeastern  Brazil,  91 
dimorphism,  plumage,  326 
Dipper,  American,  see  Cinclus  mexicanus 
dispersal  patterns,  558 

distribution,  of  Black-throated  Blue  Warbler,  461 
Diucon,  Fire-eyed,  see  Xolmis  pyrope 
diversity,  genetic,  36,  194 
diving,  capacity,  527 
depth,  527 

Dolichonyx  oryzivorus,  540 

Donehower,  Christina  E.,  Likely  predation  of  adult 
Glossy  Ibis  by  Great  Black-backed  Gulls,  420- 
422 


Double-collared  Seedeater,  see  Sporophila  caerules- 
cens 

Dove,  Diamond,  see  Geopelia  cuneata 

Emerald,  see  Chalcophaps  indica  sandwichensis 
Galapagos,  see  Zenaida  galapagoensis 
Mourning,  see  Zenaida  macroura 
Red-bellied  Fruit,  see  Ptilinopus  greyii 
Santa  Cruz  Ground,  see  Gallicolumba  sanctaecrus- 
cis 

Tanna  Fruit,  see  Ptilinopus  tannensis 
Zenaida,  see  Zenaida  aurita 
Duck,  Labrador,  see  Camptorhynchus  labradorius 
Ruddy,  see  Oxyura  jamaicensis 
Wood,  see  Aix  s pons  a 
Ducula  bakeri,  295-308 
pacifica  pacifica,  295-308 

Duffe,  Jason,  see  Elliott,  Kyle  H.,  , Sandi  L. 

Lee,  Pierre  Mineau,  and  John  E.  Elliott 
Dumetella  carolinensis,  114,  341,  357,  522 

E 

Eagle,  Bald,  see  Haliaeetus  leucocephalus 
Greater  Spotted,  see  Aquila  clanga 
Lesser  Spotted,  see  Aquila  pomarina 
Eberhard,  Jessica  R.,  see  Rodriguez  Castillo,  Angelica 

M.,  and 

Eciton  burchelli,  17 

Eckhardt,  Karen,  see  Lane,  Daniel  F.,  Thomas  Valqui 
H.,  Jose  Alvarez  A.,  Jessica  Armenta,  and 


Eclectus  infectus,  307 
Ectopistes  migratoria,  118,  427 

egg 

fertility,  23 
mass,  173 
nutrients,  173 
eggs,  abnormal,  1 14 
eggshell  removal  behavior,  59 
Egret,  Cattle,  see  Bubulcus  ibis 
Great,  see  Ardea  alba 
Snowy,  see  Egretta  thula 
Egretta  thula,  102 

Eiders,  Common,  see  Somateria  mollissima 
Eira  barbara,  88 

Elaenia,  Caribbean,  see  Elaenia  martinica 
Yellow-bellied,  see  Elaenia  flavogaster 
Elaenia  flavogaster,  222 
martinica,  222 

Elaphe  obsoleta  obsoleta,  540 
Eleothreptus  candicans,  109—112 
elk,  see  Cervus  elaphus 

Elliott,  John  E.,  see  Elliott,  Kyle  H.,  Jason  Duffe,  San- 
di L.  Lee,  Pierre  Mineau,  and 

Elliott,  Kyle  H.,  Jason  Duffe,  Sandi  L.  Lee,  Pierre  Mi- 
neau, and  John  E.  Elliott,  Foraging  ecology  of 
Bald  Eagles  at  an  urban  landfill,  380-390 
Empidonax  traillii,  540 
endangered  species,  333 

Enderson,  James,  Peregrine  Falcon:  stories  of  the  Blue 
Meanie,  reviewed,  272-275 
energetics,  316,  333,  566 


600 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Enkerlin-Hoeflich,  Ernesto,  see  Monterrubio-Rico,  Ti- 

berio  C.,  Javier  Cruz-Nieto, , Diana  Ve- 

negas-Holguin,  Lorena  Tellez-Garcia,  and  Con- 
suelo  Marin-Togo 
Erithacus  rubecula,  364-373 
Erythrura  cyaneovirens,  307 

Estabrook,  Tracy  S.,  see  Boal,  Clint  W.,  Fred  C.  Sibley, 

, and  James  Lazell 

Eudocimus  albus,  103 
Eulampis  holosericeus,  219 
Euptilotis  neoxenus,  241 
extra  individuals  at  nests,  75 

F 

Fairy-Wren,  Superb,  see  Malurus  cyaneus 
Falco  naumanni,  53 
peregrinus,  51,  53-58,  118,  307,  421 
rusticolus,  51 
sparverius,  4 1 1 
sp„  333 

tinnunculus,  425 

Falcon,  Gyrfalcon,  see  Falco  rusticolus 
Peregrine,  see  Falco  peregrinus 
Peregrine,  in  Once  Upon  a Time  in  American  Or- 
nithology, 117 

Fantail,  Gray,  see  Rhipidura  albiscapa  brenchleyi 
Streaked,  see  Rhipidura  s.  spilodera 
fat  deposition,  187,  364 
Felis  cattus,  71,  88,  298 
feral  grazers,  333 

fertilization,  extra-pair,  244,  319,  502 
Ficedula  hypoleuca,  371 

Fiehler,  Craig  M.,  William  D.  Tietje,  and  William  R. 
Fields,  Nesting  success  of  Western  Bluebirds 
( Sialia  mexicana)  using  nest  boxes  in  vineyard 
and  oak-savannah  habitats  of  California,  552- 
557 

Fields,  William  R.,  see  Fiehler,  Craig  M.,  William  D. 
Tietje,  and 

Finch,  House,  see  Carpodacus  mexicanus 
fire 

effects,  353 
management,  131,  353 
suppression,  131 
first  breeding  record,  81 
first  nesting  record,  574 

Fisher,  Ryan  J.,  and  Karen  L.  Wiebe,  Investment  in 
nest  defense  by  Northern  Flickers:  effects  of 
age  and  sex,  452-460 
Flicker,  Northern,  see  Colaptes  auratus 
flocks,  multispecies  feeding,  164 
Flycatcher,  Ash-throated,  see  Myiarchus  cinerascens 
Great  Crested,  see  Myiarchus  crinitus 
Melanesian,  see  Myiagra  caledonica  marinae 
Pied,  see  Ficedula  hypoleuca 
Social,  see  Myiozetetes  similis 
Traill’s,  see  Empidonax  traillii 
Willow,  see  Empidonax  traillii 
food 

availability,  374 
delivery,  316 


provisioning,  99 
resources,  138,  316 
selection,  64 
foraging 

attack  distances,  333 
behavior,  101 

benthic  invertebrate  prey  of  shorebirds,  152 

competition,  64 

ecology,  380 

efficiency,  64,  333,  380 

habitat,  333 

microhabitat,  152 

multispecies  flocks,  164 

opportunistic,  152 

skills,  527 

skills  and  parental  care,  527 
success  rates,  333 

Francisco,  Mercival  R.,  Breeding  biology  of  the  Dou- 
ble-collared Seedeater  ( Sporophila  caerules- 
cens),  85-90 

Fulica  americana,  208-217,  415-418 
caribaea,  208-217 

Furey,  Maria  A.,  and  Dirk  E.  Burhans,  Territory  selec- 
tion by  upland  Red-winged  Blackbirds  in  ex- 
perimental restoration  plots,  391-398 

G 

Galapagos,  194,  244 
Galictus  vittata,  88 
Gallicolumba  sanctaecruscis,  307 
Galligan,  Edward  W.,  Travis  L.  DeVault,  and  Steven 
L.  Lima,  Nesting  success  of  grassland  and  sa- 
vanna birds  on  reclaimed  surface  coal  mines  of 
the  mid  western  United  States,  537-546 
Gallus  domesticus,  425 

Galvez,  Rafael  A.,  Lexo  Gavashelishvili,  and  Zura  Ja- 
vakhisvili.  Raptors  and  owls  of  Georgia,  re- 
viewed, 582-583 

Garcelon,  David,  K.,  see  Lynn,  Suellen,  John  A.  Mar- 
tin, and 

Garcia-C.,  J.  Mauricio,  and  Rakan  A.  Zahawi,  Preda- 
tion by  a Blue-crowned  Motmot  ( Momotus 
momota)  on  a hummingbird,  261-263 
Gardali,  Thomas,  and  Nadav  Nur,  Site-specific  surviv- 
al of  Black-headed  Grosbeaks  and  Spotted  To- 
whees  at  four  sites  within  the  Sacramento  Val- 
ley, California,  178-186 
Garrulus  glandarius,  559 
gastropods,  161 
Gavia  immer,  115,  425 

Gee,  Jennifer  M.,  Natural  occurrence  of  crowing  in  a 
free-living  female  Galliform,  the  California 
Quail,  256-259 
gene  flow,  194 
Geopelia  cuneata,  65 
Geothlypis  poliocephala,  574-576 
trichas,  353-363,  574 

Gerygone,  Fan-tailed,  see  Gerygone  flavolateralis  cor- 
reiae 

Gerygone  flavolateralis  correiae,  295—308 


INDEX  TO  VOLUME  1 1 8 


601 


Girvan,  M.  Katharine,  see  Barg,  Jennifer  J.,  Jason 

Jones, , and  Raleigh  J.  Robertson 

gliding  flight,  of  Broad-winged  Hawk,  471 
Glycifohia  n.  notabilis,  295—308 
goat,  feral,  see  Capra  hircus 
Goldeneye,  Barrow’s,  see  Bucephala  islandica 
Goldfinch,  American,  see  Carduelis  tristis 
Gonzales  Rojas,  Jose  I.,  Miguel  A.  Cruz,  Oscar  Bal- 
lesteros Medrano,  and  Irene  Ruvalcaba  Ortega, 
First  breeding  record  of  a Mountain  Plover  in 
Nuevo  Leon,  Mexico,  81-84 
Goose,  Canada,  see  Branta  canadensis 
Goose,  Giant  Canada,  in  Once  Upon  a Time  in  Amer- 
ican Ornithology,  577 

Goose,  Ross’s,  in  Once  Upon  a Time  in  American  Or- 
nithology, 577 

Goshawk,  Brown,  see  Accipiter  faciatus 
Grey,  see  Accipiter  novaehollandiae 
Graber,  Allen  E.,  Craig  A.  Davis,  and  David  M.  Leslie, 
Jr.,  Golden-cheeked  Warbler  males  participate 
in  nest-site  selection,  247-251 
Gracilinanus  spp.,  95 

Grackle,  Great-tailed,  see  Quiscalus  mexicanus 
Grand,  James  B.,  see  Tucker,  James  W.,  Jr.,  W.  Douglas 

Robinson,  and 

grassland  birds,  353,  537 
grassland  loss,  70 

Grassquit,  Black-faced,  see  Tiaris  bicolor 
Gratto-Trevor,  Cheri  L.,  The  North  American  bander’s 
manual  for  banding  shorebirds  (Charadriifor- 
mes:  Suborder  Charadrii),  reviewed,  120 
Great  Abaco  Island,  508 
Grebe,  Eared,  see  Podiceps  nigricollis 
Little,  see  Tachybaptus  ruficollis 
Pied-billed,  see  Podilymbus  podiceps 
Grim,  Tomas,  and  Radim  Sumbera,  A new  record  of 
the  endangered  White-winged  Nightjar  ( Eleo - 
threptus  candicans)  from  Beni,  Bolivia,  109- 
112 

Grinnell,  George  Bird,  in  Once  Upon  a Time  in  Amer- 
ican Ornithology,  117 
grison,  see  Galictus  vittata 

Grosbeak,  Black-headed,  see  Pheucticus  melanoce- 
phalus 

ground  squirrels,  see  Spermophilus  spp. 

Ground-Dove,  Common,  see  Columbina  passerina 
Blue,  see  Passerina  caerulea 
Grouse,  Sharp-tailed,  see  Tympanuchus  phasianellus 
Grus  grus,  471-477 
Guana  Island,  218 

Gull,  Black-headed,  see  Larus  ridibundus 
Franklin,  see  Larus  pipixcan 
Great  Black-backed,  see  Larus  marinus 
Herring,  see  Larus  argentatus 
Ring-billed,  see  Larus  delawarensis 
Guris,  Paul  A.,  review  by,  124-125 
Gustafson,  Mary,  reviews  by,  123-124,  430-431, 
434-435,  583-584 
Gymnopithys  leucaspis,  17 
Gypaetus  barbatus,  254-256 


H 

habitat 

breeding,  237,  399 
degradation,  70,  333 
early-successional,  353 
edge,  399 
manipulation,  353 
nest-site,  247,  281 
preference,  353,  399 
quality,  131,  178 
restoration,  353 
wetland,  208,  399 

Haemaproteus  spp.,  in  Galapagos  Doves,  203 
Haematopus  ostralegus,  176 
palliates,  485-493 

Hagen,  Christian  A.,  see  Pitman,  James  C.,  , 

Brent  E.  Jamison,  Robert  J.  Robel,  Thomas  M. 
Loughin,  and  Roger  D.  Applegate 
Haliaeetus  leucocephalus,  53,  380-390,  569-570 
Hall,  Kimberly  R.,  see  Kearns,  Laura  J.,  Emily  D.  Sil- 
verman, and 

Harrier,  Eastern  Marsh,  see  Circus  spionotus 
Mantagu’s,  see  Circus  pygargus 
Northern,  see  Circus  cyaneus 
Pallid,  see  Circus  macrourus 
Swamp,  see  Circus  approximans 
Western  Marsh,  see  Circus  aeruginosus 
hatching  success,  23 

Hawk,  Broad-winged,  see  Buteo  platypterus 
Cooper’s,  see  Accipiter  cooperii 
Ferruginous,  see  Buteo  regalis 
Galapagos,  see  Buteo  galapagoensis 
Gray,  see  Asturina  nitidus 
Red-backed,  see  Buteo  polyosoma 
Red-shouldered,  see  Buteo  lineatus 
Red-tailed,  see  Buteo  jamaicensis 

Eastern,  see  Buteo  jamaicensis  borealis 
Harlan’s,  see  Buteo  jamaicensis  harlani 
Rough-legged,  see  Buteo  lagopus 
Swainson’s,  see  Buteo  swainsoni 
White-tailed,  see  Buteo  albicaudatus 
Hayslette,  Steven  E.,  Seed-size  selection  in  Mourning 
Doves  and  Eurasian  Collared-Doves,  64-69 
Hearne,  Samuel,  in  Once  Upon  a Time  in  American 
Ornithology,  577 
Heliodoxa  jacula,  261 
Helmitheros  vermivorum,  222 
Herman,  Steven  G.,  review  by,  273-275 
Hernandez,  Fidel,  Juan  A.  Arredondo,  Froylan  Her- 
nandez, Fred  C.  Bryant,  and  Leonard  A.  Bren- 
nan, Abnormal  eggs  and  incubation  behavior  in 
Northern  Bobwhite,  114-116 
Hernandez,  Froylan,  see  Hernandez,  Fidel,  Juan  A.  Ar- 
redondo,   , Fred  C.  Bryant,  and  Leonard 

A.  Brennan 

Heron,  Great  Blue,  see  Ardea  herodias 
Grey,  see  Ardea  cinerea 
Heterophasia  auricularis,  559 

Hilty,  Steven,  Birds  of  tropical  America:  a watcher’s 
introduction  to  behavior,  breeding,  and  diver- 
sity, reviewed,  434-435 


602 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118 , No.  4 . December  2006 


Himantopus  mexicanus,  221 
Hinds*  acacia,  see  Acacia  hindsii 
Hoffman,  Wayne,  review  by,  271-272 
home-range 

movements.  502 
size,  138.  364 

Honeyeater.  Cardinal,  see  Myzomela  cardinalis  tenuis 
Micronesian.  see  Myzomela  rubratra 
White-bellied,  see  Glycifohia  n.  notabilis 
Horobin.  David.  Falconry  in  literature,  reviewed.  585 
Houston.  Stuart.  Once  Upon  a Time  in  American  Or- 
nithology, 577-579 

Hughes.  Janice  M..  see  Debruyne,  Christine  A., 

. and  David  J.  T.  Hussell 

human  impacts.  295 

Hummingbird.  Anna's,  see  Calypte  anna 

Antillean  Crested,  see  Orthorhyncus  cristatus 
Green-crowned  Brilliant,  see  Heliodoxa  jacula 
Hung.  Hisn-Yi,  see  Yuan.  Hsiao-Wei.  Sheng-Feng 
Shen.  and 

Hunter,  John  E.,  review  by,  580-581 
hunting  at  skyscrapers,  53 

Hussell.  David  J.  T..  see  Debruyne.  Christine  A..  Janice 
M.  Hughes,  and 

hybridization,  of  Swainson’s  and  Rough-legged 
Hawks,  42 

Hylophylax  poecilinota,  17 

I 

Ibis,  Glossy,  see  Plegadis  falcinellus 
White,  see  Eudocimus  albus 
Icterus  cucullatus,  414 
galbula,  55 
spurious,  540 
Illinois  River.  152 
incubation 

abnormal.  1 14 
behavior,  485 

of  Double-collared  Seedeater.  85 
of  Northern  Bobwhite.  1 14 
prolonged.  1 14 
rhythms,  316 

Ingalls.  Victoria,  see  Staicer,  Cynthia  A.. . and 

Thomas  W.  Sherry 
insular  species.  218 
interactions 

male-female,  316 
within-pair.  316 
interbreeding.  42 
Ixobrychus  exilis,  4 1 5-4 1 8 

J 

Jaguarondi,  see  Herpailurus  yaguarondi 

Jamison.  Brent  E..  see  Pitman,  James  C..  Christian  A. 

Hagen.  . Robert  J.  Robel.  Thomas  M. 

Loughin.  and  Roger  D.  Applegate 
Jay,  Blue,  see  Cyanocitta  cristata 
Eurasian,  see  Garrulus  glandarius 
Plush-crested,  see  Cyanocorax  chrysops 
Jehl.  Joseph  R.,  Coloniality.  mate  retention,  and  nest- 


site  characterization  in  the  Semipalmated  Sand- 
piper, 478-484 

Johnson,  Steven  L..  Do  American  Robins  acquire 
songs  by  both  imitating  and  inventing?,  341- 
352 

Johnson,  W.  N.,  see  Benson,  Anna-Marie.  Brad  A.  An- 
dres,   . Susan  Savage,  and  Susan  M. 

Sharbaugh 

Jones.  H.  Lee.  Birds  of  Belize,  reviewed,  267-268 
Jones.  Jason,  see  Barg.  Jennifer  J..  . M.  Ka- 

tharine Girvan.  and  Raleigh  J.  Robertson 

K 

Kearns.  Laura  J.,  Emily  D.  Silverman,  and  Kimberly 
R.  Hall.  Black-throated  Blue  Warbler  and  Vee- 
ry  abundance  in  relation  to  understory  com- 
position in  northern  Michigan  forests,  461-470 
Kershner.  Eric  L.,  and  Eric  C.  Mruz.  Nest  interference 
by  fledgling  Loggerhead  Shrikes.  75-80 

Kershner,  Eric  L.,  see  Walk.  Jeffrey  W.. . and 

Richard  E.  Warner 

Kestrel.  American,  see  Falco  sparverius 
Common,  see  Falco  tinnunculus 
Lesser,  see  Falco  naumanni 
Killdeer.  see  Charadrius  vociferus 
Kingbird.  Eastern,  see  Tyrannus  tyrannus 
Kingery.  Hugh  E..  review  by,  268-270 
Kingfisher.  Chestnut-bellied,  see  Todiramphus  farqu- 
hari 

Collared,  see  Todiramphus  chloris  santoensis 
Ringed,  see  Megaceryle  torquata 
Kinglet,  Golden-crowned,  see  Regulus  satrapa 
Ruby-crowned,  see  Regulus  calendula 

Kirchman.  Jeremy  J.,  see  Kratter.  Andrew'  W.. . 

and  David  W.  Steadman 
Kiskadee.  Great,  see  Pitangus  sulphuratus 
Kite.  Black,  see  Milvus  migrans 

Black-shouldered,  see  Elanus  axillaris 
Red,  see  Milvus  milvus 

Knopf.  Fritz  L..  Once  Upon  a Time  in  American  Or- 
nithology, 117-119 

Kratter.  Andrew  W.,  Jeremy  J.  Kirchman.  and  David 
W.  Steadman.  Upland  bird  communities  on 
Santo,  Vanuatu,  Southwest  Pacific,  295-308 
Krementz.  David  G.,  see  Stober.  Jonathan  M..  and 


Kroodsma.  Donald  E..  The  singing  life  of  birds:  the 
art  and  science  of  listening  to  birdsong,  re- 
viewed, 125-127 

Kuehn.  Michael  J..  see  Rivers,  James  W..  and 


L 

Labidus  praedator,  17 

Labisky.  Ronald  F..  and  John  E.  Arnett.  Jr..  Pair  roost- 
ing of  nesting  Carolina  Wrens  ( Thryothorus  lu- 
dovicianus),  566-569 
Lalage  leucopyga  albiloris,  295-308 
maculosa  modesta,  295-308 
Lampropeltis  calligaster,  540 


INDEX  TO  VOLUME  118 


603 


landfill,  380 

Lane,  Daniel  F.,  Thomas  Valqui  H.,  Jose  Alvarez  A., 
Jessica  Armenta,  and  Karen  Eckhardt,  The  re- 
discovery and  natural  history  of  the  White- 
masked  Antbird  {Pithy s castaneus ),  13-22 
Lanius  ludovicianus,  70—74,  75—80,  333—340 
collurio,  457 

Larracuente,  Amanda  M.,  see  Morris,  Sara  R., 

, Kristen  M.  Covino,  Melissa  S.  Mus- 

tillo,  Kathryn  E.  Mattern,  David  A.  Liebner, 
and  H.  David  Sheets 
Larus  argentatus,  386,  420 
delawarensis,  425 
marinus,  420-422 
pipixcan,  102,  415 
ridibundus,  62 
Lasiurus  borealis,  55 

Lavers,  Jennifer  L.,  Jonathan  E.  Thompson,  Cynthia 
A.  Paszkowski,  and  C.  Davison  Ankney,  Vari- 
ation in  size  and  composition  of  Bufflehead 
(Bucephala  albeola)  and  Barrow’s  Goldeneye 
(Bucephala  islandica)  eggs,  173—177 
Lazell,  James,  see  Boal,  Clint  W.,  Fred  C.  Sibley,  Tra- 
cy S.  Estabrook,  and 

Lee,  Sandi  L.,  see  Elliott,  Kyle  H.,  Jason  Duffe, 

, Pierre  Mineau,  and  John  E.  Elliott 

Leslie,  David  M.,  Jr.,  see  Graber,  Allen  E.,  Craig  A. 
Davis,  and 

Lieber,  David  A.,  see  Morris,  Sara  R.,  Amanda  M.  Lar- 
racuente, Kristen  M.  Covino,  Melissa  S.  Mus- 
tillo,  Kathryn  E.  Mattern, , and  H.  Da- 

vid Sheets 

Liguori,  Jerry,  Hawks  from  every  angle,  reviewed, 
275-276 

Lima,  Steven  L.,  see  Galligan,  Edward  W.,  Travis  L. 

De Vault,  and 

Limicola  falcinellus,  478 
Limnothlypis  swainsonii,  249 
Liolaemus  sp.,  251 
livestock,  399 
lizard,  see  Liolaemus  sp. 

Lockwood,  Mark  W.,  review  by,  270-271 
locomotion,  571 

Loegering,  John  R,  and  Robert  G.  Anthony,  Nest-site 
selection  and  productivity  of  American  Dippers 
in  the  Oregon  Coast  Range,  281-294 
longevity,  218 

longleaf  pine  forests,  see  Pinus  palustris 
long-term  banding,  326 
Loon,  Common,  see  Gavia  immer 
Lorenz,  Stephan,  Chris  Butler,  and  Jimmy  Paz,  First 
nesting  record  of  the  Gray-crowned  Yellow- 
throat  (Geothlypis  poliocephala)  in  the  United 
States  since  1894,  574-576 
Lorikeet,  Palm,  see  Charmosyna  palmarum 

Rainbow,  see  Trichoglossus  haematodus  massena 
Loughin,  Thomas  M.,  see  Pitman,  James  C.,  Christian 
A.  Hagen,  Brent  E.  Jamison,  Robert  J.  Robel, 

, and  Roger  D.  Applegate 

Lovette,  Irby  J.,  Dustin  R.  Rubenstein,  and  Wilson 
Nderitu  Watetu,  Provisioning  of  fledgling  con- 


specifics  by  males  of  the  brood-parasitic  cuck- 
oos Chrysococcyx  klaas  and  C.  caprius,  99- 

101 

Luscinia  megarhynchos,  341 

Lynn,  Suellen,  John  A.  Martin,  and  David  K.  Garce- 
lon,  Can  supplemental  foraging  perches  en- 
hance habitat  for  endangered  San  Clemente 
Loggerhead  Shrikes?,  333-340 

M 

Macaw,  Military,  see  Ara  militaris 
Macropygia  m.  mackinlayi,  295-308 
Magpie,  Black-billed,  see  Pica  hudsonia 
Yellow-billed,  see  Pica  nuttalli 
male  detectability,  effect  of  pairing  status  on,  439 
Mallard,  see  Anas  platyrhynchos 
Malurus  cyaneus,  244 

Margalida,  Antoni,  see  Bertran,  Joan,  and 

Margarops  fuscatus,  221 
Mariana  Islands,  309 

Marin-Togo,  Consuelo,  see  Monterrubio-Rico,  Tiberio 
C.,  Javier  Cruz-Nieto,  Ernesto  Enkerlin-Hoe- 
flich,  Diana  Venegas-Holguin,  Lorena  Tellez- 
Garcia,  and 

Marshall,  James  S.,  see  Zuwerink,  David  A.,  and 


Martin,  Common  House-,  see  Delichon  urbicum 
House,  see  Delichon  urbica 

Martin,  John  A.,  see  Lynn,  Suellen, , and  David 

K.  Garcelon 

mass 

gain,  187 
loss,  3 

variation,  during  incubation,  of  Prothonotary  War- 
bler, 3 

variation,  during  nestling  stage,  of  Prothonotary 
Warbler,  3 

Massachusetts,  341,  353 
mate  retention,  478 

Mattern,  Kathryn  E.,  see  Morris,  Sara  R.,  Amanda  M. 
Larracuente,  Kristen  M.  Covino,  Melissa  S. 

Mustillo,  , David  A.  Liebner,  and  H. 

David  Sheets 

Meadowlark,  Eastern,  see  Sturnella  magna 
Medrano,  Oscar  Ballesteros,  see  Gonzales  Rojas,  Jose 

I.,  Miguel  A.  Cruz,  , and  Irene  Ruval- 

caba  Ortega 

Megaceryle  torquata,  91 
Megapodius  layardi,  295-308 
Megapodius  sp.,  307 
Melanerpes  formicivorus,  75,  244 
Melcher,  Cynthia  P,  Epilogue  to  Once  Upon  a Time 
in  American  Ornithology  ( Pinguinus ),  429 
Meleagris  gallopavo  intermedia,  259-261 
Meleagris  gallopavo  merriami,  259 
Mellink,  Eric,  see  Castillo-Guerrero,  Jose  Alfredo,  and 


Melospiza  melodia,  353-363,  540 

Melton,  Kyle  B.  see  Metz,  Steve  T.,  , Ray 

Aguirre,  Bret  A.  Collier,  T.  Wayne  Schwertner, 
Markus  J.  Peterson,  and  Nova  J.  Silvy 


604 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Mephitis  mephitis,  32 

Metz,  Steve  T.,  Kyle  B.  Melton,  Ray  Aguirre,  Bret  A. 
Collier,  T.  Wayne  Schwertner,  Markus  J.  Peter- 
son, and  Nova  J.  Silvy,  Poult  adoption  and  nest 
abandonment  by  a female  Rio  Grande  Wild 
Turkey  in  Texas,  259-261 

Mexico,  237 
Nuevo  Leon,  81 

migration,  53,  164,  471,  494,  547 
age-related  differences,  547 
arrival  dates,  326 
behavior,  471 
cost  of,  471 
differential,  547 
sex-related  differences,  547 
stopovers,  364,  513 
strategy,  471 
timing,  326,  547 

Mills,  Alexander  M.,  see  Caldwell,  Sarah  S.,  and 


Milvus  migrans,  50,  380 
milvus,  50 

Mimus  polyglottos,  319,  341 

Mineau,  Pierre,  see  Elliott,  Kyle  H.,  Jason  Duffe,  San- 

di  L.  Lee, , and  John  E.  Elliott 

mist-netting,  218 
mitochondrial  DNA,  42 
Mniotilta  varia,  169,  523 

Mockingbird,  Northern,  see  Mimus  polyglottos 
Mockingbirds,  Galapagos,  see  Nesomimus  spp. 
Molothrus  ater,  107,  146,  319,  414,  418-419,  537 
Momotus  momota,  261 

Monarch,  Buff-bellied,  see  Neolalage  banksiana 
Monterrubio-Rico,  Tiberio  C.,  Javier  Cruz-Nieto,  Er- 
nesto Enkerlin-Hoeflich,  Diana  Venegas-Hol- 
guin,  Lorena  Tellez-Garcia,  and  Consuelo  Ma- 
rin-Togo,  Gregarious  nesting  behavior  of 
Thick-billed  Parrots  ( Rhynchopsitta  pachyrhyn- 
cha)  in  aspen  stands,  237-243 
morphology,  326 

Morris,  Sara  R.,  Amanda  M.  Larracuente,  Kristen  M. 
Covino,  Melissa  S.  Mustillo,  Kathryn  E.  Mat- 
tern,  David  A.  Liebner,  and  H.  David  Sheets, 
Utility  of  open  population  models:  limitations 
posed  by  parameter  estimability  in  the  study  of 
migratory  stopover,  513-526 
Motmot,  Blue-crowned,  see  Momotus  momota 
movements,  between  breeding  and  wintering  areas, 
494 

of  Long-tailed  Duck,  494 
of  Tree  Swallows,  502 
mowing,  353 

Mukhin,  Andrey,  see  Chemetsov,  Nikita,  and 

Mustillo,  Melissa  S.,  see  Morris,  Sara  R.,  Amanda  M. 
Larracuente,  Kristen  M.  Covino, , Kath- 

ryn E.  Mattern,  David  A.  Liebner,  and  H.  Da- 
vid Sheets 

Myiagra  caledonica  marinae,  295-308 
Myiarchus  cinerascens,  553 
crinitus,  107 
Myiozetetes  similis,  564 


My otis  lucifugus,  55 
Myzomela  cardinalis  tenuis,  295-308 
rubratra  309-315 
dichromata,  309 
kobayashii,  309 
kurodai,  309 
major,  309 
rubratra,  309 
sajfordi,  309-315 

N 

Nantucket  Island,  353 
natural  history,  13 

Nderitu  Watetu,  Wilson,  see  Lovette,  Irby  J.,  Dustin  R. 

Rubenstein,  and 

Neolalage  banksiana,  295-308 

Nesomimus  spp.,  195 

nest 

first  description  of,  309 
interference,  75 
parasitism,  413,  415,  418 
placement,  309 
poaching,  225 
predation,  563 
success,  23 

nest-box  occupancy,  552 
nest  defense 

by  Northern  Flickers,  452 
influence  of  age  on,  452 
influence  of  body  size  on,  452 
influence  of  brood  size  on,  452 
influence  of  sex  on,  452 
intensity,  452 
risk  of,  452 
nesting 

behavior,  75,  237 
density,  478 
ecology,  23 
gregarious,  237 

record,  first,  of  Gray-crowned  Yellowthroat,  574 
success,  70,  85,  145,  208,  225,  316,  485,  537,  552, 
563 

nest-site 

characteristics,  478 
fidelity,  23 
selection,  247,  281 
New  Mexico,  187 

Nighthawk,  Common,  see  Chordeiles  minor 
Night-Heron,  Black-crowned,  see  Nycticorax  nycticor- 
ax 

Yellow-crowned,  see  Nycticorax  violacea 
Nightingale,  see  Luscinia  megarhynchos 
Nightjar,  Whitewinged,  see  Eleothreptus  candicans 
nocturnal  hunting,  53 
Norman,  David,  review  by,  120 
Northern  Bobwhite,  see  Colinus  virginianus 
Northern  Wheatear,  see  Oenanthe  oenanthe 
nuclear  DNA,  42 
Nuevo  Leon,  Mexico,  81 
Numenius  americanus,  83,  425 
Nur,  Nadav,  see  Gardali,  Thomas,  and 


INDEX  TO  VOLUME  1 18 


605 


Nuthatch,  Red-breasted,  see  Sitta  canadensis 
Nycticorax  nycticorax,  101  — 104,  215 
violacea,  215 

o 

oak  woodland,  552 
O’Brien,  Michael,  review  by  267-268 
Odocoileus  virginianus,  461—470 
Oenanthe  oenanthe,  10 

Ojeda,  Valerie  S.,  and  M.  Laura  Chazarreta,  Provision- 
ing of  Magellanic  Woodpecker  ( Campephiluss 
magellanicus ) nestlings  with  vertebrate  prey, 
251-254 
Oklahoma,  413 
oligochaetes,  152-163 
Oligoryzomys  nigripes,  95 

Oliveras  de  Ita,  Adan  and  Octavio  R.  Rojas-Soto,  Ant 
presence  in  acacias:  an  association  that  maxi- 
mizes nesting  success  in  birds,  563-566 
Ontario,  326,  374 
open  population  models,  513 
opossum,  see  Didelphis  spp. 

Oregon  Coast  Range,  281 
Oriole,  Baltimore,  see  Icterus  galbula 
Hooded,  see  Icterus  cucullatus 
Orchard,  see  Icterus  spurius 
Ortega,  Irene  Ruvalcaba,  see  Gonzales  Rojas,  Jose  I., 
Miguel  A.  Cruz,  Oscar  Ballesteros  Medrano, 
and 

Orthorhyncus  cristatus,  219,  422-423 
Osprey,  see  Pandion  haliaetus 

Ostrow,  Bruce  D.,  Bald  Eagle  kills  crow  chasing  a 
hawk,  569-570 
Otus  asio,  425,  457 
flammeolus,  187-193 
Ovenbird,  see  Seiurus  aurocapilla 
Owen,  Jennifer  L.,  and  James  C.  Cokendolpher,  Tail- 
less whipscorpion  ( Phrynus  longipes)  feeds  on 
Antillean  Crested  Hummingbird  ( Orthorhyncus 
cristatus ),  422-423 
Owl,  Boreal,  see  Aegolius  funereus 
Burrowing,  see  Athene  cunicularia 
Eastern  Screech-,  see  Otus  asio 
Flammulated,  see  Otus  flammeolus 
Mexican  Spotted,  see  Strix  occidentalis  lucida 
Northern  Saw-whet,  see  Aegolius  acadicus 
Oxymycterus  sp.,  95 
Oxyura  jamaicensis,  176 

Oystercatcher,  American,  see  Haematopus  palliatus 
Oystercatcher,  Eurasian,  see  Haematopus  ostralegus 

P 

Pachycephala  [pectoralis]  caledonica  intacta,  295- 
308 

pair  roosting,  of  Carolina  Wrens,  566 

Panama,  225 

Pandion  haliaetus,  53 

Parakeet,  Echo,  see  Psittacula  echo 

parameter 

estimability,  513 


uncertainty,  513 
parasitism,  23,  537 
parentage  analysis,  502 

Parker,  Patricia  G.,  see  Santiago-Alarcon,  Diego,  Su- 
san M.  Tanksley,  and 

Parrot,  Amazon,  see  Amazona  ochrocephala  panamen- 
sis 

Lilac-crowned,  see  Amazona  finschi 
Maroon-fronted,  see  Rhynchopsitta  terrisi 
Red-fan,  see  Deroptyus  accipitrinus 
Thick-billed,  see  Rhynchopsitta  pachyrhyncha 
Yellow-crowned,  see  Amazona  ochrocephala 
Parrot-Finch,  Red-headed,  see  Erythrura  cyaneovirens 
Parula  americana,  223,  523 
Parula,  Northern,  see  Parula  americana 
Parus  major,  457 
Passer  domesticus,  553 
Passerculus  sandwichensis,  353-363,  414 
Passerina  caerulea,  107—108,  540 
cyanea,  223,  540 
spp.,  575 

Paszkowski,  Cynthia  A.,  see  Lavers,  Jennifer  L.,  Jon- 
athan E.  Thompson,  , and  C.  Davison 

Ankney 
Patagonia,  251 
paternity,  extra-pair,  244 

Paz,  Jimmy,  see  Lorenz,  Stephan,  Chris  Butler,  and 


Peer,  Brian  D.,  American  Coot  parasitism  on  Least  Bit- 
terns, 415-418 

Penguin,  Chinstrap,  see  Pygoscelis  antarctica 
perch  density,  391 
Percnostola  arenarum,  17 

Peterson,  Markus  J.,  see  Metz,  Steve  T.,  Kyle  B.  Mel- 
ton, Ray  Aguirre,  Bret  A.  Collier,  T.  Wayne 

Schwertner, , and  Nova  J.  Silvy 

Petrochelidon  pyrrhonota,  414 

Phalarope,  Red-necked,  see  Phalaropus  lobatus 

Phalaropus  lobatus,  222 

Phasianus  colchicus,  27,  540 

Pheasant,  Ring-necked,  see  Phasianus  colchicus 

Pheucticus  melanocephalus,  178-186 

Phlegopsis  erythroptera,  17-19 

Phoebe,  Black,  see  Sayomis  nigricans 

Phrygilus  patagonicus,  252 

Phrynus  longipes,  422—423 

Pica  hudsonia,  532 

Pica  nuttalli,  532-536 

Picoides  lignarius,  252 

Pigeon 

Pacific  Imperial,  see  Ducula  pacifica  pacifica 
Passenger,  in  Once  Upon  a Time  in  American  Or- 
nithology, 1 1 7 

Passenger,  see  Ectopistes  migratoria 
Rock,  see  Columba  livia 
Vanuatu  Imperial,  see  Ducula  bakeri 
White-throated,  see  Clomba  vitiensis  leopoldi 
Pinguinus  impennis,  427—429 
Pintail,  White-Cheeked,  see  Anas  Bahamensis,  215 
Pinus  palustris,  131-137,  138-144 
Pipilo  erythrophthalmus,  142,  353—363 


606 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


maculatus,  178-186 
Piranga  rubra,  367 
Pitangus  sulphuratus,  88 
Pithy s albifrons  peruvianus,  17,  18 
castaneus,  13-22 

Pitman,  James  C.,  Christian  A.  Hagen,  Brent  E.  Ja- 
mison, Robert  J.  Robel,  Thomas  M.  Loughin, 
and  Roger  D.  Applegate,  Nesting  ecology  of 
Lesser  Prairie-Chickens  in  sand  sagebrush  prai- 
rie of  southwestern  Kansas,  23-35 
Pituophis  melanoleucus,  23,  27 
Plegadis  falcinellus,  420-422 
Plover,  Mountain,  see  Charadrius  montanus 
Wilson’s,  see  Charadrius  wilsonia 
Podiceps  nigricollis,  112-113 
Podilymbus  podiceps,  416 
Poecile  atricapillus,  164-172,  418 
carolinensis,  418-419 
hudsonica,  164-172 
montanus,  10 
polydactyly,  424 

Pompadour  Cotinga,  see  Xipholena  punicea 
population,  131,  194 
density,  131 
sink,  178 

Populus  tremuloides,  237 

Porras,  Pablo,  see  Careau,  Vincent,  Jean-Fran9ois 

Therrien,  , Don  Thomas,  and  Keith 

Bildstein 
Porzana  sp.,  307 
poult  adoption,  259 

Prairie-Chicken,  Attwater’s  Greater,  see  Tympanuchus 
cupido  attwateri 

Greater,  see  Tympanuchus  cupido 
Lesser,  see  Tympanuchus  pallidicinctus 
prairie  dog,  black-tailed,  see  Cynomys  ludovicianus 
prebasic  body  molt 
patterns,  374 
timing,  374 

predation,  23,  53,  59,  70,  85,  1 12,  152,  225,  261,  420, 
422,  569 

by  Black-crowned  Night  Heron,  101 
by  Great  Blue  Heron,  112 
nest,  316 

of  Amercan  Crow  by  Bald  Eagle,  569 
of  Eared  Grebe  by  Great  Blue  Heron,  112 
of  Glossy  Ibis  by  Great  Black-backed  Gulls,  420 
of  hummingbird  by  Blue-crowned  Motmot,  261 
prescribed  fire,  508 
prey 

invertebrate,  152,  251 
vertebrate,  251,  261,  420,  422 
prey  selection,  of  White-tailed  Hawk,  91 
Procyon  cancrivorus,  88 
lotor,  72 
productivity,  281 
Protonotaria  citrea,  3-12 
Pseudomyrmex  spp.,  563—566 
Psittacula  echo,  79 
Ptilinopus  greyii,  295-308 
tannensis,  295—308 


Pygarrhichas  albogularis,  252 
Pygoscelis  antarctica,  244 
Pyrrhocorax  graculus,  380 

Q 

Quail,  California,  see  Callipepla  califomica 
Common,  see  Coturnix  coturnix 
Gambel’s,  see  Callipepla  gambelii 
Japanese,  see  Coturnix  japonica 
Quebec,  411 

Quetzal,  Eared,  see  Euptilotis  neoxenus 
Quiscalus  mexicanus,  416 

R 

rabbit,  eastern  cottontail,  see  Sylvilagus  floridanus 

raccoons,  see  Procyon  spp. 

radiotelemetry,  138,  364 

Rail,  flightless,  see  Porzana  sp. 

rats,  see  Rattus  spp. 

Rattus  spp.,  102,  215 

Raven,  Chihuahuan,  see  Corvus  cryptoleucus 
Common,  see  Corvus  corax 
Rayadito,  Thom-tailed,  see  Aphrastura  spinicauda 
recapture,  178,  187 
probability,  178 

reclaimed  surface  coal  mines,  537 
recreational  disturbance,  485 
Redhead,  see  Aythya  americana 
Redstart,  American,  see  Setophaga  ruticilla 
Redwing,  see  Turdus  iliacus 
Regulus  calendula,  521,  522 
satrapa,  522 

Reidy,  Jennifer  L.,  see  Sachtleben,  Thalia, , and 

Julie  A.  Savidge 
reproductive 
behavior,  225 

ecology,  145,  208,  225,  259 
success,  145,  208,  281 
restoration  plots,  391 
Rhegmatorhina  melanosticta,  18 
Rhipidura  albiscapa  brenchleyi,  295-308 
spilodera  spilodera,  295-308 
Rhynchopsitta  pachyrhyncha,  237-243 
terrisi,  237 

Rich,  Terrell  D.,  Partners  in  flight:  North  American 
landbird  conservation,  reviewed,  123-124 
Richardson,  Michael  J.,  see  Craik,  Shawn  R.,  Rodger 

D.  Titman,  Amelie  Rousseau,  and 

Riehl,  Christina,  Widespread  cannibalism  by  fledglings 
in  a nesting  colony  of  Black-crowned  Night- 
Herons,  101-104 

Risch,  Thomas  S.,  and  Thomas  J.  Robinson,  First  ob- 
servation of  cavity  nesting  by  a female  Blue 
Grosbeak,  107-108 

Rivers,  James  W.,  and  Michael  J.  Kuehn,  Predation  of 
Eared  Grebe  by  Great  Blue  Heron,  112-113 
Robel,  Robert  J.,  see  Pitman,  James  C.,  Christian  A. 

Hagen,  Brent  E.  Jamison, , Thomas  M. 

Loughin,  and  Roger  D.  Applegate 


INDEX  TO  VOLUME  1 1 8 


607 


Robertson,  Raleigh  J.,  see  Barg,  Jennifer  J.,  Jason 

Jones,  M.  Katharine  Girvan,  and 

Robin,  American,  see  Turdus  migratorius 
European,  see  Erithacus  rubecula 
Pacific,  see  Petroica  multicolor  ambrynensis 
Robinson,  Thomas  J.,  see  Risch,  Thomas  S.,  and 


Robinson,  W.  Douglas,  see  Tucker,  James  W.,  Jr., 

, and  James  B.  Grand 

Rochepault,  Yann,  see  Buidin,  Christophe,  , 

Michel  Savard,  and  Jean-Pierre  L.  Savard 
Rodriguez  Castillo,  Angelica  M.,  and  Jessica  R.  Eber- 
hard,  Reproductive  behavior  of  the  Yellow- 
crowned  Parrot  ( Amazona  ochrocephala ) in 
western  Panama,  225-236 

Rogers,  Christopher  M.,  Nesting  success  and  breeding 
biology  of  Cerulean  Warblers  in  Michigan, 
145-151 

Rojas-Soto,  Octavio  R.,  see  Oliveras  de  Ita,  Adan,  and 


roosting 

behavior,  532,  566 
locations,  502,  532,  566 
roost-site  selection 

thermal  aspects  of,  532,  566 
vegetation  aspects  of,  532 

Rousseau,  Amelie,  see  Craik,  Shawn  R.,  Rodger  D. 

Titman, , and  Michael  J.  Richardson 

Ruback,  Patricia  A.,  see  Walley,  Harlan  D.,  and 
, review  by 

Rubenstein,  Dustin  R.,  see  Lovette,  Irby  J.,  , 

and  Wilson  Nderitu  Watetu 
Russia,  364 

s 

Saab,  Victoria  A.,  and  Hugh  D.  W.  Powell  (Eds.),  Fire 
and  avian  ecology  in  North  America,  reviewed, 
580-581 

Sachtleben,  Thalia,  Jennifer  L.  Reidy,  and  Julie  A. 
Savidge,  A description  of  the  first  Micronesian 
Honeyeater  ( Myzomela  rubratra  saffordi ) nests 
found  on  Saipan,  Mariana  Islands,  309-315 
Sacramento  River,  178 
sagebrush,  see  Artemisia  spp. 

sand,  see  Artemisia  filifolia 
Sage-Grouse,  Greater,  see  Centrocercus  urophasianus 
Gunnison,  see  Centrocercus  minimus 
Saipan,  309 

Sakai,  Walter  H.,  Polydactyly  in  a Vaux’s  Swift,  424- 
426 

Sandpiper,  Broad-billed,  see  Limicola  falcinellus 
Least,  see  Calidris  minutilla 
Pectoral,  see  Calidris  melanotos 
Semipalmated,  see  Calidris  pusilla 
Spotted,  see  Actitis  macularius 
Stilt,  see  Calidris  himantopus 
Western,  see  Calidris  mauri 
sand  sagebrush  prairie,  23 
Santa  Rosa  Mountains,  256 

Santiago- Alarcon,  Diego,  Susan  M.  Tanksley,  and  Pa- 
tricia G.  Parker,  Morphological  variation  and 


genetic  structure  of  Galapagos  Dove  ( Zenaida 
Galapagoensis)  populations:  issues  in  conser- 
vation for  the  Galapagos  bird  fauna,  194-207 
satellite  transmitters,  494 

Savage,  Susan,  see  Benson,  Anna-Marie,  Brad  A.  An- 
dres, W.  N.  Johnson,  , and  Susan  M. 

Sharbaugh 
savanna  birds,  537 
Savannah  River  Site,  138 

Savard,  Jean-Pierre  L.,  see  Buidin,  Christophe,  Yann 

Rochepault,  Michel  Savard,  and 

Savard,  Michel,  see  Buidin,  Christophe,  Yann  Roche- 
pault,   , and  Jean-Pierre  L.  Savard 

Savidge,  Julie  A.,  see  Sachtleben,  Thalia,  Jennifer  L. 

Reidy,  and 

Sayomis  nigricans,  414 
scavenging,  101 

Schwertner,  T.  Wayne,  see  Metz,  Steve  T.,  Kyle  B. 

Melton,  Ray  Aguirre,  Bret  A.  Collier, , 

Markus  J.  Peterson,  and  Nova  J.  Silvy 
Sciurus  niger,  150 
Scolopax  minor,  55 
Sedgwick,  James  A. 

Message  from  the  editor:  the  new  Wilson  Journal  of 
Ornithology,  1-2 

Once  Upon  a Time  in  American  Ornithology,  264- 
266 

review  by,  125-127 
seed-size  selection,  64 
Seiurus  aurocapilla,  169,  523 
noveboracensis,  523 
selection,  nest-site,  247,  281 
territory,  391 

Setophaga  ruticilla,  149,  223,  249,  374-379,  439-451, 
523 

sexual  dimorphism,  558 

Sharbaugh,  Susan  M.,  see  Benson,  Anna-Marie,  Brad 
A.  Andres,  W.  N.  Johnson,  Susan  Savage,  and 


Sheets,  David  H.,  see  Morris,  Sara  R.,  Amanda  M. 
Larracuente,  Kristen  M.  Covino,  Melissa  S. 
Mustillo,  Kathryn  E.  Mattem,  David  A.  Lieb- 
ner,  and 

Shen,  Sheng-Feng,  see  Yuan,  Hsiao- Wei, , and 

Hisn-Yi  Hung 

Sherry,  Thomas  W.,  see  Staicer,  Cynthia  A.,  Victoria 
Ingalls,  and 

Shew,  Justin  J.,  American  Crow  caches  rabbit  kits, 
572-573 

Shikra,  see  Accipiter  badius 
shorebirds,  152 

Shoveler,  Northern,  see  Anas  clypeata 
Shrike,  Loggerhead,  see  Lanius  ludovicianus 
Red-backed,  see  Lanius  collurio 
Shrikebill,  Southern,  see  Clytorhynchus  pachycephal- 
oides  grisescens 
shrubland  species,  353 
Sialia  currucoides,  10 
mexicana,  552-557 
sialis,  107,  114,  552-557 
Sibia,  Taiwan,  see  Heterophasia  auricularis 


608 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  1 18,  No.  4,  December  2006 


Sibley,  Fred  C.,  see  Boal,  Clint  W., , Tracy  S. 

Estabrook,  and  James  Lazell 
Sierra-Finch,  Patagonian,  see  Phrygilus  patagonicus 
Silver-eye,  see  Zosterops  lateralis  tropicus 

Silverman,  Emily  D.,  see  Kearns,  Laura  J.,  , 

and  Kimberly  R.  Hall 

Silvy,  Nova  J.,  see  Metz,  Steve  T.,  Kyle  B.  Melton, 
Ray  Aguirre,  Bret  A.  Collier,  T.  Wayne 
Schwertner,  Markus  J.  Peterson,  and 

singing 

behavior  and  pairing  status,  439 
mode,  439 

Sitta  canadensis,  164—172,  522 
size  dimorphism,  527 
skill  development,  527 
skunk,  striped,  see  Mephitis  mephitis 
Skutchia  spp.,  18 
small  mammal  abundance,  91 
snake,  black  rat,  see  Elaphe  obsolete  obsolete 
brown  tree,  see  Boiga  irregularis 
garter,  see  Thamnophis  spp. 
gopher,  see  Pituophis  melanoleucus 
prairie  king,  see  Lampropeltis  calligaster 
racer,  see  Coluber  constrictor 
Snipe,  Common  (Wilson’s),  see  Capella  gallinago 
soaring  flight,  of  Broad-winged  Hawk,  471 
Somateria  mollissima,  421 
song 

acquisition,  341 
delivery,  439 
imitation,  341 
invention,  341 
rates,  439 
repertoire,  341 
songbirds 

insectivorous,  164 
non-breeding,  164 

Sordahl,  Tex  A.,  Field  experiments  on  eggshell  remov- 
al by  Mountain  Plovers,  59-63 
South  Carolina,  138 
Southwest  Pacific,  295 

Sparrow,  Bachman’s,  see  Aimophila  aestivalis 
Field,  see  Spizella  pusilla 

Florida  Grasshopper,  see  Ammodramus  savannarum 
floridanus 

Grasshopper,  see  Ammodramus  savannarum 
House,  see  Passer  domesticus 
Savannah,  see  Passerculus  sandwichensis 
Song,  see  Melospiza  melodia 
White-throated,  see  Zonotrichia  albicollis 
Worthen’s,  see  Spizella  wortheni 
Sparrowhawk,  Levant,  see  Accipiter  brevipes 
sparrows,  131,  138,  326 
species,  new  record,  218 
rediscovery,  13 
richness,  399 
Spermophilus  spp.,  27 
Spindalis  portoricensis,  571-572 
Spindalis,  Puerto  Rican,  see  Spindalis  portoricensis 
Spiza  americana,  539 
Spizella  pusilla,  537-546 


wortheni,  83 

Sporophila  albogularis,  88 
americana,  88 
caerulescens,  85-90 
collaris,  88 
lineola,  88 
nigricollis,  88 
ruficollis,  88 
torqueola,  88 

spruce  budworm,  see  Choristoneura  fumiferana 
squirrel,  eastern  fox,  see  Sciurus  niger 
red,  see  Tamiasciurus  hudsonicus 
St.  Croix,  194 

Staicer,  Cynthia  A.,  Victoria  Ingalls,  and  Thomas  W. 
Sherry,  Singing  behavior  varies  with  breeding 
status  of  American  Redstarts  ( Setophaga  ruti- 
cilla),  439-451 

Starling,  European,  see  Sturnis  vulgaris 
Mountain,  see  Aplonis  santovestris 
Rufous-winged,  see  Aplonis  zelandicus 
Steadman,  David  W.,  see  Kratter,  Andrew  W.,  Jeremy 

J.  Kirchman,  and 

Sterna  antillarum,  215 
dougallii,  106 
fuscata,  425 
hirundo,  103,  105 
paradisaea,  105 
spp.,  421 

Stilt,  Banded,  see  Cladorhynchus  leucocephalus 
Black-necked,  see  Himantopus  mexicanus 
Stober,  Jonathan  M.,  and  David  G.  Krementz,  Varia- 
tion in  Bachman’s  Sparrow  home-range  size  at 
the  Savannah  River  Site,  South  Carolina,  138 — 
144 

Stork,  White,  see  Ciconia  ciconia 
Streptopelia  decaocto,  64-69 
Strix  occidentalis  lucida,  241 
strut,  display,  36 
rate,  36 

Sturnella  magna,  402,  539 
Sturnis  vulgaris,  173,  453,  553 
Sula  grand,  244-247 
nebouxii,  246,  527-531 

Sumbera,  Radim,  see  Grim,  Tomas,  and 

supplemental  foraging  perches,  333 
survival 
brood,  208 
site-specific,  178 

Swallow,  Cliff,  see  Petrochelidon  pyrrhonota 
Tree,  see  Tachycineta  bicolor 
Violet-green,  see  Tachycineta  thalassina 
Swift,  Chimney,  see  Chaetura  pelagica 
Common,  see  Apus  apus 
Vaux’s,  see  Chaetura  vauxi 
Swiftlet,  Glossy,  see  Collocalia  esculenta  uropygialis 
Uniform,  see  Collocalia  v.  vanikorensis 
Sylvia  atricapilla,  191,  371 
Sylvilagus  floridanus,  572-573 

T 

Tachybaptus  ruficollis,  113 
Tachycineta  bicolor,  75,  457,  553 


INDEX  TO  VOLUME  1 1 8 


609 


thalassina,  553 

Tamiasciurus  hudsonicus,  452 

Tanager,  Summer,  see  Piranga  rubra 

Tanksley,  Susan  M.,  see  Santiago- Alarcon,  Diego, 

, and  Patricia  G.  Parker 

Taxidea  taxus,  32 
taxonomy,  13 

Taylor,  Sonja  E.,  and  Jessica  R.  Young,  A comparative 
behavioral  study  of  three  Greater  Sage-Grouse 
populations,  36-41 
Tayra,  see  Eira  barbara 
Teal,  Blue- winged,  see  Anas  discors 
Cinnamon,  see  Anas  cyanoptera 
Green-winged,  see  Anas  crecca 
Tellez-Garcia,  Lorena,  see  Monterrubio-Rico,  Tiberio 
C.,  Javier  Cruz-Nieto,  Ernesto  Enkerlin-Hoe- 

flich,  Diana  Venegas-Holguin,  , and 

Consuelo  Marin-Togo 
Tern,  Arctic,  see  Sterna  paradisaea 

Black,  see  Chilidonias  niger  surinamensis 
Common,  see  Sterna  hirundo 
Least,  see  Sterna  antillarum 
Roseate,  see  Sterna  dougallii 
Sooty,  see  Sterna  fuscata 
Texas,  259 

Thamnophilidae,  molecular  phylogeny  of,  20 
Thamnophis  spp.,  540 

Therrien,  Jean-Fran§ois,  see  Careau,  Vincent, , 

Pablo  Porras,  Don  Thomas,  and  Keith  Bildstein 
Thicketbird,  Melanesian,  see  Cichlornis  whitneyi 
Thomas,  Don,  see  Careau,  Vincent,  Jean-Fran§ois 

Therrien,  Pablo  Porras,  , and  Keith 

Bildstein 

Thompson,  Jonathan  E.,  see  Lavers,  Jennifer  L., 
, Cynthia  A.  Paszkowski,  and  C.  Davi- 
son Ankney 

Thrasher,  Brown,  see  Toxostoma  rufum 
Pearly-eyed,  see  Margarops  fuscatus 
Thrush,  Austral,  see  Turdus  falcklandii 
Gray-cheeked,  see  Catharus  minimus 
Hermit,  see  Catharus  guttatus 
Island,  see  Turdus  poliocephalus 
Song,  see  Turdus  philomelos 
Swainson’s,  see  Catharus  ustulatus 
Thryothorus  felix,  563 

ludovicianus,  75,  413-415,  566-569 
sinaloa,  563 
Tiaris  bicolor,  221 

Tietje,  William  D.,  see  Fiehler,  Craig  M., , and 

William  R.  Fields 
Timaliine  babbler,  558-562 
Tit,  Great,  see  Parus  major 

Titman,  Rodger  D.,  see  Craik,  Shawn  R.,  , 

Amelie  Rousseau,  and  Michael  J.  Richardson 
Titmouse,  Tufted,  see  Baeolophus  bicolor 
Towhee,  Eastern,  see  Pipilo  erythrophthalmus 
Spotted,  see  Pipilo  maculatus 
Toxostoma  rufum,  540 

Treerunner,  White-throated,  see  Pygarrhichas  albogu- 
laris 

Trichoglossus  haematodus  massena,  295—308 


Trichomonas  gallinae,  in  Galapagos  Doves,  203 
Triller,  Long-tailed,  see  Lalage  leucopyga  albiloris 
Polynesian,  see  Lalage  maculosa  modesta 
Tringa  flavipes,  156 

Troglodytes  aedon,  10,  75,  252,  318,  414,  419,  553 
Tucker,  James  W.,  Jr.,  W.  Douglas  Robinson,  and 
James  B.  Grand,  Breeding  productivity  of 
Bachman’s  Sparrows  in  fire-managed  longleaf 
pine  forests,  131-137 
Turdus  falcklandii,  252 
iliacus,  191 
merula,  348 

migratorius,  341-352,  540 
philomelos,  114 
poliocephalus,  295-308 

Turkey,  Merriam’s  Wild,  see  Meleagris  gallopavo  mer- 
riami 

Rio  Grande  Wild,  see  Meleagris  gallopavo  inter- 
media 

Turkey  Vulture,  see  Cathartes  aura 
Tympanuchus  cupido,  attwateri,  31 
pallidicinctus,  23-35 
phasianellus,  31 
Tyrannus  tyrannus,  357,  540 

u,  v 

understory-dependent  birds,  461 
United  States  Virgin  Islands,  194 
Uropsila  leucogastra,  563 

Valqui  H.,  Thomas,  see  Lane,  Daniel  F., , Jose 

Alvarez  A.,  Jessica  Armenta,  and  Karen  Eck- 
hardt 

Van  Perlo,  Ber,  Birds  of  Mexico  and  Central  America, 
reviewed,  583-584 
Vancouver  Island,  380 
Vanuatu,  295 

Vanuatu  Megapode,  see  Megapodius  layardi 

variation,  morphological,  194 

varillal  forest,  13 

Veery,  see  Catharus  fuscescens 

Venegas-Holguin,  Diana,  see  Monterrubio-Rico,  Ti- 
berio C.,  Javier  Cruz-Nieto,  Ernesto  Enkerlin- 

Hoeflich,  , Lorena  Tellez-Garcia,  and 

Consuelo  Marin-Togo 
Vermivora  chrysoptera,  222 
peregrina,  164-172 

Vickery,  Peter  D.,  see  Zucherberg,  Benjamin,  and 


vineyards,  552 
Vireo  bellii,  540 
flavifons,  221 

olivaceus,  169,  218-224,  522 
solitarius,  522 

Vireo,  Bell’s,  see  Vireo  bellii 

Blue-headed,  see  Vireo  solitarius 
Red-eyed,  see  Vireo  olivaceus 
Yellow-throated,  see  Vireo  flavifons 
vocalizations,  145,  256,  295 
nonsong,  145,  256 

Vulture,  Bearded,  see  Gypaetus  barbatus 
Cape,  see  Gyps  coprotheres 


610 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY  • Vol.  118,  No.  4,  December  2006 


Rueppell’s,  see  Gyps  rueppellii 

w 

Wahl,  Terence  R.,  Bill  Tweit,  and  Steven  G.  Mlodinow 
(Eds.),  reviewed,  271-272 

Walk,  Jeffrey  W.,  Eric  L.  Kershner,  and  Richard  E. 
Warner,  Low  nesting  success  of  Loggerhead 
Shrikes  in  an  agricultural  landscape,  70-74 
Walley,  Harlan  D. 

and  Patricia  A.  Ruback,  review  by,  581-582 
reviews  by,  433-434,  581-582 
Warbler,  Bay-breasted,  see  Dendroica  castanea 
Black-and-white,  see  Mniotilta  varia 
Blackburnian,  see  Dendroica  fusca 
Blackpoll,  see  Dendroica  striata 
Black- throated  Blue,  see  Dendroica  cerulescens 
Black-throated  Gray,  see  Dendroica  nigrescens 
Black-throated  Green,  see  Dendroica  virens 
Cape  May,  see  Dendroica  tigrina 
Cerulean,  see  Dendroica  cerulea 
Chestnut-sided,  see  Dendroica  pensylvanica 
Eurasian  Reed,  see  Acrocephalus  scirpaceus 
Golden-cheeked,  see  Dendroica  chrysoparia 
Golden-winged,  see  Vermivora  chrysoptera 
Hermit,  see  Dendroica  occidentalis 
Hooded,  see  Wilsonia  citrina 
Magnolia,  see  Dendroica  magnolia 
Prairie,  see  Dendroica  discolor 
Prothonotary,  see  Protonotaria  citrea 
Sedge,  see  Acrocephalus  schoenobaenus 
Swainson’s,  see  Limnothlypis  swainsonii 
Tennessee,  see  Vermivora  peregrine 
Townsend,  see  Dendroica  townsendi 
Wilson’s,  see  Wilsonia  pusilla 
Worm-eating,  see  Helmitheros  vermivorum 
Yellow,  see  Dendroica  petechia 
Yellow-rumped,  see  Dendroica  coronata 
Warner,  Richard  E.,  see  Walk,  Jeffrey  W.,  Eric  L.  Ker- 
shner, and 

Waterthrush,  Northern,  see  Seiurus  noveboracensis 
Waxwing,  Cedar,  see  Bombycilla  cedrorum 
whipscorpion,  tailless  (whip  spiders),  see  Phrynus  lon- 
gipes 

Whistler,  New  Caledonian,  see  Pachycephala  [ pecto - 
ralis]  caledonica  intacta 

White-eye,  Bridled,  see  Zoserops  conspicillatus  say- 
pani 

Golden,  see  Cleptornis  marchei 
Yellow-fronted,  see  Zosterops  flavifrons  brevicauda 
white-tailed  deer,  browsing  of  understory,  461 

Wiebe,  Karen  L.,  see  Fisher,  Ryan  J.,  and 

Williams,  Ernest  H.,  Jr.,  and  Lucy  Bunkley-Williams, 
Rapid  beak-swinging  locomotion  in  the  Puerto 
Rican  Spindalis,  571-572 
Willow  Tits,  see  Poecile  montanus 
Wilson,  Alexander,  in  Once  Upon  a Time  in  American 
Ornithology,  264 


Wilsonia  citrina,  223 
pusilla,  523,  547-551 

Witt,  Christopher  C.,  see  Clark,  William  S.,  and 


Wood,  Douglas  R.,  and  William  A.  Carter,  Carolina 
Wren  nest  successfully  parasitized  by  House 
Finch,  413-415 

Woodpecker,  Acorn,  see  Melanerpes  formicivorus 
Ivory-billed,  in  Once  Upon  a Time  in  American  Or- 
nithology, 264 

Magellanic,  see  Campephilu  magellanicus 
Striped,  see  Picoides  lignarius 
Woods  wallow,  White-breasted,  see  Artamus  leucor- 
hynchus  tenuis 

Wren,  Carolina,  see  Thryothorus  ludovicianus 
Happy,  see  Thryothorus  felix 
House,  see  Troglodytes  aedon 
Marsh,  see  Cistothorus  palustris 
Rufous-naped,  see  Campylorhynchus  rufinucha 
Sedge,  see  Cistothorus  platensis 
Sinaloa,  see  Thryothorus  sinaloa 
White-bellied,  see  Uropsila  leucogastra 

x,  Y 

Xanthocephalus  xanthocephalus,  391-398,  416,  454 
Xipholena  punicea,  20 
Xiphorhynchus  ocellatus,  17 
Xolmis  pyrope,  252 
Yellowlegs,  Lesser,  see  Tringa  flavipes 
Yellowthroat,  Common,  see  Geothlypis  trichas 
Gray-crowned,  see  Geothlypis  poliocephala 
Yosef,  Reuven,  reviews  by,  582-583,  585 

Young,  Jessica  R.,  see  Taylor,  Sonja  E.,  and 

Yuan,  Hsiao-Wei,  Sheng-Feng  Shen,  and  Hisn-Yi 
Hung,  Sexual  dimorphism,  dispersal  patterns, 
and  breeding  biology  of  the  Taiwan  Yuhina:  a 
joint-nesting  passerine,  558-562 
Yuhina  brunneiceps,  558—562 
Yuhina,  Taiwan,  see  Yuhina  brunneiceps 

z 

Zahawi,  Rakan  A.,  see  Garcia-C.,  J.  Mauricio,  and 


Zenaida  aurita,  222 

galapagoensis,  194-207 
macroura,  64-69,  203,  540 
Zonotrichia  albicollis,  169,  326-332,  521,  523 
Zosterops  conspicillatus  saypani,  311 
flavifrons  brevicauda,  295-308 
lateralis  tropicus,  295-308 

Zuckerberg,  Benjamin,  and  Peter  D.  Vickery,  Effects 
of  mowing  and  burning  on  shrubland  and 
grassland  birds  on  Nantucket  Island,  Massa- 
chusetts, 353-363 

Zuwerink,  David  A.,  and  James  S.  Marshall,  Brown- 
headed Cowbird’s  fatal  attempt  to  parasitize  a 
Carolina  Chickadee  nest,  418-419 


^ Wilson  Journal 

of  Ornithology 

Published  by  the  Wilson  Ornithological  Society 
Volume  118  2006  Quarterly 


EDITOR: 
EDITORIAL  BOARD: 


REVIEW  EDITOR 
INDEX  EDITOR 
EDITORIAL  ASSISTANTS 


JAMES  A.  SEDGWICK 
KATHY  G.  BEAL 
CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 
KARL  E.  MILLER 
MARY  GUSTAFSON 
KATHY  G.  BEAL 
M.  BETH  DILLON 
ALISON  R.  GOFFREDI 
CYNTHIA  P.  MELCHER 
JULIETTE  WILSON 


The  Wilson  Ornithological  Society 
Founded  December  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  Ornithologist 

President — Doris  J.  Watt,  Dept,  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  IN 
46556,  USA;  e-mail:  dwatt@saintmarys.edu 

First  Vice-President — James  D.  Rising,  Dept,  of  Zoology,  Univ.  of  Toronto,  Toronto, 
ON  M5S  3G5,  Canada;  e-mail:  rising@zoo.utoronto.ca 

Second  Vice-President — E.  Dale  Kennedy,  Biology  Dept.,  Albion  College,  Albion,  MI 
49224,  USA;  e-mail:  dkennedy@albion.edu 

Editor — James  A.  Sedgwick,  U.S.  Geological  Survey,  Fort  Collins  Science  Center,  2150 
Centre  Ave.,  Bldg.  C,  Fort  Collins,  CO  80526,  USA;  e-mail:  wilsonbulletin@usgs.gov 

Secretary — John  A.  Smallwood,  Dept,  of  Biology  and  Molecular  Biology,  Montclair 
State  University,  Montclair,  NJ  07043,  USA;  e-mail:  smallwood@montclair.edu 

Treasurer — Melinda  M.  Clark,  52684  Highland  Dr.,  South  Bend,  IN  46635,  USA;  e-mail: 
MClark@tcservices.biz 

Elected  Council  Members — Mary  Bomberger  Brown,  Robert  L.  Curry,  and  James  R. 
Hill,  III  (terms  expire  2007);  Kathy  G.  Beal,  Daniel  Klem,  Jr.,  and  Douglas  W.  White 
(terms  expire  2008);  Carla  J.  Dove,  Greg  H.  Farley,  and  Mia  R.  Revels  (terms  expire 
2009). 


DATES  OF  ISSUE  OF  VOLUME  1 1 8 OF 

THE  WILSON  JOURNAL  OL  ORNITHOLOGY 

no.  1—3  March  2006 
no.  2 — 5 June  2006 
no.  3 — 22  September  2006 
no.  4 — 27  December  2006 


1 

3 

13 

23 

36 

42 

53 

59 

64 

70 

75 

81 

85 

91 

99 

101 

104 


CONTENTS  OF  VOLUME  118 


Message  from  the  Editor 


NUMBER  1 


Major  Articles 

Variation  in  mass  of  female  Prothonotary  Warblers  during  nesting 
Charles  R.  Blem  and  Leann  B.  Blem 

The  rediscovery  and  natural  history  of  the  White-masked  Antbird  ( Pithys  castaneus) 

Daniel  F.  Lane,  Thomas  Valqui  H.,  Jose  Alvarez  A.,  Jessica  Armenta,  and  Karen  Eckhardt 

Nesting  ecology  of  Lesser  Prairie-Chickens  in  sand  sagebrush  prairie  of  southwestern  Kansas 
James  C.  Pitman,  Christian  A.  Hagen,  Brent  E.  Jamison,  Robert  J.  Robel,  Thomas  M.  Loughin,  and  Roger 
D.  Applegate 

A comparative  behavioral  study  of  three  Greater  Sage-Grouse  populations 
Sonja  E.  Taylor  and  Jessica  R.  Young 

First  known  specimen  of  a hybrid  Buteo : Swainson’s  Hawk  ( Buteo  swainsoni)  x Rough-legged  Hawk 

{B.  lagopus)  from  Louisiana 

William  S.  Clark  and  Christopher  C.  Witt 

Nocturnal  hunting  by  Peregrine  Falcons  at  the  Empire  State  Building,  New  York  City 
Robert  DeCandido  and  Deborah  Allen 

Field  experiments  on  eggshell  removal  by  Mountain  Plovers 
Tex  A.  Sordahl 

Seed-size  selection  in  Mourning  Doves  and  Eurasian  Collared- Doves 
Steven  E.  Hayslette 

Low  nesting  success  of  Loggerhead  Shrikes  in  an  agricultural  landscape 
Jeffery  W.  Walk,  Eric  L.  Kershner,  and  Richard  E.  Warner 

Nest  interference  by  fledgling  Loggerhead  Shrikes 
Eric  L.  Kershner  and  Eric  C.  Mruz 

First  breeding  record  of  a Mountain  Plover  in  Nuevo  Leon,  Mexico 

Jose  I.  Gonzalez  Rojas,  Miguel  A.  Cruz  Nieto,  Oscar  Ballesteros  Medrano,  and  Irene  Ruvalcaba  Ortega 

Breeding  biology  of  the  Double-collared  Seedeater  (Sporophila  caerulescens ) 

Mercival  R.  Francisco 

Small  mammal  selection  by  the  White-tailed  Hawk  in  southeastern  Brazil 

Marco  A.  Monteiro  Granzinolli  and  Jose  Carlos  Motta-Junior 

Short  Communications 

Provisioning  of  fledgling  conspecifics  by  males  of  the  brood-parasitic  cuckoos  Chrysococcyx  klaas  and 
C.  caprius 

Irby  J.  Lovette,  Dustin  R.  Rubenstein,  and  Wilson  Nderitu  Watetu 

Widespread  cannibalism  by  fledglings  in  a nesting  colony  of  Black-crowned  Night-Herons 
Christina  Riehl 

First  report  of  Black  Terns  breeding  on  a coastal  barrier  island 

Shawn  R.  Craik,  Rodger  D.  Titman,  Amelie  Rousseau,  and  Michael  J.  Richardson 


107 

109 

112 

114 

117 

120 

131 

138 

143 

132 

164 

173 

178 

187 

194 

208 

218 


First  observation  of  cavity  nesting  by  a female  Blue  Grosbeak 
Thomas  S.  Risch  and  Thomas  J.  Robinson 

A new  record  of  the  endangered  White-winged  Nightjar  {Eleothreptus  candicans)  from  Beni,  Bolivia 
Tomas  Grim  and  Radim  Sumbera 

Predation  of  Eared  Grebe  by  Great  Blue  Heron 
James  W.  Rivers  and  Michael  J.  Kuehn 

Abnormal  eggs  and  incubation  behavior  in  Northern  Bobwhite 

Fidel  Hernandez,  Juan  A.  Arredondo,  Froylan  Hernandez,  Fred  C.  Bryant,  and  Leonard  A.  Brennan 

Once  Upon  a Time  in  American  Ornithology 
Ornithological  Literature 


NUMBER  2 

Major  Articles 

Breeding  productivity  of  Bachman’s  Sparrows  in  fire-managed  longleaf  pine  forests 
James  W Tucker,  Jr.,  W Douglas  Robinson,  and  James  B.  Grand 

Variation  in  Bachman’s  Sparrow  home-range  size  at  the  Savannah  River  Site,  South  Carolina 
Jonathan  M.  Stober  and  David  G.  Krementz 

Nesting  success  and  breeding  biology  of  Cerulean  Warblers  in  Michigan 
Christopher  M.  Rogers 

Migrant  shorebird  predation  on  benthic  invertebrates  along  the  Illinois  River,  Illinois 
Gabriel  L.  Hamer,  Edward  J.  Heske,  Jeffrey  D.  Brawn,  and  Patrick  W.  Brown 

Composition  and  timing  of  postbreeding  multispecies  feeding  flocks  of  boreal  forest  passerines  in 
western  Canada 

Keith  A.  Hobson  and  Steve  Van  Wilgenburg 

Variation  in  size  and  composition  of  Bufflehead  ( Bucephala  albeola ) and  Barrow’s  Goldeneye 
{Bucephala  islandica ) eggs 

Jennifer  L.  Lavers,  Jonathan  E.  Thompson,  Cynthia  A.  Paszkowski,  and  C.  Davison  Ankney 

Site-specific  survival  of  Black-headed  Grosbeaks  and  Spotted  Towhees  at  four  sites  within  the 
Sacramento  Valley,  California 
Thomas  Gardali  and  Nadav  Nur 

Pre-migratory  fattening  and  mass  gain  in  Flammulated  Owls  in  central  New  Mexico 
John  P DeLong 

Morphological  variation  and  genetic  structure  of  Galapagos  Dove  {Zenaida  galapagoensis)  populations: 

issues  in  conservation  for  the  Galapagos  bird  fauna 

Diego  Santiago-Alarcon,  Susan  M.  Tanksley,  and  Patricia  G.  Parker 

Breeding  ecology  of  American  and  Caribbean  coots  at  Southgate  Pond,  St.  Croix:  use  of  woody 
vegetation 

Douglas  B.  McNair  and  Carol  Cramer-Burke 

Insular  and  migrant  species,  longevity  records,  and  new  species  records  on  Guana  Island,  British  Virgin 
Islands 

Clint  W.  Boal,  Fred  C.  Sibley,  Tracy  S.  Estabrook,  and  James  Lazell 


225 

237 

244 

247 

251 

254 

256 

259 

261 

264 

267 

281 

295 

309 

316 

326 

333 

341 


Reproductive  behavior  of  the  Yellow-crowned  Parrot  ( Amazona  ochrocephala)  in  western  Panama 
Angelica  M.  Rodriguez  Castillo  and  Jessica  R.  Eberhard 

Gregarious  nesting  behavior  of  Thick-billed  Parrots  (. Rhynchopsitta  pachyrhyncha ) in  aspen  stands 
Tiberio  C.  Monterrubio-Rico,  Javier  Cruz-Nieto,  Ernesto  Enkerlin-Hoejlich,  Diana  Venegas-Holguin, 
Lorena  Tellez-Garcia,  and  Consuelo  Marin-Togo 

Short  Communications 

No  extra-pair  fertilization  observed  in  Nazca  Booby  ( Sula  granti)  broods 
David  J.  Anderson  and  Peter  T.  Boag 

Golden-cheeked  Warbler  males  participate  in  nest-site  selection 
Allen  E.  Graber,  Craig  A.  Davis,  and  David  M.  Leslie,  Jr. 

Provisioning  of  Magellanic  Woodpecker  ( Campephilus  magellanicus)  nestlings  with  vertebrate  prey 
Valeria  S.  Ojeda  and  M.  Laura  Chazarreta 

Reverse  mounting  and  copulation  behavior  in  polyandrous  Bearded  Vulture  ( Gypaetus  barbatus)  trios 
Joan  Bertran  and  Antoni  Margalida 

Natural  occurrence  of  crowing  in  a free-living  female  galliform,  the  California  Quail 
Jennifer  M.  Gee 

Poult  adoption  and  nest  abandonment  by  a female  Rio  Grande  Wild  Turkey  in  Texas 

Steve  T.  Metz,  Kyle  B.  Melton,  Ray  Aguirre,  Bret  A.  Collier,  77  Wayne  Schwertner,  Markus  J.  Peterson,  and 

Nova  J.  Silvy 

Predation  by  a Blue-crowned  Motmot  ( Momotus  momota)  on  a hummingbird 
J.  Mauricio  Garcia-C.  and  Rakan  A.  Zahawi 

Once  Upon  a Time  in  American  Ornithology 

Ornithological  Literature 


NUMBER  3 

Major  Articles 

Nest-site  selection  and  productivity  of  American  Dippers  in  the  Oregon  Coast  Range 
John  P Loegering  and  Robert  G.  Anthony 

Upland  bird  communities  on  Santo,  Vanuatu,  Southwest  Pacific 
Andrew  W.  Kratter,  Jeremy  J.  Kirchman,  and  David  W.  Steadman 

A description  of  the  first  Micronesian  Honeyeater  (. Myzomela  rubratra  sajfordi ) nests  found  on  Saipan, 
Mariana  Islands 

Thalia  Sachtleben,  Jennifer  L.  Reidy,  and  Julie  A.  Savidge 

Within-pair  interactions  and  parental  behavior  of  Cerulean  Warblers  breeding  in  eastern  Ontario 
Jennifer  J.  Barg,  Jason  Jones,  M.  Katharine  Girvan,  and  Raleigh  J.  Robertson 

Comparative  spring  migration  arrival  dates  in  the  two  morphs  of  White-throated  Sparrow 
Sarah  S.  A.  Caldwell  and  Alexander  M.  Mills 

Can  supplemental  foraging  perches  enhance  habitat  for  endangered  San  Clemente  Loggerhead  Shrikes 
Suellen  Lynn,  John  A.  Martin,  and  David  K Garcelon 

Do  American  Robins  acquire  songs  by  both  imitating  and  inventing? 

Steven  L.  Johnson 


353 

364 

374 

380 

391 

399 

411 

413 

415 

418 

420 

422 

424 

427 

430 

439 

452 

461 


Effects  of  mowing  and  burning  on  shrubland  and  grassland  birds  on  Nantucket  Island,  Massachusetts 
Benjamin  Zuckerberg  and  Peter  D.  Vickery 

Spatial  behavior  of  European  Robins  during  migratory  stopovers:  a telemetry  study 
Nikita  Chernetsov  and  Andrey  Mukhin 

Age-related  timing  and  patterns  of  prebasic  body  molt  in  wood  warblers  (Parulidae) 

Christine  A.  Debruyne,  Janice  M.  Hughes , and  David  J.  T.  Hussell 

Foraging  ecology  of  Bald  Eagles  at  an  urban  landfill 

Kyle  H.  Elliott,  Jason  Dujfe,  Sandi  L.  Lee,  Pierre  Mineau,  and  John  E.  Elliott 

Territory  selection  by  upland  Red-winged  Blackbirds  in  experimental  restoration  plots 
Maria  A.  Furey  and  Dirk  E.  Burhans 

The  use  of  southern  Appalachian  wetlands  by  breeding  birds,  with  a focus  on  Neotropical  migratory 
species 

Jason  E Bulluck  and  Matthew  P Rowe 

Short  Communications 

Breeding  range  extension  of  the  Northern  Saw-whet  Owl  in  Quebec 
Christophe  Buidin,  Yann  Rochepault,  Michel  Savard,  and  Jean-Pierre  L.  Savard 

Carolina  Wren  nest  successfully  parasitized  by  House  Finch 
Douglas  R.  Wood  and  William  A.  Carter 

American  Coot  parasitism  on  Least  Bitterns 
Brian  D.  Peer 

Brown-headed  Cowbird’s  fatal  attempt  to  parasitize  a Carolina  Chickadee  nest 
David  A.  Zuwerink  and  James  S.  Marshall 

Likely  predation  of  adult  Glossy  Ibis  by  Great  Black-backed  Gulls 
Christina  E.  Donehower 

Tailless  whipscorpion  (Phrynus  longipes ) feeds  on  Antillean  Crested  Hummingbird  ( Orthorhyncus 
cristatus ) 

Jennifer  L.  Owen  and  James  C.  Cokendolpher 

Polydactyly  in  a Vaux’s  Swift 
Walter  H.  Sakai 

Once  Upon  a Time  in  American  Ornithology 
Ornithological  Literature 


NUMBER  4 

Major  Articles 

Singing  behavior  varies  with  breeding  status  of  American  Redstarts  ( Setophaga  ruticilla) 

Cynthia  A.  Staicer,  Victoria  Ingalls,  and  Thomas  W.  Sherry 

Investment  in  nest  defense  by  Northern  Flickers:  effects  of  age  and  sex 
Ryan  J.  Fisher  and  Karen  L.  Wiebe 

Black-throated  Blue  Warbler  and  Veery  abundance  in  relation  to  understory  composition  in  northern 
Michigan  forests 

Laura  J.  Kearns,  Emily  D.  Silverman,  and  Kimberly  R.  Hall 


471  Soaring  and  gliding  flight  of  migrating  Broad-winged  Hawks:  behavior  in  the  Nearctic  and 
Neotropics  compared 

Vincent  Careau,  Jean-Frangois  Therrien,  Pablo  Porras , Don  Thomas , and  Keith  Bildstein 

478  Coloniality,  mate  retention,  and  nest-site  characteristics  in  the  Semipalmated  Sandpiper 
Joseph  R.  Jehl,  Jr. 

485  Effects  of  human  recreation  on  the  incubation  behavior  of  American  Oystercatchers 
Conor  P.  McGowan  and  Theodore  R.  Simons 

494  Movements  of  Long-tailed  Ducks  wintering  on  Lake  Ontario  to  breeding  areas  in  Nunavut,  Canada 
Mark  L.  Mallory,  Jason  Akearok,  Norm  R.  North,  D.  Vaughan  Weseloh,  and  Stephane  Lair 

502  Female  Tree  Swallow  home-range  movements  during  their  fertile  period  as  revealed  by  radio-tracking 
Mary  K Stapleton  and  Raleigh  J.  Robertson 

508  Effects  of  prescribed  fire  on  conditions  inside  a Cuban  Parrot  ( Amazona  leucocephala)  surrogate 
nesting  cavity  on  Great  Abaco,  Bahamas 

Joseph  J.  OBrien,  Caroline  Stahala,  Gina  P.  Mori,  Mac  A.  Callaham,  Jr.,  and  Chris  M.  Bergh 

513  Utility  of  open  population  models:  limitations  posed  by  parameter  estimability  in  the  study  of 
migratory  stopover 

Sara  R.  Morris,  Amanda  M.  Larracuente,  Kristen  M.  Covino , Melissa  S.  Mustillo,  Kathryn  E.  Mattern, 
David  A.  Liebner,  and  H.  David  Sheets 

527  Maximum  diving  depth  in  fledging  Blue-footed  Boobies:  skill  development  and  transition  to 
independence 

Jose  AIJredo  Castillo-Guerrero  and  Eric  Mellink 

532  Vegetative  and  thermal  aspects  of  roost-site  selection  in  urban  Yellow-billed  Magpies 
Scott  P.  Crosbie,  Douglas  A.  Bell,  and  Ginger  M.  Bolen 

537  Nesting  success  of  grassland  and  savanna  birds  on  reclaimed  surface  coal  mines  of  the  midwestern 
United  States 

Edward  W Galligan,  Travis  L.  DeVault,  and  Steven  L.  Lima 

547  Differential  timing  of  Wilson’s  Warbler  migration  in  Alaska 

Anna-Marie  Benson,  Brad  A.  Andres,  W.  N.  Johnson,  Susan  Savage,  and  Susan  M.  Sharbaugh 

552  Nesting  success  of  Western  Bluebirds  ( Sialia  mexicana)  using  nest  boxes  in  vineyard  and  oak-savannah 
habitats  of  California 

Craig  M.  Fiehler,  William  D.  Tietje,  and  William  R.  Fields 

558  Sexual  dimorphism,  dispersal  patterns,  and  breeding  biology  of  the  Taiwan  Yuhina:  a joint-nesting 
passerine 

Hsiao-Wei  Yuan,  Sheng-Feng  Shen,  and  Hisn-Yi  Hung 

Short  Communications 

563  Ant  presence  in  acacias:  an  association  that  maximizes  nesting  success  in  birds? 

Addn  Oliveras  de  Ita  and  Octavio  R.  Rojas-Soto 

566  Pair  roosting  of  nesting  Carolina  Wrens  ( Thryothorus  ludovicianus) 

Ronald  F.  Labisky  and  John  E.  Arnett,  Jr. 

569  Bald  Eagle  kills  crow  chasing  a hawk 
Bruce  D.  Ostrow 

571  Rapid  beak-swinging  locomotion  in  the  Puerto  Rican  Spindalis 
Ernest  H.  Williams,  Jr.  and  Lucy  Bunkley-Williams 


572  American  Crow  caches  rabbit  kits 
Justin  J.  Shew 

574  First  nesting  record  of  the  Gray-crowned  Yellowthroat  ( Geothlypis  poliocephala ) in  the  United  States 
since  1894 

Stephan  Lorenz , Chris  Butler ; and  Jimmy  Paz 

577  Once  Upon  a Time  in  American  Ornithology 
580  Ornithological  Literature 

586  Proceedings  of  the  Eighty-seventh  Annual  Meeting 
593  Reviewers  for  Volume  118 
595  Index  to  Volume  1 18 

Contents  of  Volume  118 


THE  WILSON  JOURNAL  OF  ORNITHOLOGY 


Editor  JAMES  A.  SEDGWICK 
U.S.  Geological  Survey 
Fort  Collins  Science  Center 
2150  Centre  Ave.,  Bldg.  C. 

Fort  Collins,  CO  80256-8118,  USA 
E-mail:  wjo@usgs.gov 


Managing  Editor  CYNTHIA  MELCHER 

Copy  Editors  ALISON  GOFFREDI 

JULIETTE  WILSON 


Editorial  Board  KATHY  G.  BEAL 
CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 
KARL  E.  MILLER 


Review  Editor  MARY  GUSTAFSON 

Texas  Parks  and  Wildlife  Dept. 

2800  S.  Bentsen  Palm  Dr. 

Mission,  TX  78572,  USA 
E-mail:  WilsonBookReview@aol.com 


GUIDELINES  FOR  AUTHORS 

Consult  the  detailed  “Guidelines  for  Authors”  found  on  the  Wilson  Ornithological  Society  Web  site  (http:// 
www.ummz.lsa.umich.edu/birds/wilsonbull.html).  Beginning  in  2007,  Clait  E.  Braun  will  become  the  new  editor 
of  The  Wilson  Journal  of  Ornithology.  As  of  1 July  2006,  all  manuscript  submissions  and  revisions  should  be 
sent  to  Clait  E.  Braun,  Editor,  The  Wilson  Journal  of  Ornithology,  5572  North  Ventana  Vista  Rd.,  Tucson,  AZ 
85750-7204,  USA.  The  New  Wilson  Journal  of  Ornithology  office  and  fax  telephone  number  will  be  (520)  529- 
0365,  and  the  E-mail  address  will  be  TWilsonJO@comcast.net. 

NOTICE  OF  CHANGE  OF  ADDRESS 

If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornithological 
Societies  of  North  America,  5400  Bosque  Blvd.,  Ste.  680,  Waco,  TX  76710. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  %The  Museum  of  Zoology,  The 
Univ.  of  Michigan,  Ann  Arbor,  MI  48109.  Persons  having  business  with  any  of  the  officers  may  address  them 
at  their  various  addresses  given  on  the  inside  of  the  front  cover,  and  all  matters  pertaining  to  the  journal  should 
be  sent  directly  to  the  Editor. 


MEMBERSHIP  INQUIRIES 

Membership  inquiries  should  be  sent  to  James  L.  Ingold,  Dept,  of  Biological  Sciences,  Louisiana  State  Univ., 
Shreveport,  LA  71115;  e-mail:  jingold@pilot.lsus.edu 

THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 

The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  in  the  Univ.  of 
Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  Univ.  of  Michigan  in  1930.  Until  1947 
the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  ornithological  magazines  from 
members  and  friends  of  the  Society.  Two  members  have  generously  established  a fund  for  the  purchase  of  new 
books;  members  and  friends  are  invited  to  maintain  the  fund  by  regular  contribution.  The  fund  will  be  admin- 
istered by  the  Library  Committee.  Terry  L.  Root,  Univ.  of  Michigan,  is  Chairman  of  the  Committee.  The  Library 
currently  receives  over  200  periodicals  as  gifts  and  in  exchange  for  The  Wilson  Journal  of  Ornithology.  For 
information  on  the  Library  and  our  holdings,  see  the  Society’s  web  page  at  http://www.ummz.lsa.umich.edu/ 
birds/wos.html.  With  the  usual  exception  of  rare  books,  any  item  in  the  Library  may  be  borrowed  by  members 
of  the  Society  and  will  be  sent  prepaid  (by  the  Univ.  of  Michigan)  to  any  address  in  the  United  States,  its 
possessions,  or  Canada.  Return  postage  is  paid  by  the  borrower.  Inquiries  and  requests  by  borrowers,  as  well  as 
gifts  of  books,  pamphlets,  reprints,  and  magazines,  should  be  addressed  to:  Josselyn  Van  Tyne  Memorial  Library, 
Museum  of  Zoology,  The  Univ.  of  Michigan,  1109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079,  USA.  Contri- 
butions to  the  New  Book  Fund  should  be  sent  to  the  Treasurer. 


This  issue  of  The  Wilson  Journal  of  Ornithology  was  published  on  27  December  2006. 


552 

558 

563 

566 

569 

571 

572 

574 

577 

580 

586 

593 

595 


Continued from  outside  back  cover 


Nesting  success  of  Western  Bluebirds  ( Sialia  mexicana)  using  nest  boxes  in  vineyard  and  oak-savannah 
habitats  of  California 

Craig  M.  Fiehler,  William  D.  Tietje,  and  William  R.  Fields 

Sexual  dimorphism,  dispersal  patterns,  and  breeding  biology  of  the  Taiwan  Yuhina:  a joint-nesting 
passerine 

Hsiao-Wei  Yuan,  Sheng-Feng  Shen,  and  Hisn-Yi  Hung 

Short  Communications 

Ant  presence  in  acacias:  an  association  that  maximizes  nesting  success  in  birds? 

Addn  Oliveras  de  Ita  and  Octavio  R.  Rojas -Soto 

Pair  roosting  of  nesting  Carolina  Wrens  ( Thryothorus  ludovicianus) 

Ronald  F Labisky  and  John  E.  Arnett,  Jr. 

Bald  Eagle  kills  crow  chasing  a hawk 
Bruce  D.  Ostrow 

Rapid  beak-swinging  locomotion  in  the  Puerto  Rican  Spindalis 
Ernest  H.  Williams,  Jr.  and  Lucy  Bunkley-Williams 

American  Crow  caches  rabbit  kits 
Justin  J.  Shew 

First  nesting  record  of  the  Gray-crowned  Yellowthroat  ( Geothlypis  poliocephala ) in  the  United  States 
since  1894 

Stephan  Lorenz,  Chris  Butler,  and  Jimmy  Paz 

Once  Upon  a Time  in  American  Ornithology 
Ornithological  Literature 

Proceedings  of  the  Eighty-seventh  Annual  Meeting 
Reviewers  for  Volume  ii8 
Index  to  Volume  ii8 
Contents  of  Volume  ii8 


<•923  19 


The  Wilson  Journal  of  Ornithology 

(formerly  The  Wilson  Bulletin) 

Volume  118,  Number  4 CONTENTS  December  2006 


Major  Articles 

439  Singing  behavior  varies  with  breeding  status  of  American  Redstarts  {Setophaga  ruticilla) 

Cynthia  A.  Staicer,  Victoria  Ingalls , and  Thomas  W.  Sherry 

452  Investment  in  nest  defense  by  Northern  Flickers:  effects  of  age  and  sex 
Ryan  J.  Fisher  and  Karen  L.  Wiebe 

46 1 Black-throated  Blue  Warbler  and  Veery  abundance  in  relation  to  understory  composition  in  northern 
Michigan  forests 

Laura  J.  Kearns , Emily  D.  Silverman,  and  Kimberly  R.  Hall 

471  Soaring  and  gliding  flight  of  migrating  Broad-winged  Hawks:  behavior  in  the  Nearctic  and  Neotropics 
compared 

Vincent  Careau,  Jean-Franqois  Therrien,  Pablo  Porras,  Don  Thomas,  and  Keith  Bildstein 

478  Coloniality,  mate  retention,  and  nest-site  characteristics  in  the  Semipalmated  Sandpiper 
Joseph  R.  Jehl,  Jr. 

485  Effects  of  human  recreation  on  the  incubation  behavior  of  American  Oystercatchers 
Conor  P.  McGowan  and  Theodore  R.  Simons 

494  Movements  of  Long-tailed  Ducks  wintering  on  Lake  Ontario  to  breeding  areas  in  Nunavut,  Canada 
Mark  L.  Mallory,  Jason  Akearok,  Norm  R.  North,  D.  Vaughan  Weseloh,  and  Stephane  Lair 

502  Female  Tree  Swallow  home-range  movements  during  their  fertile  period  as  revealed  by  radio-tracking 
Mary  K Stapleton  and  Raleigh  J.  Robertson 

508  Effects  of  prescribed  fire  on  conditions  inside  a Cuban  Parrot  {Amazona  leucocephala)  surrogate  nesting 
cavity  on  Great  Abaco,  Bahamas 

Joseph  J.  O'Brien,  Caroline  Stah ala,  Gina  P Mori,  Mac  A.  Callaham,  Jr.,  and  Chris  M.  Bergh 

513  Utility  of  open  population  models:  limitations  posed  by  parameter  estimability  in  the  study  of 
migratory  stopover 

Sara  R.  Morris,  Amanda  M.  Larracuente,  Kristen  M.  Covino,  Melissa  S.  Mustillo,  Kathryn  E.  Mattem, 
David  A.  Liebner,  and  H.  David  Sheets 

527  Maximum  diving  depth  in  fledging  Blue-footed  Boobies:  skill  development  and  transition  to 
independence 

Jose  Alfredo  Castillo-Guerrero  and  Eric  Mellink 

532  Vegetative  and  thermal  aspects  of  roost-site  selection  in  urban  Yellow-billed  Magpies 
Scott  P.  Crosbie,  Douglas  A.  Bell,  and  Ginger  M.  Bolen 

537  Nesting  success  of  grassland  and  savanna  birds  on  reclaimed  surface  coal  mines  of  the  midwestern 
United  States 

Edward  W.  Galligan,  Travis  L.  DeVault,  and  Steven  L.  Lima 

547  Differential  timing  of  Wilson’s  Warbler  migration  in  Alaska 

Anna-Marie  Benson,  Brad  A.  Andres,  W.  N.  Johnson,  Susan  Savage,  and  Susan  M.  Sharbaugh 


Continued  on  inside  back  cover 


:P,NST  MAYR  LIBRARY 


3 2044  114  248  230